You are on page 1of 10

Catalysis Today 197 (2012) 127136

Contents lists available at SciVerse ScienceDirect


Catalysis Today
j our nal homepage: www. el sevi er . com/ l ocat e/ cat t od
The role of oxidation catalyst in dual-catalyst bed for after-treatment
of lean burn natural gas exhaust
Preshit Gawade, Anne-Marie C. Alexander, Ryan Clark, Umit S. Ozkan

Department of Chemical and Biomolecular Engineering, The Ohio State University, 140 W. 19th Avenue, Columbus, OH 43210, United States
a r t i c l e i n f o
Article history:
Received 16 April 2012
Received in revised form1 August 2012
Accepted 2 August 2012
Available online 1 September 2012
Keywords:
NOx reduction
Sulfated zirconia
Palladium
Cobalt catalyst
Kinetics
DRIFTS
Effect of water
Lean gas engines
a b s t r a c t
A dual-catalyst bed composed of a reduction (Pd/SZ) and an oxidation (CoO
x
/CeO
2
) catalyst was investi-
gated for selective catalytic reduction (SCR) of NO
2
using hydrocarbons for lean-burn natural gas engines.
The choice of a multifunctional oxidation catalyst was seen to play a crucial role to achieve the desirable
NO
2
reduction performance. A dual-catalyst bed optimization showed that different cobalt containing
oxidation catalysts could be exploited to achieve desirable NO
x
and hydrocarbon conversions. Moreover,
cyclic experiments in the presence of water vapor were conducted to understand its implications on dual
catalyst bed performance. Finally, both kinetic and in situ diffuse reectance infrared Fourier transform
spectroscopy (DRIFTS) studies were undertaken to examine the behavior of the oxidation catalyst with
respect to hydrocarbon oxidation which consequently affect the overall NO
x
reduction activity.
2012 Elsevier B.V. All rights reserved.
1. Introduction
In the context of evermore stringent requirements for fuel ef-
ciency and CO
2
emissions, natural gas-red reciprocating engines
represent an important and increasingly popular choice within
the distributed energy market place. Lean burn natural gas recip-
rocating engines offer a simple, well-understood technology and
have several advantages over stoichiometric gasoline engines [1].
Lean burn conditions are associated with higher engine efciencies
and signicantly cleaner engine out emissions. However, despite
emissions being greatly reduced, engine exhaust still contains con-
siderable amounts of NO
x
species, carbonmonoxide andun-burned
hydrocarbons. Some NO
x
emissions can be reduced through the
modication of fuels or by altering the combustion parameters,
however in order to reach acceptable emission levels the need for
effective after-treatment is evident.
Among current catalytic NO
x
reduction control technologies,
three way catalysts and ammonia- or urea-based selective catalytic
reduction (SCR) are the most prevalent. Both methods are highly
effective for the current combustion technologies, which operate
over a wide temperature window (200400

C), yet are unsuit-


able for the next generation of lean-burn engines. For example,
three way catalysts, which are mostly used in mobile applications,

Corresponding author. Tel.: +1 614 292 6623; fax: +1 614 292 3769.
E-mail address: ozkan.1@osu.edu (U.S. Ozkan).
quickly deactivate under lean conditions, while the use of ammo-
nia in emission abatement poses hazards in itself, mainly related to
ammonia slip, direct ammonia oxidation, corrosion of equipment
due to the formation of ammoniumsalts, and the additional issues
related to ammonia storage and handling [2].
The selective reduction of NO
x
species by hydrocarbons has
attractedsignicant attentionas a promisingalternative toconven-
tional after-treatment technologies [35]. This type of application
is particularly suited to natural gas engines, in which unburned
hydrocarbons, particularly methane, are already present in the
exhaust stream and capable of acting as the reducing agent.
Although methane, which is a major component of natural gas,
is readily available, its effective use as a reducing agent in
hydrocarbon-SCR systems is somewhat limited. This is mainly due
to the competition fromthe combustion of the hydrocarbon in the
presence of excess oxygen and the inherent difculty of methane
activation [6,7].
Among the catalysts which have been reported to date,
palladium-based catalysts supported on zeolites, such as ZSM-5
and mordenite, have shown the highest activity and selectivity for
catalytic NO
x
reductionunder leanconditions [8]. Despitetheir high
activityandselectivityinHC-SCR, zeolitic catalysts suffer frompoor
hydrothermal stability [911]. Consequently dealumination of the
material may occur, resulting in the formation of PdO aggregates,
which are thought to be active for methane combustion, and loss
of metal dispersion [12]. Several authors have reported that Pd
supported on acidic materials, such as zeolites or zirconia show
0920-5861/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2012.08.009
128 P. Gawade et al. / Catalysis Today 197 (2012) 127136
a signicant improvement in NO reduction [1315]. The devel-
opment of acidic non-zeolitic catalysts, such as sulfated zirconia,
for the HC-SCR of NO
x
species, has therefore attracted much inter-
est, mainly due to an increase in hydrothermal stability compared
to zeolitic counterparts. Studies by Resasco and co-workers have
shown that acidic supports aid the stabilization of palladiumin the
form of Pd
2+
ions [16], the active sites for NO
x
reduction [17,18].
Further studies by the same authors [19] have also shown that pro-
tons associated with surface sulfate groups serve as anchoring sites
for the Pd
2+
ions in Pd/SZ. Although there has been much debate
regarding the mechanisms involved in NO
x
reduction with hydro-
carbons, NO
2
is thought to be a key intermediate species in NO
x
reduction.
Previously we have describedanintegratedoxidationandselec-
tivereductioncatalytic system[2024] whereNOwas rst oxidized
to NO
2
, then reduced selectively with hydrocarbons. This method
combines separate oxidation and reduction catalyst components
in order to perform three distinct catalytic functions, specically
NO
x
reduction, CO oxidation and hydrocarbon combustion. The
dual-catalyst approachfor lean-burnexhaust after-treatment takes
advantage of the stronger oxidizing potential of NO
2
compared
to NO, which in turn helps to utilize the reducing capability of
unburned hydrocarbons in the exhaust. Despite the exothermic
oxidation of NO to NO
2
being thermodynamically limited at high
temperatures, the dual-catalyst mechanism helps to drive the NO
oxidation reaction in the forward direction, as a result of NO
2
being
continuously removed via NO
2
SCR. Any gaseous NOthat is present
in the system, a consequence of partially reduced NO
2
, can be
re-oxidized by the oxidation catalyst. It is clear that the oxida-
tion catalyst assumes a multi-functional role in the dual-catalyst
scheme. In addition to oxidizing NO or re-oxidizing partially
reduced NO
2
species, the oxidation catalyst also plays a role in
catalyzing the combustion of un-burned hydrocarbons and the oxi-
dation of carbon monoxide, which have not been consumed during
the SCR reaction [25].
In this contribution, the multifunctional role of a metal oxide-
supported cobalt oxidation catalyst, namely CoO
x
/CeO
2
, in a
dual-catalyst systemis addressed. The effect of Co-loading in NOto
NO
2
oxidation has been investigated through steady state exper-
iments. In order to optimize the dual bed composition, the ratio
of reduction to oxidation catalyst was varied and investigated via
steady state experiments under simulated engine exhaust condi-
tions. Cyclic experiments were conducted in the presence of water
vapor to understand the effect of different cobalt loadings on the
overall hydrothermal stability of the dual-bed system. Both kinetic
and in situ DRIFTS studies are reported with the aim of under-
standing the behavior of the oxidation catalyst with respect to
hydrocarbon oxidation, which consequently affects the overall NO
x
reduction performance.
2. Experimental
2.1. Catalyst preparation
Palladiumcatalyst supported on sulfated zirconia with 0.3wt%
loading, which was employed as the NO
x
reduction catalyst for
the dual-catalyst system, was prepared using a one-pot solgel
method, details of which can be found elsewhere [20,21]. Cobalt
catalysts supported on ceria were prepared through an incipi-
ent wetness impregnation technique and were used as the NO
oxidation catalyst component in the dual-catalyst system. The
nanoparticle ceria support was prepared using a precipitation
method described previously [26]. The calcined ceria support was
impregnatedwithaqueous cobalt nitrate hexahydrate so as to yield
catalysts with Co loadings of either 2 or 10wt% [27].
2.2. Catalyst characterization
Surface area measurements were obtained using a Micromerit-
ics ASAP 2010/2020 accelerated surface area and porosimetry
instrument using a N
2
-physisorption method. Prior to analysis, the
samples (0.1g) were degassed overnight at 130

C under a vac-
uum of 3mHg to remove any adsorbed moisture. The specic
surface areas of the catalysts, 0.3% Pd/SZ, 10% CoO
x
/CeO
2
, and 2%
CoO
x
/CeO
2
, were determined by applying the Brunauer, Emmett
and Teller (BET) method to the nitrogen physisorption isotherms
which were determined at liquid nitrogen temperatures (77K) and
found to be 38m
2
/g, 84m
2
/g and 106m
2
/g, respectively.
Palladium dispersion over sulfated zirconia was calculated by
hydrogen chemisorption using a Micromeritics Autochem 2910
instrument. The technique was based on the method developed by
Maffucci et al. [28]. The experiment was conducted using 0.15-ml
hydrogen pulses at 150

C to avoid palladium hydride formation.


Nitrogen was used as a carrier gas and hydrogen uptake was mon-
itored using a thermal conductivity detector (TCD). The cobalt
dispersion over ceria was calculated using N
2
Ochemisorption. The
technique was based on the method developed by Jensen et al.
[29]. The experiment was carried out at 40

C in the presence
of 4% N
2
O/He. The nitrogen evolved during the experiment was
monitored using an MKS-Cirrus mass spectrometer. The observed
palladium dispersion over zirconia was 60% whereas the cobalt
dispersion over ceria support was around 20%, irrespective of the
metal loading. The details of surface area analysis and dispersion
studies can be found in our previous articles [21,27].
X-ray diffraction experiments over the reduction and oxidation
catalyst were conducted using either Bruker D8 Advanced diffrac-
tometer or Rigaku X-ray diffractometer both equipped with Cu
K radiation source (=1.5418

A). Pd/SZ showed the presence of
both monoclinic and tetragonal zirconia. The contribution fromthe
t-ZrO
2
was around 33% whereas, crystallize sizes of t-ZrO
2
and m-
ZrO
2
were 11.9nmand 13.7nm, respectively [21]. X-ray diffraction
analysis over 10% CoO
x
/CeO
2
and 2% CoO
x
/CeO
2
showed the pres-
ence of cerianite structure and cobalt was present in the form of
Co
3
O
4
[27].
Diffuse Reectance Infrared Fourier Transform Infrared Spec-
troscopy (DRIFTS) was performed over individual reduction (0.3%
Pd/SZ) and oxidation (2% CoO
x
/CeO
2
) catalysts and a dual-catalyst
bed (0.3% Pd/SZ: 2% CoO
x
/CeO
2
=8:1) during methane tempera-
ture programmed desorption (CH
4
-TPD). The DRIFTS instrument
(Thermo Nicolet 6700) was equipped with a liquid nitrogen-cooled
mercurycadmiumtelluride (MCT) detector and a DRIFTS cham-
ber with ZnSe window. The DRIFTS spectra were collected in the
mid-IR range with a resolution of 4cm
1
averaged over 500 scans.
Samples were pre-treated in situ at 500

C in 10% O
2
/He for 30min
at a owrate of 30ccm, followed by heliumush at the same tem-
perature for an additional 30min. The background spectra were
collected while cooling, under a He ow, at regular temperature
intervals. Adsorption of 10% CH
4
/He was performed for 30min at
50

C followed by He ush for 30min at the same temperature, to


remove physisorbed CH
4
. Finally, the spectra were acquired, in He,
at 50

C increments up to a maximumtemperature of 500

C.
2.3. Catalyst activity testing
Steady-state reaction experiments were conducted at ambient
pressure in a

O.D. stainless steel xed bed reactor system. A


physical mixture of reduction and oxidation catalysts were packed
in desired quantities inside the reactor and held in place using
quartz wool plugs. The reactor bed temperature was monitored
and controlled using an Omega K-type thermocouple, which was
upstreamof the catalyst bed and an Omega (model CS232) PIDcon-
troller. The reactor was placed upstream in a resistively heated
P. Gawade et al. / Catalysis Today 197 (2012) 127136 129
homemade furnace. Brooks mass ow controllers (5850E) were
used to regulate the gas ows to the reactor system. Water vapor
was introduced to the reactor by saturating a stream of helium
gas through a heated water bubbler containing de-ionized water. A
chemiluminescence NO
x
analyzer (Thermo-Scientic 42i-HL) was
coupled with a micro-GC (Agilent 3000A) equipped with 0.32mm
mol-sieve and PLOT Qcolumns with thermal conductivity detector
to analyze the product gases. NO
2
yield and conversions of hydro-
carbons (X
CHx
) and NO
x
(X
NOx
) were calculated using the following
relationships:
NO
2
yield =
moles of NO
2
produced
moles of NO fed
(X
CHx
) or (X
NOx
) =
[moles in] [moles out]
moles in
2.3.1. NO oxidation over CoO
x
/CeO
2
The catalyst samples were pre-treatedin10%O
2
/He (40ccm) for
30min at 450

C prior to catalytic activity testing. Steady-state NO


oxidation reactions were performed on CoO
x
/CeO
2
with different
cobalt loadings, namely 2and10%Co. These experiments were con-
ducted in the temperature range of 200400

C and at a gas hourly


space velocity (GHSV) of 100,000h
1
. The feed composition was
made up of 1000ppmNO, 10% O
2
and balance He.
2.3.2. NO
x
reduction over the dual catalyst bed: effect of
reduction-to-oxidation catalyst ratio
All the NO
x
reduction studies using hydrocarbon mixtures were
carried out at a GHSV of 32,000h
1
and under the following simu-
lated engine exhaust composition: 180ppm NO
2
, 1737ppm CH
4
,
208ppm C
2
H
6
, 104ppm C
3
H
8
, 650ppm CO, 6.5% CO
2
, 10% O
2
,
(010%) water vapor and balance He, unless otherwise stated. The
effect of the ratio of reduction to oxidation catalyst in the dual-
catalyst bed was evaluated in the temperature range of 300500

C
for NO
x
reduction performance. The ratio of reduction (0.3% Pd/SZ)
to oxidation (10% CoO
x
/CeO
2
) catalyst was varied from2:1, 4:1 and
8:1 (by wt) in the dual-catalyst bed. In order to keep the bed vol-
ume constant during reaction studies, quartz powder was mixed
with the catalyst bed. The ndings of these experiments (Section
3.2) revealed that lower amounts of oxidation catalyst in the dual-
bed assisted the overall NO
x
conversion. Therefore, the amount of
oxidation catalyst in the dual-catalyst bed was further decreased
using a lower cobalt-loading sample (2% CoO
x
/CeO
2
) and a sim-
ilar experiment was performed under simulated engine exhaust
conditions to evaluate its NO
x
reduction performance.
2.3.3. Hydrothermal stability of the dual-catalyst bed: effect of Co
loading
The hydrothermal stability of the dual-catalyst bed was evalu-
ated in the presence of 7% water vapor at 450

C under simulated
engine exhaust conditions as outlined above. Oxidation catalysts
with different cobalt loadings (10% CoO
x
/CeO
2
or 2% CoO
x
/CeO
2
)
were mixed with 0.3% Pd/SZ giving a reduction to oxidation cata-
lyst ratio of 8:1 (by wt). The cyclic experiments were conducted in
order to understand the reversibility of the water effect on the dual
catalyst bed.
2.3.4. Kinetics of hydrocarbon oxidation
The oxidation kinetics of the hydrocarbon mixture (methane,
ethane and propane) over 0.3% Pd/SZ, 2% CoO
x
/CeO
2
and the
dual catalyst bed (Pd/SZ:2% CoO
x
/CeO
2
=8:1) was studied in the
typical temperature range of 300500

C with10% O
2
. Methane,
ethane and propane concentrations were varied in the ranges of
8892583ppm, 107311ppm and 53155ppm, respectively. The
reactor was treated either as a differential or an integral reactor
Fig. 1. NO
2
yield during NO oxidation over 2% CoOx/CeO
2
() and 10% CoOx/CeO
2
(). Reaction conditions: 1000ppm NO, 10% O
2
; 1atm; GHSV: 100,000h
1
, ()
thermodynamic equilibriumconversion of NO to NO
2
.
depending upon the fractional conversion of hydrocarbons. It is
worthwhile to note that the presented oxidation kinetic study does
not include extensive experiments to examine the inhibition effect
of water. It is expected that the presence of signicant amount of
water vapor (710%) in engine exhaust would inhibit the hydrocar-
bon oxidation. This inhibition phenomenon was observed on our
dual-catalyst bed in the separate study (not shown). In these nd-
ings, hydrocarbonoxidationwas reversibly affectedinthe presence
of water vapor.
3. Results and discussion
3.1. NO oxidation over CoO
x
/CeO
2
It is known that NO
2
participates in the activation of hydrocar-
bons [30,31], therefore the efciency of the oxidation catalyst to
perform oxidation of NO to NO
2
is fundamental to the selective
reduction of NO to N
2
. NO may exist in the engine exhaust, but it
can also formthrough the partial reduction of NO
2
during NO
2
SCR,
andit is necessary toconvert it back intoNO
2.
Therefore NO-to-NO
2
oxidationover CoO
x
/CeO
2
withdifferent Co loadings was evaluated
under steady state conditions, as illustrated in Fig. 1. As a refer-
ence, the thermodynamic equilibrium limitation (dashed line) for
theconversionof NOtoNO
2
is alsoshownfor thechosenfeedcondi-
tions. Anincrease inNO
2
yieldwas observedwithanincrease inthe
reaction temperature upon going from 200

C to 300

C. However
a further increase in temperature results in a decrease in the NO
2
yield, which may be due to thermodynamic limitation imposed by
the systemat higher temperatures. As can be observed, both sam-
ples reach equilibrium at approximately 300

C, with a maximum
conversion of c.a. 83%. At higher temperatures, conversions closely
follow the equilibrium curve. These results suggest that the NO
2
yield is largely independent of cobalt loading on the ceria support.
These results differ fromthose previously reported for cobalt load-
ings on other supports, namely TiO
2
and ZrO
2
, where it was found
that increasing the cobalt concentration had a signicant effect on
the NO
2
yield [25]. The use of CeO
2
as a support appears to enhance
catalytic activity for NO conversion with reduced cobalt loadings.
This may be a direct result of the high oxygen mobility in CeO
2
at
elevated temperatures [32] and [references within]. It should also
be notedthat the mixture of oxidationcatalyst withreductioncata-
lyst in a dual-catalyst bed is capable of pushing the thermodynamic
130 P. Gawade et al. / Catalysis Today 197 (2012) 127136
Fig. 2. (a) NOx conversionand(b) CH
4
conversionover dual catalyst bedas afunction
of Pd/SZ10%CoOx/CeO
2
ratios, 1:0(), 2:1( ), 4:1( ) and8:1( ) under simulated
lean exhaust conditions. Reaction conditions: [NO
2
] =180ppm, [CH
4
] =1737ppm,
[C
2
H
6
] =208ppm, [C
3
H
8
] =104ppm, [CO] =650ppm, [CO
2
] =6.5%, [O
2
] =10%; 1atm;
GHSV: 32,000h
1
.
equilibriuminthe forwarddirectiondue to the continuous removal
of NO
2
during SCR with hydrocarbons.
3.2. NO
x
reduction over the dual catalyst bed: effect of
reduction-to-oxidation catalyst ratio
In the previous section, it was found that NO oxidation appears
to be independent of the oxidation catalyst employed. However,
the oxidation catalyst not only plays a role in NO oxidation, but
also in hydrocarbon and carbon monoxide oxidation. The extent to
which these reactions occur over the oxidation component may
eventually affect the overall NO
x
reduction performance in the
dual bed scheme. In the dual-catalyst bed, the oxidation catalyst
plays multi-functional roles by removing carbon monoxide and
un-burned hydrocarbons, which have not participated in the NO
x
reduction. Also NO formed due to the partial reduction of NO
2
can
bere-oxidizedusingtheoxidationcatalyst [22]. Thereforetheeffect
of the reduction-to-oxidation catalyst ratio in the dual-catalyst
bed was investigated under the simulated lean exhaust condi-
tions. Fig. 2a and b shows NO
x
and CH
4
conversion as a function
of reaction temperature, with various dual-catalyst bed reduction-
to-oxidation catalyst ratios. It is evident fromthese gures that the
lowest NO
x
andCH
4
conversions wereobtainedwhenthereduction
catalyst (Pd/SZ) was tested in the absence of the oxidation cata-
lyst. This result clearly indicates the importance the presence of
CoO
x
/CeO
2
has in the envisioned dual catalyst bed system.
Fig. 3. NOx (bold) and CH
4
conversion (void) over Pd/SZ: CoOx/CeO
2
=8:1 at differ-
ent cobalt loadings; 2%CoOx/CeO
2
(, ) and10%CoOx/CeO
2
(, ) under simulated
lean exhaust. Reaction conditions: 180ppm [NO
2
] =180ppm, [CH
4
] =1737ppm,
[C
2
H
6
] =208ppm, [C
3
H
8
] =104ppm, [CO] =650ppm, [CO
2
] =6.5%, [O
2
] =10%; 1atm;
GHSV: 32,000h
1
.
NO
x
conversion (Fig. 2a) was found to increase with increasing
the reduction-to-oxidation catalyst ratio (Pd/SZ:10% CoO
x
/CeO
2
),
from2:1 to 8:1 whereas, the opposite trend was observed for CH
4
conversion (Fig. 2b). These results would suggest that having the
oxidation catalyst in smaller quantities in the dual catalyst bed
could favor the NO
x
conversion at the expense of hydrocarbon
conversion. It may be expected that there is a lower hydrocarbon
combustion rate over the 2% CoO
x
/CeO
2
oxidation catalyst com-
pared to the 10% CoO
x
/CeO
2
sample, which ultimately results in
the presence of a higher hydrocarbon concentration in the reaction
medium to allow the NO
x
reduction to take place, consequently
increasing the NO
x
conversion. C
2
H
6
and C
3
H
8
conversion trends
were similar to that observed for CH
4
conversions (not shown). In
spite of this, the combustion rates for C
2
H
6
and C
3
H
8
were much
higher than CH
4
, reaching 100% conversion at temperatures above
350

C, indicating that the complete oxidation of C


2
H
6
and C
3
H
8
was much easier than that of CH
4
. Furthermore, complete carbon
monoxide conversion was observed regardless of the dual-catalyst
bedratioemployed(not shown). These results indicate that the oxi-
dation of CO to CO
2
occurs readily over the oxidation catalyst and
the percentage of metal loading on the support does not appear to
have a signicant effect on this process. A separate study was con-
ducted, (not shown), to evaluate the effect of the presence of CO,
CO
2
and higher hydrocarbons on NO
2
reduction, as well as hydro-
carbon oxidation. However, the ndings of this study are beyond
the scope of the current publication.
It was observed, from Fig. 2, that by lowering the amount of
oxidation catalyst (10% CoO
x
/CeO
2
) in the dual-catalyst bed, the
NO
x
conversion was improved. The oxidation catalyst amount in
the dual-catalyst bed was further decreased by using the lower
cobalt-loading sample (2% CoO
x
/CeO
2
). Fig. 3 shows NO
2
SCR per-
formance over the dual-catalyst bed (Pd/SZ:2% CoO
x
/CeO
2
=8:1). In
this case, NO
x
conversion was further increased due to the lower
hydrocarbon oxidation and hence having a higher concentration of
hydrocarbons available to participate in the NO
2
reduction.
3.3. Hydrothermal stability of the dual catalyst bed: effect of Co
loading
In a previous study, we demonstrated that the presence of
water in the engine exhaust signicantly affects NO
x
reduction
performance [20,21]. It was reported that the reduction catalyst,
Pd/SZ, was considerably affected by the presence of water whereas
the inhibition effect on the oxidation catalyst (Co/ZrO
2
) was not
P. Gawade et al. / Catalysis Today 197 (2012) 127136 131
Fig. 4. Reversibility of the effect of water vapor on (a) NOx and (b) CH
4
con-
version over Pd/SZ:CoOx/CeO
2
=8:1 at different cobalt loadings 2% CoOx/CeO
2
(, ) and 10% CoOx/CeO
2
(, ) under simulated lean exhaust. Reaction con-
ditions: [NO
2
] =180ppm, [CH
4
] =1737ppm, [C
2
H
6
] =208ppm, [C
3
H
8
] =104ppm,
[CO] =650ppm, [CO
2
] =6.5%, [O
2
] =10%, [H
2
O] =0% or 7%; 450

C; 1atm; GHSV:
32,000h
1
.
as apparent. Furthermore, temperature programmed desorption
(TPD) studies, in the presence of water vapor, revealed that water
vapor negatively affects the NO adsorption over Pd/SZ whereas
NO
2
adsorption over Pd/SZ was unaffected. This suggests that the
choice of oxidation catalyst in a dual bed systemcould potentially
reduce the negative effect of water vapor on the dual catalyst bed,
to improve the overall NO
x
reduction performance [20,21].
In line with above discussion and the ndings illustrated in
Figs. 2 and 3, the oxidation catalyst is likely to be associated
with improving the hydrothermal stability. Therefore, steady state
experiments were conducted with different cobalt loadings in the
dual-catalyst bed at 8:1 (by wt) reduction-to-oxidation catalyst
ratio. Simulated engine exhaust composition was used as a feed,
as described in the experimental section, and the reaction experi-
ments were performed at 450

C in the presence of 7% water vapor.


Fig. 4aandbshows NO
x
andCH
4
conversionduringthecyclic exper-
iment in the presence and absence of water vapor. It was observed
that the effect of water vapor was reversible in both dual-catalyst
bed samples, as the complete recovery of NO
x
and CH
4
conversions
was observed upon the removal of water vapor fromthe feed.
For the dual-catalyst bed (Pd/SZ:10% CoO
x
/CeO
2
), upon going
from dry to wet feeds streams the NO
x
and CH
4
conversions
dropped from 80% to 65% and 76% to 3032%, respectively. The
complete recovery of NO
x
and CH
4
conversion was observed upon
removal of the water vapor. A similar cyclic trend was observed
over the Pd/SZ:2% CoO
x
/CeO
2
dual-catalyst bed. NO
x
and CH
4
con-
versions dropped from 94% to 73% and 53% to 22%, respectively.
It is apparent that the dual-catalyst bed sample containing 2%
CoO
x
/CeO
2
showed better NO
x
conversion compared to that con-
taining 10% CoO
x
/CeO
2
whereas the hydrocarbon conversion trend
was opposite to NO
x
conversion. The higher NO
x
conversion and
lower hydrocarbon conversion in 2% CoO
x
/CeO
2
containing dual-
catalyst bed compared to 10% CoO
x
/CeO
2
containing dual-catalyst
bed could be due to lower hydrocarbon oxidation, and hence more
hydrocarbons being available for NO
2
reduction. A similar trend
was observed for higher hydrocarbons (not shown). Both C
2
H
6
and
C
3
H
8
conversions were adversely affected, but also reversible, in
the presence of water vapor. However, conversions of both C
2
H
6
and C
3
H
8
were more than 90%even the presence of water at 450

C,
while the CO conversion was 100% in the presence of water indi-
cating that CO oxidation was unaffected during the wet cycle.
AnoticeabledecreaseinCH
4
conversioninthepresenceof water
indicated that CH
4
combustion was signicantly affected by water
vapor. The inhibition effect of water on CH
4
combustion is reported
in several studies in the past over Pd-based catalysts [3335]. It
is believed that either the formation of Pd(OH)
2
phase, which is
inactive for CH
4
combustion [33], or water inhibition during CH
4
oxidation [34,35] could result in a decrease of CH
4
conversion in
the presence of water vapor.
It is worthwhile to note that time-on-streamexperiments have
been performed over the dual-catalyst bed in the presence of 10%
water (not shown). Theresults shownosignicant activityloss after
the 1520h under optimized conditions.
3.4. Kinetics of hydrocarbon oxidation
Hydrocarbon oxidation plays an indirect role in determining
NO
2
SCR performance. It is desirable to remove hydrocarbons from
the engine exhaust via combustion. However, only the remain-
ing hydrocarbons, which have not reacted during NO
2
SCR, should
be used for the oxidation process. Unselective direct oxidation of
hydrocarbons could result in a considerable drop in NO
2
SCR due to
lackof available hydrocarbons for the NO
2
SCRreaction. Inthe dual-
catalyst bed, it was expected that the oxidation catalyst would play
a major role in determining the extent of hydrocarbon oxidation;
thereby indirectly affecting the NO
2
SCR. Therefore kinetic stud-
ies of hydrocarbon oxidation were conducted to understand the
role of the oxidation catalyst, individually and in a dual-catalyst
bed composition. Figs. 57 show the fractional hydrocarbon con-
versions (CH
4
, C
2
H
6
and C
3
H
8
) in the temperature range between
300

C and 500

C over 2% CoO
x
/CeO
2
, Pd/SZ and dual-catalyst bed,
respectively. The reactor was treated either as a differential or an
integral depending upon the fractional conversion of hydrocar-
bons. It should be noted that kinetic parameters for C
3
H
8
oxidation
could not be derived, due to the fact that almost 100% C
3
H
8
con-
version was achieved under the operating conditions regardless of
the choice of the catalyst. This would indicate that the oxidation of
C
3
H
8
occurs more readily compared to both CH
4
and C
2
H
6
.
Fig. 5a shows rate of CH
4
oxidation (mol/gcat. min) as a function
of CH
4
concentration in the temperature range of 400500

C over
2% CoO
x
/CeO
2
. Fractional conversion of CH
4
(X
CH
4
) was less than
0.2, therefore differential reactor kinetics were employed and the
reaction rate was calculated using
r
CH
4
=
X
CH
4
F
CH
4,
0
W
where Wis the catalyst weight (g) and F
CH
4,0
is the initial CH
4
ow
rate (mol/min). The following Power-law model was assumed for
the kinetic analysis:
r
CH
4
= k (C
CH
4
)
a
(C
O
2
)
b
(1)
where C
CH
4
and C
O
2
are concentrations (mol/cm
3
) of CH
4
and O
2
,
respectively.
132 P. Gawade et al. / Catalysis Today 197 (2012) 127136
Fig. 5. (a) Rate of methane oxidation vs. methane concentration and (b) fractional
ethane conversion vs. W/F
A0
over 2% CoOx/CeO
2
. Inset: comparison of power-law
model predictions for hydrocarbon oxidation rate against experimentally deter-
mined rates over 2% CoOx/CeO
2
.
However, the oxygen concentration is approximately 100 times
higher thanthat of the methane concentrationand hence the above
Power-lawmodel could be reduced to
r
CH
4
= k

(C
CH
4
)
a
where k

= k (C
O
2
)
b
(2)
Using a linear regression analysis, the kinetic parameters for
CH
4
oxidation over 2% CoO
x
/CeO
2
were found as activation energy
(E
a
) =7913kJ/mol and reaction order a w.r.t. CH
4
=1.00.3.
The inset of Fig. 5a shows the parity plot comparing calculated
reaction rates (Rate
Cal
) vs. experimental reaction rates (Rate
Exp
).
For the kinetic analysis of C
2
H
6
oxidation over 2% CoO
x
/CeO
2
,
integral reaction kinetics were employed as the fractional conver-
sion of C
2
H
6
(X
C
2
H
6
) was greater than 0.2. Fig. 6b shows the plot of
X
C
2
H
6
vs. W/F
C
2
H
6,0
, where F
C
2
H
6,0
, is the initial owrate of ethane
(mol/min). The reaction rate was calculated by applying the differ-
ential analysis as follows:
r
C
2
H
6
=
dX
C
2
H
6
d(W/F
C
2
H
6
,0
)
(3)
A reduced Power-law model (Eq. (2)) was used for data tting
and linear regression analysis was applied to obtain the kinetic
parameters for C
2
H
6
oxidation over 2% CoO
x
/CeO
2
as followed,
yielding anactivationenergy (E
a
) =453kJ/mol andreactionorder
a w.r.t. C
2
H
6
=0.70.08. The inset in Fig. 5b shows the parity
Fig. 6. (a) Fractional methane conversion vs. W/F
A0
and (b) fractional ethane con-
version vs. W/F
A0
over Pd/SZ. Inset: comparison of power-lawmodel predictions for
hydrocarbon oxidation rate against experimentally determined rates over Pd/SZ.
plot comparing calculated reaction rates (Rate
Cal
) vs. experimental
reaction rates (Rate
Exp
). Kinetic analysis conrmed that the oxida-
tion of C
2
H
6
occurs more readily than the oxidation of CH
4
over 2%
CoO
x
/CeO
2
.
Fig. 6a and b illustrates the inuence of W/F
A0
on CH
4
and C
2
H
6
conversions respectively, during oxidation reaction over only the
reduction catalyst, Pd/SZ. The reactor was treated as an integral
reactor and reaction rates were obtained using Eq. (3). A reduced
Power-law model (Eq. (2)) was used for data tting and linear
regression analysis was performed to obtain the kinetic param-
eters for CH
4
and C
2
H
6
oxidation over Pd/SZ. These results have
been summarized in Table 1. Activation energies for CH
4
and
C
2
H
6
oxidation over Pd/SZ were 9514kJ/mol and 5023kJ/mol,
respectively, which was in agreement with results found by Ribeiro
et al. [35] where the kinetic study of methane oxidation over a
Pd supported catalyst was reported. The higher activation energy
barrier for hydrocarbon oxidation over Pd/SZ compared to 2%
CoO
x
/CeO
2
indicates that the direct oxidation of hydrocarbons over
Pd/SZis difcult toachieve. As will bediscussedinSection3.5, Pd/SZ
takes part primarily in the activation of CH
4
rather than its direct
oxidation. Here, CH
4
is activated via hydrogen abstraction to form
CH
3
or CH
2
species over Pd/SZ, insteadof being directly oxidized
to formCO
2
.
P. Gawade et al. / Catalysis Today 197 (2012) 127136 133
Table 1
Summary of kinetic parameters during CH
4
and C
2
H
6
oxidation over oxidation, reduction and dual catalyst bed catalyst.
CH
4
oxidation C
2
H
6
oxidation
Ko a E (kJ/mol) Ko a E (kJ/mol)
2% CoOx/CeO
2
2.7610
8
1.0 0.3 79 13 6.210
3
0.7 0.08 45 3
Pd/SZ 1.6110
7
0.7 0.19 95 14 7.510
2
0.6 0.17 50 23
Pd/SZ +2% CoOx/CeO
2
6.210
3
0.6 0.06 66 5 1.49 0.4 0.12 35 11
Fig. 7a and b shows a similar kinetic study of CH
4
and C
2
H
6
oxidation over a dual-catalyst bed sample (Pd/SZ:2% CoO
x
/CeO
2
).
The synergetic effect between reduction and oxidation catalysts
signicantly decreases the activation energy barrier. The activa-
tion energies for CH
4
and C
2
H
6
oxidation over dual catalyst bed
were found to be 665kJ/mol and 3511kJ/mol, respectively.
This observation is supported using DRIFTS data presented in Sec-
tion 3.5, where it was observed that the oxidation of CH
4
was more
achievable over 2% CoO
x
/CeO
2
after its initial activation by Pd/SZ.
3.5. Investigation of surface species using DRIFTS during CH
4
-TPD
The investigation of surface species formed as a result of
CH
4
adsorption was examined using in situ DRIFTS. Fig. 8 shows
Fig. 7. (a) Fractional methane conversion vs. W/F
A0
and (b) fractional ethane con-
version vs. W/F
A0
over Pd/SZ:2% CoOx/CeO
2
=8:1. Inset: comparison of power-law
model predictions for hydrocarbon oxidation rate against experimentally deter-
mined rates over Pd/SZ:2% CoOx/CeO
2
=8:1.
in situ DRIFT spectra collected during methane temperature
desorption (TPD) over pre-oxidized 2% CoO
x
/CeO
2
, Pd/SZ and
a 1:8 dual-catalyst bed consisting of the two afore-mentioned
catalysts. Fig. 8a illustrates the higher wavenumber region of
20003800cm
1
while Fig. 8b depicts the lower wavenumber
region of 9002000cm
1
. Band assignments for the respective
samples are summarized in Table 2.
The DRIFT spectra obtained fromboth the dual-catalyst bed and
reduction catalyst exhibited distinct bands, which were absent in
the 2%CoO
x
/CeO
2
spectra. Inbothinstances, a large negative peakis
evident at 1392cm
1
, which is indicative of sulfated zirconia and
is attributed to the asymmetric S O stretching mode [36,37]. An
additional broad band at c.a. 1250cm
1
can be assigned to S O
vibrations. Additional bands canalsobeattributedtosulfatespecies
on both Pd/SZ and dual-catalyst bed samples as shown by the nega-
tive band centered at c.a. 1175cm
1
which becomes more negative
with increasing temperature. This band corresponds to the SO
4
2
ionand symmetric stretchof O S O[38,39]. Inall catalysts a broad
Table 2
Summary of DRIFTS study during CH
4
-TPD over oxidation, reduction and dual cat-
alyst bed catalyst.
Wavelength (cm
1
)
2%
CoOx/CeO
2
Pd/SZ Dual catalyst
bed
CH stretch 2839 2835
2- (C H) 2704 2704
CH
2
stretch 2916 2918
CH species fromformate 2727 2727
CH
3
asymmetric stretch
Deformation mode
2958
1448
Gas phase CH
4
1302 1304
OH bands Type I
Type II
3720
3615
3730
3633

3622
Physically adsorbed H
2
O 36003000 36003000 36003000
Molecularly adsorbed
water
3649 3700 3649/3701
3510 3510
Bending vibration (HOH) 1638 1620 1620
Acidic terminal OH 2272
CO linear 2035 2021
Bridged 1846 1846
1130 1130
CO
2
2339/2316 2343/2330 2343/2320
O S O asymmetric 1392 1392
Symmetric 1176 1174
v (S O) 1246 1252
Bidentate carbonates 1563
1363
Monodentate carbonates 1502 1506
1331 1326
Bidentate formates 1543 1540 1540
1336
Carboxylates 1563 1559
1518
1302 1304 1304
Hydrogen carbonate 1211
1417 1427
1049 1047 1047
1022 1024 1026
134 P. Gawade et al. / Catalysis Today 197 (2012) 127136
Fig. 8. In situ DRIFT spectra collected during CH
4
-TPD over (i) 2% CoOx/CeO
2
, (ii) Pd/SZ and (iii) 1:8 dual-catalyst bed 2% CoOx/CeO
2
PdSZ; (a) high wavenumber region and
(b) lowwavenumber region.
band is apparent at around 16101630cm
1
, assigned to the bend-
ing vibration of molecularly adsorbed water (HOH). This band
becomes more intense over the Pd/SZ and the dual-catalyst bed,
signaling an increase in water formation with increasing temper-
ature. This increase coincides with the S O and O S O features
becoming more negative, which suggests a strong interaction of
water with the surface sulfate species on sulfated zirconia sup-
port, as reported previously [2224,20,4042]. A weak shoulder at
1302cm
1
was observed on the CoO
x
/CeO
2
oxidation catalyst. This
is a characteristic band of adsorbed methane, which is also evi-
denced in the reduction catalyst (1304cm
1
), but to a lesser extent
[43]. This is perhaps not surprisinginviewof the fact that Pdspecies
are known to promote H-abstraction, resulting in methane being
converted to methyl species [44].
From Fig. 8a, it is evident that all three samples have well
resolved bands in the hydroxyl region. Two negative bands at 3720
and3615cm
1
for CoO
x
/CeO
2
andat 3730and3633cm
1
for Pd/SZ,
were identiedas type I andtype II OHgroups, respectively [45,46].
These bands were not as well resolved in the dual-catalyst bed
sample, however the negative band at 3622cm
1
can be assigned
to type II OH vibrations [45]. These negative bands may repre-
sent interactions between surface hydroxyl species and adsorbed
methane. In addition to the negative features, well resolved OH
bands, due to molecularly adsorbed water, were also observed at
around 3649cm
1
and 3510cm
1
in the case of the CoO
x
/CeO
2
,
dual-catalyst bed samples and 3700cm
1
in the Pd/SZ sample.
Chen et al. [43] have previously reported similar bands occuring
in the hydroxyl region due to the adsorption of methane on a
ZSM-5 catalyst. It was reported that the strong band at 3510cm
1
may be due to a shift of the band at c.a. 3620cm
1
as a result of per-
turbationof surfacebridginghydroxyl groups byadsorbedmethane
[43]. A broad positive band was observed, for all three sam-
ples, in the range of 36003000cm
1
owing to weakly H-bonded
hydroxyl groups on the support and is characteristic of OH vibra-
tions associated with physically adsorbed H
2
O[40]. Both Pd/SZ and
dual-catalyst bed samples exhibited additional OHbands. The neg-
ative peak evident at 2272cm
1
can be attributed to terminal OH
species interacting with surface lewis acid sites on the support.
Bidentate formate species were evidenced in both CoO
x
/CeO
2
and dual-catalyst bed samples. The band observed at approxi-
mately 2835cm
1
was attributed to surface CH species while the
band at 2918cm
1
could have contributions from CH
2
stretch-
ing vibrations and
as
(OCO) vibrations. An additional small peak
was also observed at 2704cm
1
due to the overtone vibration of
2(C H) [45,47,48]. These peaks were not observed in the DRIFT
spectra of Pd/SZ; however bands arising from a CH
3
asymmetric
stretch and CH
3
deformation modes were apparent at 2958cm
1
and 1448cm
1
, respectively. A small weak feature at 2727cm
1
was also evident in both the Pd/SZ and dual-catalyst bed samples
(Fig. 8a(ii) and (iii)), which may be ascribed to the asymmetric C H
stretch of surface formate species [49,50].
The IR band observed at 2035cm
1
in spectra over Pd/SZ and
the dual-catalyst bed has previously been reported as correspond-
ingtobridged-bondedCOover reducedPd(110) [5153]. This peak
decreases with increasing temperature, however the development
of additional IR bands at 1846cm
1
and 1130cm
1
, above 250

C,
P. Gawade et al. / Catalysis Today 197 (2012) 127136 135
is observed. These peaks may be attributed to multi-coordinated
forms of CO on Pd
0
[42,54,55]. This type of carbonyl formation
and adsorption on Pd may possibly be explained by the reaction
of gasous CH
4
with the pre-oxidized surface of the Pd particles.
Pd/SZ catalyst, and to a weaker extent, the dual catalyst bed show
negative peaks around 2343 and 2320cm
1
, corresponding to CO
2
doublet. This suggests some adsorbed CO
2
species on the surface,
which leads to negative bands due to background subtraction.
These features become more positive over the dual catalyst bed
following a similar trend that is observed with temperature over
the CoO
x
/CeO
2
, signaling formation of CO
2
. However no evidence
of CO
2
formation was observed in the CH
4
-TPD reaction over the
reduction catalyst alone (Pd/SZ). These observations would sug-
gest that CH
4
is only partially oxidized to CO in the presence of the
reduction catalyst. However it may be postulated that at temper-
atures above 250

C the mobility of oxygen atoms from the CeO


2
support, in both the dual catalyst bed and oxidation catalyst itself,
is sufciently high to enable re-oxidation of the catalyst surface
allowing the complete oxidation of CO and CH
4
to occur.
The interaction between CO and/or CO
2
with surface methyl
species leads to the formation of various adspecies including,
but not limited to, carbonate and formate type intermediates
chemisorbed on the support. For this reason, the region between
1600 and 1300cm
1
is particularly difcult to interpret, which
is perhaps more apparent in the spectra of CoO
x
/CeO
2
and dual-
catalyst bed samples. The most intense bands in this zone can be
ascribed to bidentate carbonates at c.a. 1563 and 1363cm
1
which
are particularly evident inthe lower temperature spectra, Fig. 8b(i).
As the temperature is increased above 200

C, the intensity of
these bands decrease at the expense of bands at approximetaly
1540cm
1
and 1336cm
1
. These features are characteristic of
bidentate formate species [45] and correspond to asymmetric and
symmetric v(oco) modes. The intensityof these bands increase with
increasing temperature and is most likely a result of the interac-
tion between OH and CO species. These bands, however, are not
observed in the Pd/SZ or dual catalyst bed spectra (Fig. 8b(ii) and
(iii)), and are most likely masked as a result of the strong negative
band associated with sulfated zirconia. The positive peak centered
at approximately 1130cm
1
in both Pd/SZ and dual-catalyst bed
samples can be attributed to monodentate carbonate species. This
is conrmed by the weak band at 1502cm
1
and 1506cm
1
for
the respective samples, which can also be assigned to monoden-
tate carbonates [55]. The presence of carboxylate species over all
samples cannot be ruled out due to the presence of a peak at c.a.
15001510cm
1
. This band is usually accompanied by a strong
peak around c.a. 1560cm
1
, which is well evident in CoO
x
/CeO
2
.
However, the band at 1560cm
1
is not well resolved for Pd/SZ or
the dual catalyst bed due to the presence of molecularly adsorbed
water at 1620cm
1
.
The band observed at 1211cm
1
in the spectra of CoO
x
/CeO
2
,
along with those at c.a. 1420cm
1
, which are also evidenced in
the high temperature dual-catalyst bed spectra, are attributed to
hydrogen carbonate species. The presence of a strong band due
to adsorbed water makes it difcult to resolve a possible band
around 1608cm
1
, which would be expected in hydrogen carbon-
ate species. Bands at 1047 and 1024cm
1
may also be attributed to
hydrogen carbonate species [45], however it may also be possible
that these bands correspond to
S O
stretching mode(s), particu-
larly in the dual catalyst bed and Pd/SZ samples [38,56].
By studying the adsorption of CH
4
on both the reduction and
oxidation catalysts independently, it is evident that each serves a
different role in the mixed-bed system. From Fig. 8 it is clear that
in the spectra of both the oxidation catalyst and the dual cata-
lyst bed catalyst, bands attributed to surface CH and CH
2
species
are present. These bands are not, however, observed in the spec-
tra obtained from the Pd/SZ reduction catalyst. What is apparent
though is the presence of CH
3
stretching and deformation modes.
This wouldsuggest that thePdreductioncatalyst promotes theC H
bond dissociation in CH
4
, ultimately activitating CH
4
and allowing
for its further oxidation. This can be illustrated in the spectra of the
dual catalyst bed sample, Fig. 8a(i) and b(i). In these spectra there is
again no evidence of CH
3
bands, but only CHand CH
2
, which would
imply that the methyl species that are generated are immediately
oxidized to either carbonate or formate species. These, in turn, are
further oxidized to CO
2
, as evidenced by the CO
2
doublet in spectra
of both the dual catalyst bed and oxidation catalyst.
This study helps to elucidate the roles of both the oxidation and
reduction components of a dual-catalyst system. It is also clear
from this study that there is a synergystic effect in the dual cat-
alyst bed, in that the Pd/SZ reduction catalyst dissociates a C H
bond in methane and these CH
x
species subsequently participate
in NO
x
reduction or directly oxidized by the oxidation catalyst. Sec-
ondly it is evident that formation of CO
2
is over the CoO
x
/CeO
2
oxidation catalyst and the formation of CO over Pd/SZ may be a
result of surfacehydroxyl species interactingwithmethaneor other
hydrocarbons.
4. Conclusions
The investigated dual-catalyst system has been shown to
achieve high NO
x
conversions during CH
x
-SCR in lean burn con-
ditions, which can be further improved by optimizing the ratio of
reduction-to-oxidation catalyst in the bed composition. The choice
of cobalt loading in the oxidation catalyst is shown to play a crucial
role in determining the overall performance of the dual-catalyst
bed. The lower cobalt loading in the oxidation catalyst in a dual-
catalyst bed led to a higher NO
x
conversion. This improvement
in NO
x
conversion could be associated with lower hydrocarbon
oxidation and hence increased availability of hydrocarbons for
NO
x
reduction. A similar trend was also observed in the pres-
ence of water vapor. Furthermore, the water inhibition effect was
reversible on NO
x
reduction and CH
x
oxidation regardless of the
choice of oxidation catalyst used. Both the kinetic and DRIFTS stud-
ies conrmed that the reduction catalyst, Pd/SZ activated the CH
4
molecule via hydrogen abstraction to form CH
3
and CH
2
species,
which subsequently either participate in NO
x
reduction or directly
oxidized by the oxidation catalyst.
Acknowledgements
The nancial support provided by the US Department of Energy
and Caterpillar Inc. is gratefully acknowledged. The authors also
thank Dr. Ronald Silver of Caterpillar Inc. for the invaluable insight
he provided through many detailed discussions.
References
[1] S. Roy, A. Baiker, Chemical Reviews 109 (2009) 40544091.
[2] R.M. Heck, Catalysis Today 53 (1999) 519523.
[3] P. Gelin, M. Primet, Applied Catalysis B: Environmental 39 (2002) 137.
[4] V.I. Parvulescu, P. Grange, B. Delmon, Catalysis Today 46 (1998) 233316.
[5] R. Burch, J.P. Breen, F.C. Meunier, Applied Catalysis B: Environmental 39 (2002)
283303.
[6] M. Iwamoto, H. Hamada, Catalysis Today 10 (1991) 5771.
[7] M. Iwamoto, H. Yahiro, Catalysis Today 22 (1994) 518.
[8] M.D. Amiridis, T. Zhang, R.J. Farrauto, Applied Catalysis B: Environmental 10
(1996) 203227.
[9] C. Descorme, P. Gelin, C. Lecuyer, M. Primet, AppliedCatalysis B: Environmental
13 (1997) 185195.
[10] C. Descorme, P. Gelin, C. Lecuyer, M. Primet, Journal of Catalysis 177 (1998)
352362.
[11] J. Suzuki, S. Matsumoto, Topics in Catalysis 28 (2004) 171176.
[12] J.A.Z. Pieterse, H. Top, F. Vollink, K. Hoving, R.W. Brink, Chemical Engineering
Journal 120 (2006) 1723.
[13] H. Ohtsuka, Applied Catalysis B: Environmental 33 (2001) 325333.
136 P. Gawade et al. / Catalysis Today 197 (2012) 127136
[14] M. Misono, Y. Nishizaka, M. Kawamoto, H. Kato, Studies in Surface Science and
Catalysis 105 (1997) 15011508.
[15] Y. Nishizaka, M. Misono, Chemistry Letters 22 (1993) 12951298.
[16] A. Ali, W. Alvarez, C.J. Loughran, D.E. Resasco, Applied Catalysis B: Environmen-
tal 14 (1997) 1322.
[17] H. Kato, C. Yokoyama, M. Misono, Catalysis Today 45 (1998) 93102.
[18] N. Li, A. Wang, L. Li, X. Wang, L. Ren, T. Zhang, AppliedCatalysis B: Environmental
50 (2004) 17.
[19] Y.-H. Chin, A. Pisanu, L. Serventi, W. Alvarez, D.E. Resasco, Catalysis Today 54
(1999) 419429.
[20] B. Mirkelamoglu, U.S. Ozkan, Applied Catalysis B: Environmental 96 (2010)
421433.
[21] B. Mirkelamoglu, M. Liu, U.S. Ozkan, Catalysis Today 151 (2010) 386394.
[22] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Applied Catalysis B: Environmental 74
(2007) 7382.
[23] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Journal of Molecular Catalysis A:
Chemical 270 (2007) 101111.
[24] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Catalysis Letters 111 (2006) 1926.
[25] M.M. Yung, E.M. Holmgreen, U.S. Ozkan, Journal of Catalysis 247 (2007)
356367.
[26] P. Gawade, B. Mirkelamoglu, U.S. Ozkan, Journal of Physical Chemistry C 114
(2010) 1817318181.
[27] P. Gawade, B. Bayram, A.M.C. Alexander, U.S. Ozkan, Applied Catalysis B: Envi-
ronmental, http://dx.doi.org/10.1016/j.apcatb.2012.06.032, in press.
[28] L. Maffucci, P. Iengo, M. Di Serio, E. Santacesaria, Journal of Catalysis 172 (1997)
485487.
[29] J.R. Jensen, T. Johannessen, H. Livbjerg, Applied Catalysis A: General 266 (2004)
117122.
[30] G. Djega-Mariadassou, M. Boudart, Journal of Catalysis 216 (2003) 89.
[31] G. Djega-Mariadassou, Catalysis Today 90 (2004) 27.
[32] B. Rotavera, A. Kumar, S. Seal, E.L. Petersen, Proceedings of the Combustion
Institute 32 (2009) 811819.
[33] C.F. Cullis, T.G. Nevell, D.L. Trimm, Journal of the Chemical Society, Faraday
Transactions 1: Physical Chemistry in Condensed Phases 68 (1972) 14061412.
[34] K. Fujimoto, F.H. Ribeiro, M. Avalos-Borja, E. Iglesia, Journal of Catalysis 179
(1998) 431442.
[35] F.H. Ribeiro, M. Chow, R.A. Dalla Betta, Journal of Catalysis 146 (1994) 537544.
[36] E. Escalona Platero, M. Penarroya Mentruit, Catalysis Letters 30 (1995) 31.
[37] G. Larsen, E. Lotero, R.D. Parra, L.M. Petkovic, H.S. Silva, S. Raghavan, Applied
Catalysis A: General 130 (1995) 213226.
[38] C. Morterra, G. Cerrato, F. Pinna, M. Signoretto, G. Strukul, Journal of Catalysis
149 (1994) 181.
[39] F. Babou, G. Coudurier, J.C. Vedrine, Journal of Catalysis 152 (1995) 341349.
[40] C. Morterra, G. Cerato, S. Di Ciero, Applied Surface Science 126 (1998) 107.
[41] L.L. Sheu, H. Knozinger, W.M.H. Sachtler, Journal of Molecular Catalysis 57
(1989) 61.
[42] C. Schild, A. Wokaun, Journal of Molecular Catalysis 63 (1990) 223.
[43] L. Chen, L. Lin, Z. Xu, T. Zhang, Q. Xin, P. Ying, G. Li, C. Li, Journal of Catalysis 161
(1996) 107114.
[44] A. Yamaguchi, E. Iglesia, Journal of Catalysis 274 (2010) 5263.
[45] O. Pozdnyakova, D. Teschner, A. Wootsch, J. Kroehnert, B. Steinhauer, H. Sauer,
L. Toth, F.C. Jentoft, A. Knop-Gericke, Z. Paal, R. Schloegl, Journal of Catalysis
237 (2006) 116.
[46] C. Li, Y. Sakata, T. Arai, K. Domen, K.I. Maruya, T. Onishi, Journal of the Chemical
Society, Faraday Transactions 1 85 (4) (1989) 929943.
[47] M.M. Yung, Z. Zhao, M.P. Woods, U.S. Ozkan, Journal of Molecular Catalysis A:
Chemical 279 (2008) 19.
[48] H. Kalies, N. Pinto, G.M. Pajonk, D. Bianchi, Applied Catalysis A: General 202
(2000) 197.
[49] S. Bertarione, D. Scarano, A. Zecchina, V. Johnek, J. Hoffmann, S. Schauermann,
J. Libuda, G. Rupprechter, H.-J. Freund, Journal of Catalysis 223 (2004) 6473.
[50] O.S. Alexeev, D.-W. Kim, B.C. Gates, Journal of Molecular Catalysis A: Chemical
162 (2000) 6782.
[51] R. Hicks, H. Qi, M. Young, R. Lee, Journal of Catalysis 122 (1990) 280.
[52] E. Garwoski, C. Feumi-Jantou, N. Mouaddib, M. Primet, Applied Catalysis A:
General 109 (1994) 277.
[53] A.M. Bradshaw, F. Hoffman, Surface Science 72 (1978) 513.
[54] L.F. Liotta, G.A. Martin, G. Deganello, Journal of Catalysis 164 (1996) 322333.
[55] M. Daturi, C. Binet, J.C. Lavalley, G. Blanchard, Surface and Interface Analysis 30
(2000) 273277.
[56] T. Yamaguchi, T. Jin, K. Tanabe, Journal of Physical Chemistry 90 (1986)
31483152.

You might also like