You are on page 1of 8

Journal of Natural Gas Chemistry 21(2012)352359

Preparation of mesoporous alumina with large pore size and


their supported rhenium oxide catalysts in metathesis
of 1-butene and 2-butene to propene
Lei Sang, Sheng-Li Chen , Guimei Yuan, Zheng Zhou, Rui Li, Aicheng Chen, Min Zheng, Ju You
State Key Laboratory of Heavy Oil Processing, College of Chemical Engineering, China University of Petroleum, Beijing 102249, China
[ Manuscript received September 13, 2011; revised November 16, 2011 ]

Abstract
Mesoporous -aluminas with large pore size (up to 19 nm, denoted as MA19) are prepared from dispersed pseudo-boehmite using pluronic
P123 as template. It is found that these mesoporous alumina supported rhenium oxide catalysts were more active and have far longer working
life-span in gas-phase metathesis of 1-butene and 2-butene to propene than rhenium oxide on conventional alumina with small pore size
(5 nm). At 60 C and atmospheric pressure with WHSV = 1 h1 , the similar stable conversions of butene (ca. 55%) for all the 13 wt%
Re2 O7 /alumina catalysts were obtained near the chemical equilibrium, and the stable working life-spans of Re2 O7 /MA19 were far longer than
that of Re2 O7 /Al2 O3 , being about 70 h and 20 h, respectively.
Key words
mesoporous alumina; rhenium oxide; metathesis; butene; propene

1. Introduction

Alumina is one of the most widely used materials employed as catalytic support, catalyst, adsorbent and thermally
conductive materials [1,2]. Among various aluminas, mesoporous alumina attracts a lot of attention because of its excellent features such as high surface area, large void volume
and narrow pore size distribution, therefore it represents a
new advanced material with potential applications in catalytic
support and other fields [3,4]. Mesoporous alumina with tailored properties is usually synthesized by sol-gel method using anionic, cationic or nonionic surfactants as well as hyperbranched polymers as template [5,6]. The widely used templates for the formation of mesostructures are low molecular weight surfactants, and generally, the resulted mesoporous
materials have small pore sizes. The main disadvantage of
mesoporous aluminas with large pore size is that they are usually in broad pore size distribution or have small surface area.
The pore size of mesoporous alumina is mostly less than 8 nm
[7], so synthesis of mesoporous alumina with large pore size
and narrow pore size distribution is a challenge, and only a
few reports can be found [811]. The mesoporous alumi-

nas with large pore size (825 nm) have been prepared by
hydrolyzing aluminum choride in the presence of pluronic
P123 in ethanol [8,9], mixing two oppositely charged alumina
sources in the presence of pluronic F127 in aqueous media
[10] and hydrolyzing aluminum isopropoxide in the presence
of pluronic F127 [11], respectively.
Olefin metathesis is one reaction where alumina is frequently used as catalytic support. From the viewpoint of reaction mechanism and its practical application in which lessdesired olefinic products are converted into more valuable
ones, olefin metathesis is an intriguing reaction. For example, surplus butene is converted into propene to meet the increasing demand for propene. Since olefin metathesis is discovered, a variety of homogeneous and heterogeneous catalyst systems for metathesis have been developed [12]. Among
these catalyst systems, Re2 O7 catalyst supported on alumina
attracts special attention because of its high activity and selectivity under mild conditions [13]. The relatively high price
of rhenium and fast deactivation of Re2 O7 catalyst make it
important to improve the performance of this catalyst. In recent years, it has been found that rhenium oxide supported on
mesoporous alumina exhibits higher activity and selectivity in
the metathesis of higher olefins in batch reactor at room

Corresponding author. Tel: +86-10-89733396; Fax: +86-10-69724721; E-mail: slchen@cup.edu.cn


The work was financially supported by SINOPEC Jiujiang Petrochemical Company and from the National Nature Science Foundation of China (No.
20976192).
Copyright2012, Dalian Institute of Chemical Physics, Chinese Academy of Sciences. All rights reserved.
doi:10.1016/S1003-9953(11)60376-9

Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

temperature than rhenium oxide on conventional alumina


[1416]. However, the investigation of metathesis of butene
on rhenium oxide supported on large mesoporous alumina has
not been reported. In this contribution, we reported a simple
method for preparation of two types of mesoporous alumina
with large pore size (denoted as MA9 and MA19), and the
performance of Re2 O7 catalyst supported on these aluminas
in gas-phase metathesis of 1-butene and 2-butene under mild
condition was also investigated. It was found that catalyst
pore size has great effect on the catalyst life-span. Moreover,
rhenium oxide loading and catalyst regeneration were investigated, too.
2. Experimental
2.1. Materials
Pluronic P123 (Mav = 5800, EO20 PO70 EO20 ) was purchased from Sigma-Aldrich (USA). Pseudo-boehmite powder and perrhenic acid were purchased from Shanghai Xinnian Petroauxiliary Co., Ltd. (China) and Hunan Zhuzhou
Jinlai Industry Co., Ltd. (China), respectively. 1-butene
and 2-butene were provided by Qilu Petrochemical Co., Ltd.
(China) and Shanghai Secco Petrochemical Co., Ltd. (China),
respectively. All the chemicals were used as received. The
purity of the used 1-butene and 2-butene was 99.49 wt% and
79.99 wt%, respectively.
2.2. Preparation of alumina supports
The synthesis procedure of MA19 was adapted from reported paper by Liu et al. [17]. Firstly, 40.8 g pseudoboehmite powder (74.8 wt% Al2 O3 ) was added into 536 mL
deionized water and stirred at 80 C for 0.5 h. Then, 1 M nitric
acid was dropwise added into the mixture with [H+ ]/[Al3+ ]
molar ratio of 0.09, and the mixture was kept stirring at 80 C
for 6 h to obtain a boehmite sol of 1 mol/L. Subsequently,
34.8 g P123 was dissolved in the boehmite sol and stirred at
80 C for 10 h and then aged at room temperature for 3 h. The
resulting mixture was placed in a 110 C oven overnight to
form the as-synthesized sample and then calcined in a muffle
furnace, which was heated from room temperature to 550 C
at a rate of 1 C/min, at 550 C for 6 h in air to obtain MA19.
MA9 was prepared as follows: P123 was dissolved in the
boehmite sol prepared as above with [P123]/[Al3+] molar ratio of 0.01 and stirred at 80 C for 10 h. Then, 2 M ammonia
was dropwise added into the mixture under vigorous stirring
at room temperature until pH = 9 when a white precipitate was
formed, and then the mixture was aged at room temperature
for 3 h. The precipitate was separated by vacuum filtration,
washed several times with deionized water, and then dried
and calcined at the same way for MA19. The conventional
alumina (i.e. Al2 O3 ) as control support was also prepared by
the precipitation method as follows: 3 M ammonia was dropwise added into 1 M aluminium nitrate solution under stirring

353

at 60 C until pH = 7.7 when a white precipitate was formed,


and then the mixture was aged at room temperature for 3 h.
Next, the precipitate was successively filtrated, washed, dried
and calcinated at the same way for MA9.
2.3. Preparation of catalysts
The catalyst samples were prepared by incipient-wetness
impregnation of the supports MA19, MA9 and Al2 O3 with
aqueous solution of perrhenic acid, respectively. The Re2 O7
loadings were 6.5 wt%, 13 wt%, 16.7 wt% and 23 wt%, respectively. After 6 h impregnation at room temperature, these
catalysts were dried at 50 C for 5 h and then at 110 C for
2 h, subsequently calcinated at 550 C for 5 h under air atmosphere.
2.4. Characterization of supports and catalysts
Nitrogen physisorption measurements were performed
at 196 C on a Micromeritics ASAP 2020 instrument
(Micromeritics Instrument Corporation, USA). BrunauerEmmett-Teller (BET) calculation of surface area and BarrettJoyner-Halenda (BJH) calculation of pore volume and pore
size distribution were performed on the desorption branches
of the isotherms.
Powder X-ray diffraction of wide angle (10o80o ) and
small angle (0.65o6o ) was conducted using a Shimadzu XRD
6000 diffractometer with Cu K radiation (Shimadzu Corporation, Japan).
The surface morphology of the samples was observed by
scanning electron micrography (SEM, FEI Quanta200F). The
transmission electron micrographs (TEM) and selected area
electron diffraction (SAED) images of the samples were obtained on a Tecnai G2 F20 electron microscope (FEI Company, USA).
The acidity of the samples has been measured by the
temperature-programmed desorption of ammonia (NH3 -TPD)
apparatus using a quartz tube microreactor (i.d. 6 mm), with
nitrogen as the carrier gas. A 0.5 g catalyst sample was loaded
in a quartz tube microreactor supported by quartz wool and
pretreated at 500 C for 0.5 h in a N2 stream with a flow
rate of 40 mL/min. Next, the samples were cooled to 110 C
and exposed to flowing 5% NH3 -N2 (50 mL/min) for 0.5 h
and finally purged in flowing N2 for 2 h. NH3 -TPD experiment was carried out in the range of 110 C600 C using
10 C/min heating rate. The desorbed ammonia was detected
by a SP3420 gas chromatograph equipped with a thermocouple detector (TCD) and then was absorbed by HCl aqueous solution (0.02 M). The amounts of desorbed NH3 were obtained
by back titration using NaOH aqueous solution (0.01 M).
The amount of carbon in the catalyst after butene metathesis reaction was determined by a method of infrared absorption following combustion in a high-frequency induction furnace using HIR-944B type carbon and sulfur analyzer (Wuxi
high-speed analysis instrument Co., LTD, China).

354

Lei Sang et al./ Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

2.5. Investigation of butene metathesis


Metathesis of butene (with 1-butene/2-butene molar ratio of 1) was carried out in a 8350 mm continuous fixedbed downstream quartz tube microreactor at 60 C and atmospheric pressure with WHSV = 1 h1 (based on the total
feedstock). 2.0 g catalyst was placed into the middle of the
reactor with the space above and below the catalyst filled
with quartz sand. Then, the catalyst was heated at 500 C
for 1 h and then cooled to 60 C in nitrogen atmosphere.
molecBefore reaction, butene was flowed through a 4 A
ular sieve adsorber for purification. The reaction products
were analyzed by SP3420 gas chromatograph (Beijing Analytical Instrument Plant, China) with a PONA capillary column (50 m0.32 mm0.5 m) and a flame ionization detector. The conversion of butene and the reaction selectivity to
propene were calculated based on the total feedstock.
3. Results and discussion
The textural properties of the prepared alumina supports
and their 13 wt% Re2 O7 /alumina catalysts are shown in Table 1 and Figure 1. By comparing the data listed in Table 1, it can be seen that the pseudo-boehmite-based samples
(MA19 and MA9) prepared using P123 as template exhibited larger specific surface areas, average pore diameters and

pore volumes than Al2 O3 prepared without P123. Especially,


MA19 had the largest pore size, pore volume and surface area
(18.9 nm, 1.70 cm3 /g and 358 m2 /g, respectively). This result showed that using block copolymer as template can obtain alumina with relatively large mesopores. Interestingly,
the specific surface area, average pore diameter and pore volume of MA19 were far larger than those of MA9 although they
were prepared with the same [P123]/[Al3+] ratio, indicating
that the synthesis procedure greatly affects the textural properties of the resultant materials. While in the literature [17],
the mesoporous alumina was synthesized in the similar route
mentioned in Part 2.2 of this contribution when [P123]/[Al3+]
ratio was 0.01, the pore size, pore volume and surface area
were 10.4 nm, 1.00 cm3 /g and 306 m2 /g, respectively. Table 1
also shows that the textural properties of the catalysts were
fairly different from those of the original supports. It suggests
that rhenium incorporation and catalyst calcination lead to a
decrease at different extent in specific surface areas, average
pore diameters and pore volumes of the catalysts.
According to the classification of IUPAC(1985) [18],
the nitrogen adsorption-desorption isotherms exhibited typical
type IV curves with H2-shaped hysteresis loops in all the three
alumina supports (Figure 1a) and the corresponding 13 wt%
Re2 O7 /alumina catalysts (Figure 1b), indicating the mesopores materials characteristic for an aggregate of small particles. The hystersis loops of MA19, MA9 and Al2 O3 appeared
in the relative pressure (p/p0 ) range of 0.81.0, 0.60.9 and

Figure 1. Nitrogen adsorption-desorption isotherms and corresponding pore size distributions of different alumina supports (a) and Re2 O7 /alumina catalysts (b)

355

Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

0.40.9, respectively. It suggests that MA19 and MA9 have


larger mean pore diameters and narrower pore size distributions than Al2 O3 . The pore size distributions are also showed

in Figure 1. In addition, all the three alumina supports and the


corresponding catalysts were absence of micropores according to t-plot measurement.

Table 1. Properties of alumina supports and catalysts


Samples
Al2 O3
MA9
MA19
13 wt% Re2 O7 /Al2 O3
13 wt% Re2 O7 /MA9
13 wt% Re2 O7 /MA19

SBET (m2 /g)


225
234
358
126
214
208

Dp (nm)
5.2
9.2
18.9
3.4
8.2
16.2

Vp (cm3 /g)
0.29
0.54
1.70
0.11
0.44
0.85

Aa (mol/g)
228
259
275
318
326
340

Ad (mol/m2 )
1.01
1.11
0.77
2.52
1.53
1.63

C (wt%)

1.44
1.75
1.83

SBET : BET specific surface area; DP : Average pore diameter; Vp : Pore volume; Aa : Acid amount as determined using NH3 -TPD; Ad : Acid density; C:
Carbon content of catalyst after butene metathesis reaction. Reaction conditions: 1-butene/2-butene molar ratio of 1, WHSV = 1 h1 , 60 C, atmospheric
pressure

The wide-angle X-ray diffraction patterns of all the three


alumina supports and the corresponding Re2 O7 /alumina catalysts (Figure 2b) exhibited some broad crystalline peaks
at 66.8o, 46.1o, 37.3o and 39.3o, respectively, due to the
reflection of typical -alumina phase (JCPDS card 100425).
The larger mesoporous alumina supports (MA19 and MA9)
exhibited similar crystallinity with more intense and narrower diffraction peaks than that of reference support (Al2 O3 ).
The degree of crystallinity of all the Re2 O7 /alumina catalysts
was less than that of their corresponding supports. In addition, all the Re2 O7 /alumina catalysts were absence of discernible diffraction peaks of rhenium oxide, which are con-

sistent with the reference [19]. It can be inferred that rhenium oxide was uniformly deposited on the alumina support. On the other hand, it is believed that 18 wt% Re2 O7
loading represents the limit of the surface monolayer capacity (ca. 3 Re atom/nm2) for alumina [20]. Even for
23 wt% Re2 O7 /MA19, the calculation gave only 2.75 Re
atom/nm2, i.e., still lower than the capacity limit. At
small angles (Figure 2a), X-ray patterns of all the three
alumina supports did not exhibit any diffraction peaks under the experimental conditions, indicating the absence of
long-range order porous network. TEM images (Figure 3)
confirmed this result.

Figure 2. Low-angle XRD patterns and wide-angle XRD patterns of different alumina supports (a) and Re2 O7 /alumina catalysts (b)

In order to observe the morphology of Al2 O3 , MA9 and


MA19, SEM measurements were carried out and the SEM
images are shown in Figure 3(a), 3(b) and 3(c), respectively.
Obviously, Al2 O3 appeared to be random piled-up of corrugated platelets with various sizes. MA9 and MA19 displayed
the morphology of monolith composed of spherical aggregates, accordance with the result from nitrogen adsorptiondesorption measurements (Figure 1a). This result is greatly
different from that in the literature [17], where the mesoporous alumina displayed a lath-like morphology. Moreover, the size of alumina spherical particles of MA19 seemed
larger than that of MA9 and the pore space between alumina spherical aggregates of MA19 was apparently larger

than that of MA9.


TEM images (Figure 3d, 3e and 3f) of Al2 O3 , MA9 and
MA19 showed wormhole-like appearance and disordered pore
arrangement, which is in good agreement with the absence
of high order peaks in the small-angle XRD patterns. SAED
image from the same part of MA19 (Figure 3g) gave a concentric diffraction ring pattern with some less-resolved spots
superimposed upon the diffuse ring pattern, corresponding to
the strongest (400) and (440) diffraction of -Al2 O3 , as determined from the wide-angle XRD results (Figure 2). This
distinctive ring pattern demonstrates a polycrystalline property in nanometer scales. Similar SAED images (not shown)
could be obtained for MA9 and Al2 O3 .

356

Lei Sang et al./ Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

Figure 3. SEM images of (a) Al2 O3 , (b) MA9 and (c) MA19; TEM images of (d) Al2 O3 , (e) MA9 and (f) MA19; SAED pattern of (g) MA19

To investigate the properties of the large mesoporous


alumina materials (MA19 and MA9) as catalyst support, Re2 O7 /MA19, Re2 O7 /MA9 and the reference catalyst
Re2 O7 /Al2 O3 with 13 wt% Re2 O7 loading were prepared and
their catalytic performance were tested through metathesis of
1-butene and 2-butene (Figure 4). Figure 4(a) shows that the
stable conversions of butene (ca. 55%) for all the three catalysts were similar and near the chemical equilibrium, and the
stable working life-spans of Re2 O7 /MA19 and Re2 O7 /MA9
were far longer than that of Re2 O7 /Al2 O3 , being about 70 h,
50 h and 20 h for the three catalysts, respectively. The reason for this superiority has not been completely understood
yet. It may be that Re2 O7 /MA19 and Re2 O7 /MA9 have larger
surface areas, pore diameters and pore volumes than those of
the reference Re2 O7 /Al2 O3 . One of the reasons for deactivation of metathesis catalyst is the adsorption of oligomeric
or polymeric byproducts on the catalyst surface which blocks
the active sites or the whole pores. On the one hand, it needs
longer time for the larger pores to be blocked by oligomeric

or polymeric byproducts. In general, the maximal carbon


contents of the spent tungsten oxide catalysts and molybdenum oxide catalysts is comparatively high, such as 46 wt%
and 10 wt% in the literatures [21,22], respectively. While the
maximal carbon content of the spent Re2 O7 /Al2 O3 catalysts
is relatively low, such as 0.75 wt% in the literature [23]. It
can be seen from Table 1 that carbon residue was formed on
Re2 O7 /alumina catalysts after butene metathesis reaction, and
the carbon content of the used catalysts increased obviously
with increasing pore diameter. In detail, the carbon content
increased from 1.44 wt% to 1.83 wt% with the pore diameter of alumina support from 5 nm to 19 nm. It is indicated
that large mesoporous catalysts have higher carbon residue
tolerance than the conventional catalyst with small pore size.
On the other hand, the catalyst with larger surface area had
low density of acidity (Table 1), so oligomeric or polymeric
byproducts were formed more slowly. In addition, Mol [12]
reported that the oxidation state of rhenium ion is one of the
most significant parameters governing the catalytic activity,

Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

and there exists the optimal oxidation state of rhenium ion for
olefin metathesis reaction. Unfortunately, the oxidation state
of rhenium under reaction conditions is difficult to draw any
definite conclusions, although partial reduction of Re(VII) to
Re(VI), Re(IV) and Re(III) has been observed [12]. It is assumed that over-reduction of rhenium past its optimum oxidation state during olefin metathesis process is a cause for
catalyst deactivation. Herein the stable working life-span of
Re2 O7 /alumina increased significantly with the increase of

357

pore size of alumina supports, so it is possible that larger


mesoporous alumina (MA19 and MA9) make the suitable valence of active Re-species more stable than the conventional
alumina support with small pore size (Al2 O3 ). Figure 4(b)
shows that the selectivities to propene for all the three catalysts were more or less the same, around 55 wt%. In addition,
the apparent conversions of 1-butene and 2-butene were about
62%66% and 52%55%, respectively, during the period
when the catalyst activity was stable (not shown in Figure 4).

Figure 4. Metathesis of 1-butene and 2-butene over different 13 wt% Re2 O7 /alumina catalysts. (a) Conversion of butene, (b) Selectivity to propene. Reaction
conditions: 1-butene/2-butene molar ratio of 1, WHSV = 1 h1 , 60 C, atmospheric pressure

The influence of rhenium loading on activity, propene


selectivity and working life-span of Re2 O7 /MA19 catalyst
in metathesis of 1-butene and 2-butene is shown in Figure 5. It is seen that the initial butene conversion increased from 48% to 55% with increasing Re2 O7 content from 6.5 wt% to 13 wt%, but when Re2 O7 loading
was in the range from 13 wt% to 23 wt%, the change
of butene conversion was nearly negligible. This result
seems to be inconsistent with the loading-dependence activity for Re2 O7 on conventional alumina [20] and mesoporous alumina [24] reported earlier, in which the initial activity increased with increasing Re2 O7 loading up to approximately 18 wt% and 7 wt%, respectively. It is interesting to

note that the stable working life-span increased in the order


6.5 wt% Re2 O7 /MA19<<13 wt% Re2 O7 /MA19<16.7 wt%
Re2 O7 /MA19<23 wt% Re2 O7 /MA19, i.e., being 10 h, 70 h,
85 h and 100 h, respectively. This is different from the experimental result in the reported literature [25], where the deactivation rate became higher when Re2 O7 loading was more than
12 wt%. We suspect that trace polar impurities in the butene
feedstock adsorbed on the active rhenium species may be a
factor of catalyst deactivation, therefore the activity of catalyst
with higher Re2 O7 loading can maintain longer time. The dependence of propene selectivity on the catalyst loading is not
obvious regularity, and propene selectivity was in the range of
40 wt%60 wt%. It is possible that the metathesis product

Figure 5. Metathesis of 1-butene and 2-butene over Re2 O7 /MA19 catalyst with different rhenium loadings. (a) Conversion of butene, (b) Selectivity to propene.
Reaction conditions are same as those in Figure 4

358

Lei Sang et al./ Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

selectivity depends on the nature of the active Re-sites which


dont change obviously with rhenium oxide loading.
The regenerability of Re2 O7 /MA19 catalyst was also investigated. The used catalyst after butene metathesis reaction was purged with nitrogen and heated up to 500 C and
then switched-on an air flow at 500 C for 1 h for regeneration. Afterwards the system was cooled down to the reaction
temperature in nitrogen and then butene was charged again
to the reactor for activity test. Figure 6 shows the comparision of catalytic properties of fresh 13 wt%Re2 O7 /MA19 catalyst and regenerated catalyst. It indicates that there was no

significant change of catalytic properties after regeneration.


After regeneration, butene conversion and propene selectivity were slightly decreased, but the stable working life-span
was increased a little. Let us speculate that the deactivation of
Re2 O7 /alumina catalyst may be attributed to a decrease of active Re-sites by carbonaceous remainder deposition and overreduction of rhenium ion. After the spent catalyst was exposed
to air atmosphere at 500 C for 1 h, the carbon remains and
water moiety were eliminated and the rhenium ion was reoxidized to the suitable valence, so the activity almost returned
to the initial level after regeneration.

Figure 6. Metathesis of 1-butene and 2-butene over 13 wt% Re2 O7 /MA19 catalyst. (a) Conversion of butene, (b) Selectivity to propene. Reaction conditions
are same as those in Figure 4

To sum up, the best catalytic performance of


metathesis of 1-butene and 2-butene was obtained over
23 wt%Re2 O7 /MA19 under the reaction conditions of 60 C,
atmospheric pressure and WHSV = 1 h1 , and butene conversion, propene selectivity and stable working life-span were
about 60%, 50 wt% and 100 h, respectively. While in preceding study, the optimal catalytic performance of metathesis of 1-butene was obtained over 10 wt%WO3 /Al2 O3 70 wt%HY under the reaction conditions of 180 C, 0.1 MPa
and WHSV = 1.5 h1 , 1-butene conversion, propene yield and
stable working life-span were about 95%, 24 mol% and 15 h
[26], respectively. And under the best reaction conditions
of 125 C, 1 MPa and WHSV (ethene + 2-butene) = 1 h1 ,
2-butene conversion, propene selectivity and stable working
life-span were about 86%, 95 mol% and 27 h, respectively,
over 3 wt%MoO3 /(MCM-2230 wt%Al2 O3 ) in metathesis
of ethene and 2-butene [27].
4. Conclusions
In summary, mesoporous -alumina with large pore size
(9 and 19 nm), high specific surface area (358 and 234 m2 /g)
and pore volume (1.70 and 0.54 cm3 /g) have been prepared
from dispersed pseudo-boehmite using pluronic P123 as template. These mesoporous alumina show spherical aggregates
morphology. These mesoporous alumina supported rhenium
oxide catalysts, with the propene selectivity of around 55%,

are found to be more active, far longer stable working lifespan and higher tolerance to carbon residue in gas-phase
metathesis of 1-butene and 2-butene than rhenium oxide supported on conventional alumina with small pore size. At
60 C and atmospheric pressure with WHSV = 1 h1 , the stable working life-spans of Re2 O7 /MA19 and Re2 O7 /MA9 are
far longer than that of Re2 O7 /Al2 O3 , being about 70 h, 50 h
and 20 h for the three catalysts, respectively. This contribution for the first time demonstrates a positive correlation between the catalyst activity, working life-span and Re2 O7 loading amount of Re2 O7 /MA19 catalyst in the range of 13 wt%
to 23 wt%. The deactivated catalyst can easily be regenerated
by calcinations in air at 500 C for 1 h and similar catalytic
performance can be obtained.
References
[1] Trimm D L, Stanislaus A. Appl Catal, 1986, 21(2): 215
[2] Tanaka K, Imai T, Murakami Y, Matsumoto T, Sugimoto W,
Takasu Y. Chem Lett, 2002, 31(1): 110

[3] Cejka
J. Appl Catal A, 2003, 254(2): 327
[4] Boissi`ere C, Nicole L, Gervais C, Babonneau F, Antonietti M,
Amenitsch H, Sanchez C, Grosso D. Chem Mater, 2006, 18(22):
5238
[5] Magnusson H, ye G, Glomm W R, Sjoblom J. J Dispersion
Sci Technol, 2006, 27(4): 547
[6] Khalil K M S. Appl Surf Sci, 2008, 255(5): 2874

[7] Marquez-Alvarez C, Zilkov


a N, Perez-Pariente J, Cejka
J. Catal

Journal of Natural Gas Chemistry Vol. 21 No. 3 2012

Rev, 2008, 50(2): 222


[8] Yang P D, Zhao D Y, Margolese D I, Chmelka B F, Stucky G D.
Chem Mater, 1999, 11(10): 2813
[9] Yang P D, Zhao D Y, Margolese D I, Chmelka B F, Stucky G D.
Nature, 1998, 396(6707): 152
[10] Lesaint C, Kleppa G, Arla D, Glomm W R, ye G. Microporous
Mesoporous Mater, 2009, 119(1-3): 245
[11] Lesaint C, Glomm W R, Borg , Eri S, Rytter E, ye G. Appl
Catal A, 2008, 351(1): 131
[12] Mol J C. Catal Today, 1999, 51(2): 289
[13] Andreini A, Xu X D, Mol J C. Appl Catal, 1986, 27(1): 31
[14] Onaka M, Oikawa T. Chem Lett, 2002, 31(8): 850
[15] Aguado J, Escola J M, Castro M C, Paredes B. Appl Catal A,
2005, 284(1-2): 47

[16] Hamtil R, Zilkov


a N, Balcar H, Cejka
J. Appl Catal A, 2006,
302(2): 193
[17] Liu Q, Wang A Q, Wang X H, Gao P, Wang X D, Zhang T.
Microporous Mesoporous Mater, 2008, 111(1-3): 323
[18] Sing K S W, Everett D H, Haul R A W, Moscou L, Pierotti R A,
Rouquerol J, Siemieniewska T. Pure Appl Chem, 1985, 57(4):

359

603

[19] Balcar H, Hamtil R, Zilkov


a N, Cejka
J. Catal Lett, 2004, 97(12): 25
[20] Moulijn J A, Mol J C. J Mol Catal, 1988, 46(1-3): 1
[21] van Schalkwyk C, Spamer A, Moodley D J, Dube T, Reynhardt
J, Botha J M. Appl Catal A, 2003, 255(2): 121
[22] Huang H J, Liu X C, Liu S L, Liu X M, Xu L Y, Han X W, Zhang
W P, Bao X H. Chin J catal (Cuihua Xuebao), 2010, 31(2): 186
[23] Stoyanova M, Rodemerck U, Bentrup U, Dingerdissen U, Linke
D, Mayer R W, Rotgerink H G J L, Tacke T. Appl Catal A, 2008,
340(2): 242
[24] Oikawa T, Ookoshi T, Tanaka T, Yamamoto T, Onaka M. Microporous Mesoporous Mater, 2004, 74(1-3): 93
[25] Amigues P, Chauvin Y, Commereuc D, Hong C T, Lai C C, Liu
Y H. J Mol Catal, 1991, 65(1-2): 39
[26] Liu H J, Zhang L, Li X J, Huang S J, Liu S L, Xin W J, Xie S J,
Xu L Y. J Nat Gas Chem, 2009, 18(3): 331
[27] Liu S L, Li X J, Xin W J, Xie S J, Zeng P, Zhang L X, Xu L Y.
J Nat Gas Chem, 2010, 19(5): 482

You might also like