Вы находитесь на странице: 1из 11

CHEMCATCHEM

FULL PAPERS
DOI: 10.1002/cctc.201300709

Morphology Effect of CeO2 Support in the Preparation,


MetalSupport Interaction, and Catalytic Performance of
Pt/CeO2 Catalysts
Yuxian Gao, Wendong Wang, Sujie Chang, and Weixin Huang*[a]
Pt/CeO2 catalysts with various Pt loadings were prepared by
a conventional incipient wetness impregnation method that
employed CeO2 cubes (c-CeO2), rods (r-CeO2), and octahedra
(o-CeO2) as the support and Pt(NH3)4(NO3)2 as the metal precursor. Their structures and catalytic activities in CO oxidation in
excess O2 and the preferential oxidation of CO in a H2-rich gas
(CO-PROX) were studied, and strong morphology effects were
observed. The impregnated Pt precursor interacts more strongly with CeO2 rods and cubes than with CeO2 octahedra, and
the reduction/decomposition of the Pt precursor impregnated
on CeO2 octahedra is easier than that on CeO2 rods and cubes.
With the same Pt loading, the Pt/o-CeO2 catalyst contains the
largest fraction of metallic Pt, whereas the Pt/c-CeO2 catalyst
contains the largest fraction of Pt2 + species. The reducibility of
pure CeO2 and CeO2 in the Pt/CeO2 catalysts follows the order
r-CeO2 > c-CeO2 > o-CeO2, and the reducibility of CeO2 depends
on the Pt loading for the Pt/c-CeO2 catalysts but not much for
the Pt/r-CeO2 and Pt/o-CeO2 catalysts. The catalytic perfor-

mance of Pt/CeO2 catalysts in both CO oxidation and the COPROX reaction follows the order Pt/r-CeO2 > Pt/c-CeO2 >
Pt/o-CeO2. The Pt0-CeO2 ensemble is more active than the Pt2 +
-CeO2 ensemble in the catalysis of CO oxidation in excess O2.
H2-assisted CO oxidation catalyzed by the Pt/CeO2 catalysts
was observed in the CO-PROX reaction, and the Pt2 + species
and CeO2 with a large concentration of oxygen vacancies constitute the active structure of the Pt/CeO2 catalyst for the COPROX reaction. The effect of the morphology of the CeO2 support in the preparation, metalsupport interaction, and catalytic performance of Pt/CeO2 catalysts can be correlated the exposed crystal planes and surface composition/structure of the
CeO2 support with different morphologies. These results not
only demonstrate that the structure and catalytic performance
of oxide-supported catalysts can be tuned by controlling the
morphology of the oxide support but also deepens the fundamental understanding of CO oxidation reactions catalyzed by
Pt/CeO2 catalysts.

Introduction
Oxides are widely employed as catalyst supports. The surface
composition and surface structure of the oxide support greatly
affects the structure and catalytic performance of oxide-supported catalysts. Crystal planes exposed on the surface of an
oxide particle determine both the surface composition and the
surface structure. According to Wulffs rule,[1] crystal planes exposed on the surface of a crystalline material are determined
by its morphology. Thus the morphology of an oxide support
is an important macroscopic structural factor that affects the
catalyst structure and catalytic performance of oxide-supported catalysts. However, oxide particles in an oxide-supported
powder catalyst usually exhibit uneven and irregular morphologies and expose various types of crystal planes. Such morphological complexity and inhomogeneity of oxide particles
make it difficult to investigate and understand the effect of the
morphology of the oxide support in oxide-supported catalysts.
In the last decade, the controlled synthesis of oxide nanocrys[a] Y. Gao, Prof. Dr. W. Wang, S. Chang, Prof. Dr. W. Huang
Hefei National Laboratory for Physical Sciences at the Microscale
CAS Key Laboratory of Materials for Energy Conversion
Department of Chemical Physics
University of Science and Technology of China
Jinzhai Road 96, Hefei 230026 (China)
E-mail: huangwx@ustc.edu.cn

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

tals with uniform and well-defined morphologies has advanced


significantly, consequently, the effect of the morphology of
oxide nanocrystals on the structure and catalytic performance
of oxide-involved catalysts has been adequately demonstrated
and the morphology-controlled strategy has become a new
strategy to tune the catalytic performance of oxide-involved
catalysts.[213]
Ceria (CeO2) is among the oxides nanocrystals of which have
been successfully synthesized with various uniform and welldefined morphologies.[1417] Meanwhile, CeO2 is one of the
most important oxides in heterogeneous catalysis,[18] therefore,
the catalytic performance of CeO2 nanocrystals with various
morphologies has been explored extensively and morphologydependent catalytic performance has been observed in
a series of catalytic reactions.[1930] CeO2-supported catalysts
that employ CeO2 nanocrystals have also been prepared, and
the effect of the morphology of the CeO2 support was also observed on the active componentCeO2 interaction, the structure, and catalytic performance.[3142] Among CeO2 rods, cubes,
and polyhedra, CeO2 nanorods enclosed by {11 0} and {1 0 0}
crystal planes are most active for gold stabilization/activation,
and thus a Au supported on CeO2 nanorods catalyst is most
active in catalyzing the watergasshift (WGS) reaction.[31] A
Au/CeO2-rods catalyst was also reported to be more active in
ChemCatChem 2013, 5, 3610 3620

3610

CHEMCATCHEM
FULL PAPERS
catalyzing the preferential oxidation of CO in a H2-rich gas (COPROX) than Au/CeO2-polyhedra and Au/CeO2-cubes.[32, 33] We recently proposed a morphology-dependent interplay between
oxygen vacancies and Ag-CeO2 in Ag/CeO2 catalysts that controls their structure and catalytic performance.[40]
Pt/CeO2 catalysts are promising catalysts for low-temperature CO-PROX and WGS reactions and thus have attracted
great interest.[4348] Feng et al. deposited Pt clusters on CeO2
octahedra and rods by employing electron-beam evaporation
and observed the morphology-dependent structure of the deposited Pt clusters and catalytic performance in CO oxidation
and dehydrogenation reactions.[34] As far as we know, no study
of Pt/CeO2 catalysts prepared by conventional methods for
powder catalyst preparation, such as depositionprecipitation
and incipient wetness impregnation that employs CeO2 nanocrystals as the support, has been reported. Meanwhile, the
effect of metal loading on the structure and catalytic performance of metal/CeO2 catalysts that employ CeO2 nanocrystals
as the support has been seldom addressed in previous studies.
In this study, we prepared Pt/CeO2 catalysts with various Pt
loadings by a conventional incipient wetness impregnation
method that employed CeO2 cubes (c-CeO2), rods (r-CeO2), and
octahedra (o-CeO2) as the support. Their structures and catalytic activities in CO oxidation in excess O2 and the CO-PROX reaction were studied in detail. Strong morphology effects of the
CeO2 support were observed in the preparation, metalsupport
interaction, and catalytic performance of the Pt/CeO2 catalysts
and could be correlated the exposed crystal planes and surface
composition/structure of the CeO2 nanocrystals. The different
catalytic behaviors of Pt0-CeO2 and Pt2 + -CeO2 ensembles in the
catalysis of CO oxidation and the CO-PROX reaction were also
observed.

Results and Discussion


TEM and high-resolution TEM (HRTEM) images of r-CeO2,
c-CeO2, and o-CeO2 are shown in Figure 1, and their morphologies are all quite uniform. c-CeO2 has edge lengths mostly between 20 and 40 nm, r-CeO2 has a diameter distribution of
11  4 nm and a length distribution between 30 and 200 nm,
and o-CeO2 has edge lengths mostly between 100 and
300 nm. As reported previously[14, 15] and revealed by the

www.chemcatchem.org

Figure 1. TEM and HRTEM images of CeO2 rods (a, b), cubes (c, d), and octahedra (e, f). The insets schematically illustrate the crystal planes exposed on
the CeO2 rods, cubes, and octahedra.

HRTEM images, c-CeO2 has selectively exposed {1 0 0} crystal


planes, r-CeO2 has exposed {1 0 0} and {11 0} crystal planes, and
o-CeO2 has selectively exposed {111} crystal planes. The
CeO2(1 0 0) surface is a polar plane composed of CeIV layers and
O layers and its surface is terminated by the O layer. The
CeO2(11 0) surface is a nonpolar plane formed by a stack of
stoichiometric CeO2 layers and its surface has both exposed
CeIV and O atoms. The CeO2(111) surface is also a nonpolar surface formed by a stacking sequence of OCeO trilayers and
its surface has both exposed CeIV and O atoms.[49] The XRD patterns of c-CeO2, r-CeO2, and o-CeO2 are shown in Figure 2 and
they all display a typical cubic fluoride CeO2 crystal phase
(JCPDS card No. 34-0394). The diffraction peaks of o-CeO2 are
narrower than those of r-CeO2 and c-CeO2. This agrees with
the microscopic results that o-CeO2 has the largest size. The
BET specific surface areas of c-CeO2, r-CeO2, and o-CeO2 were
measured to be 23, 69, and 4 m2 g1, respectively.

Figure 2. XRD patterns of A) c-CeO2 and the Pt/c-CeO2 catalysts, B) r-CeO2 and the Pt/r-CeO2 catalysts, and C) o-CeO2 and the Pt/o-CeO2 catalysts.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemCatChem 2013, 5, 3610 3620

3611

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

For each type of CeO2 nanoTable 1. Pt loading, Pt0/Pt2 + ratio, and specific reaction rate of various Pt/CeO2 catalysts.
crystal, three Pt/CeO2 catalysts
Specific reaction rate of
Specific reaction rate of
Catalyst
Pt loading
Pt0/Pt2 +
with calculated Pt loadings of
ratio
CO oxidation (40 8C)
the PROX reaction (60 8C)
0.2, 0.5, and 1 % were prepared
[wt %]
[mmolCO molPt1 s1]
[mmolCO molPt1 s1]
by a conventional incipient wet0.2 %-Pt/r-CeO2
0.12
0
68
38
ness impregnation method that
0.33
0
118
95
0.5 %-Pt/r-CeO2
employed Pt(NH3)4(NO3)2 as the
1 %-Pt/r-CeO2
0.87
0.53
147
32
precursor followed by subse0.14
0
very low
9
0.2 %-Pt/c-CeO2
quent H2 reduction at 200 8C for
0.39
0
23
19
0.5 %-Pt/c-CeO2
0.88
0.22
40
13
1 %-Pt/c-CeO2
3 h. If we take the Pt/CeO2 cata0.14
0.35
very low
3
0.2 %-Pt/o-CeO2
lyst with a calculated Pt loading
0.32
0.4
18
7
0.5 %-Pt/o-CeO2
of 0.2 % that employs c-CeO2 as
0.72
1.58
8
very low
1 %-Pt/o-CeO2
the support as an example, the
catalyst is denoted as 0.2 %-Pt/cCeO2. The H2 reduction treatment of the catalyst precursors
was monitored by MS, and the
results (not shown) demonstrate
the complete reduction of NO3
to gaseous NH3. It was reported
previously that the reduction
temperature of NO3 can be lowered significantly from 460 to
220 8C by Rh.[50] Thus no N-containing species exist in our Pt/
CeO2 catalysts. The actual Pt
loading in Pt/CeO2 catalysts was
analyzed by using inductively
coupled plasma atomic emission
spectroscopy (ICP-AES), and the
results are summarized in
Table 1. The Pt loadings in Pt/cCeO2, Pt/r-CeO2, and Pt/o-CeO2
are similar for low Pt loadings
but the Pt loading in 1 %-Pt/oCeO2 is a little smaller than that
in 1 %-Pt/c-CeO2 and 1 %-Pt/rCeO2. The XRD patterns of the Figure 3. Representative TEM and HRTEM images and the particle size distribution of 1 %-Pt/r-CeO2 (A1, A2, A3),
1 %-Pt/c-CeO2 (B1, B2, B3), and 1 %-Pt/o-CeO2 (C1, C2, C3).
Pt/CeO2 catalysts are shown in
Figure 2. No diffraction pattern
that arises from Pt could be observed in all of the Pt/CeO2 catsupported on the faces of CeO2 nanocrystals rather than on
alysts except 1 %-Pt/o-CeO2, which exhibits a weak Pt(111) diftheir edges and truncated corners.
fraction peak at 39.88 in its XRD pattern. The morphologies
The surface compositions of Pt/CeO2 catalysts were studied
and structures of Pt/CeO2 were examined by TEM. Pt nanoparby X-ray photoelectron spectroscopy (XPS). No N signal was
detected, which also demonstrates the absence of N-containticles were only clearly identified in 1 %-Pt/r-CeO2, 1 %-Pt/cing species in the catalysts. PO43 was employed in the syntheCeO2, and 1 %-Pt/o-CeO2, the TEM and HRTEM images of which
are shown in Figure 3. The original morphology of the CeO2
sis of o-CeO2 nanocrystals, and our XPS analysis results demonstrate that the P/Ce ratio is  0.06 in o-CeO2 and the Pt/o-CeO2
supports is well preserved after the loading of Pt. By counting
more than 100 Pt nanoparticles for each catalyst, we acquired
catalysts. The Pt 4f XPS spectra of various Pt/CeO2 catalysts are
the size distributions of the supported Pt nanoparticles in 1 %shown in Figure 4. The Pt 4f XPS spectra can be adequately
Pt/r-CeO2, 1 %-Pt/c-CeO2 and 1 %-Pt/o-CeO2. The mean sizes of
fitted with two components with the Pt 4f7/2 binding energy at
the Pt nanoparticles in 1 %-Pt/r-CeO2, 1 %-Pt/c-CeO2, and 1 % 72.6 and  71.5 eV. The former can be assigned to Pt2 + spePt/o-CeO2 do not differ much and are 2.0, 1.5, and 1.7 nm, recies and the latter can be assigned to Pt0 (Pt nanoparticles).[47]
2+
0
spectively. As shown below, the Pt /Pt ratios in 1 %-Pt/rIt can be seen that the Pt species in the Pt/CeO2 catalysts
CeO2, 1 %-Pt/c-CeO2, and 1 %-Pt/o-CeO2 differ significantly,
depend on the Pt loading and the morphology of the CeO2
which will affect the sizes of the supported Pt nanoparticles.
support. The Pt0/Pt2 + ratios in the Pt/CeO2 catalysts estimated
Almost all of the counted Pt nanoparticles in the catalysts are
from the Pt 4f XPS results are summarized in Table 1. For the
 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemCatChem 2013, 5, 3610 3620

3612

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

Figure 4. Pt 4f XPS spectra of A) Pt/c-CeO2, B) Pt/r-CeO2, and C) Pt/o-CeO2.

Figure 5. H2-TPR profiles of the catalyst precursors of A) Pt/c-CeO2, B) Pt/r-CeO2, and C) Pt/o-CeO2 catalysts.

same CeO2 support, the Pt0/Pt2 + ratio increases with the Pt


loading. For the same Pt loading, the Pt0/Pt2 + ratio follows the
order Pt/o-CeO2 > Pt/r-CeO2 > Pt/c-CeO2. However, only Pt2 + is
present in Pt/r-CeO2 and Pt/c-CeO2 with low Pt loadings, and
Pt0 appears in 1 %-Pt/r-CeO2 and 1 %-Pt/c-CeO2. As the catalysts
were prepared by H2 reduction of the catalyst precursor at
200 8C for 3 h, these results indicate a morphology-dependent
Pt2 + CeO2 interaction in the catalyst precursor. Pt2 + , which has
a stronger Pt2 + CeO2 interaction, is more resistant to reduction. Thus the Pt2 + CeO2 interaction follows the order Pt2 + cCeO2 > Pt2 + r-CeO2 > Pt2 + o-CeO2.
The H2 temperature-programmed reduction (TPR) results of
various catalyst precursors (Figure 5) also demonstrate that the
Pt2 + CeO2 interaction follows the order of Pt2 + c-CeO2 > Pt2 +
r-CeO2 > Pt2 + o-CeO2. The H2-TPR of the catalyst precursors
includes the reduction/decomposition of impregnated Pt(NH3)4(NO3)2 (Pt2 + and NO3) on CeO2 and the reduction of
CeO2. The reduction of the bare CeO2 supports starts around
or above 300 8C (Figure 6), thus the signals observed below
300 8C are initiated by the reduction/decomposition of impregnated Pt(NH3)4(NO3)2. The precursors of Pt/o-CeO2 undergo the
reduction/decomposition process far below 200 8C (Figure 5 C),
consistent with the observation of Pt nanoparticles in Pt/o 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

CeO2 catalysts prepared by the reduction of catalyst precursors


at 200 8C for 3 h. The multiple peaks in the TPR spectra were
tentatively attributed to the reduction/decomposition processes at different stages or of precursors with different environments. Compared with those of the precursors of Pt/o-CeO2,
the profiles of the precursors of Pt/c-CeO2 and Pt/r-CeO2 (Figure 5 A and B) are simple but their reduction/decomposition
temperatures depend on the Pt loading. The reduction/decomposition process of Pt precursors in 0.2 %-Pt/c-CeO2 and 0.2 %Pt/r-CeO2 initiates at  215 8C. This agrees with the observation
of only Pt2 + species in both catalysts prepared by the reduction of the catalyst precursors at 200 8C for 3 h. With increased
Pt loading, the reduction/decomposition process shifts to
 200 8C. However, as a result of the strong Pt2 + c-CeO2 and
Pt2 + r-CeO2 interactions, only Pt2 + species form in 0.5 %-Pt2 +
/c-CeO2 and 0.5 %-Pt2 + /r-CeO2, and Pt2 + and Pt0 species coexist
in 1 %-Pt2 + /c-CeO2 and 1 %-Pt2 + /r-CeO2. The reduction/decomposition rate of the precursor of 1 %-Pt/r-CeO2 is much faster
than that of 1 %-Pt/c-CeO2. This might explain the larger Pt0/
Pt2 + ratio in 1 %-Pt/r-CeO2 than in 1 %-Pt/c-CeO2. We found
that the H2-TPR profiles of the precursors of Pt/r-CeO2 and Pt/
c-CeO2 are similar at low Pt loadings but differ at the 1% Pt
loading. As c-CeO2 has exposed {1 0 0} crystal planes and
ChemCatChem 2013, 5, 3610 3620

3613

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

Figure 6. H2-TPR profiles of A) c-CeO2 and as-prepared Pt/c-CeO2, B) r-CeO2 and as-prepared Pt/r-CeO2, and C) o-CeO2 and as-prepared Pt/o-CeO2 catalysts subjected to reoxidation treatment at 300 8C in 20 % O2-Ar for 2 h.

r-CeO2 has exposed {1 0 0} and {11 0} crystal planes, it is likely


that the Pt precursor preferentially interacts with the {1 0 0}
crystal planes of r-CeO2.
The XRD, TEM, XPS, and precursor-TPR results show clearly
that the interaction and reduction/decomposition of the Pt
precursor with the CeO2 support vary with the morphology
and exposed crystal planes of the CeO2 support. DFT calculations show that the surface energy of low-indexed CeO2 surfaces follows the order (111) < (11 0) < (1 0 0).[49] Thus it is reasonable that the Pt2 + precursor preferentially interacts with the
CeO2(1 0 0) surface, and the Pt2 + CeO2 interaction follows
the
order
Pt2 + c-CeO2
(CeO2(1 0 0)) >
Pt2 + r-CeO2
2+
(CeO2(1 0 0)+(11 0)) > Pt o-CeO2 (CeO2(111)), as we observed
experimentally. If we compare Pt/r-CeO2 and Pt/c-CeO2, the
poor dispersion of Pt nanoparticles in Pt/o-CeO2 results from
both the very low specific surface area of o-CeO2 and the weak
Pt2 + o-CeO2 (CeO2(111)) interaction. The transformation of the
catalyst precursor into an active supported catalyst is very
complicated but plays a key role in the catalyst composition
and structure. Our results reveal a strong effect of the morphology of the oxide support in such a process and shed light
on the fundamental understanding of the preparation of
oxide-supported catalysts.
The reduction behavior of CeO2 in various Pt/CeO2 catalysts
was also studied by H2-TPR (Figure 6). The Pt/CeO2 catalysts
were oxidized at 300 8C in 20 % O2/Ar for 2 h and then cooled
to room temperature in Ar prior to the measurements. The
starting surface reduction temperature of bare CeO2 nanocrystals follows the order r-CeO2 (  290 8C) < c-CeO2 (  330 8C) <
o-CeO2 (  400 8C). This indicates that the surface reducibility
follows the order r-CeO2 < c-CeO2 < o-CeO2. As the reduction of
the surface leads to the formation of surface oxygen vacancies,
the formation energy of surface oxygen vacancies should play
an important role in the reducibility of the CeO2 surface. DFT
calculations show that the oxygen vacancy formation energy
of low-indexed CeO2 surfaces follows the order (11 0) < (1 0 0) <
(111),[5154] which supports our experimental results. The surface reduction of CeO2 supports is promoted significantly in
 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Pt/CeO2 catalysts. Interestingly, the promotion effect does not


depend much on the Pt loading and Pt species. For the same
CeO2 support, all Pt/CeO2 catalysts exhibit similar H2-TPR profiles. These observations suggest that a small amount of Pt,
either as Pt2 + or Pt0, is enough to activate the surface CeO
bond and promote the surface reduction of CeO2 to the maximum extent. The only exception is the 0.2 %-Pt/c-CeO2 catalyst,
the low-temperature surface reduction peak of which is much
stronger than that of 0.5 %-Pt/c-CeO2 and 1 %-Pt/c-CeO2. This
result is reproducible but we do not yet understand the origin.
However, the promotion effect of Pt on the surface reduction
of CeO2 in Pt/CeO2 catalysts depends on the morphology of
the CeO2 support. The surface reduction of CeO2 in the Pt/oCeO2 catalysts starts at  100 8C and evolves into a main peak
at  306 8C with a shoulder at  438 8C, whereas a low-temperature surface reduction feature that starts at  40 8C occurs for
CeO2 in the Pt/r-CeO2 and Pt/c-CeO2 catalysts. This low-temperature surface reduction feature evolves into a broad peak between  40 and  285 8C for the Pt/r-CeO2 catalysts but into
a main peak at  90 8C with a shoulder at  65 8C and a long
tail extending to  250 8C for the Pt/c-CeO2 catalysts. The absence of the low-temperature surface reduction peak in the Pt/
o-CeO2 catalysts could be attributed the weak Pto-CeO2 interaction and the stable CeO2(111) surface. Notably, r-CeO2 and
Pt/r-CeO2 exhibit the largest surface reduction peaks. This
could be associated with their large specific surface areas.
These H2-TPR results clearly demonstrate the morphology-dependent surface reduction of CeO2 nanocrystals and the morphology-dependent promotion effect of Pt on the surface reduction of CeO2 nanocrystals in the Pt/CeO2 catalysts. The distinctly different surface reduction behaviors of the CeO2(111)
surface in the Pt/o-CeO2 catalysts and the CeO2(11 0) and (1 0 0)
surfaces in Pt/r-CeO2 and Pt/c-CeO2 catalysts provide an alternative explanation for the TPR profile of the oxide support in
oxide-supported powder catalysts in which various reduction
peaks in the TPR profile of the catalyst might result from the
reduction of the different crystal planes exposed on the oxide
support.
ChemCatChem 2013, 5, 3610 3620

3614

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

Figure 7. Visible Raman spectra of various CeO2 and Pt/CeO2 catalysts.

The concentrations of the oxygen vacancies of various CeO2


and Pt/CeO2 catalysts were further studied by visible and UV
Raman spectroscopy (Figure 7 and 8). As a result of the stronger absorption of CeO2 in the UV region than in the visible
region,[40] UV Raman spectroscopy is more surface sensitive in
this case than visible Raman spectroscopy. Two peaks were observed in the Raman spectra, a strong peak at  465 cm1 and
a weak peak at  598 cm1, which correspond to the F2g and
defect-induced (D) modes of the cubic CeO2 fluoride phase, respectively.[55] Peak-fitting analysis was performed for the acquired Raman spectra and the ID(598 cm1)/IF2g(465 cm1)
ratio was calculated to approximate the defect concentration
(Figures 7 and 8). For bare CeO2 nanocrystals, the ID/IF2g ratio estimated from the visible Raman spectra follows the order rCeO2 > c-CeO2 > o-CeO2, which agrees with the order of their
reducibility; however, the ID/IF2g ratio estimated from the UV
Raman spectra follows the order c-CeO2 > r-CeO2 > o-CeO2,
which indicates that defects in the surface region are more
stable in c-CeO2 than in r-CeO2. The Pt loading affects the ID/IF2g
ratios of c-CeO2, r-CeO2, and o-CeO2 in different ways. For the
Pt/o-CeO2 catalysts, the ID/IF2g ratio in the visible Raman spectra

does not vary much, whereas that in the UV Raman spectra increases slowly with the Pt loading. This could be associated
with both the weak Pto-CeO2 interaction and the stability of
the CeO2(111) surface. The ID/IF2g ratio in the UV Raman spectra
increases from 0.39 for 0.2 %-Pt/c-CeO2 to 0.64 for 0.5 %-Pt/cCeO2, whereas that in the visible Raman spectra does not
change. Meanwhile, the ID/IF2g ratio in the visible Raman spectra
increases from 0.04 for 0.2 %-Pt/r-CeO2 to 0.18 for 0.5 %-Pt/rCeO2, whereas that in the UV Raman spectra does not increase
much. These observations again indicate that defects in the
surface region are more stable in c-CeO2 than in r-CeO2. With
the increase of Pt loading to 1 %, the ID/IF2g ratios in both the
UV and visible Raman spectra increase significantly for Pt/cCeO2, which implies that the Pt nanoparticles are stronger to
create oxygen vacancies within c-CeO2 than the Pt2 + species;
however, the ID/IF2g ratio in the UV Raman spectra does not
change much for Pt/r-CeO2 but that in the visible Raman spectra decreases to 0.07 from 0.18 for 0.5 %-Pt/r-CeO2, which suggests a decrease of the concentration of oxygen vacancies. As
discussed above, in the 0.2 %-Pt/r-CeO2 and 0.5 %-Pt/r-CeO2
catalysts, Pt mainly interacts with the (1 0 0) surface exposed

Figure 8. UV Raman spectra of various CeO2 and Pt/CeO2 catalysts.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemCatChem 2013, 5, 3610 3620

3615

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

Figure 9. Catalytic performance of Pt/CeO2 catalysts in CO oxidation. The reactants consist of 1 % CO balanced with dry air.

on r-CeO2 and exists as Pt2 + ; in 1 %-Pt/r-CeO2, Pt interacts with


both the (1 0 0) and (11 0) surfaces exposed on r-CeO2 and
exists as both Pt2 + species and Pt nanoparticles. Therefore, the
decrease of the concentration of oxygen vacancies in Pt/r-CeO2
with the increase of Pt loading from 0.5 to 1 % could be associated with the presence of Pt nanoparticleCeO2(11 0) interactions in 1 %-Pt/r-CeO2.

The catalytic performances of CeO2 and the Pt/CeO2 catalysts


were examined for CO oxidation in excess O2 (Figure 9) and
the CO-PROX reaction (Figure 10). The bare CeO2 supports do
not show any catalytic activity in either of the reactions under
the investigated experimental conditions. For the Pt/CeO2 catalysts with the same Pt loading, the catalytic performance follows the order Pt/r-CeO2 > Pt/c-CeO2 > Pt/o-CeO2 in CO oxidation, and the same order holds in the CO-PROX reaction at re-

Figure 10. CO conversion (A1C1) and O2 conversion and selectivity (A2C2) of the CO-PROX reaction catalyzed by Pt/CeO2 catalysts. The reactants consist of
1 % CO, 1 % O2, and 50 % H2 balanced with N2.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemCatChem 2013, 5, 3610 3620

3616

CHEMCATCHEM
FULL PAPERS
action temperatures below 80 8C. This order is consistent with
the order of reducibility of the various CeO2 supports. In CO
oxidation reactions catalyzed by CeO2-involved catalysts, the
surface lattice oxygen of CeO2 is an active oxygen species.[18]
Thus, the easier the reduction of the employed CeO2 support,
the weaker the surface CeO bond, and the more active the
catalyst. Therefore, among the CeO2 nanocrystals investigated
herein, as a result of its strong reducibility and large specific
surface area, r-CeO2 is the best support to prepare efficient Pt/
CeO2 catalysts for CO oxidation and the CO-PROX reaction. At
high reaction temperatures in the CO-PROX reaction at which
the O2 conversion approaches 100 %, the CO conversion and
O2 selectivity of Pt/CeO2 catalysts do not depend much on the
Pt loading and CeO2 morphology and decrease with the reaction temperature. This is because high reaction temperatures
increase the surface coverage of H(a) (chemisorbed H adatoms)
on Pt surfaces at the expense of the surface coverage of CO(a)
(chemisorbed CO) on Pt surfaces.[48]
Interestingly, Pt/CeO2 catalysts with the same type of CeO2
support exhibit a morphology-dependent catalytic behavior on
the Pt loading in CO oxidation and the CO-PROX reaction at
low temperatures. In CO oxidation, the CO conversions of the
Pt/c-CeO2 and Pt/r-CeO2 catalysts increase with the Pt loading,
whereas those of the Pt/o-CeO2 catalysts follow the order
0.5 %-Pt/o-CeO2 > 1 %-Pt/o-CeO2 > 0.2 %-Pt/o-CeO2. In the COPROX reaction, the CO and O2 conversion of the Pt/c-CeO2 and
Pt/o-CeO2 catalysts follow the same orders as in CO oxidation
but those of Pt/r-CeO2 do not. The CO conversion in CO oxidation follows the order 1 %-Pt/r-CeO2 > 0.5 %-Pt/r-CeO2 > 0.2 %Pt/r-CeO2 but the CO and O2 conversion in the CO-PROX reaction follow the order 0.5 %-Pt/r-CeO2 > 1 %-Pt/r-CeO2 > 0.2 %-Pt/
r-CeO2. We further compared the specific reaction rates of various Pt/CeO2 catalysts normalized to the Pt loading in CO oxidation (40 8C) and the CO-PROX reaction (60 8C) at low temperatures, and the results are summarized in Table 1. The active
structure of Pt/CeO2 catalysts in CO oxidation reactions is generally believed to be the PtCeO2 interface, but it is still argued
about whether Pt0 or Pt2 + is more active.[4449] 0.2 %-Pt/c-CeO2
(0.2 %-Pt/r-CeO2) and 0.5 %-Pt/c-CeO2 (0.5 %-Pt/r-CeO2) only
contain Pt2 + species, whereas 1 %-Pt/c-CeO2 (1 %-Pt/r-CeO2)
contains both Pt2 + species and Pt nanoparticles. In CO oxidation, the specific reaction rate of 1 %-Pt/c-CeO2 (1 %-Pt/r-CeO2)
is higher than those of 0.2 %-Pt/c-CeO2 (0.2 %-Pt/r-CeO2) and
0.5 %-Pt/c-CeO2 (0.5 %-Pt/r-CeO2). This suggests that the Pt0CeO2 ensemble is more active than the Pt2 + -CeO2 ensemble in
the catalysis of CO oxidation in excess O2. In Pt/o-CeO2 catalysts, which contains both Pt2 + and Pt0 species, the reduced
catalytic activity of 1 %-Pt/o-CeO2 than 0.5 %-Pt/o-CeO2 can be
attributed to the aggregation of supported Pt nanoparticles, as
evidenced by both XRD and TEM results. This reduces the
number of active Pt-CeO2 ensemble sites and thus the catalytic
activity.
In the CO-PROX reaction, the specific reaction rate of 0.5 %Pt/c-CeO2 (0.5 %-Pt/r-CeO2) is higher than that of 1 %-Pt/c-CeO2
(1 %-Pt/r-CeO2), which is in contrast to the case in CO oxidation. This suggests that CO oxidation in the CO-PROX reaction
does not follow the same reaction mechanism as CO oxidation
 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemcatchem.org
in excess O2. It has been established that H2-assisted CO oxidation occurs in the CO-PROX reaction catalyzed by Pt/CeO2 catalysts.[4446] In the presence of H2, H(a) formed on Pt surface
spills over to the CeO2 support to form oxygen vacancies in
CeO2 as well as to hydroxylate and hydrate the CeO2 surface.
The presence of oxygen vacancies in CeO2 promotes its ability
to activate O2, meanwhile, the hydroxylated and hydrated
CeO2 surface supplies additional oxidizing species to oxidize
CO to CO2. It was also found that the presence of a significant
density of oxygen vacancies in CeO2 is beneficial for the H2-assisted CO oxidation mechanism. Recently model catalyst studies have also revealed that hydroxyl groups on an oxide surface can react with CO(a) on the neighboring Pt surface to
form CO2 at  300 K and this reaction is favored by the presence of oxygen vacancies in oxides if a large amount of H(a)
coexists with CO(a) on a Pt surface.[5658] Therefore, both reaction mechanisms of CO oxidation in excess O2 and H2-assisted
CO oxidation should coexist in the CO-PROX reaction catalyzed
by Pt/CeO2 catalysts. Moreover, as the Pt0-CeO2 ensemble is
more active than the Pt2 + -CeO2 ensemble in the catalysis of
CO oxidation following the CO oxidation mechanism in excess
O2, the higher specific reaction rate of 0.5 %-Pt/c-CeO2 (0.5 %Pt/r-CeO2) over 1 %-Pt/c-CeO2 (1 %-Pt/r-CeO2) in the CO-PROX
reaction indicates that the Pt2 + -CeO2 ensemble should be
more active than the Pt0-CeO2 ensemble in the catalysis of CO
oxidation following the H2-assisted CO oxidation mechanism.
Notably, the CO conversion of 0.5 %-Pt/r-CeO2 with a lower Pt
loading is higher that of 1 %-Pt/r-CeO2. This cannot be merely
attributed to the Pt2 + species because 1 %-Pt/r-CeO2 contains
both Pt2 + and Pt0, but can be reasonably associated with the
large concentration of oxygen vacancies in CeO2 of 0.5 %-Pt/rCeO2 that benefits the H2-assisted CO oxidation. The ID/IF2g ratio
in the visible Raman spectra is 0.18 for 0.5 %-Pt/r-CeO2 and
0.07 for 1 %-Pt/r-CeO2. Thus the Pt2 + species and CeO2 with
large concentration of oxygen vacancies constitute the active
structure of the Pt/CeO2 catalyst for the CO-PROX reaction.
The above results provide solid experimental evidence for
the effect of the morphology of the CeO2 support in Pt/CeO2
catalysts from the catalyst preparation process to the catalytic
performance. The effect of the morphology of the CeO2 support can be reasonably correlated with different exposed crystal planes on the CeO2 supports with different morphologies
and their surface composition/structure. c-CeO2 has selectively
exposed {1 0 0} crystal planes, r-CeO2 has exposed {1 0 0} and
{11 0} crystal planes, and o-CeO2 has selectively exposed {111}
crystal planes. DFT calculations show that the surface energy of
low-indexed CeO2 surfaces follows the order (111) < (11 0) <
(1 0 0)[49] but the oxygen vacancy formation energy follows the
order (11 0) < (1 0 0) < (111) because of the restructuring of the
(11 0) surface.[5154] During the Pt/CeO2 catalyst preparation process, the least stable CeO2(1 0 0) surface (c-CeO2) interacts most
strongly with the Pt2 + precursor, followed by CeO2(11 0)
(r-CeO2) and CeO2(111) (o-CeO2). Reduced at the same temperature, Pt/o-CeO2 catalysts with the weakest Pt2 + precursor
CeO2 interaction contain the largest fraction of metallic Pt,
whereas Pt/c-CeO2 catalysts with the strongest Pt2 + precursor
CeO2 interaction contain the largest fraction of Pt2 + species.
ChemCatChem 2013, 5, 3610 3620

3617

CHEMCATCHEM
FULL PAPERS
The reducibility and concentration of oxygen vacancies of
CeO2 in the Pt/CeO2 catalysts vary with the CeO2 morphology
and the PtCeO2 interaction. Thus, prepared by the same
method, Pt/CeO2 catalysts with the same Pt loading that
employ CeO2 cubes, rods, and octahedra as the supports have
different structures and exhibit different catalytic activities. On
one hand, the effect of the morphology of the CeO2 support in
the Pt/CeO2 catalysts adequately reveals the structural complexity of oxide-supported catalysts. Different crystal planes exposed on the catalyst nanoparticle surface exhibit different surface reactivities and catalytic activities, thus their contributions
to the observed overall catalytic activity must be unequal. It
has been reported that the coupling between adjacent crystal
planes with different catalytic activities can occur by the migration of surface species during the catalytic reaction.[13, 59, 60] The
structural complexity at the level of the crystal plane and morphology is a major hindrance for the correlation between the
structure and catalytic activity of powder catalysts. On the
other hand, the effect of the morphology of the CeO2 support
in the Pt/CeO2 catalysts proves that the structure and catalytic
performance of Pt/CeO2 catalysts for CO oxidation and the COPROX reaction can be tuned by controlling the morphology of
the CeO2 support. Among the various morphologies of the
CeO2 support investigated in this study, CeO2 rods are the best
support for the preparation of active Pt/CeO2 catalysts. Therefore, the morphology-controlled strategy of oxides will contribute significantly not only to the fundamental understanding of
heterogeneous catalysis by oxide-involved catalysts but also to
the innovation of efficient oxide-involved catalysts.[5]

Conclusions
We have comprehensively studied the structures and catalytic
performances of Pt/CeO2 catalysts with different Pt loadings
and different morphologies of the CeO2 support (cubes, rods,
and octahedra) in CO oxidation and the preferential oxidation
of CO in a H2-rich gas (CO-PROX) reaction. The effect of the
morphology of the CeO2 support exists in Pt/CeO2 catalysts
from the catalyst preparation process to the catalytic performance. The impregnated Pt precursor interacts more strongly
with CeO2 rods and cubes than with CeO2 octahedra and its reduction/decomposition is easier on CeO2 octahedra than on
CeO2 rods and cubes. The Pt/CeO2-octahedra catalyst contains
the largest fraction of metallic Pt, whereas the Pt/CeO2-cubes
catalyst contains the largest fraction of Pt2 + species. The reducibility of pure CeO2 and CeO2 in the Pt/CeO2 catalysts follows the order CeO2-rods > CeO2-cubes > CeO2-octahedra, and
the promotion effect of Pt on the reducibility of CeO2 is stronger in Pt/CeO2-rods and Pt/CeO2-cubes than in Pt/CeO2-octahedra. The concentration of oxygen vacancies of CeO2 in Pt/CeO2
catalysts varies with the CeO2 morphology and the PtCeO2 interaction. The catalytic performance of the Pt/CeO2 catalysts in
both CO oxidation and the CO-PROX reaction follows the order
Pt/CeO2-rods > Pt/CeO2-cubes > Pt/CeO2-octahedra. The Pt0CeO2 ensemble is more active than the Pt2 + -CeO2 ensemble in
the catalysis of CO oxidation in excess O2. H2-assisted CO oxidation catalyzed by the Pt/CeO2 catalysts was observed in the
 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemcatchem.org
CO-PROX reaction, and the Pt2 + species and CeO2 with a large
concentration of oxygen vacancies constitute the active structure of the Pt/CeO2 catalyst for this reaction. These results not
only demonstrate that the structure and catalytic performance
of oxide-supported catalysts can be tuned by controlling the
morphology of the oxide support but also greatly deepens the
fundamental understanding of CO oxidation reactions catalyzed by Pt/CeO2 catalysts.

Experimental Section
All the chemicals were analytical grade reagents and were purchased from Sinopharm Chemical Reagent Co., Ltd. and used as received in our experiments. The synthesis of CeO2 cubes and rods
followed the procedure of Mai et al.[14] Typically, Ce(NO3)36 H2O
(1.96 g) was dissolved in ultrapure water (40 mL; resistance >
18 MW), and NaOH (16.88 g) was dissolved in ultrapure water
(30 mL). The NaOH solution was added dropwise into the Ce(NO3)3
solution under stirring at RT. The solution was stirred for an additional 30 min at RT and then transferred into a 100 mL Teflon
bottle. The Teflon bottle was tightly sealed and hydrothermally
treated in a stainless-steel autoclave at 180 8C for 24 h. After cooling, the obtained white precipitate was collected, washed with ultrapure water, and dried in vacuo at 80 8C for 16 h. Then the acquired yellow powder was calcined in a muffle oven at 500 8C for
4 h to synthesize CeO2 cubes. The synthesis procedure for CeO2
rods was the same as that for CeO2 cubes except that the hydrothermal treatment temperature was 100 8C. The synthesis of CeO2
octahedra followed the procedure of Yan et al.[15] Typically, Ce(NO3)36 H2O (2 mmol) was dissolved in ultrapure water (79 mL),
and then Na3PO4 (1 mL, 0.02 m) was added. The mixed solution
was stirred for 1 h at RT and then transferred into a 100 mL Teflon
bottle. The Teflon bottle was tightly sealed and hydrothermally
treated in a stainless-steel autoclave at 170 8C for 10 h. After cooling, the obtained white precipitate was collected, washed with ultrapure water and ethanol several times, and dried in vacuo at
80 8C for 16 h. Then, the acquired white powder was calcined in
a muffle oven at 500 8C for 4 h to synthesize CeO2 octahedra.
The Pt/CeO2 catalysts were prepared by a conventional incipient
wetness impregnation method. Typically, the desired amount of Pt(NH3)4(NO3)2 solution was added dropwise to CeO2 nanocrystals
(0.3 g) and ultrasonicated for 10 min. After the impregnation, the
sample was kept at RT for 24 h and then dried at 80 8C for 12 h to
prepare the catalyst precursor. The catalyst precursors were reduced in pure H2 at 200 8C for 3 h to prepare the Pt/CeO2 catalysts.
Three Pt/CeO2 catalysts with calculated Pt loadings of 0.2, 0.5, and
1 % (Pt/CeO2 weight ratio) were prepared.
The loading of Pt in the Pt/CeO2 catalysts was determined by using
an Optima 7300 DV inductively coupled plasma atomic emission
spectrometer (ICP-AES). BET specific surface areas were acquired by
using a Beckman Coulter SA3100 surface area analyzer, and the
sample was degassed at 300 8C for 5 h in a N2 atmosphere before
the measurement. XRD patterns were recorded by using a Philips
XPert PRO diffractometer using a Ni-filtered CuKa (wavelength =
0.15418 nm) radiation source with the operation voltage and operation current of 40 kV and 40 mA, respectively. XPS measurements
were performed by using an ESCALAB 250 high-performance electron spectrometer using monochromatized AlKa (hn = 1486.7 eV) as
the excitation source. The likely charging of samples was corrected
by setting the binding energy of the adventitious carbon (C 1s) to
284.8 eV. Laser Raman spectra were obtained in back-scattering
ChemCatChem 2013, 5, 3610 3620

3618

CHEMCATCHEM
FULL PAPERS

www.chemcatchem.org

configuration by using a LABRAM-HR Confocal Laser Raman spectrometer. Ar + (514.5 nm) and HeCd (325 nm) lasers were employed as the excitation source to obtain visible Raman and UV
Raman spectra, respectively. TEM and HRTEM images were acquired by using Jeol-2010 and Jeol-2100F high-resolution transmission electron microscopes with the electron acceleration energy of
200 kV.
H2-TPR experiments were performed by using a Micromeritics
ChemiSorb 2750 instrument, in which the sample (50 mg) was
heated at a rate of 10 8C min1 in a 5 % H2-Ar mixture with a flow
rate of 20 mL min1. During the TPR measurements of the Pt/CeO2
catalysts, the catalyst precursors were reduced at 200 8C with pure
H2 for 3 h, reoxidized at 300 8C in 20 % O2-Ar for 2 h, and then
cooled to RT in Ar prior to the measurements.
The catalytic activity of the Pt/CeO2 catalysts in CO oxidation and
the CO-PROX reaction were evaluated by using a fixed-bed flow reactor. In CO oxidation, Pt/CeO2 (25 mg) diluted with SiO2 (50 mg)
was used, and the reaction gas that consisted of 1 % CO balanced
with dry air was fed at a rate of 30 mL min1. In the CO-PROX reaction, Pt/CeO2 (100 mg) diluted with SiO2 (100 mg) was used, and
the feed gas stream that consisted of 1 % CO, 1 % O2, and 50 % H2
balanced with N2 was fed at a total flow rate of 100 mL min1. The
catalyst was heated to the desired reaction temperature at a rate
of 2 8C min1 and then kept for 35 min until the catalytic reaction
reached the steady state. Then the composition of the effluent gas
was analyzed by using an online GC-14 gas chromatograph. The
conversion of CO and O2 were calculated from the change in their
concentrations of the inlet and outlet gases. O2 selectivity to CO2
was as calculated according to Equation (1):
SO2

X CO
 100 %
2 X O2

in which X and S are the conversion and selectivity, respectively.

Acknowledgements
This work was financially supported by the National Basic Research Program of China (2013CB933104, 2010CB923301), the
National Natural Science Foundation of China (21173204), and
the Fundamental Research Funds for the Central Universities.
Keywords: cerium metal-support interaction oxidation
platinum supported catalysts structureactivity relationships

[1] G. Wulff, Z. Kristallogr. Mineral. 1901, 34, 449 530.


[2] J. A. van Bokhoven, ChemCatChem 2009, 1, 363 364.
[3] K. B. Zhou, Y. D. Li, Angew. Chem. 2012, 124, 622 635; Angew. Chem.
Int. Ed. 2012, 51, 602 613.
[4] Y. Li, W. J. Shen, Sci. China Chem. 2012, 55, 2485 2496.
[5] W. X. Huang, Top. Catal. 2013, DOI: 10.1007/s11244-013-0139-6.
[6] B. M. Choudary, R. S. Mulukutla, K. J. Klabunde, J. Am. Chem. Soc. 2003,
125, 2020 2021.
[7] L. H. Hu, Q. Peng, Y. D. Li, J. Am. Chem. Soc. 2008, 130, 16136 16137.
[8] X. W. Xie, Y. Li, Z. Q. Liu, M. Haruta, W. J. Shen, Nature 2009, 458, 746
749.
[9] H. Z. Bao, W. H. Zhang, D. L. Shang, Q. Hua, Y. S. Ma, Z. Q. Jiang, J. L.
Yang, W. X. Huang, J. Phys. Chem. C 2010, 114, 6676 6680.
[10] Q. Hua, D. L. Shang, W. H. Zhang, K. Chen, S. J. Chang, Y. S. Ma, Z. Q.
Jiang, J. L. Yang, W. X. Huang, Langmuir 2011, 27, 665 671.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

[11] H. Z. Bao, W. H. Zhang, Q. Hua, Z. Q. Jiang, J. L. Yang, W. X. Huang,


Angew. Chem. 2011, 123, 12502 12506; Angew. Chem. Int. Ed. 2011, 50,
12294 12298.
[12] Q. Hua, K. Chen, S. J. Chang, Y. S. Ma, W. X. Huang, J. Phys. Chem. C
2011, 115, 20618 20627.
[13] Q. Hua, K. Chen, S. J. Chang, H. Z. Bao, Y. S. Ma, Z. Q. Jiang, W. X. Huang,
RSC Adv. 2011, 1, 1200 1203.
[14] H. X. Mai, L. D. Sun, Y. W. Zhang, R. Si, W. Feng, H. P. Zhang, H. C. Liu,
C. H. Yan, J. Phys. Chem. B 2005, 109, 24380 24385.
[15] L. Yan, R. Yu, J. Chen, X. Xing, Cryst. Growth Des. 2008, 8, 1474 1477.
[16] D. Zhang, X. Du, L. Shi, R. Gao, Dalton Trans. 2012, 41, 14455 14475.
[17] C. Sun, H. Li, L. Chen, Energy Environ. Sci. 2012, 5, 8475 8505.
[18] A. Trovarelli, Catal. Rev. Sci. Eng. 1996, 38, 439 520.
[19] K. Zhou, X. Wang, X. Sun, Q. Peng, Y. Li, J. Catal. 2005, 229, 206 212.
[20] W. Q. Han, L. J. Wu, Y. M. Zhu, J. Am. Chem. Soc. 2005, 127, 12814
12815.
[21] K. B. Zhou, Z. Q. Yang, S. Yang, Chem. Mater. 2007, 19, 1215 1217.
[22] Tana, M. L. Zhang, J. Li, H. J. Li, Y. Li, W. J. Shen, Catal. Today 2009, 148,
179 183.
[23] Y. Guan, E. J. M. Hensen, Phys. Chem. Chem. Phys. 2009, 11, 9578.
[24] X. Liu, K. Zhou, L. Wang, B. Wang, Y. Li, J. Am. Chem. Soc. 2009, 131,
3140 3141.
[25] J. Lv, Y. Shen, L. Peng, X. Guo, W. Ding, Chem. Commun. 2010, 46,
5909 5911.
[26] Z. Wu, M. Li, J. Howe, H. M. Meyer, S. H. Overbury, Langmuir 2010, 26,
16595 16606.
[27] Z. Wu, M. Li, S. H. Overbury, J. Catal. 2012, 285, 61 73.
[28] S. Agarwal, L. Lefferts, B. L. Mojet, ChemCatChem 2013, 5, 479 489.
[29] L. Torrente-Murciano, A. Gilbank, B. Puertolas, T. Garcia, B. Solsona, D.
Chadwick, Appl. Catal. B 2013, 132 133, 116 122.
[30] T. Dsaunay, G. Bonura, V. Chiodo, S. Freni, J. P. Couzini, J. Bourgon, A.
Ringued, F. Labat, C. Adamo, M. Cassir, J. Catal. 2013, 297, 193 201.
[31] R. Si, M. Flytzani-Stephanopoulos, Angew. Chem. 2008, 120, 2926 2929;
Angew. Chem. Int. Ed. 2008, 47, 2884 2887.
[32] G. Yi, N. Xu, G. Guo, K.-i. Tanaka, Y. Yuan, Chem. Phys. Lett. 2009, 479,
128 132.
[33] G. Yi, H. Yang, B. Li, H. Lin, K.-i. Tanaka, Y. Yuan, Catal. Today 2010, 157,
83 88.
[34] L. Feng, D. T. Hoang, C. K. Tsung, W. Y. Huang, S. H. Y. Lo, J. B. Wood,
H. T. Wang, J. Y. Tang, P. D. Yang, Nano Res. 2011, 4, 61 71.
[35] Y. Guan, D. Ligthart, . Pirgon-Galin, J. Pieterse, R. van Santen, E.
Hensen, Top. Catal. 2011, 54, 424 438.
[36] X. Han, H. J. Kim, S. Yoon, H. Lee, J. Mol. Catal. A 2011, 335, 82 88.
[37] L. Liu, Z. Yao, Y. Deng, F. Gao, B. Liu, L. Dong, ChemCatChem 2011, 3,
978 989.
[38] X. Du, D. Zhang, L. Shi, R. Gao, J. Zhang, J. Phys. Chem. C 2012, 116,
10009 10016.
[39] Z. Wu, M. Li, S. H. Overbury, ChemCatChem 2012, 4, 1653 1661.
[40] S. Chang, M. Li, Q. Hua, L. Zhang, Y. Ma, B. Ye, W. Huang, J. Catal. 2012,
293, 195 204.
[41] Z. Wu, V. Schwartz, M. Li, A. J. Rondinone, S. H. Overbury, J. Phys. Chem.
Lett. 2012, 3, 1517 1522.
[42] D. Gamarra, A. L. Cmara, M. Monte, S. B. Rasmussen, L. E. Chinchilla,
A. B. Hungra, G. Munuera, N. Gyorffy, Z. Schay, V. C. Corbern, J. C.
Conesa, A. Martnez-Arias, Appl. Catal. B 2013, 130 131, 224 238.
[43] C. Ratnasamy, J. P. Wagner, Catal. Rev. 2009, 51, 325 440.
[44] O. Pozdnyakova, D. Teschner, A. Wootsch, J. Krhnert, B. Steinhauer, H.
Sauer, L. Toth, F. Jentoft, A. Knop-Gericke, Z. Pal, R. Schlgl, J. Catal.
2006, 237, 1 16.
[45] O. Pozdnyakova-Tellinger, D. Teschner, J. Krhnert, F. C. Jentoft, A. KnopGericke, R. Schlgl, A. Wootsch, J. Phys. Chem. C 2007, 111, 5426 5431.
[46] D. Teschner, A. Wootsch, O. Pozdnyakovatellinger, J. Krhnert, E. Vass,
M. Hvecker, S. Zafeiratos, P. Schnrch, P. Jentoft, A. Knop-Gericke, R.
Schlgl, J. Catal. 2007, 249, 318 327.
[47] Y. Zhai, D. Pierre, R. Si, W. Deng, P. Ferrin, A. U. Nilekar, G. Peng, J. A.
Herron, D. C. Bell, H. Saltsburg, M. Mavrikakis, M. Flytzani-Stephanopoulos, Science 2010, 329, 1633 1636.
[48] C. S. Polster, R. Zhang, M. T. Cyb, J. T. Miller, C. D. Baertsch, J. Catal.
2010, 273, 50 58.
[49] M. Nolan, S. C. Parker, G. W. Watson, Surf. Sci. 2005, 576, 217 229.
[50] J. Barbier Jr., F. Marsollier, D. Duprez, Appl. Catal. A 1992, 90, 11 23.

ChemCatChem 2013, 5, 3610 3620

3619

CHEMCATCHEM
FULL PAPERS
[51] M. Nolan, S. C. Parker, G. W. Watson, Surf. Sci. 2005, 595, 223 232.
[52] M. Nolan, Chem. Phys. Lett. 2010, 499, 126 130.
[53] D. O. Scanlon, N. M. Galea, B. J. Morgan, G. W. Watson, J. Phys. Chem. C
2009, 113, 11095 11103.
[54] M. V. Ganduglia-Pirovano, J. L. F. Da Silva, J. Sauer, Phys. Rev. Lett. 2009,
102, 026101.
[55] W. H. Weber, K. C. Hass, J. R. McBride, Phys. Rev. B 1993, 48, 178 185.
[56] L. Xu, Y. Ma, Y. Zhang, Z. Jiang, W. Huang, J. Am. Chem. Soc. 2009, 131,
16366 16367.
[57] L. Xu, Z. Wu, Y. Zhang, B. Chen, Z. Jiang, Y. Ma, W. Huang, J. Phys. Chem.
C 2011, 115, 14290 14299.

 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemcatchem.org
[58] L. Xu, Z. Wu, Y. Jin, Y. Ma, W. Huang, Phys. Chem. Chem. Phys. 2013, 15,
12068 12074.
[59] V. Gorodetskii, J. Lauterbach, H. H. Rotermund, J. H. Block, G. Ertl, Nature
1994, 370, 276 279.
[60] W. X. Huang, R. S. Zhai, X. H. Bao, Langmuir 2001, 17, 3629 3634.

Received: August 25, 2013


Published online on September 23, 2013

ChemCatChem 2013, 5, 3610 3620

3620

Вам также может понравиться