You are on page 1of 6

Microporous and Mesoporous Materials 164 (2012) 3–8

Contents lists available at SciVerse ScienceDirect

Microporous and Mesoporous Materials


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

One-pot synthesis of mesoporous Cu–c-Al2O3 as bifunctional catalyst for direct


dimethyl ether synthesis
Heqing Jiang, Hans Bongard, Wolfgang Schmidt, Ferdi Schüth ⇑
Max-Planck-Institut für Kohlenforschung, Kaiser-Wilhelm-Platz 1, 45470 Mülheim an der Ruhr, Germany

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

Article history: Mesoporous copper–alumina (Cu–Al2O3) with different copper contents was synthesized in a one-pot
Available online 23 August 2012 reaction via the evaporation-induced self-assembly of Pluronic P123 and the corresponding metal precur-
sors in ethanolic solution in the presence of nitric acid. Mesoporous Cu–Al2O3 calcined at 400 °C exhibits
Dedicated to Prof. Jens Weitkamp on the
occasion of his 70th birthday a large BET surface area of 265 m2/g and a pore volume of 0.48 cm3/g. XRD results indicate that the wall of
mesoporous Cu–Al2O3 calcined at 400 °C is amorphous, and that it is transformed to crystalline material
by further thermal treatment at 800 °C. Copper was formed as very small particles in the composite under
Keywords:
Evaporation-induced self-assembly 5% H2 flow at high temperature. Moreover, the mesoporous structure did not collapse after the sample
Mesoporous materials was reduced at 650 °C for 4 h, and the copper particles with sizes of around 6 nm were well distributed
Synthesis gas through the entire mesoporous c-Al2O3 network. Using the mesoporous Cu/c-Al2O3 as a bifunctional cat-
Cu/c-Al2O3 alyst for one-step dimethyl ether synthesis from synthesis gas, a CO conversion of 72% and a DME selec-
Dimethyl ether synthesis tivity of 69% were obtained at 50 bar and 310 °C.
Ó 2012 Elsevier Inc. All rights reserved.

1. Introduction Compared to methanol dehydration (2), methanol synthesis (1)


is more equilibrium-controlled. The equilibrium limitation of
In addition to being the building block for several industrially methanol synthesis (1) can be overcome by coupling it with the
important chemicals [1,2], dimethyl ether (DME) is exploited as a methanol dehydration step (2). In the single step process, the
sulfur-free alternative diesel fuel due to its low NOx emissions formed methanol can be quickly consumed by the consecutive
and almost no particulate formation during combustion [3–5]. methanol dehydration (2) if two types of active site are very close
Moreover, it may also be used as a source of hydrogen for fuel cell to each other in the bifunctional catalyst system. To favor the cat-
applications [6,7]. At present, DME is mainly produced from syn- alytic reaction, it is crucial to obtain a very fine distribution of two
thesis gas by two possible approaches. One is a two-step method types of active site in the composite catalyst. Previous studies
including methanol synthesis from synthesis gas (1) over copper- mostly focus on the physical mixture of the two types of catalysts,
based catalysts [8–10] and the subsequent methanol dehydration improving the activity and lifetime [16,17]. Very recently, Yang
to DME (2) on a solid acid catalyst, such as gamma alumina (c- et al. reported a well-designed H–ZSM-5/Cu–ZnO–Al2O3 capsule
Al2O3) [11,12] or different zeolites [13–15]. The other is the direct catalyst with core–shell structure, which exhibited a higher selec-
synthesis of DME from synthesis gas in one step in the presence of tivity for DME compared with the conventional hybrid catalyst
a bifunctional catalyst or physically mixed catalysts. Water formed [18]. In this work, we explore a bifunctional catalyst with mesopor-
by methanol dehydration (2) will partly be consumed by the water ous structure for one-step DME synthesis.
gas shift reaction (3). Thus, the overall reaction is described by Eq. c-Al2O3 has been widely used as methanol dehydration catalyst
(4). due to its moderate acidity, low price, and easy availability [11,12].
Recently, Yuan et al. reported a facile route to synthesize ordered
2CO þ 4H2
2CH3 OH ð1Þ
mesoporous c-Al2O3 by self-assembly of Pluronic P123 triblock
2CH3 OH
CH3 OCH3 þ H2 O ð2Þ copolymer and alumina precursors in ethanolic solution [19]. Later,
this method was further extended to synthesize mesoporous alu-
CO þ H2 O
H2 þ CO2 ð3Þ mina-supported noble metals or metal oxides such as CeO2, NiO,
TiO2, Cr2O3, etc. [20–22]. Compared to the conventional wet
3CO þ 3H2
CH3 OCH3 þ CO2 ð4Þ impregnation method, the mesoporous alumina supported metals
by one-pot method were reported to have high-quality meso-
structures that exhibit strong metal-support interactions and a
⇑ Corresponding author. Tel.: +49 208 306 2373; fax: +49 208 306 2995.
homogeneous distribution of active sites [23]. A bifunctional cata-
E-mail address: schueth@mpi-muelheim.mpg.de (F. Schüth).

1387-1811/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.micromeso.2012.08.004
4 H. Jiang et al. / Microporous and Mesoporous Materials 164 (2012) 3–8

lyst for one-step DME synthesis can be obtained by loading copper 2.3. Catalytic synthesis of DME
on the surface of c-Al2O3 using the conventional impregnation
method or the co-precipitation and deposition method reported Catalytic activity of mesoporous Cu/c-Al2O3 was evaluated in a
by Baek et al. [24]. In order to get a high dispersion of copper, fixed bed stainless-steel reactor with an inner diameter of 6 mm. A
we therefore extended the one-pot method described above for mixture of CO (6 ml/min), H2 (12 ml/min) and N2 (2 ml/min) was
the synthesis of mesoporous c-Al2O3 supported copper (Cu/c- sent to the top section of the reactor that acted as a pre-heater.
Al2O3), and explored its application as bifunctional catalyst for All gas flows were controlled by mass flow controllers (Brooks,
one-step DME synthesis from synthesis gas after reduction of the Model 5850), which were calibrated by using a soap bubble meter
copper species to result in a supported catalyst Cu/ c-Al2O3. before measurement. 1.0 g of catalyst (grain size 250–500 lm) was
loaded in the middle section of the reactor. The catalyst was re-
duced at 250 °C for 4 h in flowing 5% H2 in N2 prior to the catalytic
measurement. The catalytic reaction was conducted at a pressure
2. Experimental
of 50 bar and temperatures from 280 to 325 °C. Before the first data
point was taken, the catalyst was equilibrated under reaction con-
2.1. Synthesis of mesoporous Cu–c-Al2O3
ditions for 15 h, after this time typically a stationary state was
reached. Then the temperature was increased stepwise to assess
The mesoporous Cu–c-Al2O3 was prepared based on the evapo-
the performance of the catalysts at different temperature. Repro-
ration-induced self-assembly procedure using Pluronic P123 as a
ducibility of the performance for different catalysts and experi-
structure directing agent, following the previous reports [19,21].
ments was found to be highly satisfactory, with differences in
In a typical synthesis, 2.0 g of Pluronic P123 (EO20PO70EO20 from
conversions amounting to less than 3%. The effluent gas, after
BASF, Co.) was dissolved in 40 mL of ethanol, and the mixture
reducing its pressure by a back pressure regulator, was directly
was stirred at 30 °C for 4 h. Then 3.2 mL of 65 wt.% nitric acid
sent to an on-line gas chromatograph (Agilent 7890). All gas lines
(J.T. Baker) and the appropriate quantity of copper(II) nitrate trihy-
to the reactor and the gas chromatograph were heated to 160 °C
drate (99–104%, Sigma–Aldrich) and aluminum isopropoxide (98%,
to avoid the possible condensation of water and methanol. The
Aldrich) were added into the above solution with vigorous stirring.
gas chromatograph was equipped with two parallel channels to
The total amount of metal species was kept constant (20 mmol),
two detectors: a thermal conductivity detector (TCD) and a flame
and the molar ratio of copper to aluminum was adjusted accord-
ionization detector (FID). The TCD channel was used with packed
ingly. The beaker (volume: 100 mL) containing the mixture was
columns (HayeSep Q and molecular sieve 5A) to separate and ana-
covered with PE film and stirred at 30 °C for about 12 h. Then the
lyze CO, N2 and CO2, whereas the FID channel with a capillary col-
final solution was put into a drying oven and underwent the sol-
umn (HP-PLOT Q) was employed to analyze DME, methanol, and
vent evaporation at 60 °C for two days. Calcination was carried
possible hydrocarbons. The conversion of CO (XCO) is defined as
out by slowly increasing temperature from room temperature to
follows:
400 °C (1 °C/min ramping rate) and holding at 400 °C for 4 h under
air. The Cu–c-Al2O3 was obtained by further treating the above Fðtotal; inÞ  cðCO; inÞ  Fðtotal; outÞ  cðCO; outÞ
X co ¼  100%
sample at 800 °C for 4 h. The final samples were labeled in the gen- Fðtotal; inÞ  cðCO; inÞ
eral form xCu–Al-T, starting with copper molar percentage (x), Cu
indicating copper, Al indicating aluminum, followed by calcination The selectivities for DME (SDME) and CO2 (SCO2) are calculated
temperature (T). For example, 5Cu–Al-800 refers to the mesopor- based on the number of carbon atoms in the feed and product
ous aluminum oxides with 5% molar fraction of copper, which streams, as indicated by the following equations:
was calcined at 800 °C for 4 h in air.
2  Fðtotal; outÞ  cðDME; outÞ
SDME ¼  100%
Fðtotal; inÞ  cðCO; inÞ  Fðtotal; outÞ  cðCO; outÞ

2.2. Characterization
Fðtotal; outÞ  cðCO2 ; outÞ
SCO2 ¼  100%
Powder X-ray diffraction (XRD) measurements were performed Fðtotal; inÞ  cðCO; inÞ  Fðtotal; outÞ  cðCO; outÞ
on a Stoe STADI P diffractometer operating in reflection mode with
Cu Ka radiation at room temperature from 0.7° to 5° (small angle) where F (total, in) is the total flow rate of the mixing gases at the
and 20° to 80° (wide angle). Scanning transmission electron micro- inlet of the reactor, and F (total, out) represents the total flow rate
scope (STEM) images were taken on a Hitachi S-5500 microscope of the effluent gases, which was obtained by using N2 as an internal
with an EDS detector operating at an acceleration voltage of standard. c (i, out) is the concentration of the corresponding gas at
30 kV. The samples were prepared on carbon-coated nickel grid. the exit of the reactor, which was determined by the calibrated gas
Nitrogen adsorption–desorption isotherms were obtained on a chromatograph.
NOVA 4200e instrument at 196 °C. Prior to the measurements,
all samples were degassed under vacuum for around 12 h at 3. Results and discussion
300 °C. The BET surface area was calculated from the data in a rel-
ative pressure range from 0.04 to 0.2. By using the Barrett–Joyner– Fig. 1 presents the small angle XRD patterns of samples with
Halenda (BJH) model, the pore volumes and pore size distributions different copper contents prepared by the one-pot method. The
were derived from the adsorption branches of isotherms. samples 2Cu–Al-400 and 5Cu–Al-400 exhibit a very strong diffrac-
Temperature-programmed reduction (TPR) was performed in a tion peak at 1.0° and one weak peak at around 1.7°, indicating the
homemade apparatus loaded with 100 mg of sample, using a gas presence of uniform mesoporosity, which could either be ordered
mixture of 5% H2 in nitrogen at a flow of 30 ml/min. The tempera- or worm-like. With further increasing copper content up to 15%,
ture was raised from room temperature to 650 °C with a heating the small angle reflection becomes less defined, until at 15% it is
rate of 10 °C /min. The water produced during the reduction was not present any more. This indicates an increasing effect of the
trapped on a 5A molecular sieve. The amount of hydrogen con- introduction of copper into the alumina matrix on the formation
sumed by the reduction was measured by a thermal conductivity of the mesoporous structure. At a copper content of 15%, the order
detector (TCD). of the pore system is obviously lost. The wide angle XRD pattern of
H. Jiang et al. / Microporous and Mesoporous Materials 164 (2012) 3–8 5

Fig. 1. Small angle XRD patterns of mesoporous Cu–Al2O3 with different molar
fractions of copper and aluminum calcined at different temperatures.

the sample 5Cu–Al-400 (Fig. 2) indicates that the wall of the mes-
oporous Cu–Al2O3 is amorphous, which normally results in poor
hydrothermal stability. After further treating at 800 °C for 4 h,
the amorphous Cu–Al2O3 was transformed into crystalline Cu–c-
Al2O3. As shown in Fig. 2, the XRD pattern of 5Cu–Al-800 displays
three reflections at 37.5°, 45.8° and 66.9°, indicating the formation
of crystalline Cu–c-Al2O3 [19,25,26]. The absence of isolated cop-
per oxide or other copper-related compounds indicates that copper
was incorporated into the alumina matrix, thus leading to a fine
distribution of copper on the atomic level. It is noted that the good
mesoscopic order is still maintained after the treatment at 800 °C.
As shown in Fig. 1c, the peak at 1.0° is still present in the small an-
gle XRD pattern of 5Cu–Al-800, indicating the good thermal stabil-
ity of the mesoporous Cu–Al2O3.
The formation of mesoporous Cu–Al2O3 is also supported by Fig. 3. Nitrogen adsorption isotherms (top) and the corresponding pore size
distribution (bottom) for Cu–Al2O3 calcined at different temperatures.
nitrogen adsorption measurements. Fig. 3 shows the nitrogen
adsorption–desorption isotherms and the corresponding pore size
distribution for the samples treated at 400 and 800 °C. Both iso-
and the sample 5Cu–Al-800 has a BET surface area of 198 m2/g
therms are type IV with H1 hysteresis loops, indicating the forma-
and a pore volume of 0.38 cm3/g. The corresponding pore size dis-
tion of uniform mesopores. The sample 5Cu–Al-400 exhibits a large
tribution curves in Fig. 3 reveal a narrow pore size distribution cen-
BET surface area of 265 m2/g and a pore volume of 0.48 cm3/g. To
tered at around 8 nm for 5Cu–Al-400 and 6 nm for 5Cu–Al-800.
get the catalytically active component c-Al2O3 for methanol dehy-
The reduction of pore size is probably due to the shrinkage of the
dration, the treatment at higher temperature is necessary accord-
structure after the thermal treatment.
ing to the XRD studies. It can be seen from Fig. 3, that after
In the mesoporous Cu–Al2O3 catalyst system, copper after its
transforming the catalysts to crystalline Cu–c-Al2O3 by further
reduction is expected to work as active component for methanol
treatment at 800 °C, the mesoporous structure does not collapse,
synthesis and the crystalline c-Al2O3 matrix as an acid component
for methanol dehydration. According to previous TPR studies, the
reduction temperature of copper oxide under hydrogen is around
250 °C [27–29]. In this work, in order to obtain a homogeneous dis-
tribution, copper first was incorporated into the alumina matrix,
forming Cu–Al–O crystalline network that is normally reduced at
320–390 °C [29–31]. Fig. 4 presents the reduction profile of the
sample 5Cu–Al-800. The onset of reduction is around 170 °C and
a broad reduction peak centered at around 330 °C was observed
in the TPR patterns. The higher reduction temperature of the crys-
talline Cu–Al2O3 indicates that copper had really been incorporated
in the crystalline alumina network, which is in accordance with the
XRD results. Fig. 5 shows the XRD patterns of 5Cu–Al-800 before
and after reduction under 5% H2 flow at different temperatures
for 4 h. There is no obvious change in the XRD patterns of the sam-
ples reduced at 280 or 500 °C, and no reflections characteristic of
metallic copper particles were detected.
STEM was used to observe the morphology change of the sam-
ples before and after reduction. It was found that both ordered and
Fig. 2. Wide angle XRD patterns of Cu–Al2O3 calcined at different temperatures. worm-like mesoporous regions are present in the Cu–Al2O3 pre-
6 H. Jiang et al. / Microporous and Mesoporous Materials 164 (2012) 3–8

Fig. 4. TPR profile of the sample 5Cu–Al-800.

Fig. 6. STEM bright-field (a) and dark-field (b and c) images of the sample 5Cu–Al-
800 after reduction at 280 °C for 4 h. EDS measurements from several selected areas
Fig. 5. XRD patterns of 5Cu–Al-800 before (a) and after reduction under 5% H2 flow (see image c) indicate the homogeneous distribution of copper through the
at different temperatures for 4 h: (b) 280 °C, (c) 500 °C, and (d) 650 °C. mesoporous matrix.

This makes such solids promising bifunctional catalysts for the


pared by the one-pot method, which is a result of the synthesis one-step DME synthesis. Fig. 8 presents the CO conversion and
conditions applied. The synthesis of especially exclusively the or- the selectivities of DME and CO2 at different operating tempera-
dered phase is difficult for alumina, and even more so in the pres- tures. At 295 °C, a CO conversion of 63% was obtained, which is
ence of copper species. However, for the catalytic performance, the lower than the equilibrium conversion (78% for the overall reaction
presence of a mix of ordered and worm-like mesopores is not ex- (Eq. (4)) at 295 °C) [32]. With increasing temperature to 310 °C, CO
pected to be crucial. Fig. 6 shows the STEM images of the sample conversion almost reaches the equilibrium conversion, which is
5Cu–Al-800 after reduction at 280 °C for 4 h. The mesoporous caused by the enhanced rate of methanol synthesis and increasing
structure is still maintained after the reduction, and the pore size activity of the c-Al2O3 functionality at higher temperature. Also
is around 6 nm. EDX measurements of several selected areas CO2 was found in the effluent, and its selectivity is around 25%.
(Fig. 6c) indicate a homogeneous distribution of copper through As can be seen from Eq. (2), in addition to DME, water will form
the whole mesoporous matrix. However, metallic copper particles by methanol dehydration. The formed water will react with CO,
were not detected after the reduction at these temperatures, which leading to the formation of CO2. The selectivity of DME is around
could be attributed to their small size. When the sample was fur- 73% when the operating temperature is below 300 °C, indicating
ther treated under hydrogen flow at higher temperature, the small the almost all the CO was converted into DME and CO2. Further
particles grew and became detectable. From the STEM images increasing the temperature to 325 °C led to the slight decrease of
(Fig. 7) of the sample 5Cu–Al-800 after reduction at 650 °C for both CO conversion and DME selectivity. At higher temperature,
4 h, small copper particles (light dots) with the size of around more hydrocarbons were detected. Compared to the published
6 nm were detected. They are well distributed over the entire alu- study on a related catalyst system [24], in which a selectivity to
mina network in the ordered (Fig. 7a and c) or worm-like (Fig. 7b methanol exceeding 30% in the best case was observed, no metha-
and d) mesoporous structures. The presence of the metallic copper nol was detected here at all investigated temperatures. Methanol
particles was also confirmed by the corresponding wide angle XRD selectivity is most probably governed by the balance between
pattern in samples reduced at high temperatures (Fig. 5d). CO-hydrogenation functionality and acid functionality, which need
The results discussed above show that copper particles are to be optimized to achieve the desired performance. In the de-
formed as small particles distributed through the whole crystalline signed mesoporous Cu/c-Al2O3 bifunctional catalyst, all the copper
c-Al2O3 network with large surface area and high thermal stability. sites are well surrounded by the acid sites. Methanol is rapidly con-
H. Jiang et al. / Microporous and Mesoporous Materials 164 (2012) 3–8 7

Fig. 7. STEM dark-field images of the sample 5Cu–Al-800 reduced at 650 °C for 4 h. (a) and (c): ordered structure, (b) and (d): worm-like mesoporous structure.

surface and good thermal stability did not collapse after the reduc-
tion, and the resulting copper nanoparticles are well distributed
through the entire mesoporous c-Al2O3 network, which makes
the material a promising bifunctional catalyst for one-step DME
synthesis. Preliminary studies show that a CO conversion of 72%
and DME selectivity of 69% can be obtained at 50 bar and 310 °C.
The synthetic route described here could provide the blueprint
also for the one-step synthesis of other metal and metal oxide spe-
cies supported on mesoporous c-Al2O3. This is interesting, since c-
Al2O3 is one of the most important commercial catalyst support
materials. Moreover, an additional conclusion can be drawn with
respect to the required catalytic functionality of the catalysts.
While it is obvious, that both CO hydrogenation activity and acidity
for the coupling step are required, the presence of ZnO, which is
normally assumed to be necessary for methanol synthesis activity,
is not mandatory. The catalyst described by us is highly active
without ZnO. This could open additional options for the prepara-
Fig. 8. CO conversion and selectivity of DME and CO2 over 5Cu–Al-800 at different tion of novel catalysts for this reaction.
temperatures. FH2 = 12 mL/min, FCO = 6 mL/min, and FN2 = 2 mL/min. The catalyst
was reduced at 250 °C for 4 h in flowing 5% H2 in N2 prior to the catalytic
measurement. Acknowledgments

sumed by methanol dehydration once it forms, leading to the ab- The authors acknowledge the financial support by BMBF
sence of methanol in the final products. through DMEEXCO2 project (Grant agreement No. 01RC1108)
and of the Max Planck Society and Fraunhofer-Gesellschaft through
the joint project ‘‘Heterogeneous catalysis – the production of base
4. Conclusions chemicals and fuels from renewable resources’’.

In summary, mesoporous Cu–c-Al2O3 with crystalline walls was


References
successfully synthesized by a one-pot method via the cooperative
self-assembly of Pluronic P123 and the corresponding metal pre- [1] D.M. Brown, B.L. Bhatt, T.H. Hsiung, Catal. Today 8 (1991) 279.
cursors in ethanolic solution under acidic conditions. Copper was [2] Q. Ge, Y. Huang, F. Qiu, S. Li, Appl. Catal. A 167 (1998) 23.
formed as very small particles from the crystalline Cu–c-Al2O3 net- [3] V. Vishwanathan, K.W. Jun, J.W. Kim, H.S. Roh, Appl. Catal. A 276 (2004) 251.
[4] J. Xia, D. Mao, B. Zhang, Q. Chen, Y. Zhang, Y. Tang, Catal. Commun. 7 (2006)
work after its reduction under 5% H2 at high temperature. More- 362.
over, it was found that the mesoporous structure with large [5] G.R. Moradi, S. Nosrati, F. Yaripor, Catal. Commun. 8 (2007) 598.
8 H. Jiang et al. / Microporous and Mesoporous Materials 164 (2012) 3–8

[6] K. Faungnawakij, K. Eguchi, Catal. Surv. Asia 15 (2011) 12. [20] Q. Yuan, H.H. Duan, L.L. Li, Z.X. Li, W.T. Duan, L.S. Zhang, W.G. Song, C.H. Yan,
[7] C. Ledesma, J. Llorca, J. Phys.Chem. C 115 (2011) 11624. Adv. Mater. 22 (2010) 1475.
[8] X.M. Liu, G.Q. Lu, Z.F. Yan, J. Beltramini, Ind. Eng. Chem. Res. 42 (2003) 6518. [21] S.M. Morris, P.F. Fulvio, M. Jaroniec, J. Am. Chem. Soc. 130 (2008) 15210.
[9] C. Baltes, S. Vukojeviu, F. Schuth, J. Catal. 258 (2008) 334. [22] W.Q. Cai, J.G. Yu, C. Anand, A. Vinu, M. Jaroniec, Chem. Mater. 23 (2011) 1147.
[10] S. Vukojevic, O. Trapp, J.D. Grunwaldt, C. Kienre, F. Schuth, Angew. Chem. Int. [23] L.B. Sun, W.H. Tian, X.Q. Liu, J. Phys. Chem. C 113 (2009) 19172.
Ed. 44 (2005) 7978. [24] S.C. Baek, S.H. Kang, J.W. Bae, Y.J. Lee, D.H. Lee, K.Y. Lee, Energy Fuels 25 (2011)
[11] J.W. Bae, H.S. Potdar, S.H. Kang, K.W. Jun, Energy Fuels 22 (2008) 223. 2438.
[12] S.H. Kang, J.W. Bae, H.S. Kim, G.M. Dhar, K.W. Jun, Energy Fuels 24 (2010) 804. [25] Z.R. Zhang, T.J. Pinnavaia, J. Am. Chem. Soc. 124 (2002) 12294.
[13] S.R. Blaszkowski, R.A. van Santen, J. Phys. Chem. B 101 (1997) 2292. [26] S.X. Zhou, M. Antonietti, M. Niederberger, Small 3 (2007) 763.
[14] S.H. Kang, J.W. Bae, K.W. Jun, H.S. Potdar, Catal. Commun. 9 (2008) 2035. [27] T. Kawabata, H. Matsuoka, T. Shishido, D.L. Li, Y. Tian, T. Sano, K. Takehira,
[15] Y. Fu, T. Hong, J. Chen, A. Auroux, J. Shen, Thermochim. Acta 434 (2005) 22. Appl. Catal. A 308 (2006) 82.
[16] D.S. Mao, W.M. Yang, J.C. Xia, B. Zhang, Q.Y. Song, Q.L. Chen, J. Catal. 230 (2005) [28] M. Behrens, I. Kasatkin, S. Kuhl, G. Weinberg, Chem. Mater. 22 (2010) 386.
140. [29] Y. Xing, Z.X. Liu, S.L. Suib, Chem. Mater. 19 (2007) 4820.
[17] K.S. Yoo, J.H. Kim, M.J. Park, S.J. Kim, O.S. Joo, K.D. Jung, Appl. Catal. A 330 [30] Y. Xing, Z.X. Liu, S. Gomez, S.L. Suib, J. Phys. Chem. C 112 (2008) 1446.
(2007) 57. [31] P. Ammendola, R. Chirone, L. Lisi, G. Ruoppolo, G. Russo, J. Mol. Catal. A 266
[18] G.H. Yang, N. Tsubaki, J. Shamoto, Y. Yoneyama, Y. Zhang, J. Am. Chem. Soc. 132 (2007) 31.
(2010) 8129. [32] T. Ogawa, N. Inoue, T. Shikada, Y. Ohno, J. Nat. Gas Chem. 12 (2003) 219.
[19] Q. Yuan, A.X. Yin, C. Luo, L.D. Sun, Y.W. Zhang, W.T. Duan, H.C. Liu, C.H. Yan, J.
Am. Chem. Soc. 130 (2008) 3465.

You might also like