You are on page 1of 6

Oxidative bromination of methane on silica-supported non-noble metal

oxide catalysts
Ronghe Lin
a,c
, Yunjie Ding
a,b,
*, Leifeng Gong
a,c
, Jingwei Li
a,c
, Weimiao Chen
a
, Li Yan
a
, Yuan Lu
a
a
Applied Catalysis Laboratory, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, PR China
b
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, PR China
c
Graduate School of Chinese Academy of Sciences, Beijing 100049, PR China
1. Introduction
Methane, the primary component of natural gas, is one of the
most abundant and low-cost feedstocks available. In spite of this,
the direct conversion of methane to a transportable liquid such as
methanol, DME, or other hydrocarbons is still of intense interest,
and the reaction is challenging due to the difculty of CH bond
cleavage, which is supposed to be the rst step for methane
activation. One-step partial oxidation of methane to methanol and
formaldehyde on MoO
3
/SiO
2
and V
2
O
5
/SiO
2
has been extensively
studied for many years, but the combination of low conversion,
low yield, and poor catalyst stability renders it less competitive
with the classic reforming process [1,2]
.
Another approach for the
direct conversion was explored by Fujiwara and coworkers [3]
utilizing H
2
SO
4
as an oxidant. According to the literature, methane
and CO were selectively and quantitatively transformed to acetic
acid by the function of CaCl
2
and H
2
SO
4
at 80 8C, 3040 atm. Since
then, other similar systems were also proposed [4,5]. Low-
temperature direct conversion of methane to liquids is an excellent
idea and an economical route, but the drawbacks are obvious,
considering the presence of concentrated H
2
SO
4
.
While one step from methane to fuels or chemicals remains a
great challenge to the catalysis community, various multi-step
reactions provide us alternative routes. Industrially, methane is
transformed to syngas by steaming reforming. The syngas is then
subsequently converted to oxygenated compounds or higher
parafns by Fischer-Tropsch catalysts [1]. Due to the high cost of
syngas production, partial oxidation of methane to syngas has been
intensively studied in the past decades [6]. Being slightly
exothermic, it is more energy efcient than methane steam
reforming. However, the formation of carbon deposits tends to
poison the noble metal catalysts [7]. The CO
2
reforming of methane
over supported noble metal catalysts has also been patented [8],
but still carbon deposition occurs.
Halogenations of methane towards higher hydrocarbons have
also been disclosed in many publications. These routes take at least
two steps, and the common drawback is the corrosion of feedstock
and products. Among the halogens, bromine was believed to be the
Applied Catalysis A: General 353 (2009) 8792
A R T I C L E I N F O
Article history:
Received 9 July 2008
Received in revised form 20 October 2008
Accepted 20 October 2008
Available online 28 October 2008
Keywords:
Methane oxidative bromination
Non-noble metal oxide
BaO/SiO
2
(CH
3
OH + CH
3
Br): CO
Metathesis
Deactivation mechanism
A B S T R A C T
Oxidative bromination of methane, mediated by HBr/H
2
O (solution), provides an original route for the
production of oxygenates and gasoline directly from natural gas. However, the reported catalysts for this
reaction all involved noble metals. From the consideration of replacing noble metal catalysts with
cheaper oxides, various silica-supported oxide catalysts were surveyed. It was found that the redox
ability of different metals had a strong impact on the product distribution. On the catalysts with metals
that lack facile redox ability, such as BaO/SiO
2
, both CH
3
Br and CO were main products. Otherwise, deep
oxidation proceeded. Moreover, methanol was rstly reported in this system. In order to obtain a molar
ratio of (CH
3
OH + CH
3
Br): CO = 1, which can provide a perfect feedstock for the synthesis of acetic acid,
the process variables were optimized on BaO/SiO
2
. It was demonstrated that 44.0% methane conversion
and 95% total selectivity of CH
3
Br, CH
3
OH and CO could be achieved at 620 8C, and the molar ratio of
(CH
3
Br + CH
3
OH): CO was close to 1. Time-on-stream tests showed declined catalytic performance after
30 h. The results fromN
2
-adsorption, XRF and XRDimplied that aggregation of bariumparticles occurred,
making the metathesis between BaO and BaBr
2
difcult. The functions of BaO were also proposed.
Additionally, the activity data were compared with those of noble metal catalysts.
2008 Elsevier B.V. All rights reserved.
* Corresponding author at: Applied Catalysis Laboratory, Dalian Institute of
Chemical Physics, Chinese Academy of Sciences, Dalian 116023, PR China.
Tel.: +86 411 84379143; fax: +86 411 84379143.
E-mail address: dyj@dicp.ac.cn (Y. Ding).
Contents lists available at ScienceDirect
Applied Catalysis A: General
j our nal homepage: www. el sevi er . com/ l ocat e/ apcat a
0926-860X/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2008.10.029
best choice, for the slightly exothermic reaction of bromine and
methane has the advantage that one can control the degrees of
halogenation [9]. In 1985, Olah [10,11] found that super-acid
catalyzed oxidative condensation of methane showed excellent
selectivity and could be carried out under milder conditions than
radical reactions. In his experiments, the CH
4
/X
2
(X = Cl, Br)
mixture was rstly catalyzed on a series of supported super-acid
catalysts, and then the products, mainly methyl halide, were
catalytically hydrolyzed to methanol and DME over g-alumina-
supported metal oxide/metal hydroxide catalysts. Lorkovic [12
15] also described an integrated multi-step process for the partial
oxidation of methane to alcohols, ethers and olens using O
2
as
oxidant. The rst step was the homogeneous brominationof alkane
to alkyl bromides and HBr. This mixture was then allowed to react
with a supported metal oxide, generating oxygenates or alkenes
depending upon the conditions. Lastly, the metathesis capacity of
metal oxide catalysts was recovered by the oxygenation of the
spent solid, and bromine was quantitatively regenerated for use.
One problemof this route is the signicant buildup of CH
2
Br
2
, even
of some CHBr
3
, at appreciable methane conversion. More recently,
Zhou et al. utilized HBr/H
2
O (solution) and O
2
instead of bromine
for methane activation. They found that Ru/SiO
2
[16,17] and Rh/
SiO
2
[18] were active catalysts for oxidative bromination of
methane (OBM) reaction [Eq. (1)], and the good activity was
ascribed to the bromine radical generating ability. By adjustment
of reaction variables, the products with a molar ratio of CH
3
Br:
CO = 1 were attainable. The aim products of this reaction (CH
3
Br
and CO) could be readily carbonylated to acetic acid. The CH
3
Br
could also be successfully hydrolyzed to DME [19] or converted to
gasoline [18]. Since both CH
3
Br and COare the main products of the
OBM reaction, it would be more efcient for acetic acid synthesis
[Eq. (2)].
2CH
4
2O
2
HBr ! CH
3
Br CO 3H
2
O; DH
0
685:8kJ=mol
(1)
CH
3
Br CO H
2
O ! CH
3
COOH HBr; DH
0
89:0kJ=mol
(2)
As an extension of the OBM reaction, this paper focuses on
replacing noble catalysts (Ru/SiO
2
, Rh/SiO
2
) with supported oxides,
especially on BaO/SiO
2
. The role and importance of various reaction
variables were examined with the aim of getting a product
distribution of (CH
3
Br + CH
3
OH): CO = 1 (mol/mol). Moreover, the
stability test was carried out, and the deactivation mechanismwas
studied using N
2
-adsorption, X-ray uorescence (XRF) and powder
X-ray diffraction (XRD) techniques. Additionally, the activity data
were compared with the reported data. It is expected that such
experiments would provide alternative ideas for the OBM reaction
and feedstock for acetic acid as well.
2. Experimental
2.1. Catalyst preparation
Various supported oxide catalysts were prepared based on the
procedures described elsewhere [16]. Firstly, silica gel (60100
mesh) was impregnated with corresponding aqueous solutions of
metal nitrates or ammoniumfor a fewhours. After drying at 120 8C
overnight, the residual was calcined for 12 h at 800, 600 or 450 8C
for W, Mo, and the other samples respectively for the generation of
the catalysts. The catalysts were used hereafter without further
treatments.
2.2. Kinetic evaluation
The OBM reaction was conducted in a quartz-tube micro-
reactor (15 550 mm, hot zone 300 mm) at atmospheric pressure.
For a typical test 2.0 g catalyst was inserted between two plugs of
quartz wool, and the empty space inside the tube was packed with
quartz sand. Ahead of each reaction, the catalyst bed was heated
from room temperature to 450 8C under 20 ml/min N
2
ow. Then,
the gaseous reactants (CH
4
and O
2
) were introduced into the
reactor by two separate mass ow controllers, together with the
feed of liquid reactants by a syringe. After a few minutes for
stabilizing of the liquid feeds, the temperature was nally raised to
a settled one at a rate of 5 8C/min. All the activity data in this paper
were rstly collected after 12 h unless illuminated otherwise. For
blank reaction test, the reactor was packed with only quartz sand.
Efuent exiting the reactor was immediately cooled down by an
ice-water mixture before passing through two CCl
4
traps (each
5.0 g collected for 1 h) which were kept at 0 8C for the full
dissolution of CH
3
Br (b.p. 3.6 8C). After sampling, known CH
2
Cl
2
was added in as an internal standard to quantify liquid products on
an HP4890 GC tted with 30 m0.53 mm i.d. SE-54 capillary
column and FIDdetector. The gaseous products downstreamof the
traps were directly analyzed on-line using an HP4890 GC with
2 m2 mm i.d. TDX-01 column and TCD detector.
2.3. Characterization of catalysts
The structure properties of the samples were determined by N
2
-
adsorption using a physical adsorption instrument (Quanta-
chrome, USA). The specic surface area and pore volume were
determined by N
2
-adsorption at 77 K, after samples were
pretreated at 623 K under vacuum for 3 h.
Elemental composition of the catalysts was analyzed on a
Philips Magix XRF instrument.
XRD measurements of the samples were performed over a
Philips Xpert MPD X-ray diffraction spectrometer equipped with a
graphite monochromator and Cu Ka (40 kV, 40 mA) irradiation,
covering 2u between 10 and 558.
3. Results and discussion
3.1. Screening of different silica-supported oxide catalysts on the OBM
reaction
In order to nd a noble metal-free catalyst that is both active
and selective, various supported metal oxide catalysts were tested.
The activity data are summarized in Table 1. It is clear that, in the
absence of catalysts, the OBM reaction was poor in activity and
selectivity, with multiple substitution products like di-, tri-
bromomethane formed. However, when catalysts were employed,
methane conversion increased by roughly 10%, and the products
distribution changed greatly. One can see that the selectivity of CO
Table 1
Oxidative bromination of methane on different supported metal oxide catalysts.
Catalysts Conversion (%) Selectivity (C%)
MO
x
/SiO
2
CH
4
CH
3
OH CH
3
Br CH
2
Br
2
CHBr
3
CO CO
2
Blank 20.0 13.6 55.4 21.4 1.9 7.6 0
5 V 25.8 3.5 9.2 0.5 0 86.7 0
10 Ce 28.1 0 3.2 0.2 0 93.6 3.0
5 Mo 29.6 18.0 34.3 3.4 0 43.0 1.3
10 Ba 30.6 11.2 47.6 1.9 0 38.3 2.0
5 W 31.4 14.7 26.5 0.4 0 58.4 0
Reaction variables: T 650 8C; 40 wt.% HBr/H
2
O ow 8.0 ml/h; gas ow 25 ml/min
(CH
4
:O
2
= 4.0).
R. Lin et al. / Applied Catalysis A: General 353 (2009) 8792 88
increased while simultaneously the selectivity of Br-substituted
products decreased.
Interestingly, the product distribution fell into two groups on
different catalysts. On Mo, Ba and W catalysts, the main products
were CH
3
OH, CH
3
Br and CO, while CO was predominant on V and
Ce catalysts (selectivity exceeding 85%). Originally, we hoped to
nd a relation between product distribution and the acidic or basic
nature for the selected catalysts, but the data showed no clear
trends. As V [20] and Ce [21,22] are good redox catalysts, we
deduced that the oxidation of bromomethanes mainly occurred on
the surface of catalysts and that the metals with facile redox ability
accelerate this process. This is in accord with the viewpoints of
Lorkovic, who observed deep oxidation on facile redox oxide
catalysts when converting CH
3
Br to higher hydrocarbons [14].
On most catalysts, methanol was detected, which is contrary to
Zhous experimental results [1619]. In an earlier report [10] Olah
found that 6% methanol and DME were the observed products
when methane was reacted with bromine in the presence of steam
and oxygen over 0.5Pt/Al
2
O
3
catalyst at 250 8C, but no data or
further reports were published. CH
3
Br hydrolyzation was sup-
posed to be the possible pathway of methanol formation in this
system. In fact, methanol selectivity, to some extent, was
irrespective of reaction conditions, but showed a certain propor-
tion to the selectivity of CH
3
Br. It should be noted that methanol
selectivity on Mo and W catalysts was higher, and this can be
accounted for by the stability of methanol on these oxide catalysts
[23].
Since for the OBM reaction, CH
3
OH, CH
3
Br and CO were the
main products on non-noble oxide catalysts, it should be much
easier to include CO as one of target-products rather than focus on
high CH
3
Br selectivity. Among the surveyed catalysts, BaO/SiO
2
showed the best total selectivity of CH
3
Br, CH
3
OH and CO,
therefore, it was chosen for further experiments. We hoped to
achieve high methane conversion and a molar ratio of
(CH
3
Br + CH
3
OH): COclose to 1 by optimizing the process variables
on the catalyst.
3.2. Effect of process variables on the OBM reaction for BaO/SiO
2
catalyst
3.2.1. Feed rate of liquid reactants
The ow rate of HBr/H
2
O (40 wt.%) was varied from 5.0 to
8.0 ml/h in order to study the effect of weight hourly space velocity
(WHSV) of liquid feeds. The results obtained are given in Table 2. It
was observed that increasing WHSV dramatically enhanced CH
3
Br
selectivity (from 22.7 to 47.6) and decreased the selectivity of CO
(from 67.3 to 38.3), while methane conversion increased slightly.
Zhou and coworkers [17] found that CH
3
Br was the primary
product, while CH
2
Br
2
and CO were secondary products. This is in
accord with our results. To obtain an acceptable selectivity of
CH
3
Br and CO, 8.0 ml/h was set hereafter for the liquid feed rate.
3.2.2. Temperature effect
The effect of temperature on the OBMreaction was investigated
from580 to 650 8C. The results are presented in Table 3. Obviously,
raising temperature reduced CH
2
Br
2
selectivity (from 13.4 to 1.9)
and increased the selectivity of CO (from 24.1 to 38.3) at the same
time, but it had little impact on methane conversion under the set
conditions. By calculating the Gibbs free energies of CH
2
Br
2
formation [Eq. (3)], Degirmenci et al. [9] claimed that CH
2
Br
2
formation became non-spontaneous when the temperature is
above 450 8C. Since both CH
2
Br
2
and CO are secondary products of
CH
3
Br [Eq. (3), Eq. (4)], one can control the product distribution by
adjusting the reaction temperature. Obviously, increasing tem-
perature inhibits CH
2
Br
2
formation and facilitates CH
3
Br oxidation.
The results in Table 3 also indicate that, when considering
feedstock for acetic acid, it is meaningless when the temperature is
above 620 8C.
CH
3
Br Br
2
! CH
2
Br
2
HBr; DH
0
70:6kJ=mol (3)
CH
3
Br O
2
! CO 3H
2
O HBr; DH
0
89:0kJ=mol (4)
3.2.3. Effects of CH
4
:O
2
ratio and HBr concentration
Though we had failed to enhance methane conversion by
increasing the temperature, we turned to adjust CH
4
:O
2
ratio and
HBr concentration. The results obtained are summarizedin Table 4.
As one can see, raising HBr concentration from 40 to 48 wt.% did
increase methane conversion (A1 vs. A2), but it also greatly
improved the selectivity of all Br-substituted products, especially
for the selectivity of CH
2
Br
2
, a presumed byproduct for acetic acid
synthesis. However, when CH
4
:O
2
ratio was reduced from 4.0 to
2.5, methane conversion increased by 14% even at a lower
temperature. It is self-evident that reduction in CH
4
:O
2
ratio is
in favor of generating bromine radicals and bromomethane
oxidation.
3.3. Catalytic performance on the OBM reaction with time-on-stream
Fig. 1 shows the whole 40 h time-on-stream performance of
BaO/SiO
2
on the OBM reaction. One should note that methane
conversion and the products distribution reached a steady and
optimal state after 8 h initial testing. Namely, 44.0% methane
conversion and 95% total selectivity of CH
3
Br, CH
3
OH and CO were
Table 2
Inuence of HBr/H
2
O (40 wt.%) ow rate on the OBM reaction.
HBr/H
2
O (ml/h) Conversion (%) Selectivity (C%)
CH
4
CH
3
OH CH
3
Br CH
2
Br
2
CO CO
2
5.0 28.7 6.0 22.7 0.7 67.3 3.3
6.5 29.6 12.7 33.7 1.2 50.0 2.4
8.0 30.6 11.2 47.6 1.9 38.3 2.0
Reaction variables: T 650 8C; gas ow 25 ml/min (CH
4
:O
2
= 4.0).
Table 3
Inuence of temperature on the OBM reaction.
T (8C) Conversion (%) Selectivity (C%)
CH
4
CH
3
OH CH
3
Br CH
2
Br
2
CO CO
2
580 26.0 12.7 49.8 13.4 24.1 0
600 29.4 13.0 47.7 8.3 29.9 1.1
620 30.0 12.0 46.1 2.5 38.1 1.3
650 30.6 11.2 47.6 1.9 38.3 2.0
Reaction variables: 40 wt.% HBr/H
2
O ow 8.0 ml/h; gas ow 25 ml/min
(CH
4
:O
2
= 4.0).
Table 4
Inuence of CH
4
:O
2
ratio and HBr concentration on the OBM reaction.
Group Conversion (%) Selectivity (C%)
CH
4
CH
3
OH CH
3
Br CH
2
Br
2
CO CO
2
A1 29.6 12.7 33.7 1.2 50.0 2.4
A2 33.8 15.0 47.9 12.0 25.2 0
B1 30.6 11.2 47.6 1.9 38.3 2.0
B2 44.0 12.3 37.0 2.1 46.2 2.3
Reaction variables: T 650 8C; 40 wt.% HBr/H
2
O ow 8.0 ml/h; gas ow 25 ml/min
(CH
4
:O
2
= 4.0). A1: 40 wt.% HBr/H
2
O ow 6.5 ml/h; A2: 48 wt.% HBr/H
2
O ow
6.5 ml/h; B2: T 620 8C; gas ow 28 ml/min (CH
4
:O
2
= 2.5).
R. Lin et al. / Applied Catalysis A: General 353 (2009) 8792 89
achieved, and the molar ratio of (CH
3
Br + CH
3
OH): CO was close to
1. After 25 h duration of the test, however, there was a decline in
the catalytic performance. Methane conversion and the selectivity
of deep oxidation products began to decline, while the selectivity
of liquid products changed otherwise. 5 h later, this trend became
even more obvious, and the product distribution got closer to the
results of the blank test. This poor activity indicated the
deactivation of the catalyst.
3.4. Characterization of catalysts and deactivation mechanism
To get a better understanding of the deactivationmechanism, we
employed N
2
-adsorption, XRF and XRDtechniques. The results from
XRF and XRD are presented in Table 5 and Fig. 2, respectively. The
loading of Ba in the fresh catalyst was 9.9936 wt.%, which was close
to the value set in the catalyst design. But this index dropped to
8.2682 wt.% in the rst 20 h, and declined to 5.6884 wt.% after 40 h.
The rst reduction in Ba loading was well explained by the XRD
pattern, which showed that a mixture of Ba(OH)
2
and SiO
2
phases
existed for the fresh catalyst after calcination, but no diffractions
corresponding to BaO were present. Thus, the BaO must be well
dispersed onthe support, withsizes smaller thanthe detection limit
of the instrument. However, the diffractions of Ba(OH)
2
disappeared
in the rst 20 h. Since HBr is a strong acidic reactant, which is very
reactive with Ba(OH)
2
, the reduction in Ba loading in the rst period
was attributed to the leaching of Ba(OH)
2
. On the other hand, with
the extinction of Ba(OH)
2
, more bromine radicals were generated,
resulting in better activity, as reected in Fig. 1. One can see that
there were rises of methane conversion and of the proportion of
(CH
3
Br + CH
3
OH): CO during the initial phase. As time went by, the
reaction gradually became stable. The second phase reduction in Ba
loading might be caught by the leaching of BaBr
2
.
In order to conrmwhether the deactivation was resulted from
the transformation from BaO to BaBr
2
, we re-oxidized the used
catalyst in a gas ow of O
2
/N
2
(N
2
:O
2
= 4, 25 ml/min) for 4 h at
620 8C, then cooled it down to room temperature. It was observed
that the catalyst changed fromlight yellow(used catalyst) to white
(re-oxidized catalyst). Thus, BaBr
2
must be partially re-oxidized
back to BaO. When the reactants were fed in on this treated
catalyst, however, bromine was detected immediately in the
efuent and the initial catalytic performance was as poor as that of
the deactivated one. So the metathesis between BaOand BaBr
2
was
hard to proceed.
To our surprise, XRD analysis demonstrated the existence of
BaSi
2
in the used catalyst. Though the formation of BaSi
2
is still a
puzzle, one thing we are certain about is that it must have some
connections to the strong reductive atmosphere, as well as the
intense exothermic property of the OBM reactions. It also
reminded us that the deactivation might be caught by the
aggregation of barium particles. Table 6 shows the physical
properties of the samples. N
2
-adsorption revealed that the specic
surface area of the fresh and used catalysts were 293 and 226 m
2
/g,
respectively. Meanwhile, the mesopore volume decreased from
1.0290 to 0.7635 ml/g. The diffractions of BaBr
2
also indicated that
the aggregation of barium particles happened. Since no BaBr
2
diffractions were observed after 20 h, but the results from XRF
(Br:Ba = 1.5) indicated a co-existence of BaO and BaBr
2
. 20 h later,
Br:Ba reached 2.0 and BaBr
2
diffractions appeared. Based on these
results, we deduce that the deactivation was caused by the
abundant heat from the OBM reaction which resulted in severe
aggregation of barium particles, thus making the metathesis
between BaO and BaBr
2
more difcult.
3.5. Further understanding on the function of BaO/SiO
2
Obviously, the radical reactions between methane and Br
2
in
gas phase are poor in selectivity. The huge change in products
distribution indicated that, besides generating more bromine
radicals, there must be other functions when catalysts were
employed. As for BaO/SiO
2
, a satisfying proportion of
(CH
3
Br + CH
3
OH): CO was achieved, which is quite in contrast to
the blank test in which CO selectivity was only 7.6%. The dramatic
Fig. 1. Stability test of BaO/SiO
2
as a function of 40 h time-on-stream. Reaction
variables: T 620 8C; 40 wt.% HBr/H
2
O = 8.0 ml/h; gas ow28 ml/min (CH
4
:O
2
= 2.5).
Table 5
Elemental composition of BaO/SiO
2
from XRF.
BaO/SiO
2
Element/concentration (wt.%) Br/Ba
(mol/mol)
Ba Si O Br Others
Fresh 9.9936 41.0990 48.3180 0 0.5894 0
After 20 h 8.2682 39.4044 44.1530 7.4623 0.7121 1.5
After 40 h 5.6884 39.9350 46.7760 6.7570 0.8436 2.0
Fig. 2. XRD patterns of the fresh and used BaO/SiO
2
catalysts.
Table 6
Physical properties of fresh and used BaO/SiO
2
catalysts.
BaO/SiO
2
Surface
area (m
2
/g)
Mesopore
volume (ml/g)
Average pore
size (nm)
Fresh 293 1.0290 12.4
Used (after 40 h) 226 0.7635 12.6
R. Lin et al. / Applied Catalysis A: General 353 (2009) 8792 90
increase in CO selectivity indicated that the oxidation of CH
3
Br
mainly occurred on barium oxides via surface reactions. Fig. 3
shows O
2
consumption as a function of reaction time. O
2
was
almost fully consumed during the initial phase, and no Br
2
was
found in the efuent. However, a curve appeared in the 30th h, and
the consumption of O
2
declined to 77.5% in the next 10 h.
Meanwhile, the efuent changed from light yellow to red and Br
2
was detected. The decline of O
2
consumption and the detection of
Br
2
could be related to the depletion of BaO, which not only
prevented the oxidation of CH
3
Br but also led to the leaching of Br
2
.
The metathesis dynamics between BaO and BaBr
2
played an
important role in the stability of the catalyst. As the combination of
XRF and XRDresults indicated a co-existence of well dispersed BaO
and BaBr
2
during the initial 20 h, the metathesis between BaO and
BaBr
2
might proceed well, which was reected on the stable
activity performance. This metathesis dynamic was nally broken
due to the gradual transformation fromBaOto BaBr
2
, together with
the particle size growth of BaBr
2
, as reected on the BaBr
2
diffractions in Fig. 2.
The strong surface basicity of BaO might have an impact on the
product distribution, although the functions of acidbase proper-
ties of catalysts are still unclear as a whole. As a basic oxide, BaO is
well employed in methane dehydrogenation reactions [2426]. It
was suggested that highly basic sites are benecial for the efcient
activation of methane [26]. While in the OBM system, radical
reactions are predominant. The moderate selectivity of CO might
indicate that the strong basicity of BaO, to some extent, promotes
the deep oxidation of methane, which has been observed in
methane dehydrogenation reactions [24].
We propose that BaO functions mainly in the following two
aspects. Firstly, HBr was partially converted to Br
2
on its surface and
it actedas a bromine absorber (Br
2
bothonsurface andingas) bythe
metathesis between BaO to BaBr
2
, which prevented the leaching of
Br
2
. Thus, the radical reaction in the gas phase was restricted.
Secondly, Br-substituted methane formed on the catalyst surface
was muchmore easilyoxidizedtoCO
x
(x = 1, 2) onBaOcomparingto
the gaseous oxidation. The hypothesis above was in accord with the
activity performance, as the mainreaction products were CH
3
Br and
CO. The exhausted oxide could be regenerated in-situ through
metathesis reaction, and the catalytic cycle retained. The overall
steps of the OBM reaction are presented in Scheme 1.
3.6. Comparison with the activity of different catalysts during the
OBM reaction
The performance of BaO/SiO
2
was comparedwithtypical activity
data of noble metal-including catalysts from published reports, as
listedinTable7. Fromthistable, it is obvious that theOBMreactionis
more benecial, as regards providing feedstock for acetic acid
synthesis. Although high CH
3
Br selectivity could be achieved on Rh/
SiO
2
catalyst, it required more severe conditions, namely higher
reaction temperature and HBr/H
2
O concentration, and still deep
oxidation and further Br-substitution proceeded, leading to a large
amount of byproducts (17%). It should be noted that Rh loading was
0.406%in this catalyst and the gas hourly space velocity (GHSV) was
Fig. 3. O
2
consumption as a function of reaction time. Reaction variables: T 620 8C;
40 wt.% HBr/H
2
O = 8.0 ml/h, gas ow 28 ml/min (CH
4
:O
2
= 2.5).
Scheme 1. Pathways for the OBM reaction on BaO/SiO
2
.
Table 7
Comparison between BaO/SiO
2
and noble metal-including catalysts.
Catalysts T (8C) GHSV ml/(g h) CH
4
:O
2
HBr/H
2
O (ml/h) Conversion% Selectivity (C%)
CH
4
CH
3
Br CO CH
2
Br
2
CO
2
Rh [18] 660 300 4.0 6.5
a
30 83 7 7 2
RuNiBaLa [27] 660 300 1.0 8.0
b
70.0 44.7 44.3 3.0 8.0
BaO 620 840 2.5 8.0
b
44.2 45.8
c
50.1 1.7 2.4
BaO 620 750 4.0 8.0
b
30.0 58.1
c
38.1 2.5 1.3
Note: all the catalysts were supported on SiO
2
.
a
48 wt.% HBr/H
2
O.
b
40 wt.% HBr/H
2
O.
c
CH
3
OH included.
R. Lin et al. / Applied Catalysis A: General 353 (2009) 8792 91
only 300 ml/(g h). While both are catalysts for achieving a desired
product distribution for acetic acid synthesis, BaO/SiO
2
showed
superior activity performance to RuNiBaLa/SiO
2
. In spite of the
production of fewer byproducts (less that 5%), high conversion of
methane (44.2%) was obtained in a higher CH
4
: O
2
ratio (2.5) feed
when the GHSV was more than doubled. Moreover, it was
demonstrated that 620 8C was sufcient for this catalysis system,
and further increase in temperature had little benet for methane
conversion. As a matter of fact, a relativelyhigher CO
2
selectivitywas
observed in our experiments when Ru was involved in the catalysts.
4. Conclusions
In this study, various silica gel-supported non-noble metal
oxide catalysts were surveyed as regards the OBM reaction. It was
found that the product distribution was greatly inuenced by the
redox ability of the metals involved in the catalyst. It was also
demonstrated that the OBM reaction can be favorable on BaO/SiO
2
catalyst in terms of providing feedstock for acetic acid. High
methane conversion and desired product distribution with higher
GHSV were achieved at a lower temperature compared to the
earlier results. Even though bromine radicals were generated, we
propose that BaO functioned as a bromine absorber and promoted
the oxidation of CH
3
Br on its surface. As to the deactivation, one
reasonable interpretation is: the aggregation of barium particles
caused by abundant reaction heat prevented the metathesis
between BaO and BaBr
2
.
Acknowledgements
We are grateful for the nancial support of the Ministry of
Science and Technology of the Peoples Republic of China
(2005CB221406).
References
[1] R.H. Crabtree, Chem. Rev. 95 (1995) 9871007.
[2] K. Otsuka, M. Hatano, J. Catal. 108 (1987) 252255.
[3] M. Asadullah, T. Kitamura, Y. Fujiwara, Angew. Chem. Int. Ed. 39 (2000) 2475
2478.
[4] R.A. Periana, O. Mironov, D. Taube, G. Bhalla, C.J. Jones, Top. Catal. 32 (2005) 169
174.
[5] M. Zerella, S. Mukhopadhyay, A.T. Bell, Chem. Commun. 17 (2004) 19481949.
[6] Q.G. Yan, T.H. Wu, W.Z. Weng, H. Toghiani, R.K. Toghiani, H.L. Wan, C.U. Pittman, J.
Catal. 226 (2004) 247259.
[7] A.T. Ashcroft, A.K. Cheetham, J.S. Foord, M.L.H. Green, C.P. Grey, A.J. Murrell, P.D.F.
Vernon, Nature 344 (1990) 319321.
[8] B.L. Gustafson, J.V. Walden, US Patent 5, 068, 057.(1991).
[9] V. Degirmenci, D. Uner, A. Yilmaz, Catal. Today 106 (2005) 252255.
[10] G.A. Olah, et al. J. Am. Chem. Soc. 107 (1985) 71057109.
[11] G.A. Olah, Acc. Chem. Res. 20 (1987) 422428.
[12] X.P. Zhou, A. Yilmaz, G.A. Yilmaz, I.M. Lorkovic, L.E. Laverman, M. Weiss, J.H.
Sherman, E.W. Mcfarland, G.D. Stucky, P.C. Ford, Chem. Commun. 18 (2003)
22942295.
[13] I.M. Lorkovic, A. Yilmaz, G.A. Yilmaz, X.P. Zhou, L.E. Laverman, S. Sun, D.J. Schaefer,
M. Weiss, M.L. Noy, C.I. Cutler, J.H. Sherman, E.W. Mcfarland, G.D. Stucky, P.C.
Ford, Catal. Today 98 (2004) 317322.
[14] I.M. Lorkovic, M. Noy, M. Weiss, J. Sherman, E. Mcfarland, G.D. Stucky, P.C. Ford,
Chem. Commun. 5 (2004) 566567.
[15] A. Breed, M.F. Doherty, S. Gadewar, P. Grosso, I.M. Lorkovic, E.W. Mcfarland, Catal.
Today 106 (2005) 301304.
[16] K.X. Wang, H.F. Xu, W.S. Li, X.P. Zhou, J. Mol. Catal. A: Chem. 225 (2005) 6569.
[17] K.X. Wang, H.F. Xu, W.S. Li, C.T. Au, X.P. Zhou, Appl. Catal. A: Gen. 304 (2006) 168
177.
[18] Z. Liu, L. Huang, W.S. Li, F. Yang, C.T. Au, X.P. Zhou, J. Mol. Catal. A: Chem. 273
(2007) 1420.
[19] K.X. Wang, H.F. Xu, W.S. Li, X.P. Zhou, Catal. Lett. 100 (2005) 5357.
[20] T.C. Watling, G. Deo, K. Seshan, I.E. Wachs, J.A. Lercher, Catal. Today 28 (1996)
139145.
[21] W. Liu, M. Flytzanistephanopoulos, J. Catal. 153 (1995) 304316.
[22] P. Bera, S.T. Aruna, K.C. Patil, M.S. Hegde, J. Catal. 186 (1999) 3644.
[23] S.H. Taylor, J.S.J. Hargreaves, G.J. Hutchings, R.W. Joyner, Appl. Catal. A 126 (1995)
287296.
[24] X. Wang, Y.C. Xie, Catal. Lett. 72 (2001) 5157.
[25] X. Wang, Y.C. Xie, New J. Chem. 25 (2001) 964969.
[26] R. Burch, D.J. Crittle, M.J. Hayes, Catal. Today 47 (1999) 229234.
[27] K.X. Wang, H.F. Xu, S.W. Li, X.P. Zhou, CN Patent 1,724,503.(2005).
R. Lin et al. / Applied Catalysis A: General 353 (2009) 8792 92

You might also like