You are on page 1of 13

Journal of Catalysis 344 (2016) 365–377

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Structure dependence and reaction mechanism of CO oxidation: A model


study on macroporous CeO2 and CeO2-ZrO2 catalysts
Yane Zheng a,b, Kongzhai Li a,b,⇑, Hua Wang a, Yuhao Wang a,b, Dong Tian a,b, Yonggang Wei a,b, Xing Zhu a,b,
Chunhua Zeng a, Yongming Luo c
a
State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, Kunming University of Science and Technology, Kunming 650093, China
b
Faculty of Metallurgical and Energy Engineering, Kunming University of Science and Technology, Kunming 650093, China
c
Faculty of Environmental Engineering, Kunming University of Science and Technology, Kunming 650093, China

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

Article history: Three-dimensionally ordered macroporous (3DOM) CeO2 and CeO2-ZrO2 oxides with different particle
Received 1 July 2016 sizes, oxygen vacancy concentrations, oxygen mobility and preferentially exposed surface planes were
Revised 24 September 2016 synthesized, which, as model catalysts, give a new approach for understanding the structural dependence
Accepted 12 October 2016
and reaction mechanism of CO oxidation over CeO2-based catalysts. The prepared 3DOM CeO2 and CeO2-
ZrO2 catalysts exhibited much higher catalytic activity for CO oxidation than the nonporous samples.
Although the textural and reducible features of catalysts could affect the catalytic performance, the pref-
Keywords:
erentially exposed surface plane is crucial for determining the catalytic activity. The {1 1 0} plane of CeO2
3DOM CeO2-ZrO2
CO oxidation
could create more active sites for CO adsorption, resulting in relatively high activity for CO oxidation.
Oxygen vacancy Higher concentration of oxygen vacancy would also enhance the reactivity for CO oxidation. Langmuir-
In situ DRIFTs Hinshelwood mechanism should be the crucial reaction pathway for CO oxidation over the 3DOM
Reaction mechanism CeO2-based catalysts, although the Mars-van Krevelen mechanism cannot be ignored.
Ó 2016 Elsevier Inc. All rights reserved.

1. Introduction peratures, which leads to the production of large numbers of oxy-


gen vacancies. The partly reduced FeOx is involved in the CO
Catalytic oxidation of carbon monoxide (CO), one of the most oxidation, acting as an oxygen donor. On the other hand, the inter-
important prototype reactions in heterogeneous catalysis, has face sites between Al2O3 and Pt/Pd are insignificant. Ceria-based
attracted considerable attention due to its extensive application supports as reducible oxides owned higher redox capacity, and
in environmental and energy areas (such as stationary and vehicle the high dispersion of Pt on ceria-based oxides plays an important
exhaust control, air purification and CO removal in H2-rich feed) role for its superior catalytic activity [9]. The interaction between
[1]. Among various heterogeneous catalysts involved, supported Pt and ceria-based oxides can inhibit the Pt growth and sintering
metals and reducible oxides show prominent properties. Currently, at high temperature, which is absent on the Pt/Al2O3 or Pt/ZrO2
noble metals, especially Au and platinum group metals, with the catalyst.
presence of supports present high activity for CO oxidation at For the Au on inert supports system, such as Au/Al2O3 [3], the
low temperature [2]. Metallic oxides used as supports can be clas- noble metals are the only active sites, which determined the activ-
sified as inert or active according to their redox properties. Thus, ity for CO oxidation. As for the Au-reducible supports system, the
Al2O3 and SiO2 fit within the inert supports [3,4], while reducible support plays an important role in the catalytic process, particu-
transition metal oxides such as, MnO2, Fe2O3, Co3O4 and CeO2 have larly for CeO2 and other transition metal oxides [10–14]. CeO2
to be considered as active supports [5–8]. Liu et al. [6] prepared Pt- was proposed to enhance the dispersion and stability of metal
group metals on both the inert supports and reducible supports, components which behave as active sites [12–14]. More impor-
such as Al2O3- or Fe2O3-supported Pt or Pd. It was found that Pt tantly, since CO oxidation usually takes place at the metal-
or Pd on Fe2O3 can facilitate Fe3+ reduction at relatively low tem- support interface, the ‘‘active” support CeO2 is also involved in
the reaction owing to the abundant active oxygen species present
⇑ Corresponding author at: State Key Laboratory of Complex Nonferrous Metal on the surface [15]. Kim et al. [8] reported that, for Au/CeO2 cata-
Resources Clean Utilization, Kunming University of Science and Technology, lysts, the chemical modification on the CeO2 support is promising
Kunming 650093, China. for the optimization of oxidation catalysis.
E-mail addresses: kongzhai.li@aliyun.com, kongzhai.li@foxmail.com (K. Li).

http://dx.doi.org/10.1016/j.jcat.2016.10.008
0021-9517/Ó 2016 Elsevier Inc. All rights reserved.
366 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

As for the independent reducible oxide catalysts, Venkatas- ity, oxygen mobility and preferential exposed crystal planes of
wamy et al. [5] suggested that the mixed reducible CeO2-MnOx CeO2. Investigations on the physicochemical properties of these
oxides showed remarkable enhancement on the activity for CO oxi- catalysts were performed, which were associated with their cat-
dation at lower temperatures, which can be attributed to a highly alytic activity for CO oxidation. Particularly, the nature of the inter-
dispersed state of Mn2+/Mn3+ in the ceria matrix, facile redox beha- action of CO with 3DOM catalysts using in situ infrared technology
viour, and a synergistic Mn-Ce interaction. Luo et al. [7] found that was also studied. Based on the experimental data, the possible
in Co3O4-CeO2 catalyst, Co3O4 crystallites are considered to be reaction mechanisms are discussed in detail. The results showed
encapsulated by nanosized CeO2, with only a small fraction of Co that the crystal plane of CeO2 preferentially exposed and oxygen
ions exposing on the surface and strongly interacting with CeO2. vacancy concentration are the critical factors for catalytic CO oxi-
They proposed that the CO oxidation over Co3O4-CeO2 should take dation over 3DOM CeO2-based oxide catalysts. These results will
place preferentially at the interface of Co3O4-CeO2 instead of the bring new insights into the determinant factors and reaction
surface of Co3O4. Experimental results also reveal that CeO2 and mechanisms.
CeO2-ZrO2 oxides show prominent catalytic activity for CO oxida-
tion [16–18]. CeO2-ZrO2 mixed oxides with better oxygen storage
capacity have been considered as an outstanding material in 2. Experimental
three-way catalysts (TWCs), which aim at the simultaneous purifi-
cation of CO, HC and NOx in automotive exhaust. The introduction 2.1. Catalyst preparation
of Zr could retard the sintering of oxide particles and improve the
thermal stability, which are required for highly active TWCs. The 3DOM CeO2 and CeO2-ZrO2 solid solutions were fabricated
Currently, the interaction between CO and CeO2 during the via PMMA-templating route. The precursor solutions of catalysts
reaction has emerged as one of the most important issues to were obtained by mixing stoichiometric amounts of Ce(NO3)36H2-
understand the catalytic mechanism. The reaction models include O and ZrOCl28H2O with deionized water at room temperature
rapid adsorption of CO molecules on low-coordination surface (RT). The atomic ratio of Ce:Zr was 8:2. And then citric acid was
noble metal sites and its subsequent oxidation at the perimeter added into the precursor and dissolved at 60 °C for 1 h under stir-
of metal nano-particles where oxygen is activated [19–21]. For ring. After that, dried PMMA colloidal crystal templates were
the independent ceria based catalysts, CO oxidation is believed to added in the solutions for 4 h. The precursors were then dried at
proceed via Mars-van Krevelen mechanism over ceria [22–24]. It 60 °C for 24 h. Finally, the CeO2 and CeO2-ZrO2 samples were
involves a complex series of redox reactions: the removal of sur- obtained by calcination at 350 °C for 2 h with a ramp rate of
face oxygen by CO with formation of oxygen vacancies and conse- 2 °C/min. The obtained powders were then heated at 2 °C/min up
quent annihilation of vacancies by gas-phase oxygen. Since the to 450 °C with a dwell time of 2 h. These as-synthesized CeO2
oxidation of reduced cerium oxides is very fast, the reduction pro- and CeO2-ZrO2 samples were abbreviated accordingly as C450
cess, i.e., removal of surface oxygen, is the rate determination step. and CZ450, respectively. Both the C450 and CZ450 samples were
Moreover, it is also found that CO interaction with ceria is structure further calcined at 600 and 800 °C in air for 2 h, respectively,
dependent and oxygen vacancy formation energy is sensitive to the labelled as C600, C800, CZ600 and CZ800.
exposing crystal planes of CeO2 nano particles [24–27]. The activity The nonporous CeO2-ZrO2 sample was prepared via a co-
of CeO2 can be improved through predominantly exposing their precipitation method. The stoichiometric amounts of Ce(NO3)3-
{1 0 0} and {1 1 0} planes [24,25,27]. Further investigation suggests 6H2O and ZrOCl28H2O were dissolved in deionized water under
that the interaction of CeO2 with CO also involves the formation of stirring at RT. The atomic ratio of Ce:Zr was also 8:2. The hydrox-
carbonate-like complexes on the {1 0 0} and {1 1 0} surfaces [28]. ides were precipitated by adding drops of ammonia solution. The
The carbonate species is competitive to CO2 formation and desorp- mixtures were stirred for one hour, when the pH increased to
tion, although CO could directly react with the surface active oxy- 8.0. The precipitates were dried at 60 °C for 24 h, and then calcined
gen on CeO2 [29]. This is similar with the characteristics of CO at 350 °C for 2 h with a ramp rate of 2 °C/min. The obtained solids
oxidation over noble metal catalysts. Despite the intensive studies were then heated at 2 °C/min up to 450 °C with a dwell time of 2 h.
of CO oxidation on ceria based catalysts, there is still a lack of clear The as-synthesized nonporous CeO2-ZrO2 sample was labelled as
recognition as to the relationship among different factors (i.e., CO N-CZ450. Similarly with the 3DOM samples, the N-CZ450 was also
adsorption, formation/concentration of oxygen vacancies, oxygen calcined at 600 and 800 °C in air for 2 h, respectively, named as N-
mobility and the crystal planes of CeO2 exposed) in the catalytic CZ600 and N-CZ800.
process.
Strategy for investigating the catalytic mechanism usually
involves preparing a series of catalysts with variable characteristics 2.2. Physical and chemical characterizations
(e.g., particle size, morphology, externally-exposed facets, or defect
concentration) and observing their catalytic behaviours using Specific surface area of samples was performed on a Quan-
in situ instruments. Three-dimensionally ordered macroporous tachrome Autosorb-iQ instrument. The data were calculated
(3DOM) materials, which possess periodic features with uniform according to the BET method by the N2 adsorption isotherm at liq-
large pore size (>50 nm), are proposed to benefit the catalytic uid N2 temperature (77 K).
activity resulting from the more abundant active sites on the acces- X-ray powder diffraction (XRD) patterns of the catalysts were
sibly macroporous wall surface [30]. The effect of catalyst features measured by a Rigaku diffractometer using Cu Ka radiation
on the catalytic behaviour may be magnified over the 3DOM cata- (k = 0.15406 nm). The operating voltage and current were 40 kV
lysts, which would make the processes able to respond to nuances and 200 mA. The scanning rate is 5°/min.
in its physicochemical properties. It is therefore reasonable to Raman spectra of the catalysts were recorded at RT in a Ren-
believe that the 3DOM CeO2-based catalysts would be ideal candi- ishaw Invia Raman imaging microscope. The exciting wavelength
dates to thoroughly investigate the relationship between their tex- was 514.5 nm from an Ar ion laser with a powder of ca. 10 mW
tural/reducible properties and catalytic activity. on the catalysts.
In the present work, two series of 3DOM CeO2 and CeO2-ZrO2 The morphology of the catalysts was observed by a scanning
solid solutions were prepared, which present different characteris- electron microscopy (SEM) by a Hitachi S4800 instrument with
tics in CeO2 particle size, oxygen vacancy concentration, reducibil- accelerating voltages of 3 kV.
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 367

The morphology of the catalysts was observed by transmission 2.3. Catalytic activity testing
electron microscopy (TEM) characterization, using a JEOL-2100
microscope operating at 200 kV. Catalytic activity evaluation was performed in a plug-flow,
The X-ray photoelectron spectroscopy (XPS) analysis was per- temperature-controlled microreactor system. The sample
formed on a PHI 5000 Versaprobe II system using a monochromatic (100 mg), loaded into the quartz tube with inner diameter of
Al-Ka X-rays source. Spectra were recorded at RT in vacuum 6 mm, was pretreated in flowing 10%O2/Ar (25 mL/min) at 400 °C
(<107 Pa). The C 1s signal at 284.8 eV was used for calibration of for 1 h. And then, it was cooled down to RT in O2/Ar atmosphere
the XP-signals. and switched to argon purging for 30 min thereafter. After that,
Temperature-programmed reduction (TPR) and temperature- the gas mixture (1% of CO, 10 vol% O2 balance Ar) was introduced
programmed desorption (TPD) profiles were obtained from RT to into the reactor with a flow of 50 mL/min at RT, and the system
900 °C. Experimental data were collected from Quantachrome was ramped up to 600 °C and cooled back to RT at a rate of
Instruments. The amount of catalyst either for TPR or for TPD test- 10 °C/min. The outlet gaseous products were analysed by a gas
ing was 100 mg using a heating rate 10 °C/min in all cases. After a chromatograph (Agilent 7890A GC System, produced by Agilent
standard cleaning pretreatment, TPR was carried out in a flow of Co.), which is equipped with HP-Plot 5A and HP-Plot-Q column.
10% H2/Ar (25 mL/min) up to 900 °C. For TPD testing, the pre-
treated sample was exposed to 5% O2/N2 (25 mL/min) at 100 °C 3. Results
for 30 min and then purged with N2 for 30 min. Then the sample
was heated up to 900 °C using a heating rate of 10 °C/min in all 3.1. Catalytic activity
cases.
CO-TPR over different catalysts (100 mg) was also conducted on Fig. 1 exhibits the CO oxidation activity over 3DOM CeO2 and
the CATLAB catalyst characterization system (produced by Hiden CeO2-ZrO2 and nonporous CeO2-ZrO2 samples. It is convenient to
Analytical Co., England). Before each of the experiments, the sam- compare the catalytic activities of the samples by adopting the
ple was pretreated in flowing 10% O2/Ar (25 mL/min) at 400 °C for reaction temperatures T10, T50, and T90 (corresponding to CO con-
1 h, cooled down to RT and then switched to argon purging for version of 10%, 50% and 90%, respectively), as summarized in
30 min. After that, the pretreated sample was exposed to 1%CO/ Table 1. For the series of 3DOM CeO2 samples, C450 is the most
Ar (25 mL/min) at RT for 30 min and then ramped (10 °C/min) up active, followed by C600 and C800. For example, the temperature
to 600 °C. The gas was analysed by an online mass spectrometer for T50 is 300 °C over C450, 340 °C over C600, and 366 °C over
(MS). C800. For the series of 3DOM CeO2-ZrO2 samples, CZ450 shows a
Diffuse reflectance infrared Fourier transform spectroscopy very low catalytic activity, while a drastic increase in the activity
(DRIFTS) analysis was used to evaluate the adsorbed species on occurs over the CeO2-ZrO2 sample preheated at 600 °C (CZ600
the catalyst under reaction conditions. Infrared spectra were sample). Further increasing the ageing temperature to 800 °C
recorded by an FTIR spectrometer (vertex 70, Bruker, Germany) would reduce the activity. It can be seen that a T50 is achieved at
equipped with a liquid N2 cooled Mercury-Cadmium-Telluride 255 °C for CZ600, while only 4% and 13% CO conversion is obtained
(MCT) detector. A Pike Technologies HC-900 DRIFTS cell with nom- at the same temperature for CZ450 and CZ800 samples, respec-
inal cell volume of 6 cm3 was used. Scans were collected from 4000 tively. The macroporous CeO2-ZrO2 samples present totally differ-
to 1000 cm1 at a resolution of 4 cm1. Catalyst samples were ent phenomenon compared with the 3DOM CeO2-ZrO2 samples, as
placed in a DRIFTS cell equipped with SeZn windows. Before each shown in Fig. 1B. Their activity for CO oxidation linearly decreases
of the following in situ DRIFTS experiments, the sample was pre- with increasing the calcination temperature. The N-CZ450 catalyst
treated in the DRIFTS cell in flowing 10%O2/Ar (25 mL/min) at oxidizes 50% of CO at a temperature of 345 °C, while the T50 values
400 °C for 1 h and then cooled to RT before switching to Ar. In for N-CZ600 and N-CZ800 catalysts are 370 and 420 °C, respec-
CO adsorption experiments, the pretreated sample was purged tively. This is a usual case for the CeO2-ZrO2 catalyst, because cal-
with Ar at RT before switching to 1%CO/Ar (25 mL/min) flow for cination at high temperature may result in sintering of the catalyst,
30 min. reducing the catalytic activity.
For the In situ DRIFTS CO-TPR experiments, the pretreated sam- It should be stressed that the CZ450 sample exhibits the lowest
ple was exposed to 1%CO/Ar (25 mL/min) at RT for 30 min and then catalytic activity, while the CZ600 sample shows the best catalytic
purged with argon at RT for 30 min. Then the sample was heated performance among the six samples. This suggests that some evo-
(10 °C/min) up to 600 °C in flowing argon (50 mL/min). lution of the structural and/or surface properties of the material

100 100
A B
T90 T90
80 80
CO conversion%
CO conversion%

60 60
T50 T50

40 C450 40
C600
C800 N-CZ450
20 CZ450 20 N-CZ600
T10 CZ600 T10 N-CZ800
CZ800
0 0
200 300 400 500 600 200 300 400 500 600
Temperature (oC) Temperature (oC)

Fig. 1. CO conversion as a function of reaction temperature over macroporous (A) and nonporous (B) samples.
368 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

Table 1
Catalytic activities, activation energies (Ea) and BET specific surface areas (SBET) of different samples.

Samples T10 (°C) T50 (°C) T90 (°C) Ea (kJ/mol) SBET (m2/g)
C450 240 300 355 62 42
C600 260 340 425 92 40
C800 300 366 473 112 38
CZ450 308 445 534 137 50
CZ600 200 255 343 57 46
CZ800 225 310 394 83 43
N-CZ450 270 345 445 73 87
N-CZ600 295 370 545 75 70
N-CZ800 325 420 600 94 50

-4 -4
C450
A C600
B CZ450
CZ600
-5 C800 -5
CZ800
CZ450
CZ600
-6 CZ800 -6
ln k

ln k
-7 -7

-8 -8

-9 -9
1.4 1.6 1.8 2.0 2.2 2.4 1.5 1.6 1.7 1.8 1.9 2.0
1000/T (K -1) 1000/T (K -1)

Fig. 2. Arrhenius plots for CO oxidation over macroporous (A) and nonporous (B) catalysts under the conditions of CO concentration = 1 vol%, CO/O2 molar ratio = 1/10, and
SV = 30,000 mL/(g h).

may occur when increasing the calcination temperature from 450 CeO2-ZrO2 samples. In addition, the stability of the CZ600 catalyst
to 600 °C. On the other hand, it is an exceptional phenomenon that is examined at 300 °C with a CO conversion at ca. 80%, and the
the CeO2-ZrO2 solid solutions (i.e., CZ450 sample) possess much results shown in Fig. S1. As can be seen, the CZ600 catalyst main-
lower catalytic activity than single CeO2 (i.e., C450 sample), tained its activity during 100 h testing.
because CeO2-ZrO2 solid solutions prepared in the same conditions
usually have smaller particle size and a higher oxygen vacancy 3.2. Structural properties of catalysts
concentration which prove to be beneficial to improving catalytic
activity. Further investigation of the difference among the above X-ray diffraction patterns of the different 3DOM CeO2 and CeO2-
different samples may lead to a better understanding of the CO oxi- ZrO2 catalysts are shown in Fig. 3. In the case of the bare CeO2, all
dation mechanism (involving the key factor to determine the activ- the diffraction peaks can be indexed to (1 1 1), (2 0 0), (2 2 0), (3 1 1),
ity and the reaction path) over CeO2-based catalysts. (4 0 0), (3 1 1) and (4 2 0) crystal faces, corresponding to a face-
The apparent activation energy (Ea) over different samples was centred cubic (fcc) fluorite structure of CeO2 [37]. No ZrO2 crystal
also obtained based on the catalytic activity. It is reported that the phase was detected in the XRD profiles of all the CeO2-ZrO2
reaction order was related to the concentration of CO and O2. samples, which indicates the formation of solid solutions. For
When the CO concentration is relatively low, the CO should be
first-order [31,32]. Jia et al. [33] also claimed that the oxidation
of CO obeys a reaction mechanism of first order towards CO con-
(111)

centration. Some works suppose that the oxidation of CO in the


presence of excess oxygen (CO/O2 molar ratio P1/2) would obey
(220)

(311)
(200)

a first-order reaction [34–36]. In the present work, the amount of


(331)
(420)
(222)

(400)

O2 is excess (CO/O2 molar ratio = 1/10). Therefore, it is reasonable


CZ600
Intensity (a.u.)

to propose that the oxidation of CO should obey a first-order reac-


tion with respect to CO concentration(c). r = kc = (A exp(Ea/ CZ450
RT))c, where r, k, A, and Ea are the reaction rate (mol/s), rate con-
stant (s1), pre-exponential factor, and apparent activation energy
(kJ/mol), respectively.
The Arrhenius plots for CO oxidation over the macroporous and
C600
nonporous samples are shown in Fig. 2, and their apparent activa-
tion energies are summarized in Table 1. It can be observed that
C450
the Ea value for CO oxidation decreased in the sequence CZ450
(137 kJ/mol) > C800 (112 kJ/mol) > N-CZ800 (94 kJ/mol) > C600 10 20 30 40 50 60 70 80
(92 kJ/mol) > CZ800 (83 kJ/mol) > N-CZ600 (75 kJ/mol) > N-CZ450 2 Theta (deg.)
(73 kJ/mol) > C450 (62 kJ/mol) > CZ600 (57 kJ/mol). This suggests
that CO oxidation proceeds more readily over the macroporous Fig. 3. XRD patterns of different 3DOM CeO2 and CeO2-ZrO2 samples.
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 369

Table 2 tions. This explanation should be that some isolated Zr species


Average crystal parameters of different 3DOM CeO2 and CeO2-ZrO2 samples. may exist in an amorphous state on the sample heated at low tem-
Samples Phase Crystallite size (nm) Lattice constant (nm) peratures, which could be incorporated into CeO2 lattice after ther-
CeO2 CeO2 mal treatment at relatively high temperatures.
C450 Cubic 15.2 0.5401 Raman spectra of the 3DOM CeO2 and CeO2-ZrO2 samples are
C600 Cubic 23.5 0.5410 shown in Fig. 4. The pure CeO2 sample exhibits a strong peak cen-
CZ450 Cubic 5.9 0.5365 tred at 462 cm1 corresponding to F2g Raman active mode in metal
CZ600 Cubic 7.8 0.5347
oxides with fluorite-like structure [39]. This peak broadens and
slightly shifts to 470 cm1 for the CeO2-ZrO2 samples. It has been
CeO2, the diffraction peaks become sharper and stronger after ther- reported that the broadening of the ceria main line is related either
mal treatment at 600 °C. This is a consequence of the better to a decrease of ceria crystallite size or to the presence of defects
arrangement of the atoms into the framework of pure CeO2 and such as oxygen vacancies [37]. For the CeO2-ZrO2 samples, a broad
the decreased number of lattice defects [38]. However, for the band centred at 601 cm1 can be observed. This band could be
CeO2-ZrO2, these changes are less remarkable, indicating a good linked to oxygen vacancies due to the substitution of Zr4+ into
thermal stability. the ceria lattice [40]. The appearance of a very weak band at
The average crystal parameters of the samples are listed in 300 cm1 could be related to the displacement of the oxygen atoms
Table 2. It is obvious that both the two CeO2-ZrO2 samples show from their ideal fluorite lattice positions [41]. To make a quantita-
much lower lattice constants for CeO2 compared with pure CeO2. tive analysis of the concentration of oxygen vacancies, the I601/I462
This lattice shrinkage should result from the smaller Zr4+ ion enter- ratios were calculated from the Raman spectra, as shown in Fig. 4B.
ing into the fluorite-type lattice of CeO2. It is also noted that the The I601/I462 ratio is only 0.04 and 0.02 for the C450 and C600
lattice constant for CeO2-ZrO2 samples (CZ450 vs CZ600) further respectively, while it is 0.24 for the CZ600 sample.
decreases when increasing the calcination temperature from 450 Fig. 5 shows the SEM images of the 3DOM CeO2 and CeO2-ZrO2
to 600 °C, suggesting greater formation of CeO2-ZrO2 solid solu- samples. It can be seen that the C450 and CZ450 samples have

0.25
A B
601
0.20
CZ600
Intensity (a.u.)

I 601 /I 462

CZ450 0.15
462
0.10

C600 0.05

C450
0.00
200 300 400 500 600 700 800 900 C450 C600 CZ450 CZ600
-1
Raman shift (cm ) Samples

Fig. 4. Raman spectra (A) and the I601/I462 ratios calculated from Raman spectra (B) of different 3DOM CeO2 and CeO2-ZrO2 samples.

Fig. 5. SEM images of 3DOM C450 (a1), CZ450 (b1), C600 (a2) and CZ600 (b2) samples.
370 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

open, periodic and interconnected three-dimensionally (3D) pore sizes of macroporous samples are 100 ± 20 nm, and the voids
frameworks. The pore sizes estimated from the SEM images are are interconnected through open windows which agrees with the
in the range of 100–150 nm. When the samples were further pre- observed by SEM images. The CeO2 particle size of C450 is in the
heated at 600 °C, there is no great difference in the morphologies range of 4–10 nm with a narrow distribution (Fig. 6-a2) and a
and textural structure among the two samples (C600 and CZ600). mean diameter of 6.0 nm by statistical analysis of more than 300
It indicates that the 3DOM ceria-based materials have higher ther- CeO2 particles. The 3DOM C600, CZ450 and CZ600 samples reveal
mal and structural stability, which are required for the further a mean particle diameter of 8.0, 5.1 and 5.9 nm, respectively (see
development of durable, highly active and versatile catalysts for Fig. 6-b2, c2, d2). In addition, the macroporous walls are composed
CO oxidation. of closely packed particles with limited microporous or meso-
Fig. 6 shows the TEM and HRTEM images of different 3DOM porous. The lattice fringes with a width of 3.1 Å and 1.9 Å indexed
CeO2 and CeO2-ZrO2 samples. The macroporous structure with as {1 1 1} and {1 1 0} planes of CeO2, respectively, can be found in all
overlapped pores can be clearly observed in the TEM images. The the samples (see Fig. 6-a3, b3, c3, d3). To identify the percentage of

Fig. 6. TEM and HRTEM images of 3DOM C450 (a1-3), C600 (b1-3), CZ450 (c1-3) and CZ600 (d1-3).
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 371

the exposed {1 1 0} plane, the ratio of {1 1 0}/{1 1 0 + 1 1 1} was cerium species calculated by the area ratios of the Ce3+/(Ce + Zr)
obtained by statistical analysis of more than 300 CeO2 particles and two oxygen species. The peaks labelled as V correspond to
for each sample. Parts of the TEM images used for statistical anal- Ce 3d5/2 contributions, and those of labelled as U represents the
ysis were plotted in Fig. S2. The ratio of {1 1 0}/{1 1 0 + 1 1 1} for Ce 3d3/2 contributions. The peaks labelled as V, V0 0 and V000 are
C450, C600, CZ450 and CZ600 samples is 23%, 20%, 8% and 45%, assigned to Ce(IV) final states, and the peak V0 is denoted as Ce
respectively. This suggests that there are more {1 1 0} planes on (III) final states. The same assignment can be applied to the U
the surface of CZ600 sample. structures, which correspond to the Ce 3d3/2: the line U0 is assigned
Fig. 7 shows the N2 adsorption-desorption isotherms of differ- to Ce(III) final state coordinated to oxygen vacancies and the other
ent 3DOM CeO2 and CeO2-ZrO2 samples. The CZ450 sample dis- lines indicative of the presence of Ce4+ ions [42–44].
plays a type II isotherm and a type H2 hysteresis loop in the The quantitative analysis (see Table 3) shows that the atom
relative pressure (P/P0) range of 0.2–0.85, which could be linked ratio of Ce/Zr for CZ450 is lower than that for the CZ600 sample
to capillary condensation taking place in mesopores, indicating (2.34 vs. 3.25), indicating that the CZ450 sample is enriched with
that textural mesopores existed within the skeletons. However, ZrO2 on the surface. Since CeO2 is the active component for CO oxi-
some of the mesopores are lost after calcination at higher temper- dation, this can be used to explain the relatively lower activity of
atures, which can be explained by grain growth. The two 3DOM CZ450 sample. In addition, it is also found that the average ratios
CeO2 samples presented a nonporous model, indicating that there of Ce3+/(Ce + Zr) decreased (see Table 3) in the order of
is almost no mesopores existing within the macroporous walls. CZ600 > C450 > C600 > CZ450 (which ranges from 0.13 to 0.18),
Compared with the CeO2 samples (surface area = 40–42 m2/g), which is completely consistent with the order of the catalytic
the macroporous CeO2-ZrO2 samples showed a slightly higher sur- activity over different samples. This indicates that Ce3+ concentra-
face area (46–50 m2/g). tion plays a very important role for determining the catalytic
XPS was performed to determine the chemical state of Ce and O property.
species on the surface of different samples. The oxidation states of To further confirm the importance of the oxygen mobility, oxy-
Ce were analysed by fitting the curves of Ce 3d XPS bands, as gen storage capacity (OSC) measurements were also performed by
shown in Fig. 8. Table 3 presents the relative percentages of the the successive CO pulse mode (see Fig. S3). The quantitative data
on the TOSC for different samples are shown in Table 3. The TOSC
values decrease in the order of CZ600 > C450 > C600 > CZ450
which is completely agreement with the order of catalytic activity.
C450
The O 1s spectra of pure CeO2 samples (C450 and C600) can be
C600
Volume Adsorbed(cm3/g)

fitted with three peaks at ca. 529.6 eV (main band, labelled as OIII),
CZ450
531.1 eV (labelled as OII) and 532.2 eV (labelled as OI), respec-
CZ600
tively. It is generally accepted that the 529.6 eV peak is character-
istic of the lattice oxygen (O2) in cerium oxide with the peak at
531.1 and 532.2 eV attributed to surface oxygen (Os) and/or surface
hydroxyl species OHs, respectively [45]. However, the O 1s spectra
of CeO2-ZrO2 samples (CZ450 and CZ600) can only be fitted with
two peaks (OII and OIII), as indicated by the disappearance of the
surface hydroxyl species (OI). It is noted that the lattice oxygen
(OIII) is the most abundant species for each sample. Since surface
oxygen species usually play a very important role in the catalytic
process, the surface-Oads.(OI + OII)/bulk-Olatt.(OIII) molar ratios
0.0 0.2 0.4 0.6 0.8 1.0 are calculated based on the XPS data. As shown in Table 3, the ratio
Relative Pressure (P/P0) decreases in the order of C450 > CZ600 > C600 > CZ450. However,
by ignoring the effect of surface hydroxyl species (OI), the
Fig. 7. Nitrogen adsorption-desorption isotherms of different 3DOM CeO2 and sequence of the Oads./Olatt. ratios among the four samples would
CeO2-ZrO2 samples. be as follows: CZ600 > C450 > CZ450 > C600.

Fig. 8. Ce 3d and O 1s XPS patterns of different 3DOM CeO2 and CeO2-ZrO2 samples.
372 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

Table 3
XPS-derived characteristics for different 3DOM CeO2 and CeO2-ZrO2 samples.

Samples Ce3+ Ce4+ Ce3+/Ce4+ Atom ratio of Ce/Zr Ce3+/(Ce+Zr) OI OII OIII (OI + OII)/OIII TOSCa mmol/g
C450 17 83 0.20 – 0.17 10 19 71 0.41 1.01
C600 15 85 0.18 – 0.15 7 13 80 0.25 0.93
CZ450 19 81 0.23 2.34 0.13 – 16 84 0.19 0.77
CZ600 24 76 0.32 3.25 0.18 – 22 78 0.28 1.12
a
TOSC is determined by summarizing OSC of 20 CO pulses at T = 400 °C.

Fig. 9. H2-TPR profiles (A) and cumulative H2 uptake (B) and initial H2 consumption rate (C) of different 3DOM CeO2 and CeO2-ZrO2 samples.

3.3. Reducibility and oxygen mobility

Fig. 9A shows the H2-TPR profiles of the four samples, and O2-TPD
Fig. 9B exhibits the cumulative hydrogen uptake associated with
the phases present in the samples as a function of reduction tem- CZ600
perature. For pure CeO2, two reduction peaks are detected in the
H2-TPR profile: a low temperature peak at 580 °C and a high tem- CZ450
perature peak in a range higher than 800 °C corresponding to the
TCDsignal (a.u.)

reduction of surface and bulk ceria, respectively. The CeO2-ZrO2


samples only show one obvious reduction peak, centred at about
C600
656 °C, and the intensity of this peak is much higher than either
of the two peaks over the pure CeO2 sample. This suggests that C450
oxygen located within the subsurface or bulk of CeO2-ZrO2 samples
could migrate to the surface immediately as surface oxygen is con-
sumed, indicative of relatively high lattice oxygen mobility [38].
The total hydrogen consumption of CeO2-ZrO2 samples
(123 mmol/g for CZ450 and 134 mmol/g for CZ600) is much higher
100 200 300 400 500 600 700 800 900
than that of CeO2 (68 mmol/g for C450 and 42 mmol/g for C600). It
is generally accepted that the low-temperature reducibility of a Temperature (οC)
catalyst can be conveniently evaluated using the initial H2 con-
Fig. 10. O2-TPD profiles of different 3DOM CeO2 and CeO2-ZrO2 samples.
sumption rate where less than 25% oxygen in the sample is
removed for the first reduction peak [46]. As shown in Fig. 9C,
CZ600 presents the highest H2 consumption rate among the four profiles, which indicates that there was no clear distinction
samples. between the desorption of surface and bulk oxygen. Because the
To further investigate the adsorption and activation of oxygen oxygen in the subsurface can quickly migrate to the surface as
on CeO2 and CeO2-ZrO2, O2-TPD measurements were carried out, the surface oxygen is released indicates a relatively high lattice
as shown in Fig. 10. The O2 desorption peaks of the pure CeO2 oxygen mobility [38].
samples (C450 and C600) can be divided into three parts in the Fig. 11A shows the CO2 profiles evolved during CO-TPR over dif-
temperature ranges of 80–250 °C (low-temperature), 250–500 °C ferent 3DOM CeO2 and CeO2-ZrO2 samples. Both the two CeO2
(middle-temperature), and 500–900 °C (high-temperature), samples exhibit three weak reduction peaks with the increase of
respectively. The low-temperature peak is assigned to the the temperature. For the two CeO2-ZrO2 samples, only a strong
desorption of physically adsorbed oxygen (O2), and the middle- band is observed in the range of 240–600 °C. The production tem-
temperature peak can be associated with the consumption of che- perature of CO2 in the CO-TPR is in general alignment with the
misorbed surface-active oxygen species [47]. The high temperature trend of the catalytic activity (Fig. 1A), which is an indication that
must be related to the release of lattice oxygen. For the CeO2-ZrO2 the lattice oxygen participates in CO oxidation on 3DOM CeO2 and
samples, the peaks at temperatures lower than 300 °C are very CeO2-ZrO2 samples. H2, as a byproduct, is also observed during
weak and only one main peak in the range of 300–900 °C is CO-TPR over the four samples, as shown in Fig. 11B. This is similar
observed. This is similar to the phenomena observed in the TPR to the observation by Wu et al. [24]. It is proposed that the forma-
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 373

Fig. 11. CO2 (A) and H2 (B) evolution during CO-TPR over different 3DOM CeO2 and CeO2-ZrO2 samples. (C) In situ DRIFT spectra collected during CO-TPR of C450 at different
temperature.

Fig. 12. In situ DRIFT spectra of CO adsorption on different 3DOM CeO2 and CeO2-ZrO2 samples at RT.

tion of H2 is mainly due to the reaction between adsorbed surface OH groups on the surface of CeO2. On the other hand, the 3DOM
OH groups and CO via water-gas shift type reactions [48]: CeO2-ZrO2 samples (CZ450 and CZ600) produce much more CO2
than the pure CeO2 in the CO-TPR testing. This suggests that the
CO þ OH ! 1=2H2 þ CO2 ð1Þ
CO2 formation over the 3DOM CeO2-ZrO2 samples should be due
to direct CO reaction with reactive lattice oxygen instead of the
CO þ 2OH ! H2 þ CO2 þ OL ð2Þ
water-gas shift reaction between CO and hydroxyl groups. This
To investigate the evolution of surface OH groups during CO- must be associated with relatively high reducibility and oxygen
TPR of the samples up to 600 °C, the in situ DRIFT spectra of C450 mobility of the 3DOM CeO2-ZrO2 samples.
were employed, as shown in Fig. 11C. It can be seen that, bands
characteristic of formate (2842 cm1) derived from reaction 3.4. Interaction with CO
between CO and surface OH groups start to appear at 300 °C. For-
mate species is an intermediate in water-gas shift type reactions. DRIFT spectroscopy technology was used to investigate CO
With the temperature increase, the formate decomposes to H2 adsorption and oxidation behaviour on these 3DOM CeO2 and
and CO2. CeO2-ZrO2 samples. Fig. 12 shows the in situ DRIFT spectra from
It should be noted that the intensity of H2 peaks over pure CeO2 CO adsorption on different samples at RT as a function of time after
is much higher than that over CeO2-ZrO2 samples, indicating more switching in the CO feed. Based on the investigations of CO and CO2
374 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

Table 4
Assignments of in situ DRIFT spectra observed in the CO adsorption at room temperature over the 3DOM CeO2 and CeO2-ZrO2 samples.

Species v(CO3) p(CO3) d(OH) v(OH)


Bicarbonate 1420, 1626 864 1216 3619
Unidentate carbonate 1420, 1360 – – –
Bidentate carbonate 1036, 1270, 1560 864 – –
Bridged carbonate 1036, 1310, 1695 – – –

adsorption on ceria [24,48–51], CO interacts strongly with the sur- shifted towards higher frequencies. However, the p-back bond is
face oxygen and forms a variety of carbonate species in the range formed at the expense of electron transfer from the d electrons
from 800 to 1800 cm1, and the assignments of the observed vibra- of the metal cation to the 2p⁄ antibonding orbitals of the CO mole-
tional bands are summarized in Table 4. Bicarbonate species, cule [58,59]. As a result, the metalAC bond becomes stronger (i.e.
unidentate carbonate species, bidentate carbonate species, bridged the carbonyls are more stable), which reflects in a red shift of the
carbonate species are formed upon the interaction of CO with the CAO stretching vibrations [57]. Therefore, the lower band at
surface oxygen of CeO2. Similar carbonate species are also observed 2007 cm1 can be attributed to the enhanced CeACO bond, indicat-
on the surface of the two CeO2-ZrO2 samples, while the corre- ing a stronger interaction between CeO2 and CO. It can be seen in
sponding peaks are stronger for the CZ600 sample. It indicates that Fig. 12 that the CZ600 sample shows relatively high intensity of
CZ600 possesses more CO active sites [10]. peaks in the ranges of both 800–1800 cm1 and 2007 cm1, which
In addition to the carbonate species (wave numbers in the range indicates that this sample could strongly interact with the CO
from 800 to 1800 cm1), the bands at 2000–2200 cm1 were also molecules.
distinguished. Bands in the 2000–2200 cm1 region can be associ- Fig. 13 shows the in situ DRIFT spectra obtained as the samples
ated with linear or bridge-bonded CO interacting with sites on the are heated in the reactant mixture. Three major types of carbonate
surfaces. A bridged carbonyl species will adsorb at lower wave species were detected over the samples: bicarbonate species (864,
numbers than a linear carbonyl species [52]. The vibrational peaks 1440, 1626 and 3619 cm1), unidentate carbonate species (1440–
at ca. 2172 and 2117 cm1 are ascribed to CO adsorbed on the Ce 1360 cm1), and bidentate carbonate species (864, 1036, 1260–
sites [53]. 1290 and 1563 cm1). The bicarbonate species (major peak at
The bands with a wave number lower than 2100 cm1 are often 1626 cm1) gradually disappeared, while the bidentate carbonate
detected on noble metal catalysts (Au or Pt catalysts), which can be (major peak at 1523–1563 cm1) species began to increase to a
assigned to negatively charged metal clusters or metal atoms [54]. great extent when the reaction temperature increased to 200 °C.
The presence of the band at 2007 cm1 should be part of the wave On the other hand, gaseous CO2 peaks (2293 and 2356 cm1)
number at 2117 cm1 shifted to the lower wave number. The shift started to appear at 200 °C and kept increasing in the intensity
of CO vibrational frequency is determined by the d and p-back with temperature. This is consistent with the increasing of CO con-
bonds orbitals [55–57]. When only a d bond is formed, the stability version (see Fig. 1A). These phenomena can be summarized as fol-
of the carbonyls increases with the effective positive charge of the lows: bicarbonate species first transformed into bidentate
cations, and, correspondingly, the CO vibrational frequency is carbonate species (below 200 °C), followed by the decomposition

Fig. 13. In situ DRIFT spectra of CO oxidation on different 3DOM CeO2 and CeO2-ZrO2 samples.
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 375

120 50 120 100


A B

Relative intensity (a.u.)


CO conversion at 673 K

Relative intensity (a.u.)


CO conversion at 673 K surface areas
ratio of (110) oxygen vacancy Crystallite Size

CO conversion / %
CO conversion / %
100 40 100 80

80 30 80 60

60 20 60 40

40 10 40 20

20 0 20 0
C450 C600 CZ450 CZ600 C450 C600 CZ450 CZ600
Samples Samples

Fig. 14. The relationship between the catalytic activities of CO oxidation and physicochemical properties of different catalysts (the relative intensity is calculated from that
amplify the same multiple on the original data).

of bidentate carbonate species into gaseous CO2 with increasing showed that the TOF (converted CO per surface O per second) of CO
temperature. The band intensity of carbonate species over CeO2 oxidation follows the order of {1 1 0} > {1 0 0}  {1 1 1}. The TEM
samples, especially C600, is much higher than that over CeO2- images show that all the 3DOM CeO2 and CeO2-ZrO2 samples
ZrO2 samples in the range of 200–400 °C. It is reported that the expose both {1 1 0} and {1 1 1} surfaces (see Fig. 6). The statistical
accumulation of carbonate species could cover the CO active sites, analysis based on the HRTEM images suggests that the ratios of
leading to a reversible deactivation [10,60]. Therefore the C600 and {1 1 0}/{1 1 0 + 1 1 1} decrease in the order of CZ600 > C450 >
C450 samples show lower activity than CZ600. C600 > CZ450 (see Figs. 6 and S2). This order is consistent with
the sequence of catalytic activity (see Fig. 1).
The relationship between the catalytic activities (CO conver-
4. Discussion sion) for CO oxidation and physicochemical properties of different
catalysts is summarized in Fig. 14. It can be seen that, for pure CeO2
4.1. Dominant factors for CO oxidation over CeO2-based catalysts samples (C450 and C600), the increase in particle size and the
decrease in oxygen vacancy concentration would reduce the cat-
It is generally accepted that the importance of ceria in catalysis alytic activity. We also note that this influence occurs with the pre-
originates from its extraordinary efficiency for reversible oxygen condition that there are no changes of the preferentially exposed
release via the rapid formation and elimination of oxygen vacan- surface plane over the two pure CeO2 samples (ratios of {1 1 0}/
cies [61]. The presence of oxygen vacancies provides relative free- {1 1 0 + 1 1 1} for C450 and C600 are 23% and 20%, respectively, as
dom for the movement of lattice oxygen and thus increases the shown in Fig. 6). For the two CeO2-ZrO2 samples (CZ450 and
mobility of oxygen [62]. O2 can adsorb on the vacancies, which CZ600), the sample heated at 600 °C (CZ600 sample) showed more
could reduce surface Gibbs energy, balance surface charge, and abundant oxygen vacancy and higher catalytic activity. This can be
form surface active oxygen species [11]. This would promote oxy- ascribed to the relatively higher ratios of {1 1 0}/{1 1 0 + 1 1 1} over
gen reactivity to CO in the catalytic process [15]. The results in the the CZ600 sample. CZ600 also revealed lower activation energies
present work reveal a similar phenomenon. The XRD, Raman spec- for CO oxidation (see Table 1), which should be linked to the bind-
tra and XPS results (see Figs. 3, 4 and 8 and Tables 2 and 3) indicate ing energies of CO or O (O2) with the exposed surface plane. Chen
that more oxygen vacancies are detected on the surface of CZ600 et al. [29] suggested that the barrier for the CO to react with a sur-
compared with the other samples, resulting in relatively high face O at CeO2 {1 1 0} is 0.25 eV, while it is 0.61 eV for the CO to
reducibility and oxygen mobility (observed by H2-TPR and react with a surface O at CeO2 {1 1 1}. In other words, the O at
O2-TPD, as shown in Figs. 9 and 10, respectively). The CZ600 shows CeO2 {1 1 0} is more active compared to that at CeO2 {1 1 1}. The fact
much higher catalytic activity for CO oxidation than the other three that the {1 1 0} surface is more reactive for promoting CO oxidation
samples (see Fig. 1). This suggests that both the oxygen vacancies can also be seen from the overall CO oxidation energy, which was
and oxygen mobility are important for catalytic CO oxidation. calculated to be exothermic by 1.1 eV at the {1 1 0} surface and
The textural properties such as the crystallite size and specific only 0.4 eV at the {1 1 1} surface (with respect to the noninteract-
surface area also influence the catalytic performance for CO oxida- ing CO-CeO2 systems). Nolan and Watson [28] reported that the
tion. Fan et al. [18] reported that the higher specific surface area interaction with CO molecule on the CeO2 {1 1 1} surface leads to
and smaller crystallite size of the catalyst can enhance the catalytic a weak adsorption, with an adsorption energy of 6 kcal/mol.
performance for CO oxidation. The CZ450 is also an exception in However, for the {1 1 0} surfaces, the adsorption energy is much
this respect. It can be seen in Table 1 that the specific surface areas higher (45 kcal/mol), indicating a relatively strong interaction
of 3DOM CZ450 are higher than or equal to those of the other between adsorbed CO and CeO2 {1 1 0} surface. Fig. 14A also
3DOM CeO2 or CeO2-ZrO2 samples. In addition, it also possesses a showed that the ratios of {1 1 0}/{1 1 0 + 1 1 1} and oxygen vacancy
smaller crystallite size of CeO2, as shown in Table 2. However, concentration over different samples followed the same trend as in
the CZ450 sample shows the worst catalytic performance for CO the values of CO conversion for CO oxidation. This phenomenon
oxidation among all the samples. demonstrates that the preferentially exposed surface plane and
Surface structure dependence of CeO2 catalysts for CO oxidation oxygen vacancy concentration are crucial factor in determining
is also confirmed by intensive experimental and theoretical studies catalytic activity.
[63–66]. To get a greater understanding of the nature of different
CeO2 surface planes in the catalytic process, Wu et al. [24] system- 4.2. CO adsorption and reaction mechanism
atically investigated CO oxidation over a series of ceria nanocrys-
tals with defined surface planes (nanoshapes) including rods As observed in the CO-TPR process (Fig. 11), surface water-gas
({1 1 0} + {1 0 0}), cubes ({1 0 0}), and octahedra ({1 1 1}). The results shift reaction between CO and surface OH groups was detected.
376 Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377

Scheme 1. Reaction mechanism model of CO oxidation over 3DOM CeO2-ZrO2 catalysts.

The results of in situ DRIFTS during CO-TPR (Fig. 11C) show the On the other hand, the CO-TPR result (see Fig. 11A) reveals that,
presence of formate species, which is indentified as an intermedi- in the absence of O2, the 3DOM CeO2 or CeO2-ZrO2 can be reduced
ate from CO to carbonate [48]. Jardim et al. [53] also reported that by CO during the similar temperature region with the catalytic CO
the presence of surface OH groups on the catalyst could enhance oxidation. This indicates that a Mars-van Krevelen (M-K) mecha-
the catalytic activity for CO oxidation. In our results, there are nism based on the redox property of CeO2 or CeO2-ZrO2 catalysts
abundant OH groups on the surface of 3DOM CeO2 sample (Figs. 8 also cannot be ignored. In a Mars-van Krevelen process, CO reacts
and 11), while low catalytic activity for CO oxidation observed over with lattice oxygen of ceria to form CO2 and leave an oxygen
the CZ600 sample. By contrast, the CZ600 shows the best catalytic vacancy, and then the replenishing of the vacancy by gas-phase
performance, although limited OH groups can be detected on the oxygen completes the cycle (see Scheme 1). However, it should
surface of CeO2-ZrO2 samples. In addition, Fig. S5 shows the curves not be crucial compared with the L-H mechanism, because there
of CO conversion over the C450 catalyst both in the heating and is no obvious relationship between the reducibility and the cat-
cooling processes in CO oxidation. As seen, there are no obvious alytic activity among the different samples.
differences between the heating and cooling processes. Since the
fresh catalyst has abundant OH groups, these phenomena indicate
that the surface OH groups should not be a crucial factor in deter- 5. Conclusions
mining the catalytic activity.
The results of in situ DRIFTS during CO adsorption (Fig. 12) show A series of 3DOM CeO2 and CeO2-ZrO2 catalysts with varying
the presence of two types of CO species: adsorbed CO on surface Ce particle sizes, oxygen vacancy concentrations, oxygen mobility
sites (the band at ca. 2007 cm1) and CO chemically bonded to sur- and preferentially exposed surface planes were synthesized by a
face oxygen forming carbonate species. The former species were colloidal crystal template method. Their physicochemical proper-
often detected on the supported noble metal catalysts, indicating ties were associated with the catalytic performances to identify
a strong interaction between CO and catalysts [54]. The high inten- the most crucial factor in determining catalytic activity. The inter-
sity of the band at 2007 cm1 on the CZ600 sample suggests strong actions of catalysts with CO and reaction mechanism were further
interaction between CO and CZ600. This can be related to the fact investigated in detail. The prepared 3DOM CeO2 and CeO2-ZrO2
that there are more {1 1 0} surfaces exposed on this sample, materials were highly ordered and interconnected with each other
because Ce is more active on {1 1 0} surface than on {1 1 1} [67]. by small pore windows, and the good macroscopic order still
On the other hand, the Raman (Fig. 4) and XPS (Fig. 8) results retained even after ageing at 800 °C, indicating a high thermal sta-
indicate that the CZ600 sample also possesses abundant oxygen bility of the macroporous structure. Such 3DOM framework is ben-
vacancies, which can adsorb O2 molecules. In this case, a eficial for CO oxidation. One of the most important phenomena is
Langmuir-Hinshelwood (L-H) mechanism should occur over the that CZ450 sample, which possesses higher surface area and smal-
CZ600 catalyst. According to this mechanism, the CO molecules ler particle size compared with the pure CeO2, showed much
firstly adsorb on the catalyst surface and react with the adsorbed poorer catalytic activity for CO oxidation. By contrast, the CZ600
oxygen thereafter, followed by the formed CO2 desorbing into the sample presented the highest catalytic activity among all the cata-
gas phases (see Scheme 1). It should be noted that large amounts lysts due to the highest ratio of {1 1 0}/{1 1 0 + 1 1 1} and oxygen
of carbonate species were detected on the C600 sample during vacancy concentration. The ratios of {1 1 0}/{1 1 0 + 1 1 1} and oxy-
the CO catalytic oxidation (see Fig. 13), indicating the accumula- gen vacancy concentration over different samples follow the same
tion of carbonate species. The accumulation should be related to trend as the CO conversion. These results suggest that the prefer-
the relatively low concentration of oxygen vacancy on this sample, entially exposed surface plane and oxygen vacancy concentration
because less O2 can be adsorbed on the vacancies to form active are more crucial for determining the catalytic activity of CeO2-
oxygen to remove the adsorbed CO species. This results in rela- based catalysts when compared with the other parameters (e.g.,
tively lower catalytic activity of this sample. crystallite size and surface area).
Y. Zheng et al. / Journal of Catalysis 344 (2016) 365–377 377

Langmuir-Hinshelwood mechanism should be the crucial reac- [24] Z.L. Wu, M.J. Li, S.H. Overbury, J. Catal. 285 (2012) 61.
[25] E. Aneggi, J. Llorca, M. Boaro, A. Trovarelli, J. Catal. 234 (2005) 88.
tion pathway for CO oxidation over the 3DOM CeO2-based cata-
[26] K.B. Zhou, X. Wang, X.M. Sun, Q. Peng, Y.D. Li, J. Catal. 229 (2005) 206.
lysts, although the Mars-van Krevelen mechanism cannot be [27] H.X. Mai, L.D. Sun, Y.W. Zhang, R. Si, W. Feng, H.P. Zhang, H.C. Liu, C.H. Yan, J.
ignored. It is also observed that the two CeO2 samples possess large Phys. Chem. B 109 (2005) 24380.
amounts of carbonate species, while they show a low catalytic [28] M. Nolan, G.W. Watson, J. Phys. Chem. B 110 (2006) 16600.
[29] F. Chen, D. Liu, J. Zhang, P. Hu, X.Q. Gong, G.Z. Lu, Phys. Chem. Chem. Phys. 14
activity for CO oxidation. This can be attributed to the accumula- (2012) 16573.
tion of carbonate species on the samples which may cover the [30] S.H. Xie, J.G. Deng, S.M. Zang, H.G. Yang, G.S. Guo, H. Arandiyan, H.X. Dai, J.
CO active sites, leading to a reversible deactivation. Catal. 322 (2015) 38.
[31] M. Boaro, F. Giordano, S. Recchia, V.D. Santo, M. Giona, A. Trovarelli, Appl.
Catal. B 52 (2004) 225.
Acknowledgments [32] C.S. Polster, H. Nair, C.D. Baertsch, J. Catal. 266 (2009) 308.
[33] A.P. Jia, G.S. Hu, L. Meng, Y.L. Xie, J.Q. Lu, M.F. Luo, J. Catal. 289 (2012) 199.
[34] V.P. Santos, S.A.C. Carabineiro, J.J.W. Bakker, O.S.G.P. Soares, X. Chen, M.F.R.
The authors acknowledge the fruitful discussions with Prof. Pereira, J.J.M. Órfão, J.L. Figueiredo, J. Gascon, F. Kapteijn, J. Catal. 309 (2014)
Robert (Bob) Farrauto, Columbia University. This work was 58.
supported by the National Natural Science Foundation of China [35] S.H. Xie, H.X. Dai, J.G. Deng, Y.X. Liu, H.G. Yang, Y. Jiang, W. Tan, A.S. Ao, G.S.
Guo, Nanoscale 5 (2013) 11207.
(Nos. 51604137, 51374004 and 51204083), the Candidate Talents [36] X.W. Li, H.X. Dai, J.G. Deng, Y.X. Liu, S.H. Xie, Z.X. Zhao, Y. Wang, G.S. Guo, H.
Training Fund of Yunnan Province (Nos. 2012HB009, Arandiyan, Chem. Eng. J. 228 (2013) 965.
2014HB006), an Applied Basic Research Program of Yunnan Pro- [37] G. Zhang, Z. Zhao, J. Xu, J. Zheng, J. Liu, G. Jiang, A. Duan, H. He, Appl. Catal. B
107 (2011) 302.
vince (No. 2014FB123), a School-Enterprise Cooperation Project
[38] I. Atribak, A. Buenolopez, A. Garciagarcia, J. Catal. 259 (2008) 123.
from Jinchuan Corporation (No. Jinchuan 201115) and a Talents [39] Y.C. Wei, J. Liu, Z. Zhao, A.J. Duan, G.Y. Jiang, J. Catal. 287 (2012) 13.
Training Program of Kunming University of Science and Technol- [40] B.M. Reddy, A. Khan, J. Phys. Chem. B 107 (2003) 11475.
ogy (No. KKZ3201352038). [41] K.Z. Li, H. Wang, Y.G. Wei, D.X. Yan, Appl. Catal. B 97 (2010) 361.
[42] G.L. Markaryan, L.N. Ikryannikova, G.P. Muravieva, A.O. Turakulova, B.G.
Kostyuk, E.V. Lunina, V.V. Lunin, E. Zhilinskaya, A. Aboukais, Colloid. Surface A
Appendix A. Supplementary material 151 (1999) 435.
[43] R.C.R. Neto, M. Schmal, Appl. Catal. A 450 (2013) 131.
[44] M.D. Hernández-Alonso, A.B. Hungría, A. Martínez-Arias, M. Fernández-García,
Supplementary data associated with this article can be found, in J.M. Coronado, J.C. Conesa, J. Soria, Appl. Catal. B 50 (2004) 167.
the online version, at http://dx.doi.org/10.1016/j.jcat.2016.10.008. [45] J.D. Grunwaldt, C. Kiener, C. Wögerbauer, A. Baiker, J. Catal. 181 (1999) 223.
[46] H.X. Dai, A.T. Bell, E. Iglesia, J. Catal. 221 (2004) 491.
[47] C. Ma, Z. Mu, J. Li, Y. Jin, J. Cheng, G. Lu, Z. Hao, S. Qiao, J. Am. Chem. Soc. 132
References (2010) 2608.
[48] H.Q. Zhu, Z.F. Qin, W.J. Shan, W.J. Shen, J.G. Wang, J. Catal. 225 (2004) 267.
[1] S. Najafishirtari, P. Guardia, A. Scarpellini, M. Prato, S. Marras, L. Manna, M. [49] O. Pozdnyakova, D. Teschner, A. Wootsch, J. Krohnert, B. Steinhauer, H. Sauer,
Colombo, J. Catal. 338 (2016) 115. L. Toth, F.C. Jentoft, A. Knop-Gericke, Z. Paal, R. Schlogl, J. Catal. 237 (2006) 1.
[2] Y. Xu, L. Chen, X.C. Wang, W.T. Yao, Q. Zhang, Nanoscale 7 (2015) 10559. [50] C. Li, Y. Sakata, T. Arai, K. Domen, K.I. Maruya, T. Onishi, J. Chem. Soc., Faraday.
[3] K. Ding, A. Gulec, A.M. Johnson, N.M. Schweitzer, G.D. Stucky, L.D. Marks, P.C. Trans. I 85 (1989) 1451.
Stair, Science 385 (2015) 189. [51] T. Shido, Y. Iwasawa, J. Catal. 136 (1992) 493.
[4] C.K. Costello, J. Guzman, J.H. Yang, Y.M. Wang, M.C. Kung, B.C. Gates, H.H. Kung, [52] A. Dandekar, M.A. Vannice, J. Catal. 178 (1998) 621.
J. Phys. Chem. B 108 (2004) 12529. [53] E.O. Jardim, S. Rico-Francés, F. Coloma, J.A. Anderson, J. Silvestre-Albero, A.
[5] P. Venkataswamy, K.N. Rao, D. Jampaiah, B.M. Reddy, Appl. Catal. B 162 (2015) Sepúlveda-Escribano, J. Colloid. Interf. Sci. 443 (2015) 45.
122. [54] F. Romero-Sarria, L.M. Martínez, T.M.A. Centeno, J.A. Odriozola, J. Phys. Chem. C
[6] L.Q. Liu, F. Zhou, L.G. Wang, X.J. Qi, F. Shi, Y.Q. Deng, J. Catal. 274 (2010) 1. 111 (2007) 14469.
[7] J.Y. Luo, M. Meng, X. Li, X.G. Li, Y.Q. Zha, T.D. Hu, Y.N. Xie, J. Zhang, J. Catal. 254 [55] A. Hornés, P. Bera, A.L. Cámara, D. Gamarra, G. Munuera, A.M. Arias, J. Catal.
(2008) 310. 268 (2009) 367.
[8] H.Y. Kim, G. Henkelman, J. Phys. Chem. Lett. 3 (2012) 2194. [56] D. Scarano, S. Bordiga, C. Lamberti, G. Spoto, G. Ricchiardi, A. Zecchina, C. Otero
[9] M.Q. Shen, L.F. Lv, J.Q. Wang, J.X. Zhu, Y. Huang, J. Wang, Chem. Eng. J. 255 Areán, Surf. Sci. 411 (1998) 272.
(2014) 40. [57] K.I. Hadjiivanov, M. Kantcheva, D.G. Klissurski, J. Chem. Soc. Faraday Trans. 92
[10] S. Zhang, X.S. Li, B.B. Chen, X.B. Zhu, C. Shi, A.M. Zhu, ACS Catal. 4 (2014) 3481. (1996) 4595.
[11] L. Meng, A.P. Jia, J.Q. Lu, L.F. Luo, W.X. Huang, M.F. Luo, J. Phys. Chem. C 115 [58] A. Davydov, I R Spectroscopy Applied to Surface Chemistry of Oxides, Nauka,
(2011) 19789. Novosibirsk, 1984.
[12] P. Bera, A. Gayen, M.S. Hegde, N.P. Lalla, L. Spadaro, F. Frusteri, F. Arena, J. Phys. [59] N.S. Ahmetov, General and Inorganic Chemistry, Nauka, Moscow, 1981.
Chem. B 107 (2003) 6122. [60] E.D. Río, S.E. Collins, A. Aguirre, X.W. Chen, J.J. Delgado, J.J. Calvino, S. Bernal, J.
[13] Y. Nagai, T. Hirabayashi, K. Dohmae, N. Takagi, T. Minami, H. Shinjoh, S. Catal. 316 (2014) 210.
Matsumoto, J. Catal. 242 (2006) 103. [61] F. Esch, S. Fabris, L. Zhou, T. Montini, C. Africh, P. Fornasiero, G. Comelli, R.
[14] U. Oran, D. Uner, Appl. Catal. B 54 (2004) 183. Rosei, Science 309 (2005) 752.
[15] C. Bozo, N. Guilhaume, J.M. Herrmann, J. Catal. 203 (2001) 393. [62] Y. Madier, C. Descorme, A.M.L. Govic, D. Duprez, J. Phys. Chem. B 103 (1999)
[16] Z. Wang, Q. Wang, Y.C. Liao, G.L. Shen, X.Z. Gong, N. Han, H.D. Liu, Y.F. Chen, 10999.
Chem. Phys. Chem. 12 (2011) 2763. [63] I. Florea, C. Feral-Martin, J. Majimel, D. Ihiawakrim, C. Hirlimann, O. Ersen,
[17] H.J. Wu, L.D. Wang, Catal. Commun. 12 (2011) 1374. Cryst. Growth Des. 13 (2013) 1110.
[18] J. Fan, D. Weng, X.D. Wu, X.D. Wu, R. Ran, J. Catal. 258 (2008) 177. [64] G.Q. Yi, Z.N. Xu, G.C. Guo, K. Tanaka, Y.Z. Yuan, Chem. Phys. Lett. 479 (2009)
[19] G.C. Bond, D.T. Thompson, Gold Bull. 33 (2000) 41. 128.
[20] D.R. Schryer, B.T. Upchurch, B.D. Sidney, G.B. Hoflund, R.K. Herz, J. Catal. 130 [65] X.S. Huang, H. Sun, L.C. Wang, Y.M. Liu, K.N. Fan, Y. Cao, Appl. Catal. B 90 (2009)
(1991) 314. 224.
[21] M. Sheintuch, J. Schmidt, Y. Lecthman, G. Yahav, Appl. Catal. 49 (1989) 55. [66] Y. Chen, P. Hu, M.H. Lee, H.F. Wang, Surf. Sci. 602 (2008) 1736.
[22] A. Trovarelli, Catal. Rev. - Sci. Eng. 38 (1996) 439. [67] M.V. Ganduglia-Pirovano, A. Hofmann, J. Sauer, Surf. Sci. Rep. 62 (2007) 219.
[23] G.S. Zafiris, R.J. Gorte, J. Catal. 143 (1993) 86.

You might also like