Вы находитесь на странице: 1из 9

Chemical Engineering Journal xxx (2012) xxxxxx

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric
acid based mesoporous catalysts
Aysegul Ciftci a, Dilek Varisli b, Kenan Cem Tokay a, N. Asl Sezgi a, Timur Dogu a,
a
Chemical Engineering Department, Middle East Technical University, 06800 Ankara, Turkey
b
Advanced Technologies, Gazi University, Teknikokullar-Ankara, Turkey

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

Article history: Tungstophosphoric acid (TPA) incorporated silicate structured new mesoporous catalysts were synthe-
Available online xxxx sized following one-pot hydrothermal and impregnation procedures. Surface area of TPA@MCM-41,
which was prepared by impregnating TPA into MCM-41, was two orders of magnitude higher than the
Keywords: surface area of pure TPA and this catalyst showed very high activity in dehydration reactions of both
Tungstophosphoric acid ethanol and methanol. Ethanol fractional conversion values reaching to 1.0 was obtained at 300 C at a
Mesoporous catalyst space time of 0.27 s.g/cm3, over TPA@MCM-41. Diethyl ether selectivity showed a decreasing trend by
Dehydration
increasing temperature from 180 to 400 C in ethanol dehydration reaction. Ethylene yield values
Ethylene
Diethyl ether
approaching to 100% were obtained at temperatures over 250 C. DME yield passed through a maximum
Dimethyl ether at about 200 C with this catalyst, over which coke formation caused catalyst deactivation. One-pot
Fuel hydrothermal synthesis procedure was very successful to synthesize a catalyst (TRC-W40) which did
not lose any activity after repeated washing steps. This catalyst gave highly stable catalytic performance
in dehydration of both ethanol and methanol. Well dispersed WOx clusters were formed within the mes-
oporous silicate matrix of this material. This catalyst showed very good activity in dehydration reactions
of alcohols, giving 100% conversion in ethanol dehydration at 400 C and 100% DME selectivity in meth-
anol dehydration at temperatures less than 300 C.
2012 Elsevier B.V. All rights reserved.

1. Introduction Mesoporous silicate structures (M41S family) having ordered


pore structures, have been considered as highly promising catalyst
Heteropolyacid (HPA) compounds have very strong Bronsted supports, due to their high surface area and larger pore diameters
acidity, with very high proton mobility. They also have good redox than zeolites, which allow easy penetration of the reacting mole-
properties. These properties make them suitable catalysts for acid cules to the active sites [18]. However, these materials are quite
catalyzed and selective oxidation reactions [13]. Recent research inert and have low catalytic activity in pure form, for most catalytic
on etherication reactions to produce gasoline blending oxygen- applications. For the enhancement of catalytic activity of these
ates [47] and on dehydration of alcohols [814] showed high materials, metals or metal oxides were incorporated into their
activity of HPAs in these reactions. Some ethanol dehydration structure by direct (one-pot) hydrothermal procedures or by post
reaction results obtained using pure silicotungstic acid (STA), synthesis methods [1922]. For the improvement of the acidic
tungstophosphoric acid (TPA) and molybdophosphoric acid characteristics of such silicate structured materials, heteropolyac-
(MPA) catalysts, were reported in our recent publication [10]. Re- ids were incorporated into these materials and their catalytic per-
sults of that work showed that ethylene was the main product at formances were tested in different reactions [16,17,2326]. In a
high temperatures while diethyl ether (DEE) was formed at lower recent study of our research group, acidic characteristics of such
temperatures. Although HPAs are considered to have good poten- silicate structured mesoporous materials were improved by the
tial as solid acid catalysts, their very low surface area (<1 m2/g) and impregnation of silicotungstic acid (STA) into these materials and
high solubility in polar solvents limit their use as catalysts in prac- their catalytic performances were tested in ethanol dehydration
tical applications. In order to improve these limitations of HPAs, reaction [11]. That study showed that, STA impregnated MCM-41
recent studies were focused on the synthesis of supported catalysts was very active in ethanol dehydration to produce ethylene and
and/or on the production of their water insoluble salts [4,5,1517]. diethyl ether. Characterization studies of STA incorporated silicate
structured mesoporous materials had also indicated the formation
Corresponding author. Tel.: +90 312 2102631; fax: +90 312 2102600. of well dispersed WOx nano-rods within the mesoporous silicate
E-mail address: tdogu@metu.edu.tr (T. Dogu). matrix of the synthesized materials [30,12].

1385-8947/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.04.016

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
2 A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx

Alcohols have been considered as being among the best alter- from their lattice [11,1822]. However, in the case of synthesis of
nates to petroleum for the production of environmentally benign temperature sensitive materials, lower calcination temperatures
transportation fuels, fuel additives and petrochemicals [27]. Ethyl- or use of supercritical or alcohol-acid extraction procedures should
ene, which is one of the main feed stocks of petrochemical industry be applied [10,11,31,35]. Heteropolyacids are known to be highly
and diethyl ether (DEE), which has excellent transportation fuel temperature sensitive, as far as their acidity and catalytic perfor-
properties (cetane number > 125) may be produced by dehydration mances are considered [5,6,31,36] As it was reported in our earlier
of ethanol over solid acid catalysts [28,10]. Good burning properties publication, calcination temperature caused signicant variation
of DEE make it a promising alternative fuel or fuel additive for diesel on the catalytic performance of STA incorporated materials and
fuel [27,29]. Also, blending of DEE with ethanol was reported to im- some activity loss was observed in the materials calcined at tem-
prove the cold start problem of ethanol fueled cars. STA incorpo- peratures over 400 C [31]. Loss of acidity of TPA was reported to
rated silicate structured mesoporous materials, which were take place at lower temperatures than STA [36]. Consequently, a
synthesized by a one-pot hydrothermal method and containing set of TGA-DTA experiments were made to decide about the calci-
various W/Si ratios, were also shown to give very good catalytic nation temperature of the novel TPA incorporated material, which
performance in dehydration reaction of ethanol to produce was synthesized in this work following the one-pot procedure.
ethylene and DEE [29,30].
Another highly attractive environmentally clean diesel fuel 2.2. Synthesis of TPA@MCM-41 by impregnation
alternate is dimethyl ether (DME), which is commercially produced
by dehydration of methanol. DME has higher cetane number and TPA@MCM-41 catalyst was synthesized following an impregna-
much lower NOx emissions than conventional diesel fuel. Also, tion procedure. In the synthesis of TPA@MCM-41, rst MCM-41
near-zero smoke formation was reported in DME derived engines. was synthesized following a hydrothermal synthesis procedure.
Our recent studies showed that mesoporous aluminosilicate, nano- Sodium silicate and cetyltrimethylammonium bromide (CTMABr)
composite Naon-silica and STA incorporated mesoporous silicate were used as the silica source and the surfactant, respectively.
structured catalysts were highly active and selective for the syn- After the hydrothermal synthesis step, the synthesized material
thesis of DME by dehydration of methanol [12,13,32,33]. was calcined at 550 C, before the application of the impregnation
Most of our earlier studies with heteropolyacid catalysts were process. Details of the MCM-41 synthesis procedure are available
performed with silicotungstic acid (STA) based materials. The main in the literature [18,22]. Predetermined amount of tungstophos-
objectives of the present study were the synthesis of new water phoric acid (TPA) (Acros Organics) was then dissolved in deionized
insoluble and high surface area tungstophosphoric acid (TPA) water and MCM-41 was added to this solution. The resulting mix-
incorporated silicate structured mesoporous catalysts, following a ture was stirred at 30 C for 65 h. Evaporation of excess water was
new one-pot hydrothermal procedure, and comparison of catalytic achieved at 70 C. Then, the temperature of the oven was in-
performance of such materials with the performance of TPA creased, rst to 96 C and later to 120 C, for completion of the dry-
impregnated MCM-41 type catalytic materials, in dehydration ing process.
reactions of methanol and ethanol, to produce DME, DEE and eth-
ylene, respectively. 2.3. Characterization of the synthesized catalysts

TPA@MCM-41 and TRC-W40 catalysts were characterized by


2. Experimental
nitrogen adsorptiondesorption (Quantachrome Autosorb-1-C/MS
instrument), X-ray diffraction (Rigaku D/MAX2200 diffractometer
In this study, TPA incorporated silicate structured mesoporous
with a CuK radiation source with a 2h scanning range between 1
catalytic materials were synthesized following two different
and 50), SEM-EDS (JSM-6400 (JEOL) equipped with NORAN sys-
routes, namely by a new one-pot hydrothermal synthesis route
tem Six), FTIR (Bruker FTIR-IFS66/S instrument), pyridine adsorp-
and by impregnation of TPA into MCM-41. The catalysts prepared
tion-DRIFTS (Perkin Elmer Spectrum One instrument) and
by these two procedures were denoted as TRC-W40 and
transmission electron microscopy (JEOL Model JEM 1010 TEM)
TPA@MCM-41, respectively. Details of the synthesis procedures
techniques. TEM analysis of the synthesized materials was per-
of these materials are given in the following sections.
formed at the Electron Microscopy Center of Babes-Bolyai Univer-
sity, in Romania.
2.1. Synthesis of mesoporous TRC-W40 catalyst by one-pot Calcined materials were further degassed at 100 C for 16 h in
hydrothermal procedure vacuum before nitrogen adsorptiondesorption analysis. Degas-
sing procedure was decided basing on our earlier tests for com-
Tungstophosphoric acid (TPA) incorporated mesoporous TRC- plete removal of moisture from heteropolyacid incorporated
W40 catalyst was synthesized following a one-pot hydrothermal silicate structured mesoporous materials. Catalysts were also dried
route. Synthesis mixture was prepared using tetraethylorthosilicate at 110 C before the SEM-EDS and diffuse reectance FT-IR
(TEOS) (Merck) as the silica source and cetyltrimethylammonium (DRIFTS) analyses. For the SEM-EDS analyses, samples were cov-
bromide (CTMABr), (99% pure, Merck) as the surfactant. TPA, which ered with a very thin layer of gold. DRIFTS analysis of the pyridine
was obtained from Acros Organics, was used as the active compo- adsorbed catalysts gave information about the relative strengths of
nent. The W/Si molar ratio in the synthesis solution was adjusted Bronsted and Lewis acid sites of the synthesized materials. DRIFTS
to 0.40 [34]. The pH of the clear solution, which contained TEOS, analysis, was made at room temperature in a wave number inter-
CTMABr and TPA, was below 0.5. This solution was continuously val of 4004500 cm 1. Difference of the DRIFT spectra of pyridine
mixed and transferred into a Teon-lined stainless steel autoclave, adsorbed and fresh catalysts was analyzed for the acidity measure-
in which hydrothermal synthesis was performed at 120 C for a per- ments. FTIR analysis of the synthesized materials (mixed with KBr)
iod of 96 h. Solid product obtained by this procedure was crushed was also performed at a spectral resolution of 4 cm 1.
and then thoroughly washed with deionized water, until the pH of
the wash water was close to neutral. Then, it was dried at 40 C 2.4. Activity tests of the synthesized catalysts
and calcined in a tubular furnace.
Calcination temperature of MCM-41-like materials was gener- Catalytic activities of TPA@MCM-41 and TRC-W40 catalysts
ally selected as being over 500 C, for the removal of the surfactant were tested in vapor phase dehydration of ethanol and methanol.

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx 3

DrTGA TGA TRC-W40 Uncalcined DTA


mg/min % uV
378.88C
1.00 100.0 Weight Loss -5.367m
Weight Loss -3.840mg -48.583 40.00
-34.761%
80.0
Weight Loss -0.481mg
-4.354%
20.00
493.64 C
60.0
0.00

40.0 349.64C 505.10 C


110.12 C Weight Loss -0.861mg 0.00
-7.794%
236.20 C
20.0
-20.00
-1.00 0.0 233.06C

0.0 100.0 200.0 300.0 400.0 500.0 600.0


Temp[C]

Fig. 1. TGADTA analysis of uncalcined form of TRC-W40.

For each experiment 0.2 g of fresh catalyst was placed into the and the characterization results of these materials were compared.
xed bed ow reactor and supported by quartz wool from both Basing on the literature results, which indicated loss of acidity of
ends. Liquid ethanol (99.98% purity, Sigma Aldrich) was pumped TPA at higher temperatures, calcination temperature was not se-
at a ow rate of 2.9 ml/h, into the evaporator by a syringe pump. lected over 400 C. Further discussion about the effects of calcina-
Evaporator was placed before the reactor and its temperature tion temperature on the catalyst structure will be given in the
was kept at 150 C. In the evaporator, vaporized alcohol was mixed following sections.
with helium at a ratio of 1:1. Total ow rate of the gas mixture fed Chemical compositions of the TPA incorporated mesoporous
to the reactor was 44 ml/min. In order to control the reaction zone materials were evaluated by EDS analysis and W/Si ratio in the so-
temperature, reactor was placed into a temperature controlled lid products was compared with the W/Si ratio in the synthesis
tubular furnace. All the lines, before and after the reactor, were solution. EDS analyses made at different locations of the materials
heated to 150 C, to avoid condensation of any chemicals in the were highly consistent, indicating well dispersion of TPA in the
experimental system. Composition of the reactor exit system was synthesized material. For TPA@MCM-41 catalyst, the molar ratio
analyzed by a gas chromatograph (Varian 3800), which was con- of W/Si was found as 0.18. This result was consistent with the ratio
nected on-line to the reactor. Gas chromatograph was equipped in the synthesis solution, indicating that TPA was successfully
with a Porapak T column and a TCD detector. impregnated into the mesoporous MCM-41 support. The material
synthesized by the one-pot hydrothermal synthesis procedure,
namely TRC-W40, was prepared by adjusting the W/Si molar ratio
3. Results and discussions as 0.40 in the synthesis solution. This ratio was found as 0.52 from
EDS analysis of the synthesized material. This result indicated suc-
3.1. Results of catalyst characterization tests cessful incorporation of TPA into the catalyst structure and loss of
some silica during the hydrothermal synthesis and washing steps
Thermal analysis (TGA-DTA) of the uncalcined material, which of the synthesis procedure. As it was mentioned in Section 2.1, this
was synthesized by the one-pot procedure, was performed by a material was repeatedly washed with deionized water after the
Dupont 951 thermal analysis system, at a heating rate of 10 C/ hydrothermal synthesis step. As it was found by the EDS analysis,
min, under the ow of dry air. The weight loss (about 18%), which W/Si ratio of the nal product was even higher than the W/Si ratio
was observed between 200270 C (Fig. 1), corresponds to the re- in the synthesis solution, indicating no leaching of the incorporated
moval of most of the surfactant from the synthesized material. The tungstophosphoric acid from the lattice of the synthesized material
minimum of the corresponding endothermic DTA peak was at as a result of washing steps.
236 C. This result indicated that calcination temperature should Physical and chemical properties of the synthesized materials
be over 250 C. As shown in Fig. 1, there are two other DTA peaks are summarized in Table 1. According to the nitrogen physisorp-
at about 380 C and 495 C. Both of these peaks correspond to the tion analysis, TRC-W40(L) has multipoint BET surface area, pore
exothermic processes, indicating the removal of some oxidized volume and average pore diameter values of 143 m2/g, 0.23 cm3/
species from the synthesized material. Basing on the TGA/DTA g and 5.6 nm, respectively. This material was calcined at 270 C.
results, synthesized materials were calcined either at 270 C Corresponding values of the material, which was calcined at
(denoted as TRC-W40(L)) or at 400 C (denoted as TRC-W40(H)), 400 C (TRC-W40(H)), showed signicantly higher surface area

Table 1
Physical and chemical properties of the synthesized materials.

Sample ID W/Si (atomic) from EDS Surface area (m2/g) BET Pore volume (cm3/g) BJH des. Average pore diameter (nm) BJH des.
MCM-41 1000 1.00 2.9
TRC-W40(L) 0.52 143 0.23 5.6
TRC-W40(H) 0.52 263 0.51 5.6
TPA@MCM-41 0.18 183 0.15 1.4

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
4 A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx

Fig. 4. Pore size distributions of TRC-W40(L) and TRC-W40(H).


Fig. 2. Pore size distribution of TPA@MCM-41.

desorption branches and the closure of the hysteresis loops was


and pore volume values (Table 1). By increasing the calcination
at a relative pressure of nitrogen, as high as 0.85. Such a hysteresis
temperature from 270 C to 400 C, surface area of the material in-
loop is generally considered to indicate signicant interconnectivi-
creased from 143 m2/g to 263 m2/g. Similarly, pore volume in-
ty in the mesoporous network [13]. Major difference of the adsorp-
creased from 0.23 cm3/g to 0.51 cm3/g, by increasing the
tiondesorption isotherms of TRC-W40(L) and TRC-W40(H) is the
calcination temperature. These results showed that 270 C was
higher quantity of adsorbed nitrogen in the material calcined at
not sufcient for the removal of all of the surfactant from the struc-
higher temperature (TRC-W40(H)). This is consistent with the high-
ture of the synthesized material. These results also indicated that
er pore volume of this material than the pore volume of TRC-
the second TGA-DTA peak (Fig. 1), observed in the temperature
W40(L) (Table 1). Pore size distributions of TRC-W40(L) and TRC-
range of 350380 C, is probably due to the removal of carbon
W40(H) showed the presence of mesopores, mostly in the range
containing surfactant remains from the catalyst pores through an
of 210 nm (Fig. 4). In fact, the average pore diameter of both of
exothermic oxidation process.
these materials was 5.6 nm, which is about twice the average pore
Results of nitrogen physisorption analysis of TPA impregnated
diameter of MCM-41.
material (TPA@MCM-41) showed that surface area evaluated by
XRD pattern of the synthesized material, which was calcined at
the BET method was 183 m2/g. Its pore volume and average pore
270 C (TRC-W40(L)), indicated an amorphous structure (Fig. 5a).
diameter values were 0.15 cm3/g and 1.4 nm, respectively. Pore
Typical XRD peaks of the Keggin structure of TPA were not ob-
size distribution of this material indicated the presence of pores
served in this XRD spectrum. In the case of the material calcined
in the range of 110 nm (Fig. 2). Surface area and pore volume val-
at 400 C (TRC-W40(H)), XRD pattern indicated the presence of
ues of this supported catalyst were much lower than that of MCM-
broad diffraction peaks at 2h values of 23.2 and 33, which were
41 support itself, which were 1000 m2/g and 1.0 cm3/g, respec-
tively. These results indicated plugging of signicant fraction of
the pores present in the support material (MCM-41), with the
impregnated TPA. The average pore diameter of the TPA impreg-
nated material (1.4 nm) was also lower than the support material
MCM-41 (Table 1).
The nitrogen adsorptiondesorption isotherms of TRC-W40(L)
and TRC-W40(H) (Fig. 3) were both Type IV, according to the IUPAC
classication, with H2-like hysteresis loops. These isotherms indi-
cated the formation of the ordered mesoporous structures. For
MCM-41-like materials having long range ordered mesopores,
sharp hysteresis loops with parallel adsorption and desorption
branches were expected. The desorption branches of the isotherms
of TRC-W40(L) and TRC-W40(H) were quite sharp. However, the
adsorption branches of these isotherms were not as sharp as the

Fig. 3. Nitrogen adsorptiondesorption isotherms of TRC-W40(L) and TRC-W40(H). Fig. 5. XRD patterns of the synthesized materials.

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx 5

Fig. 6. SEM photographs of TPA@MCM-41 catalyst.

assigned to WOx clusters (Fig. 5b). Similar results were reported in for pure TPA and TPA@MCM-41. However, in the case of TPA
our earlier study with STA incorporated mesoporous materials impregnated MCM-41 (TPA@MCM-41), intensity of the bands cor-
[30,31]. A small but a rather sharp peak observed at a 2h value of responding to Bronsted acid sites was somewhat less. Heteropoly-
25.8 is consistent with the main XRD peak of Keggin structure acids are known to have very high Bronsted acidity. In fact, the
of TPA. These results indicated that Keggin structure was mostly Hammet acidity function of TPA (Ho = 13.2) is even higher than
destroyed and WOx clusters were formed within the matrix of the Hammet acidity function of 100% sufuric acid. While TPA and
the silicate structured mesoporous material prepared by the one- TPA@MCM-41 have high Bronsted acidity, bands observed at
pot hydrothermal procedure, which was calcined at 400 C. Cluster 1447 cm 1 and at about 1600 cm 1 in the spectra also indicated
size of WOx was estimated from the Scherrer equation as 5.4 nm. the presence of some Lewis acid sites. However, intensity of the
Main peaks corresponding to pure TPA, which were observed at bands corresponding to the Lewis acid sites was much lower than
2H = 25.88 and 2H = 34.72, were also seen in the X-ray diffraction the intensity of the bands corresponding to the Bronsted acid sites.
patterns of the TPA impregnated MCM-41 (TPA@MCM-41). As it DRIFT spectra of the pyridine adsorbed catalysts prepared by the
can be seen in Fig. 5a, peaks observed at 2H values of 29.52 and one-pot hydrothermal procedure (TRC-W40(L) and TRC-W40(H))
37.92 also correspond to the Keggin structure of TPA. These results have three distinct IR absorption bands at 1640, 1537 and
indicated that, in the case of TPA impregnated MCM-41 catalyst, 1488 cm 1 (Fig. 10). The bands at 1537 cm 1 and 1640 cm 1 corre-
Keggin structure was essentially conserved. MCM-41 is a mesopor- spond to Bronsted acid sites, indicating strong acidity of the synthe-
ous material having a narrow ordered pore size distribution. The sized materials. The broad band at 1442 cm 1, which belonged to
characteristic peaks of the ordered pore structure of MCM-41 were the Lewis acid sites, was observed to be quite low in intensity.
expected at 2h values of about 2.5, 4.2, 4.8 and 6.2 [22]. Among The band observed at about 1488 cm 1 is due to adsorbed pyridine
these peaks, only the main peak corresponding to d100 was ob- on both Bronsted and Lewis acid sites [11].
served in the low angle range of XRD patterns of TPA@MCM-41 FTIR spectra of pure TPA, TRC-W40(L) and TRC-W40(H) catalyst
(Fig. 5a). Also, the intensity of this peak was quite low. These are given in Fig. 11. Pure TPA showed IR peaks at 1080, 981, 889,
results are consistent with the nitrogen adsorptiondesorption 796, 594, 522 cm 1. First ve of these bands are due to the stretch-
results, which indicated closure of signicant fraction of mesop- ing vibrations POa, WOd, WObW, WOcW, and to the bending
ores of MCM-41 with the impregnated TPA. vibration OaPOa, respectively [8]. The subscripts signify oxygen
SEM photographs of TPA@MCM-41 indicated the presence of bridging the W and the heteroatom (a), corner-sharing (b) and
agglomerated particles, in a particle-size range of 38 lm edge-sharing oxygen (c), belonging to octahedral WO6 and terminal
(Fig. 6). However, in the case of the material synthesized by the oxygen (d). Most of the peaks belonging to pure TPA were also ob-
one-pot procedure, the resulting solid product was not in powder served in the case of TRC-W40(L), but there was a decrease in their
form. After the synthesis, the solid product was crushed to the de- intensities. This is partly due to small quantity of TPA in
sired size. SEM images of TRC-W40 showed that the average size of TRC-W40(L) and also due to some distortion of the TPA phase in
the crushed catalyst particles was about 110 lm (Fig. 7). the lattice of the synthesized material. It seems that the highest
TEM images showed well dispersed nanoballs of about 15 nm deformation was observed in the band corresponding to PO at
in size, which were due to the incorporated TPA within the amor- 1080 cm 1. In the case of TRC-W40(H), which was calcined at
phous silicate matrix of the synthesized material (Fig. 8). These re- 400 C, characteristic peaks of the Keggin structure were mostly de-
sults were consistent with the predicted WOx cluster size from the formed. This is consistent with the XRD results of this material,
XRD analysis. which had indicated deformation of TPA and the formation of
In order to get some information about the relative strengths of WOx clusters within the mesoporous silicate lattice of the synthe-
Bronsted and Lewis acid sites of the synthesized catalyst, diffuse sized material after calcination.
reectance FT-IR spectroscopic (DRIFTS) analysis of pyridine ad-
sorbed samples was performed. By taking the difference of the 3.2. Results of dehydration reaction experiments
spectra of fresh and pyridine adsorbed samples, relative signi-
cance of Lewis and Bronsted acid sites on the surface of these cat- Catalytic performances of TPA incorporated mesoporous cata-
alysts was determined. Bands observed at 1540 and 1640 cm 1 in lysts synthesized in this study were tested in dehydration reac-
the DRIFT spectra of pure TPA and TPA@MCM-41 (Fig. 9) corre- tions of ethanol and methanol. Ethylene, which is a an important
spond to the pyridinium ion adsorbed on Bronsted acid sites. Very feedstock of petrochemical industry and DEE (diethyl ether), which
strong bands corresponding to Bronsted acid sites were observed is an environmentally clean transportation fuel alternate, are the

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
6 A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx

Fig. 7. SEM photographs of TRC-W40. Fig. 8. TEM images of TRC-W40.

main products obtained from dehydration of ethanol. On the other of TPA@MCM-41 catalyst, fractional conversion values reaching
hand, the main product of dehydration of methanol is dimethyl to 0.75 and 0.80 were obtained, respectively, in the present study.
ether, which is considered as one of the best alternatives of con- The main reason of this increase in activity of TPA is much higher
ventional diesel fuel. Fractional conversion of alcohols and the surface area of the supported TPA in comparison to its pure form.
product selectivities were calculated using the data obtained from While the surface area of pure TPA was around 1 m2/g, this new
the online analysis of the reactor exit stream. Data points reported supported catalyst had a surface area of 183 m2/g.
in this paper correspond to the averages of at least three successive The NMR study of Thomas et al. [36] had indicated the loss of
steady state experimental results obtained within a reaction period protons of TPA between 200300 C. However, the work of Obali
of 23 h. and Dogu [4], had indicated stabilization of TPA at higher temper-
atures when supported on activated carbon. In the present study,
activity of the TPA@MCM-41 catalyst was also quite stable in eth-
3.2.1. TPA@MCM-41 in ethanol and methanol dehydration reaction anol conversion, even at temperatures over 350 C, indicating sta-
Experiments were performed using 0.2 g of fresh catalyst in a bilization of TPA when supported on MCM-41.
temperature range of 200400 C and at a space time of
0.27 s.g.cm 3. TPA impregnated MCM-41 catalyst showed very
high activity in dehydration reactions of both ethanol and metha-
nol. Fractional conversion of ethanol sharply increased when reac-
tion temperature was approached to 200 C. As shown in Fig. 12,
fractional conversion of ethanol was approximately 0.75 at
200 C and an increase in activity of TPA@MCM-41 was observed
with an increase of temperature, up to 300 C. At this temperature,
ethanol conversion approached to 100%.
Heteropolyacid catalysts have very high activity in dehydration
reaction of alcohols due to their high Bronsted, acidity. However,
their activity is limited due to their very low surface area. Catalytic
activities of pure heteropolyacid catalysts had been tested in our
earlier work [10] in ethanol dehydration reaction. In that work,
ethanol fractional conversion values of 0.40 and 0.55 were re-
ported at 200 and 250 C, respectively, with pure TPA being the
catalyst. At the same reaction conditions, with the same amount Fig. 9. DRIFT spectra of pyridine adsorbed TPA and TPA@MCM-41 samples.

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx 7

Fig. 10. DRIFT spectra of TRC-W40(H) and TRC-W40(L).


Fig. 13. Selectivity of products in ethanol dehydration over TPA@MCM-41 (space
time: 0.27 s.g/cm3).

Fig. 11. FTIR spectra of pure TPA, TRC-W40(L) and TRC-W40(H).

Fig. 14. Methanol fractional conversion over TPA@MCM-41 (space time 0.27
s.g/cm3).

at temperatures over 300 C. Increase of temperature is expected


to increase thermodynamic limitations on conversion. However,
the decrease observed in methanol dehydration over 220 C was
much higher than it would be predicted from equilibrium consid-
erations. Some decrease of methanol conversion was also reported
at high temperatures, in our recent publication on STA incorpo-
rated mesoporous catalysts [12] and this decrease was attributed
to catalyst deactivation due to coke formation. Similarly the
decrease of catalyst activity observed in this work in methanol
dehydration is considered to be mainly due to coke formation.
Decomposition of produced DME is expected to cause coke forma-
Fig. 12. Ethanol fractional conversion over TPA@MCM-41 (space time 0.27 s.g/cm3). tion at high temperatures. In the case of DME production by meth-
anol dehydration over TPA@MCM-41, operating temperatures
should be kept below 220 C.
Product distributions of ethanol dehydration reaction indicated The main product of methanol dehydration reaction was DME,
the formation of diethyl ether (DEE) as well as ethylene. Selectivity with a selectivity value of approximately 1.0, in the reaction tem-
of DEE was found to be higher than that of ethylene up to 220 C perature range of 150300 C. Apart from DME, small amounts of
(Fig. 13). Above this temperature ethylene became the main prod- formaldehyde and methane formation was also observed as side
uct and its selectivity approached to 1.0 at 250 C. At temperatures products (Fig. 15). Increase in the selectivity of methane was recog-
higher than 220 C, decomposition of DEE was also expected to nized at temperatures higher than 300 C. This result also conrms
contribute to the formation of ethylene. As it was also discussed decomposition of produced DME at high temperatures.
in our recent publication [11], Bronsted acid sites are the main con-
tributor to DEE formation, while Lewis acid sites are also expected 3.2.2. Dehydration reaction results of alcohols with TRC-W40
to contribute to the formation of ethylene at higher temperatures. Although the activity of TPA impregnated MCM-41 (TPA@MCM-
Methanol dehydration reaction test experiments also showed a 41) was very high in dehydration reactions of alcohols, TPA was
sharp increase in methanol conversion from 0.17 to 0.79 by not expected to be strongly attached to the support. Through
increasing reaction temperature from 150 C to 200220 C washing of the supported catalyst with water caused leaching of
(Fig. 14). It was interesting to observe the decrease of methanol some of the impregnated TPA. Consequently, this TPA impregnated
conversion over this temperature. In the case of ethanol dehydra- MCM-41 catalyst was more suitable for vapor phase reactions.
tion, only a slight decrease was observed in fractional conversion However, for the TPA incorporated catalyst prepared by the one

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
8 A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx

Fig. 17. Variation in product selectivities in ethanol dehydration over TRC-W40.

Fig. 15. Variation in product selectivities in methanol dehydration over TPA@MCM-


41 (space time: 0.27 s.g/cm3).
the main product at temperatures lower than 250 C. DEE selectiv-
ity was close to 0.9 at 200 C. DEE selectivity decreased and ethyl-
ene selectivity increased with an increase in temperature.
pot procedure described in this manuscript (TRC-W40), no leach-
Selectivity of ethylene exceeded 0.9 by increasing reaction temper-
ing of heteropolyacid from the catalyst matrix was observed after
ature over 350 C. These results are consistent with the results
it was repeatedly washed with water. In fact, in the synthesis pro-
obtained with TPA@MCM-41 catalyst.
cedure of these catalysts, washing of the solid product obtained
In the case of methanol dehydration, main product was DME. In
after the hydrothermal synthesis step was continued until the pH
fact, even at 180 C, DME selectivity was 0.97. This value further
of the wash water was close to neutral. No loss of tungsten from
increased with an increase in temperature, approaching to 1.0 at
the catalyst lattice was observed during this washing step. The
250 C. Besides DME, some formaldehyde formation was also
W/Si ratio of the synthesized material (after the washing steps)
observed, at lower temperatures. These results proved that the
was found to be even greater than the W/Si ratio in the synthesis
dehydration activity of the synthesized catalysts were much higher
solution. This is due to stable distribution of the WOx clusters with-
than their dehydrogenation activity.
in the silicate structured mesoporous matrix of the catalyst.
Experimental results for ethanol dehydration over TRC-W40
showed that fractional conversion increased with temperature, 4. Conclusions
approaching to 1.0 at around 400 C (Fig. 16). Although the ethanol
conversion values obtained with TRC-W40 were somewhat less High surface are TPA based mesoporous catalysts were success-
than the values obtained with TPA@MCM-41, this material was fully synthesized using one-pot hydrothermal and impregnation
quite stable, even at temperatures as high as 400 C. Higher activity procedures. In the synthesis of TPA incorporated silicate structured
of TPA@MCM-41 than TRC-W40 is consistent with the higher acid- mesoporous catalysts by the one-pot hydrothermal route
ity of this catalyst, as indicated by DRIFTS analysis of pyridine (TRC-W40), calcination temperature was found to be quite impor-
adsorbed materials. tant on the properties of the synthesized material. Well dispersion
A similar trend was observed in methanol dehydration (Fig. 16). of WOx clusters was achieved within the silicate structured meso-
Ethanol and methanol conversion values were quite close in a tem- porous matrix of the catalytic material, which was calcined at
perature range between 200300 C. Experiments were not per- 400 C. Incorporated TPA was not washed away from the structure
formed at temperatures higher than 300 C, in methanol of these materials. This catalyst showed highly stable catalytic per-
dehydration tests. As it was observed with TPA@MCM-41, methane formance both in ethanol and methanol dehydration reactions.
formation cause a decrease in DME selectivity at higher tempera- Highly stable ethanol dehydration results, with ethanol fraction
tures. At 180 C somewhat higher conversion was obtained with conversion values approaching to 1.0, were obtained with an in-
methanol than ethanol. It was also quite promising to observe no crease of temperature up to 400 C. At such high temperatures, eth-
deactivation of TRC-W40 due to coke formation in methanol dehy- ylene selectivity also approached to 1.0, while the main product was
dration, in the studied temperature range. DEE at temperatures lower than 250 C. Very high DME selectivity
Product distributions obtained in ethanol dehydration experi- values of about 1.0 were highly promising in the case of methanol
ments are illustrated in Fig. 17. As shown in this gure, DEE is dehydration performed in the temperature range of 180300 C.
Very high and stable ethanol fractional conversion values,
approaching to 1.0, were obtained at 300 C with the TPA impreg-
nated MCM-41-like mesoporous catalyst (TPA@MCM-41). Pro-
duced DEE was concluded to be further dehydrated to ethylene
over 220 C. In the case of methanol dehydration, DME yield passed
through a maximum at about 220 C, above which coke formation
caused catalyst deactivation. Best operating temperature was con-
cluded as about 200 C, in DME production from methanol over
TPA@MCM-41.
MCM-41 impregnated catalyst (TPA@MCM-41), was shown to
have higher activity than pure TPA, due to its much higher surface
area. Activity of this catalyst and its Bronsted acidity was also high-
er than the material prepared by the one-pot procedure (TRC-
W40). However, the material prepared by the one-pot procedure
was more stable and did not cause any coke formation up to
Fig. 16. Ethanol and methanol fractional conversions over TRC-W40. 300 C. Besides, TPA incorporated into this material was quite

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016
A. Ciftci et al. / Chemical Engineering Journal xxx (2012) xxxxxx 9

stable and did not leach out during the repeated washing steps [15] S. Soled, S. Miseo, G. McVicker, W.E. Gates, A. Gutierrez, J. Paes, Chem. Eng. J. 64
(1996) 235254.
with water. Results proved that TPA incorporated silicate struc-
[16] Y. Liu, L. Xu, B.B. Xu, Z.K. Li, L.P. Jia, W.H. Guo, J. Mol. Catal. A: Chem. 297 (2009)
tured mesoporous materials prepared in this work were highly 8692.
promising solid acid catalysts for dehydration reaction of alcohols. [17] G. Kamalakar, K. Komura, Y. Sugi, Appl. Catal A: Gen. 310 (2006) 155163.
[18] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck, Nature 359 (1992)
710712.
Acknowledgements [19] X. Guo, M. Lai, Y. Kong, W. Ding, Q. Yan, Langmuir 20 (2004) 28792882.
[20] A. Taguchi, F. Schuth, Microporous Mesoporous Mater. 77 (2005) 145.
Financial support of TUBITAK through Project No. 107M184 and [21] O. Terasaki, Stud. Surf. Sci. Catal. 148 (2004) 241260.
[22] Y. Gucbilmez, T. Dogu, S. Balci, Ind. Eng. Chem. Res. 45 (2006) 34963502.
contributions of Prof. Gulsen Dogu are gratefully acknowledged. [23] J. Toufaily, M. Soulard, J.L. Guth, J. Patarin, L. Delmonte, T. Hamieha, M. Kodeih,
D. Naoufal, H. Hamada, Colloid. Surf. A 316 (2008) 285291.
References [24] G. Karthikeyan, A. Pandurangan, J. Mol. Catal. A: Chem. 311 (2009) 3645.
[25] R. Palcheva, A. Spojakina, L. Dimitrov, K. Jiratova, Microporous Mesoporous
Mater. 122 (2009) 128134.
[1] M. Misono, I. Ono, G. Koyano, A. Aoshima, Pure Appl. Chem. 72 (2000) 1305
[26] H.J. Kim, Y.G. Shul, H. Han, Appl. Catal A: Gen. 299 (2006) 4651.
1311.
[27] T. Dogu, D. Varisli, Turk. J. Chem. 31 (2007) 551567.
[2] T. Okuhara, N. Mizuno, M. Misono, Appl. Catal. A-Gen. 222 (2001) 6377.
[28] A. Takahara, M. Saito, M. Inaba, K. Murata, Catal. Lett. 105 (2005) 249252.
[3] F. Cavani, Catal. Today 41 (1998) 7386.
[29] T. Kito-Borsa, D.A. Pacas, S. Selim, S.W. Cowley, Ind. Eng. Chem. Res. 37 (1998)
[4] Z. Obali, T. Dogu, Chem. Eng. J. 138 (2008) 548555.
33663374.
[5] L. Degirmenci, N. Oktar, G. Dogu, Ind. Eng. Chem. Res. 48 (2009) 25662576.
[30] D. Varisli, T. Dogu, G. Dogu, Ind. Eng. Chem. Res. 48 (2009) 93949401.
[6] L. Degirmenci, N. Oktar, G. Dogu, Fuel Process. Technol. 91 (2010) 737742.
[31] D. Varisli, T. Dogu, G. Dogu, Chem. Eng. Sci. 65 (2010) 153159.
[7] S. Shikata, T. Okuhara, M. Misono, J. Mol. Catal. A: Chem. 100 (1995) 4559.
[32] D. Varisli, K.C. Tokay, A. Ciftci, T. Dogu, G. Dogu, Turk. J. Chem. 33 (2009) 355
[8] P. Vzquez, L. Pizzio, C. Cceres, M. Blanco, H. Thomas, E. Alesso, L.
366.
Finkielsztein, B. Lantano, G. Moltrasio, J. Aguirre, J. Mol. Catal. A: Chem. 161
[33] K.C. Tokay, T. Dogu, G. Dogu, Chem. Eng. J. doi: 10.1016/j.cej.2011.12.034, in
(2000) 223232.
press.
[9] J. Haber, K. Pamin, L. Matachowski, B. Napruszewska, J. Poltowicz, J. Catal. 207
[34] A. Ciftci, Nanocomposite Naon and Heteropolyacid Incorporated Mesoporous
(2002) 296306.
Catalysts for Dimethyl Ether Synthesis from Methanol, MS Thesis, Middle East
[10] D. Varisli, T. Dogu, G. Dogu, Chem. Eng. Sci. 62 (2007) 53495352.
Technical University, Ankara, Turkey, 2009.
[11] D. Varisli, T. Dogu, G. Dogu, Ind. Eng. Chem. Res. 47 (2008) 40714076.
[35] M. Fujiwara, K. Shiokawa, Y.C. Zhu, J. Mol. Catal. A: Chem. 264 (2007) 153161.
[12] A. Ciftci, D. Varisli, T. Dogu, Int. J. Chem. React. Eng. 8 (2010) A45.
[36] A. Thomas, C. Dablemont, J.M. Basset, F. Lafebre, C.R. Chimie 8 (2005) 1969
[13] A. Ciftci, N.A. Sezgi, T. Dogu, Ind. Eng. Chem. Res. 49 (2010) 67536762.
1974.
[14] H. Atia, U. Armboster, A. Martin, J. Catal. 128 (2009) 307312.

Please cite this article in press as: A. Ciftci et al., Dimethyl ether, diethyl ether & ethylene from alcohols over tungstophosphoric acid based mesoporous
catalysts, Chem. Eng. J. (2012), http://dx.doi.org/10.1016/j.cej.2012.04.016

Вам также может понравиться