Anda di halaman 1dari 11

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.elsevier.com/locate/he

An experimental study on single-step dimethyl


ether (DME) synthesis from hydrogen and carbon
monoxide under various catalysts
Men-Han Huang a, How-Ming Lee a, Kuo-Chao Liang a,
Chin-Ching Tzeng a, Wei-Hsin Chen b,*
a
b

Institute of Nuclear Energy Research, Taoyuan 325, Taiwan


Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan

article info

abstract

Article history:

Dimethyl ether (DME) is a potential and green substitute to diesel and liquefied petroleum

Received 18 April 2015

gas. In this study, DME synthesized from hydrogen and carbon monoxide (i.e. syngas)

Received in revised form

through a single step process is investigated to figure out the reaction characteristics. The

28 July 2015

influences of reaction temperature, H2/CO molar ratio, and prepared catalyst on DME

Accepted 30 July 2015

synthesis are investigated, while the space velocity is fixed at 15,000 mL (gcat h)1. The

Available online 5 September 2015

results indicate that the optimal reaction temperature for DME synthesis develops at 220  C
where the chemical kinetics and thermodynamic equilibrium are in a comparable state.

Keywords:

Increasing H2/CO ratio increases the CO conversion but lowers the H2 conversion. The

Dimethyl ether (DME)

maximum DME yield is 2.3 g (gcat h)1 which occurs at H2/CO 1. The addition of palladium

Single step and direct synthesis

(Pd) into a CueZnOeAl2O3 catalyst intensifies the CO conversion and DME yield of a gas

Physically mixed catalyst

mixture with syngas and 10 vol% of CO2. This is the consequence of hydrogen spillover

Catalyst acidity

which is able to increase the stability of active Cu against CO2 oxidation. The results also

Syngas and CO2

suggest that the dehydration catalyst with higher acidity gives lower DME selectivity and
yield. The higher the CO2 concentration in syngas, the lower the CO conversion and DME
yield. The present study has provided comprehensive insights into DME synthesis which is
conducive to DME production in industry.
Copyright 2015, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

Introduction
On account of growth in global population and demand for
higher living standards, the energy resources are being
consumed at a more rapid rate. Currently, most of the utilized
energy resources come from fossil fuels, which are mainly
consumed by means of combustion. This leads to the

significant deterioration of environmental quality. For these


reasons, the development of green substitutes to fossil fuels
has become an important task for energy and environmental
sustainability.
Dimethyl ether (DME) is a potential alternative fuel to
diesel used in compression ignition engines [1,2]. This is
attributed to the higher cetane number (55e60) and lower

* Corresponding author. Tel.: 886 6 2004456; fax: 886 6 2389940.


E-mail address: weihsinchen@gmail.com (W.-H. Chen).
http://dx.doi.org/10.1016/j.ijhydene.2015.07.168
0360-3199/Copyright 2015, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.

13584

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

auto-ignition temperature (235  C) of DME when compared to


those of diesel (40e50 and 250  C) [3,4]. The physical properties
of DME are close to those of liquefied petroleum gas (LPG) [5,6];
therefore, DME can also be used as a LPG substitute for
household and industrial purposes. Other advantages
accompanied by DME utilization as a fuel or feedstock include:
(1) higher oxygen content and lower boiling point; (2) less air
pollutants emissions from its burning; (3) more efficient
reforming for hydrogen production; and (4) lower risk in
damaging health during its usage [3,7].
As far as DME production is concerned, it can be categorized into a two-step process and a single step process [8]. The
former and the latter are termed the indirect method and the
direct method, respectively. In the conventional indirect
method, methanol is synthesized from synthesis gas (or
syngas) on a metallic catalyst in the first step where the reaction of CO hydrogenation is expressed as:
CO 2H2 4CH3 OH DH0 90:6 kJ=mol

(1)

In the second step, DME is subsequently produced through


methanol dehydration on an acid catalyst, and the chemical
reaction is given by [9e12]:
2CH3 OH4CH3 OCH3 H2 O DH0 23:4 kJ=mol

(2)

In the single step process [10,13], not only methanol synthesis and dehydration but also a water gas shift reaction
(WGSR) is involved. WGSR is written as [14]:
CO H2 O4CO2 H2

DH0 41:2 kJ=mol

(3)

Accordingly, the net reaction in the single step process is


given by
3CO 3H2 4CH3 OCH3 CO2

DH0 245:8 kJ=mol

(4)

Syngas is the feedstock in DME synthesis. If syngas is


produced from biomass gasification [15], biogas (methane)
steam reforming, or biogas partial oxidation [16,17], DME
synthesized from the syngas, so called bio-DME, can be
thought of as a renewable and green fuel.
In the direct method, bifunctional catalysts play a key role
in carrying out DME synthesis. Two kinds of active site for
simultaneously triggering methanol synthesis and dehydration are integrated in a bifunctional catalyst [18,19]. After
methanol is generated from syngas in the catalyst, it is
dehydrated immediately, thereby enhancing the forward reaction and abating the reverse reaction of methanol synthesis.
Meanwhile, water produced from methanol dehydration is
consumed by CO via WGSR to generate H2. The aforementioned two mechanisms intensify the forward reactions of
methanol synthesis and dehydration; this results in a high
degree of synergistic effect for effective conversion of CO to
DME [18]. In recent years, more industrial interests have been
paid to the single step process because of its lower thermodynamic limitation, lower investment, and higher theoretical
significance [8,20,13].
In bifunctional catalysts, Cu-based catalysts, say, CuOeZnOeAl2O3 catalysts [21,22], are the most commonly adopted
catalysts to drive methanol synthesis, owing to their high
activity and selectivity. Alternatively, a number of acid catalysts, such as g-Al2O3, silicaealumina, and HZSM-5 zeolites,

HY, HMCM-49, HMCM-22, SAPOs, and ferrierite [7], have been


employed to fulfill methanol dehydration. In the single step
process, methanol synthesis (i.e. Eq (1)) is the limiting mechanism and the methanol yield is proportional to the surface
area of active Cu [23]. However, the reaction of CO with H2
appears to shift the active site Cu toward Cu0 so as to decay
the catalyst activity [24]. When CO2 is contained in syngas, the
following methanol synthesis reaction is involved [25]:
CO2 3H2 4CH3 OH H2 O DH0 49:5 kJ=mol

(5)

This implies that CO2 utilization can be achieved in


methanol synthesis. Seeing that less heat is liberated from Eq.
(5) when compared to Eq. (1), adding CO2 into syngas is able to
suppress the increase in temperature and avoid overheating
phenomenon in the catalyst, thereby enhancing its thermal
stability. Moreover, weakly oxidative CO2 in syngas is conducive to Cu formation and stabilizes methanol synthesis reaction [24]. Despite the aforementioned merits of CO2 addition
into syngas, H2O is also generated, as expressed in Eq. (5), and
will stay on the catalyst surface. This lessens the catalyst activity [26] and DME formation [7].
The net reaction of the single step process, namely, Eq. (4),
reveals that a high CO2 concentration disadvantages DME
formation, in accordance with the thermodynamic equilibrium or Le Chatelier's principle. In spite of many outstanding
studied conducted, the information of DME synthesis under
various operating conditions and prepared catalysts remains
insufficient. For this reason, the present study aims to: (1)
develop a number of physically mixed catalysts to achieve the
single step DME synthesis; (2) investigate DME formation at
various operating conditions under the prepared catalysts; (3)
evaluate the impact of CO2 on DME synthesis. The obtained
results in the present study can provide comprehensive insights into DME synthesis and are conducive to DME production in industry.

Experimental
Catalyst preparation
The physically mixed catalysts for synthesizing DME were the
physical mixtures of a syngas-to-methanol catalyst (i.e.,
CueZnOeAl2O3 or PdeCueZnOeAl2O3) and a dehydration
catalyst (i.e., g-Al2O3, ferrierite, or HZSM-5). The CueZnOeAl2O3 catalyst was prepared by a coprecipitation method. A
1.0 M solution contained Cu(NO3)2$2.5H2O, Zn(NO3)2$6H2O and
Al(NO3)3$9H2O with a Cu:Zn:Al molar ratio of 6:3:1 was prepared. The solution was slowly titrated with 1.0 M Na2CO3 to
maintain the pH at 7.0 0.1, in a heated beaker containing
deionized water of 70 0.5  C. A continuous-stirring titration
process took a total time of about 1 h where coprecipitations
took place. The titration solution was kept stirring for additional 3 h for aging, followed by filtering, washing, and drying
at 120  C for 12 h. The precipitate was calcined in air at a
heating rate of 1  C min1 up to 350  C and held for 2 h. Finally,
the prepared catalyst particles were physically mixed with
commercized g-Al2O3 or HZSM-5 dehydration catalyst to form
the physically mixed catalyst. All catalysts tested in this study

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

were pressed, crashed, and sieved to 250e500 mm (i.e., 35 and


60 mesh). The set of PdeCuO/ZnO/Al2O3 catalysts were prepared with the impregnation method, by impregnating
appropriate amounts of Pd(NO3)2 solution onto the CuO/ZnO/
Al2O3 catalyst powders prepared previously, followed by drying at 120  C for 12 h, calcined in air at a heating rate of
1  C min1 from the room temperature to 350  C, and kept for
2 h.
The weight percentages of Cu, Zn and Al in the prepared
CueZnOeAl2O3 catalyst were 47.3, 23.9 and 2.9 wt%, respectively. Accordingly, the atomic ratio between the three metals

13585

Cu, Zn, and Al was 1: 0.49: 0.14. The BET surface area, pore
volume, and pore radius of the prepared catalyst were
approximately 69 m2 g1, 0.4 mL g1 and 9.8 nm, respectively.
With regard to the adopted ZSM5, it was a commercial zeolite
supplied by Zeolyst (CBV5524G). The SiO2/Al2O3 molar ratio of
the ZSM-5 was 50 and the nominal cation form was ammonium. The weight percentage of Na2O in the ZSM-5 was
0.05 wt% and its BET surface area was 425 m2 g1. It should be
illustrated that the ZSM-5 was not treated or modified with an
alkali. The ZSM-5 was reduced to HZSM-5 in a gas environment containing 50% H2 and 50% N2 with the total flow rate of

Fig. 1 e (a) SEM image ( 3000), EDS maps of (b) Cu, (c) Zn, and (d) Al, and (e) XRD pattern of fresh CuOeZnOeAl2O3 catalyst.

13586

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

150 mL min1 and at a ramping rate of 4  C min1 from the


room temperature to 320  C. Then it was kept for 60 min and
naturally cooled down for 120 min.
In general, the formation temperature of aromatics with
high molecular weight (e.g., 350e450  C) is higher than the
temperature for DME synthesis (i.e. 200e300  C) by around
100  C. In the experiments, the signals detected by the microGC for hydrocarbon compounds above C3 were relatively
small. Therefore, the formation of hydrogen compounds
above C3 was ignored in this study.

Reaction system
The entire reaction system comprised: (1) a reactant feeding
unit; (2) a preheating and reaction unit; and (3) a gas analysis
unit. In the feeding unit, four gases of H2, N2, CO, and CO2
were individually stored in steel cylinders, and their flow
rates were controlled by mass flow controllers (MKP, Model
TSC-100) to produce a desired gas mixture or feed gas. In the
preheating and reaction unit, a preheater, a reactor, and a
back pressure valve were installed in series. The feed gas was
preheated by a heating tape in the preheater. The catalyst bed
reactor was made of a 316 stainless steel tube (12.7 mm o.d.)
packed with a mixture of quartz sands and catalysts. The
tube was placed and heated in an oven. Four K-type thermocouples were installed in the system; one was at the exit of
the catalyst bed and the other three thermocouples were
aligned along the tubular reactor. The former and the latter
were employed for measuring and controlling the bed temperature, respectively. The pressure in the reactor was

regulated by the back pressure valve. The gas analysis unit


was composed of an online micro gas chromatography
(Micro-GC, Varian, CP-4900) and a computer. The concentrations of H2, CO, CO2, methanol, and DME in the product gas
were detected by the GC. The signals from the GC were displayed and recorded by the computer.

Experimental procedure
Prior to performing experiments, the prepared catalysts were
sieved to particle sizes ranging from 250 to 500 mm (i.e.,
35e60 mesh). Then, 1 g of methanol synthesis catalyst was
mixed with 0.2 g of methanol dehydration catalyst and packed
in the reaction tube. The physically mixed catalysts were
reduced by a gas mixture (H2/N2 1/1) at a flow rate of
150 mL min1 and 1 atm where the catalyst bed was heated
from room temperature to 320  C at a heating rate of
4  C min1, and naturally cooled down for 2 h. The reaction
temperature in the range of 200e260  C was considered. The
space velocity is defined as the volumetric flow rate of the feed
gas per unit weight catalyst, whereby a high space velocity
stands for a short residence time of the feed gas in the catalyst
bed. In general, the space velocity for DME synthesis is between 800 and 6000 mL (gcat h)1 [7]. In this study, a high space
velocity of 15,000 mL (gcat h)1 was adopted for all tests and the
reaction pressure was kept at 40 bar. The feed gas was preheated to 120  C by the preheater. The product gas was online
analyzed by the micro-GC. In the experiments, an inert gas
(i.e., N2) was used as an internal standard for calculation of the
flow rate of effluent, and the volumetric concentration of N2 in

Fig. 2 e (a) SEM image ( 3000) and EDS maps of (b) Cu, (c) Zn, and (d) Al of reacted CuOeZnOeAl2O3 catalyst.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

the feed gas was 4 vol%. N2 was assumed to be conserved in


the reaction system. Once the concentrations of H2, CO, CO2,
methanol, and DME at the exit of the reaction system were
measured, their molar flow rates were calculated using Excel
based on the nitrogen tracer method [27]. To ascertain the
experimental quality, the reproducibility of experiments had
been tested, and the relative error of experiment was below
5%.

Indexes of reactions

H2 conversion %

CO conversion %

n_CO;in  n_CO;out
 100
n_CO;in

(6)

n_H2 ;in  n_H2 ;out


 100
n_H2 ;in

(7)

DME selectivity %

2n_DME;out
 100
2n_DME;out n_MeOH;out n_CO2 ;out

(8)

DME yield g=gcat ,h

n_DME;out g=h
mcat gcat

(9)

In DME synthesis, CO and H2 are employed as reactants.


Hence their conversions are important indicators to evaluate
the performance of the catalysts. Higher CO and H2 conversions imply the better the reaction of CO with H2. In addition
to DME, methanol is also produced in the reactions. Therefore,
the DME selectivity plays an important role in evaluating DME
synthesis from the competition between methanol and DME
formations. As a whole, the higher the CO and H2 conversions
as well as DME selectivity and yield, the better is the performance of the catalyst. The four indexes of CO conversion, H2
conversion, DME selectivity, and DME yield [7,28] are defined
as follows.

13587

where n and m designate the molar flow rate (mol h1) and
mass flow rate (g h1), respectively, and the subscripts in and
out represent the inlet and outlet of the reactor, respectively.

Results and discussion


Catalyst characterization
The scanning electron microscope (SEM, Hitachi S3000N) images, energy dispersive spectrometer (EDS, Horiba EX-220)
maps, and X-ray diffraction (XRD, Bruker AXS, D8) of the
fresh CuOeZnOeAl2O3 catalyst are shown in Fig. 1. The SEM
image of the catalyst (Fig. 1a) with the amplification of 3000
indicates that there are many tiny holes on the catalyst surface. In fact, the sizes of the holes are in nanoscale, and this

Fig. 3 e (a) SEM image ( 3000) and (b) EDS map of Pd of the catalyst with 5 wt% of Pd and (c) SEM image ( 3000) and (d) EDS
maps of Pd of the catalyst with 10 wt% of Pd.

13588

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

Effect of reaction temperature


The reaction temperature plays a pivotal role on the performance of DME synthesis. From the kinetic point of view, a
higher temperature facilitates the reaction rates of methanol
and DME syntheses. On the other hand, from the viewpoint of
thermodynamics, a low temperature is conducive to the
methanol and DME formations due to the exothermic reactions involved; that is, a lower reaction temperature favors
the forward reactions of Eqs. (1) and (2), in accordance with the
Le Chatelier's principle. Fig. 4 demonstrates the distributions
of H2 and CO conversions as well as DME yield at the feed gas
of H2/CO/CO2/N2 61/30/5/4 (vol%). The reaction temperature
of 180  C is not high enough to counteract the activation energy of methanol synthesis; hence the H2 and CO conversions
as well as the DME yield are very low. When the temperature
reaches 200  C, methanol and DME syntheses are triggered,
revealing that this temperature behaves as a threshold to
activate the syntheses. When the temperature is increased to
220  C, the H2 and CO conversions as well as DME yield are
further pushed up, and their values are approximately 65%,
36%, and 2.1 g (gcat h)1, respectively. With further increasing
temperature to 240  C, however, a decrease rather than an
increase in the conversions and yield is exhibited. When the
temperature increases from 240 to 260  C, there is an obvious

Fig. 4 e Distributions of (a) H2 and CO conversions and (b)


DME yield with time and temperature.

can increase the specific surface area of the catalyst. The EDS
maps of Cu, Zn, and Al (Fig. 1bed) depict that these element
are uniformly distributed on the catalyst surface, reflecting
the successful preparation of the CuOeZnOeAl2O3 catalyst.
The diffraction peak breadth of CuO is broadened (Fig. 1e),
implying that ZnO is able to disperse CuO [29]. The crystallite
size of CuO can be evaluated from the Scherrer equation [30].

lK
L  cosq

(10)

where b is the crystallite size contribution to the peak width


(or full width at half maximum, FWHM) in radians
(0.0253 rad), l is wavelength (0.15406 nm), K is a shape
factor (0.94 for spherical crystallites), L is the average
thickness of the crystal in a direction normal to the diffracting
plane or crystallite size (nm), and q is Bragg angle (0.34 rad).
Accordingly, the crystallite size (L) of CuO is 6.1 nm. The SEM
image of the reacted CuOeZnOeAl2O3 catalyst (Fig. 2a) suggests that its surface is not as dispersive as the fresh one.
Nevertheless, the elements Cu, Zn, Al, and O are still uniformly distributed on the catalyst surface (Fig. 2bed). Meanwhile, the SEM images and the dispersity of Pd on the surface
of the PdeCueZneAl catalyst with two different Pd weight
percentages (5 and 10 wt%) have been examined by EDS and
shown in Fig. 3. As can been seen in the figure, the Pd is well
dispersed on the surface of the catalyst.

Fig. 5 e Distributions of (a) H2 conversion and (b) CO


conversion with time and temperature. (H2/CO//N2 64/32/
4).

13589

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

Chatelier's principle. In light of Eq. (4), the theoretically selectivities of DME and CO2 are 66.7% and 33.3%, respectively.
Fig. 6a depicts that the DME selectivity is around 58%, which is
close to the theoretical one. The selectivity is insensitive to the
reaction temperature and H2/CO ratio. However, Fig. 6b re1
veals that the DME yield at H2/CO 1 (2.3 g g1
cat h ) is higher
1
1
than at H2/CO 2 (2.1 g gcat h ), mainly stemming from a
higher CO amount provided at the former.

Influence of Pd in catalyst
It has been reported that the addition of palladium (Pd) into
catalysts could facilitate the performance of the catalysts for
DME synthesis [32]. To realizing the influence of Pd in the
physically mixed catalyst on DME synthesis in a high CO2
environment (10 vol%), three different physically mixed catalysts with Pd contents of 0, 5, and 10 wt% are tested where
the composition of the feed gas is H2/CO/CO2/N2 61/25/10/4.
Fig. 7a reveals that the H2 conversion is characterized by 5%
Pd > 0% Pd > 10% Pd, while Fig. 7b depicts that the CO conversion is featured by 10% Pd > 5% Pd > 0% Pd. Fig. 8a depicts
that the DEM selectivity is almost invariant when the catalyst
and temperature are changed and its value is approximately
60%. Fig. 8b indicates that the DME yield is ranked as: 10%
Pd > 5% Pd > 0% Pd. It is thus recognized that the addition of Pd

(a)

Effect of H2/CO ratio


The stoichiometric ratios between H2 and CO in the two-step
and single step processes are 2 and 1, respectively, as
expressed in Eqs. (1) and (4). Therefore, two H2/CO molar ratios
of 2 and 1 are experimentally tested to figure out the influence
of the ratio on DME synthesis. In Figs. 5 and 6, two different
feed gases are tested; they are H2/CO/N2 64/32/4 (H2/CO 2)
and 48/48/4 (H2/CO 1). Fig. 5 depicts that the H2 conversion at
H2/CO 2 is lower than that at H2/CO 1, whereas the adverse
behavior on CO conversion is observed. Compared to H2/
CO 1, more H2 supplied at H2/CO 2 results in a lower H2
conversion and a higher CO conversion, in light of Le

H2 conversion (%)

220

200

240

220

120

0% Pd
5% Pd
10% Pd

40
30
20
10
0

(b)
80

CO conversion (%)

drop in the DME yield, implying that this temperature is fairly


unfavorable to DME formation.
When the temperature decreases from 260 to 240 or 220  C,
the values of H2 and CO conversions as well as DME yield are
the same as those in the rising temperature process. That is, no
hysteresis effect [31] develops in the ascending and descending
temperature processes. This means that the performance of
the catalyst is stable and also reversible within the testing
period of 14 h. Overall, the optimal reaction is implemented at
220  C. It follows that chemical kinetics dominates DME synthesis for the reaction temperature no greater than 220  C.
Once the temperature is higher than 220  C, thermodynamics
becomes the predominant factor in DME synthesis.

180

50

Fig. 6 e Distributions of (a) DME selectivity and (b) DME


yield with time and temperature.

Temp. (oC)
260
240

200

180

Time (h)

10

Temp. (oC)
260
240

220

12

240

14

220

16

120

60

40
0% Pd
5% Pd
10% Pd

20

Time (h)

10

12

14

Fig. 7 e Distributions of (a) H2 conversion and (b) CO


conversion with time and temperature under the
CueZnOeAl2O3 catalyst with different Pd contents.

16

13590

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

Temp. (oC)
240
260

(a)

DME selectivity (%)

200

180

100

220

240

220

Table 1 e Acid properties of three dehydration catalysts


tested in the study.

120

Catalyst
80

Weak acid site (a.u.)a Strong acid site (a.u.)


HZSM-5
ferrierite (FER)
g-Al2O3

60
40

0% Pd
5% Pd
10% Pd

20
0

Desorption temperature ( C)

(b)

200

180

2.5

Time (h)

10

Temp. (oC)
240
260

220

12

240

14

220

16

215
260
190

475
555
e

An arbitrary unit (a.u.) of a TCD signal.

of active Cu by CO2 and, thereby, reduces the production rate


of methanol [33,34]. The CO conversion featured by 10%
Pd > 5% Pd > 0% Pd (Fig. 7b) suggests that the addition of Pd
into the catalyst is able to intensify CO conversion. This can be
explained by hydrogen spillover, which is induced by Pd, from

120

DME (g/gcat.h)

2
1.5
1
0% Pd
5% Pd
10% Pd

0.5
0

Time (h)

10

12

14

16

Fig. 8 e Distributions of (a) DME selectivity and (b) DME


yield with time and temperature under the CueZnOeAl2O3
catalyst with different Pd contents.

into a physically mixed catalyst can enhance the tolerance of


the catalyst to CO2.
A higher CO2 concentration leads to a higher coverage of
oxygen on the catalyst surface, which results in the oxidation

Fig. 9 e NH3-TPD profiles of three different solid acid


catalysts.

Fig. 10 e Profiles of (a) DME selectivity and (b) DME yield


under three different hydration catalysts.

13591

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

Table 2 e Influence of g-Al2O3 to CuOeZnOeAl2O3 ratio on


DME selectivity.
g-Al2O3 to CuOeZnOeAl2O3
ratio (g/g)

DME selectivity (%)

0.1
0.2
0.4

57.5
61.6
58.5

metallic Pd to metallic Cu [33]. This spillover increases the


stability of active Cu against CO2 oxidation.

Influence of dehydration catalyst


Fig. 9 shows the ammonia temperature-programmed
desorption (NH3-TPD) profiles of three dehydration catalysts,
including HZSM-5, ferrierite (FER), and g-Al2O3. The NH3-TPD
tests were carried out on a catalyst analyzer (BELCAT, B-85) by
the following procedure: (1) each catalyst (0.1 g) was preheated
to 500  C for 1 h in a ultra-pure He stream; (2) the catalyst was
cooled to 100  C and saturated with 5% NH3 in He for 30 min;
(3) the physisorbed NH3 was desorbed under a constant He
flow of 30 mL min1 at a heating rate of 10  C min1 from 100 to
650  C; and (4) the acidity of the catalyst was analyzed via the
NH3 concentration in the exhaust stream which was online
monitored by a thermal conductivity detector (TCD). As
shown in Fig. 9, the profiles of HZSM-5 and ferrierite are
characterized by two peaks; one is at 150e300  C and the other
is at 400e600  C, representing the weak and strong acid sites,
respectively. The acid properties of the three dehydration
catalysts are summarized in Table 1. The unit of the acid sites
is an arbitrary unit (a.u.) of a TCD signal. HZSM-5 and FER have
both weak and strong acid sites, and the intensities are
abundantly high. Alternatively, g-Al2O3 only has weak acid
sites and its intensity is relatively low. Overall, the acidity of
the catalysts is in the following order: FER > HZSM-5 >> gAl2O3.
The profiles of DME selectivity and yield under the
CuOeZnOeAl2O3 catalyst individually blended with the three
acid catalysts are sketched in Fig. 10 where the feed gas
composition is H2/CO/CO2/N2 63/30/3/4. Fig. 10a depicts that
g-Al2O3 has the highest DME selectivity among the three acid
catalysts and its value is around 59%. On the other hand,
HZSM-5 gives a higher DME yield as compared to FER and gAl2O3. It was reported that g-Al2O3 exhibited only Lewis
acidity, while HZSM-5 had both Lewis and Brnsted acid sites
[35]. Water produced from the reverse WGSR has a strongly
inhibiting effect on the Lewis acidity, implying that it

markedly retards the activity of g-Al2O3, whereas the effect is


less significant over HZSM-5. This is the reason why the DME
yield under HZSM-5 is higher than that under g-Al2O3.
The values of DME selectivity and yield at three different
amounts of the dehydration catalyst g-Al2O3 in 1 g CuOeZnOeAl2O3 catalyst are given in Table 2 where the feed gas
composition is H2/CO/CO2/N2 63/30/3/4. It is noteworthy
that there exists an optimum g-Al2O3 to CuOeZnOeAl2O3
ratio. Theoretically, the more the dehydration catalyst used,
the higher the DME selectivity is. For instance, the DME
selectivity increases from 57.5% to 61.6% when the g-Al2O3 to
CuOeZnOeAl2O3 ratio increases from 0.1 to 0.2. However, the
selectivity declines to 58.5% when the ratio is further
increased to 0.4. A similar trend has also been found in the
experimental work of Ramos et al. [36]. This issue deserves
further studies in the future.

Influence of CO2 concentration in feed gas


Table 3 summarizes the influence of CO2 content in the feed
gas on the H2 and CO conversions, DME selectivity, and DME
yield. As a whole, an increase in CO2 concentration leads to an
obvious decline in CO conversion and DME yield, whereas its
impact on the H2 conversion and DME selectivity are not so
significant. In particular, when the CO2 concentration is as
high as 32%, there is a substantial drop in the CO and H2
conversions as well as the DME selectivity and yield. It is
known that H2, CO, and CO2 compete with each other to be
absorbed on the active sites of the catalyst surface. A higher
CO2 concentration disadvantages methanol formation and
thereby DME production. Meanwhile, the reverse WGSR will
be intensified in the course of DME synthesis when the CO2
concentration in syngas rises. The produced water will
accelerate the crystallization of Cu and ZnO contained in a Cu/
ZnO-based catalyst and lead to the deactivation of the catalyst
[27]. This is the reason why a CO2-rich gas is unfavorable to
DME synthesis, as observed.

Conclusions
The reaction characteristics of DME synthesis using a single
1
step process at a high space velocity (15,000 mL g1
cat h ) and
different reaction temperatures, H2/CO ratios, and physically
mixed catalysts have been investigated. The results indicate
that DME synthesis is governed by chemical kinetics when the
reaction temperature is lower than 220  C, whereas the synthesis is dominated by thermodynamics once the temperature is higher than 220  C. As a result, the optimal reaction

Table 3 e A list of CO and H2 conversions as well as DME selectivity and yield at various contents of CO2 in feed gas.
CO2 content (vol%)
0
3
5
10
16
32

H2/CO/CO2/N2 (vol%)

CO conversion (%)

H2 conversion (%)

DME selectivity (%)

DME yield (g/gcat h)

64/32/0/4
63/30/3/4
61/30/5/4
61/25/10/4
48/32/16/4
64/0/32/4

63
63
56
55
39
e

34
36
34
29
35
12

60
59
62
62
66
10

2.12
2.06
1.93
1.56
1.43
0.13

13592

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

temperature is located at 220  C. The H2 conversion and DME


yield at H2/CO 1 is higher than those at H2/CO 2. Within the
investigated ranges of reaction temperature and H2/CO ratio,
the maximum DME yield is 2.3 g (gcat h)1, revealing that the
high performance of the physically mixed catalyst (CuOeZnOeAl2O3 and HZSM-5) is achieved. In examining the acidity
of three dehydration catalysts, the acidity is ranked as:
ferrierite > HZSM-5 >> g-Al2O3. Water produced from the
reverse WGSR has a strongly inhibiting effect on the Lewis
acidity, which leads to a higher DME yield under HZSM-5 than
under g-Al2O3. Increasing CO2 concentration in syngas reduces the activity of the physically mixed catalyst so as to
lessen the CO conversion and DME yield. The addition of
palladium into the CuOeZnOeAl2O3 catalyst induces
hydrogen spillover from metallic Pd to metallic Cu. This reduces the oxidation of active Cu by CO2 and increases the
stability of active Cu. As a consequence, when the Pd content
in the CuOeZnOeAl2O3 catalyst is lifted, the CO conversion is
intensified. This increases the methanol and DME yield. The
findings in the present study have provided a comprehensive
insight into the catalyst preparation and operation for DME
production.

Acknowledgments
The authors acknowledge the financial support of the Ministry
of Science and Technology, Taiwan, ROC, in this research.

references

[1] Namasivayam AM, Korakianitis T, Crookes RJ, BobManuel KDH, Olsen J. Biodiesel, emulsified biodiesel and
dimethyl ether as pilot fuels for natural gas fuelled engines.
Appl Energy 2010;87:769e78.
[2] Xu S, Wang Y, Zhang Z, Zhen X, Tao C. Development of a
novel common-rail type Dimethyl ether (DME) injector. Appl
Energy 2012;94:1e12.
[3] Arcoumanis C, Bae C, Crookes R, Kimoshita E. The potential
of di-methyl ether (DME) as an alternative fuel for
compression-ignition engines: a review. Fuel
2008;87:1014e30.
[4] Vakili R, Pourazadi E, Setoodeh P, Eslamloueyan R,
Rahimpour MR. Direct dimethyl ether (DME) synthesis
through a thermally coupled heat exchanger reactor. Appl
Energy 2011;88:1211e23.
[5] Adachi Y, Komoto M, Watanabe I, Ohno Y, Fujimoto K.
Effective utilization of remote coal through dimethyl ether
synthesis. Fuel 2000;79:229e34.
[6] Marchionna M, Patrini R, Sanfilippo D, Migliavacca G.
Fundamental investigations on di-methyl ether (DME) as LPG
substitute or make-up for domestic uses. Fuel Process
Technol 2008;89:1255e61.
[7] Chen WH, Lin BJ, Lee HM, Huang MH. One-step synthesis of
dimethyl ether from the gas mixture containing CO2 with
high space velocity. Appl Energy 2012;98:92e101.
~ a J, Mier D, Arandes JM, Olazar M, Bilbao J.
[8] Aguayo AT, Eren
Kinetic modeling of dimethyl ether synthesis in a single step
on a CuO-ZnO-Al2O3/g-Al2O3 catalyst. Ind Eng Chem Res
2007;46:5522e30.

[9] Zhu Y, Wang S, Ge X, Liu Q, Luo Z, Cen K. Experimental study


of improved two step synthesis for DME production. Fuel
Process Technol 2010;91:424e9.
[10] Bae JW, Kang SH, Lee YJ, Jun KW. Effect of precipitants during
the preparation of Cu-ZnO-Al2O3/Zr-ferrierite catalyst on the
DME synthesis from syngas. J Ind Eng Chem 2009;15:566e72.
[11] Stelmachowski M, Nowicki L. Fuel from the synthesis
gasdthe role of process engineering. Appl Energy
2003;74:85e93.
[12] Pellegrini LA, Soave G, Gamba S, Lange S. Economic analysis
of a combined energyemethanol production plant. Appl
Energy 2011;88:4891e7.
[13] Wang D, Han Y, Tan Y, Tsubaki N. Effect of H2O on Cu-based
catalyst in one-step slurry phase dimethyl ether synthesis.
Fuel Process Technol 2009;90:446e51.
[14] Chen WH, Hsieh TC, Jiang TL. An experimental study on
carbon monoxide conversion and hydrogen generation from
water gas shift reaction. Energy Convers Manag
2008;49:2801e8.
[15] Chen WH, Chen CJ, Hung CI. A comparison of gasification
phenomena among raw biomass, torrefied biomass and coal
in an entrained-flow reactor. Appl Energy 2013;112:421e30.
[16] Chen WH, Chiu TW, Hung CI, Lin MR. Hysteresis and reaction
characterization of methane catalytic partial oxidation on
rhodium catalyst. J Power Sources 2009;194:467e77.
[17] Chen WH, Lin MR, Yu AB, Du SW, Leu TS. Hydrogen
production from steam reforming of coke oven gas and its
utility for indirect reduction of iron oxides in blast furnace.
Int J Hydrogen Energy 2012;37:11748e58.
[18] Sun K, Lu W, Qiu F, Liu S, Xu X. Direct synthesis of DME over
bifunctional catalyst: surface properties and catalytic
performance. Appl Catal A General 2003;252:243e9.
[19] Hadipour A, Sohrabi M. Synthesis of some bifunctional
catalysts and determination of kinetic parameters for direct
conversion of syngas to dimethyl ether. Chem Eng J
2008;137:294e301.
[20] Ge Q, Huang Y, Qiu F, Li S. Bifunctional catalysts for
conversion of synthesis gas to dimethyl ether. Appl Catal A
General 1998;197:23e30.
[21] Zhang Y, Sun Q, Deng J, Wu D, Chen S. A high activity Cu/
ZnO/A12O3 catalyst for methanol synthesis: preparation
and catalytic properties. Appl Catal A General
1997;158:105e20.
[22] Yang R, Yu X, Zhang Y, Li W, Tsubaki N. A new method of
low-temperature methanol synthesis on Cu/ZnO/Al2O3
catalysts from CO/CO2/H2. Fuel 2008;87:443e50.
[23] Baltes C, Vukojevic S, Schuth F. Correlations between
synthesis, precursor, and catalyst structure and activity of a
large set of CuO/ZnO/Al2O3 catalysts for methanol synthesis.
J Catal 2008;258:334e44.
[24] Li Q, Meng ZL, Du SQ, Liao ZF, Zhao YX. Prospect of
innovatory process for methanol syngas preparation and
methanol isopiestic synthesis. Sino Global Energy
2008;13:31e7.
[25] Qi GX, Fei JH, Zheng XM, Hou ZY. DME synthesis from carbon
dioxide and hydrogen over Cu-Mn/HZSM-5 catalyst. Catal
Lett 2001;72:121e4.
[26] Wu J, Saito M, Takeuchi M, Watanabe T. The stability of Cu/
ZnO-based catalysts in methanol synthesis from a CO2-rich
feed and from a CO-rich feed. Appl Catal A General
2001;218:235e40.
[27] Chen WH. A simplified model of predicting coal reaction in a
partial oxidation environment. Int Commun Heat Mass
Transf 2007;34:623e9.
[28] Xing LH, Gao YZ, Wang ZB, Du CY, Yin GP. Effect of anode
diffusion layer fabricated with mesoporous carbon on the
performance of direct dimethyl ether fuel cell. Int J Hydrogen
Energy 2011;36:11102e7.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 0 ( 2 0 1 5 ) 1 3 5 8 3 e1 3 5 9 3

[29] Mao D, Xia J, Zhang B, Lu G. The direct synthesis of dimethyl


ether from syngas over hybrid catalysts with sulfatemodified g-alumina as methanol dehydration components.
Energy Convers Manag 2010;51:1134e9.
[30] Burton AW, Ong K, Rea T, Chan IY. On the estimation of
average crystallite size of zeolites from the Scherrer
equation: a critical evaluation of its application to zeolites
with one-dimensional pore systems. Microporous
Mesoporous Mater 2009;117:75e90.
[31] Chen WH, Cheng YC, Hung CI. Enhancement of heat
recirculation on the hysteresis effect of catalytic partial
oxidation of methane. Int J Hydrogen Energy
2013;2013(38):10394e406.
[32] Sun K, Lu W, Wang M, Xu X. Low-temperature synthesis of
DME from CO2/H2 over Pd-modified CuO-ZnO-Al2O3-ZrO2/
HZSM-5 catalysts. Catal Commun 2004;5:367e70.

13593

[33] Chinchen GC, Waugh KC, Whan DA. The activity and state of
the copper surface in methanol synthesis catalysts. Appl
Catal 1986;25:101e7.
[34] Melian-Cabrera I, Lopez Granados M, Fierro JLG. Effect of Pd
on CueZn catalysts for the hydrogenation of CO2 to
methanol: stabilization of Cu metal against CO2 oxidation.
Catal Lett 2002;79:165e70.
[35] Xu M, Lunsford JH, Goodman DW, Bhattacharyya A.
Synthesis of dimethyl ether (DME) from methanol over solidacid catalysts. Appl Catal A General 1997;149:289e301.
[36] Ramos FS, Duarte de Farias AM, Borges LEP, Monteiro JL,
Fraga MA, Sousa-Aguiar EF, et al. Role of dehydration catalyst
acid properties on one-step DME synthesis over physical
mixtures. Catal Today 2005;101:39e44.

Anda mungkin juga menyukai