You are on page 1of 16

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/311153562

Rapid Microwave Assisted Sol-Gel Synthesis of


CeO2 and CexSm1-xO2 Nanoparticle Catalysts
for CO Oxidation

Article in Journal of Molecular Catalysis A Chemical · November 2016


DOI: 10.1016/j.molcata.2016.11.039

CITATIONS READS

7 209

8 authors, including:

Kyriaki Polychronopoulou Marios S. Katsiotis


Khalifa University of Science Technology & R… TITAN Cement S.A.
100 PUBLICATIONS 1,374 CITATIONS 49 PUBLICATIONS 201 CITATIONS

SEE PROFILE SEE PROFILE

Ayesha Khouri Siham Alqaradawi


Khalifa University of Science Technology & R… Qatar University
1 PUBLICATION 7 CITATIONS 44 PUBLICATIONS 499 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Three-dimensional manifestation of 2D materials for sensing and energy storage applications View
project

Energy Conversion View project

All content following this page was uploaded by Siham Alqaradawi on 30 September 2017.

The user has requested enhancement of the downloaded file.


Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


journal homepage: www.elsevier.com/locate/molcata

Rapid microwave assisted sol-gel synthesis of CeO2 and Cex Sm1-x O2


nanoparticle catalysts for CO oxidation
K. Polychronopoulou a,∗ , Abdallah F. Zedan b , M.S. Katsiotis c , M.A. Baker d , A.A. AlKhoori a ,
Siham Y. AlQaradawi b , S.J. Hinder d , Saeed AlHassan c
a
Department of Mechanical Engineering, Khalifa University of Science, Technology and Research, Abu Dhabi, P.O. Box 127788, UAE
b
Department of Chemistry and Earth Sciences, Qatar University, Doha 2713, Qatar
c
Chemical Engineering Department, The Petroleum Institute, Abu Dhabi, UAE
d
The Surface Analysis Laboratory, Faculty of Engineering and Physical Sciences, University of Surrey, Guildford GU2 4DL, UK

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

Article history: CeO2 and Cex Sm1-x O2 nanoparticle mixed oxides have been synthesized by microwave assisted sol-
Received 8 August 2016 gel (MW sol-gel) and conventional sol-gel synthesis carried out at 60 ◦ C (typical sol-gel) and 100 ◦ C
Received in revised form (approaching the MW temperature). Different characterization techniques, namely, XRD, BET, Raman,
15 November 2016
SEM, FTIR, TEM, XPS, H2 -TPR, CO2 -TPD, and XPS have been employed to understand the process-structure-
Accepted 29 November 2016
properties relationship of the catalysts. The CO oxidation performance has been determined both in the
Available online 30 November 2016
absence and in the presence of H2 in the feed gas stream. Microwave heating yields a more thermally
stable precursor material, which preserves 75% of its mass up to 600 ◦ C, attributable to the different
Keywords:
Microwave
chemical nature of the precursor, compared to the typical sol-gel material with the same composition.
Sol-gel Varying the synthesis method has no profound effect on the surface area of the materials, which is in
CeO2 the range 4–35m2 /g. Conventional sol-gel synthesis performed at 60 and 100 ◦ C yields CeO2 particles
Cex Sm1-x O2 with a crystallite size of 29 nm and 24 nm compared to 21–27 nm for MW sol-gel synthesis (at different
Sm-doped ceria power values). The MW sol-gel Cex Sm1-x O2 catalysts exhibit a smaller crystallite size (12–18 nm). The
CO oxidation pure ceria nanoparticles were shown to have a stoichiometry of approximately CeO1.95 . The presence of
XPS Ce3+ and Sm3+ in the mixed oxide particles facilitates the presence of oxygen vacant sites, confirmed by
XRD
Raman. Oxygen mobile species have been traced using H2 -TPR studies and a compressive lattice strain
H2 -TPR
in the 0.45–1.9% range of the cubic Cex Sm1-x O2 lattice were found to be strongly correlated with the CO
Redox properties
Kinetc studies oxidation performance in the presence and absence of H2 in the oxidation feed stream. MW sol-gel syn-
thesis led to more active CeO2 and Ce0.5 Sm0.5 O2 catalysts, demonstrated by T50 (temperature where 50%
CO conversion is achieved), being reduced by 131 ◦ C and 47 ◦ C, respectively, compared to typical sol-gel
catalysts. Conventional synthesis performed at 100 ◦ C leads to a CeO2 catalyst of initially higher activity
at a certain temperature window (220–420 ◦ C), though with a slower increase of XCO as a function of
temperature compared to the MW synthesized catalyst. MW sol-gel synthesized Ce0.8 Sm0.2 O2 exhibited
a high performance (∼90%) for CO oxidation over a period of more than 20 h in stream. In addition the
effect of reaction temperature and contact time (W/F) on the activity of the CeO2 -based materials for CO
oxidation kinetics were investigated. The activation energy of the reaction was found to be in the range
36–43 kJ/mole depending on the catalyst composition.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction terial due to its wide range of applications, including but not limited
to catalysis [2], fuel cells [3] and sensors [4], due to its unique redox
Metal oxides are of interest to many scientific fields and techno- capacity (Ce4+ /Ce3+ couple). However, CeO2 suffers from limited
logical areas, such as environmental remediation, capacitors, fuel thermal stability as CeO2 is subjected to sintering equally along
cells, catalysis, etc. [1]. In particular, CeO2 is an important nanoma- the x, y and z axes [5]. Several studies have reported improve-
ments in the thermal stability and reduction/oxidation behavior
of CeO2 , especially in the lower temperature regime (<300 ◦ C); the
latter being strongly facilitated by the presence of isovalent (Zr4+ )
∗ Corresponding author.
or trivalent (e.g. La3+ , Eu3+ , Mn3+ , Sm3+ ) dopants [6–10]. The ceria-
E-mail address: kyriaki.polychrono@kustar.ac.ae (K. Polychronopoulou).

http://dx.doi.org/10.1016/j.molcata.2016.11.039
1381-1169/© 2016 Elsevier B.V. All rights reserved.
42 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

0.12
1.0
Ce0.50Sm0.50O2 1360
0.10
1550 1420 0.8
(a)

Normalized Mass
Absorption (a.u)

0.08
0.6
0.06

MW 1036 0.4 MW_CeO2


0.04 CONV_Ce0.5Sm0.5O2
MW_Ce0.8Sm0.2O2
CONV 0.2
0.02 MW_Ce0.5Sm0.5O2
CONV2_Ce0.5Sm0.5O2
0.00 0.0
3000 2000 1000 0 150 300 450 600
Temperature (oC)
-1
Wavelength (cm )

Fig. 1. FTIR absorption spectra obtained over the Ce0.5 Sm0.5 O2 materials synthesized
by using Conv and MW sol-gel methods.
40
(b)
based mixed metal oxides were found to perform much better
compared to pure ceria due to the doping induced modifications

Heat Flow
of the fluorite-type lattice (increased lattice strain, more oxygen
vacancy sites, etc.). In particular, after introduction of such dopants, 0
the change in the oxygen local environment causes a change in
MW_CeO2
the Oxygen Storage Capacity (OSC) of the mixed oxide [11] which CONV_Ce0.5Sm0.5O2
is directly linked with the material’s performance in technologi- MW_Ce0.8Sm0.2O2
cally important catalytic reactions such as Water Gas Shift (WGS) MW_Ce0.5Sm0.5O2
[12,13], hydrocarbon steam reforming [14–16] and CO oxidation -40
CONV2_Ce0.5Sm0.5O2
[17]. There are many studies regarding the preferred elements for
doping ceria. Among them, [18,19] reported that high radii def- 125 250 375 500 625
erence between Ce and the dopant can enhance the reducibility Temperature (oC)
and hence boost the CO oxidation. Sm seems to be one of the best
dopants for inducing changes in the structure and chemical prop- Fig. 2. (a) TGA and (b) DSC profiles obtained over the Cex Sm1-x O2 , (x = 0.5, 0.8, 1)
erties of ceria because of the values of the ionic radii (107 pm for catalysts synthesized using MW sol-gel method. Ce0.5 Sm0.5 O2 Conv and Conv2 sol-
Sm3+ and 97 pm for Ce4+ ) and electronegativity (1.17 for Sm3+ gel catalysts (synthesized at 60 ◦ C and 100 ◦ C) TGA/DSC profiles are provided for
comparison.
and 1.12 for Ce4+ ) of the two elements. It has been reported that
Sm induces the largest increase in ionic conductivity in ceria [20].
The proximity in ionic radii is essential criterion for solid solution interest to the scientific community to explore non-conventional
formation. Also, Sm cations are soluble to ceria lattice. Sm-doped preparation methods of active and selective catalytic materials.
ceria has been extensively studied for oxidation reactions such as Microwave synthesis has also been applied for the synthesis of
allylic oxidation, methane oxidation and diesel soot combustion CeO2 and Ce-Sm-O materials. In particular, Riccardi et al. [36]
[21]. Several synthesis methods of CeO2 or CeO2 -related materials reported on a microwave-hydrothermal method of forming CeO2
have been proposed in the literature, such as hydrothermal [22], rhombus-shape microplates employing a temperature of 150 ◦ C for
sol-gel [23], co-precipitation [24], microwave-hydrothermal [25], 30 min [36]. Dos Santos et al. [37] discussed the rapid microwave
urea-nitrate combustion [26], microemulsion [27], metal organic preparation of CeO2 . However, their research work is confined
chemical vapor deposition (MOCVD) [28], electrospinning [29], only to preparation and the characterization of the material. The
hard-template mediated methods [30] etc. The surface area of the authors discussed the potential applications of these materials in
mixed metal oxides obtained from the above techniques does not catalysis. Li et al. [38] reported on the preparation of Ce1-x Smx O2-␦ -
exceed 150 m2 /g. Other synthesis methods involving surfactants attapulgite nanocomposites using a low power (160 W) microwave
may result in a higher surface area (235 m2 /g) [31]; however, the synthesis and showed its capability in the methylene blue degrada-
requirement for longer aging times and the removal of surfactants tion. They reported an optimum amount of Sm dopant to increase
by a number of washing cycles renders these methods unattractive the catalytic activity, but after introducing Sm to an amount corre-
for industrial implementation. sponding to x = 0.4, formation of agglomerates causes a drop in the
Microwave-hydrothermal synthesis belongs to the family of activity.
low-temperature methods for the preparation of nanophase mate- CO removal is a highly valuable process both in the automo-
rials with a wide range of size and shape [32]. These methods tive industry (three way catalytic converters) and the technology
are considered energy and time efficient as well as environmen- of fuel cells [39]. For the latter application, the methods proposed
tally friendly. Microwave chemistry has been proved an excellent for CO removal from the hydrogen stream include purification with
approach to achieving higher reaction rates (hence significantly hydrogen selective membranes, CO methanation, pressure swing
shorter reaction times of several minutes to hours) compared adsorption, and preferential oxidation (PROX) of CO. From these
to conventional heating hydro/solvothermal methods. The proce- methods, CO methanation and CO-PROX are considered to be the
dures and benefits of microwave synthesis have been presented most promising for small-scale fuel processor applications [39]. In
in a number of good reviews [33–35]. Given the prominent role addition, CO oxidation can be considered as a benchmark catalytic
of the synthesis method on the intrinsic properties of a cata- reaction due to its simplicity in terms of stoichiometry and thus
lyst (including OSC, mobile oxygen species, defects, crystallite can be used as a diagnostic tool for the evaluation of new materi-
size and shape, surface acidity/basicity and others), it is of great als or new material preparation routes. In addition, the CO-PROX
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 43

(a) CeO2 Conv


MW-130-800
MW-150-1200

(220)

(311)
(111)
Intensity (a.u)

(200)

(331)

(422)
(420)
(400)
(222)

20 30 40 50 60 70 80 90
2theta (degrees)
2theta
Fig. 4. XRD profiles of the (a) Cex Sm1-x O2 (x = 0.2, 0.5, 0.8, 1) catalysts synthesized
Ce0.5Sm0.5O2_Conv
(b) Ce0.5Sm0.5O2_MW
using MW sol-gel method and Ce0.5 Sm0.5 O2 Conv2 for comparison.
Intensity (a.u)

catalysts were characterized in terms of their microstructure, mor-


phology, texture, surface properties (acidity, basicity, reducibility)
Conv and the catalytic performance towards CO oxidation both in the
absence and presence of H2 . A correlation of their catalytic perfor-
mance with their intrinsic properties was sought. Kinetic studies
were performed and Eact values are reported for the MW catalysts
MW in order to show their behavior under reaction conditions.

2. Experimental

20 30 40 50 60 70 80 90 100 2.1. Catalyst preparation


2theta
(a) Conventional sol-gel synthesis: The precursor salts,
Fig. 3. (a) XRD profiles of the CeO2 catalyst synthesized using sol-gel (Conv) and MW Ce(NO3 )3 ·6H2 O and Sm(NO3 )3 ·6H2 O (Sigma-Aldrich) were
sol-gel (130 ◦ C–800 W and 150 ◦ C–1200 W) methods, (b)Ce0.5 Sm0.5 O2 synthesized
dissolved with the appropriate molar ratios in distilled water. The
using Conv sol-gel and MW sol-gel.
total metal loading [Ce + Sm] was kept at 0.03 mol in all cases. Citric
acid (the complexing agent) was dissolved in distilled water, main-
process is one of the most effective methods for the removal of CO taining the total metal loading to citric acid ratio (Mtot /citric acid)
trace from reformate stream. Therefore, preferential oxidation is at 0.75, yielding an excess of citric acid. The metal salt and citric
an indispensable step to reduce the concentration of CO to 10 ppm acid solutions were then mixed and the final solution subjected to
level in a H2 generation process [39]. The types of catalysts which conventional heating (60 ◦ C) and stirring until a yellowish gel was
have been used for CO oxidation and CO-PROX include Pt- and Pd- formed (duration ∼2 h, materials’ code: Conv). This conventional
based catalysts [40], nano-gold catalysts and mixed oxide catalysts synthesis protocol was used to prepare CeO2 and Ce0.5 Sm0.5 O2
with the main focus being on the CeO2 -CuO system [41]. Though it mixed oxide compositions only, for critical comparison purposes.
has to be mentioned here that noble metal catalysts can operate in However, in order to achieve a temperature closer to that one used
the presence of moisture but need temperatures higher than 100 ◦ C, in the MW synthesis, the conventional synthesis was selectively
whereas gold catalysts can function only in the form of supported performed for CeO2 and Ce0.5 Sm0.5 O2 catalysts, at using the same
nanoparticles on a metal oxide support. Finding an active catalyst conditions, except the temperature was raised to 100 ◦ C (duration
for CO oxidation at ambient and sub-ambient conditions is chal- ∼20 min, materials’ code: Conv2).
lenging. Xie et al. [42] reported relatively recently for CO oxidation, (b) Microwave Assisted Synthesis. A Microwave accelerated reac-
that the active and stable Co3 O4 system in the form of nanorods out- tion system (MARS-6) is a microwave power system with a
performs at temperatures as low as −77 ◦ C and the system exposes power output of 0–1800 W ± 5% (IEC 705 Method-1988). The major
the 110 planes where Co3+ is the active center for the CO reaction. components of the reaction system include a magnetron, isola-
Lou et al. [43,44] reported an enhancement in the catalytic activ- tor, waveguide, cavity, and mode stirrer. Microwave energy is
ity in the case of In2 O3 -Co2 O3 . In particular, they demonstrated generated by the magnetron, propagated down the waveguide
that In2 O3 -Co2 O3 catalyst presents high CO conversion activity at and introduced into the cavity. The mode stirrer causes multi-
temperature as low as −105 ◦ C, whereas single Co2 O3 is an active directional distribution of the energy and the cavity acts as a
catalyst at only −40 ◦ C. It was discussed by the authors that tun- containment housing for the energy until it is absorbed by the
ing the CO adsorption sites along with tuning the oxygen vacant sample load within the cavity. The isolator protects the mag-
sites were the main factors contributing to the outstanding catalytic netron from energy not absorbed by the sample load. The isolator
performance. allows energy to propagate from the magnetron to the cavity
In the present study, CeO2 and Cex Sm1-x O2 mixed metal oxides but will not allow it to propagate from the cavity back to the
have been synthesized by microwave assisted sol-gel (MW sol- magnetron. This feature is desirable because the magnetron can
gel) synthesis and compared to materials of the same compositions be damaged when exposed to this reflected energy. A turntable
prepared by conventional synthesis. The latter was carried out at is used to rotate the sample load within the cavity to ensure a
temperatures chosen in a way to simulate both a typical sol-gel pro- uniform energy distribution. The materials were prepared by dis-
cedure (60 ◦ C) and the MW temperature (100 ◦ C). The synthesized solving the precursor salts (Ce(NO3 )3 ·6H2 O and Sm(NO3 )3 ·6H2 O)
44 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

Table 1
Textural and structural parameters of the CeO2 and Cex Sm1-x O2 catalysts prepared by Conv and MW sol-gel methods.

Oxide System BET (m2 /g) Dd (nm) Pore size (nm) ´


Lattice Parameter (Å) Lattice Strain, ␧e

Sm2 O3 -MW 0.5 – 10–300 – –


Ce0.2 Sm0.8 O2 -MW 4.1 12.5 3.4–184 5.32
Ce0.5 Sm0.5 -MW 12.2 14.8 3.1–206 5.29 −0.01068
Ce0.80 Sm0.2 O2 -MW 17.8 18.8 3.19–181 5.29 −0.01964
CeO2 -MWa 14.8 27 2.8–318 5.26 −0.00457
CeO2 -MWb 21 5.19
CeO2 CONV 35.1 29 2.9–198 5.22
CeO2 CONV(2)c 6.7 24 3.5–193 5.21
Ce0.50 Sm0.50 O2 CONV 11.3 3–89 5.29 −0.00511
Ce0.50 Sm0.50 O2 CONV(2) 35 17 9–205 5.29
a
130 ◦ C–800 W.
b
150 ◦ C–1200 W.
c
Conventional sol-gel synthesis carried out at 100 ◦ C.
d
Calculated based on Scherrer formula (XRD studies).
e
Calculated based on Williamson-Hall method.

in distilled water with the molar ratios for final compositions 2.2.5. FT-IR studies
of Cex Sm1-x O2 , where x = 0, 0.2, 0.5, 0.8 and 1. As for the Conv FITR analysis was performed on the as-synthesized catalysts
sol-gel route, an aqueous citric acid solution was prepared in using a Nicolet 740 FT-IR spectrometer at ambient conditions with
molar excess (Mtot /citric acid = 0.75). The two solutions were a nominal resolution of 4 cm−1 and averaging of 100 spectra.
mixed and then heated in the microwave reactor. Two microwave
heating conditions were employed, namely 130 ◦ C–800 W (mild) 2.2.6. Thermal studies
and 130 ◦ C–1200 W (severe). Following microwave heating (dura- The thermal stability of the mixed oxides catalyst precursors
tion <10 min, materials’code: MW), all synthesized materials were was investigated by a TA SDT 600 simultaneous thermogravimetry
calcined at 500 ◦ C for 6 h under atmospheric conditions to form the and differential scanning calorimetry (TG-DSC) instrument.
mixed oxide catalyst.
2.2.7. H2 temperature-Programmed reduction (TPR) studies
2.2. Characterization H2 temperature-programmed reduction (H2 -TPR) experiments
were conducted by passing a 10 vol%/H2 /Ar gas mixture
2.2.1. X-ray diffraction (30 NmL/min) over ∼0.2 g of the pre-calcined (20 vol% O2 /He,
The crystal structure of the calcined catalysts was analyzed by 500 ◦ C, 2 h) solid and using a temperature ramp of 30 ◦ C/min while
®
X-ray diffraction (XRD) with a D2-Phaser apparatus (Bruker, MA, the Thermal Conductivity Detector (TCD) signal was recorded con-
USA), using an excitation wavelength Cu K␣ of 1.5418 Å. The X-ray tinuously. The material was mounted on a U-shaped quartz sample
generator was operated at 30 kV and 20 mA. A -2 scan range of tube plugged with quartz wool and placed in a flow through reac-
10–100◦ with a 0.02◦ step size was employed. The average crystal- tor (Autochem 2920, Micromeritics, Atlanta, USA). The H2 (m/z = 2)
lite size was calculated from line broadening of (111) peak using the and H2 O (m/z = 18) signals were monitored using an online Cirrus
Scherrer equation. The lattice parameter was estimated by apply- mass spectrometer to follow the kinetics of the process.
ing a standard cubic indexation method, using the predominant
(111) peak. Williamson-Hall plots (ˇcos /␭ vs. sin /␭) were used
2.2.8. CO2 temperature-Programmed desorption (TPD) studies
to estimate the lattice strain [45].
CO2 -TPD experiments were conducted using Autochem 2920,
(Micromeritics, Atlanta, USA). In particular, the material was
2.2.2. Porosimetry
mounted on a U-shaped quartz sample tube plugged with quartz
Porosimetry measurements were carried out using a high-
wool and placed in a flow through reactor. CO2 -TPD experiments
resolution 3Flex Micromeritics (Atlanta, USA) adsorption instru-
were conducted by passing a 5 vol%/CO2 /Ar gas mixture for 30 min
ment equipped with high-vacuum system and three 0.1 Torr
(30 NmL/min) over ∼0.2 g of the pre-calcined (20 vol%O2 /He,
pressure transducers. Before performing the analysis, the materials
were degassed at 150 ◦ C for 2 h to remove any residual moisture.
The BET surface area was calculated from the adsorption data in 1800
CeO2_MW
the relative pressure range of P/Po = 0.04–0.25, where the Barrett- F2g band of CeO2 Ce0.8Sm0.2O2 _MW
Joyner-Halenda (BJH) method was applied so to estimate the pore
size and pore volume.
Raman Intensity (a.u)

1200
2.2.3. Scanning electron microscopy (SEM)
The morphology of the catalysts was examined with JSM 7610F-
Field Emission Scanning Electron Microscope (JEOL Ltd., Tokyo,
600
JPN), using both secondary and low-angle-backscattered imaging.

2.2.4. Transmission electron microscopy (TEM)


Information regarding the crystallinity and elemental com-
0
position of the catalysts was obtained with an FEI Tecnai G20
Transmission Electron Microscope (TEM) coupled with a post 200 300 400 500 600 700 800
-1
Raman Shift (cm )
column energy filtered camera (Gatan Quantum 963–energy reso-
lution 0.9 eV full width at half maximum at zero loss peak), and an Fig. 5. Raman spectra of the CeO2 single oxide and Ce0.8 Sm0.2 O2 mixed oxide syn-
EDAX energy dispersive X-ray (EDX) analyser. thesized using MW sol-gel method.
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 45

Fig. 6. (a) TEM microphotograph and (b) HRTEM microphotograph obtained over the MW sol-gel synthesized Ce0.8 Sm0.2 O2 catalyst.

Fig. 7. (a) TEM microphotograph, (b) HRTEM microphotograph, obtained over the MW sol-gel synthesized CeO2 catalyst.

500 ◦ C, 2 h) solid and using a temperature ramp of 30 ◦ C/min while and 1% CO, 1% O2 , 4% H2 and balance He for CO oxidation in the
the TCD signal was recorded continuously. The mass numbers (m/z) absence and presence of H2 measurements, respectively. The reac-
28 and 44 were used for CO and CO2 , respectively, during CO2 -TPDs. tor was equipped with a 9 mm ID quartz tube placed in the middle
of a programmable split tube furnace (Lindberg/Blue M Mini-Mite
2.2.9. X-ray photoelectron spectroscopy Tube Furnace − Thermo). In all experiments, unless otherwise men-
Surface elemental concentrations and chemical states were tioned, the feed gas was passed over 50 mg of the catalysts at a
studied via X-ray Photoelectron Spectroscopy (XPS). XPS analy- flow rate of 50 ml/min (60,000 cm3 g−1 h−1 WHSV). The flow rate
ses were performed on a ThermoFisher Scientific (East Grinstead, was controlled by a set of digital mass flow controllers (HI-TEC
UK) Theta Probe spectrometer. XPS spectra were acquired using a Model F-201CV-10K-AGD-22-V Multi-Bus DMFC; Bronkhorst). The
monochromated Al K␣ X-ray source (h = 1486.6 eV). An X-ray spot temperature was ramped at a heating rate of 5 ◦ C/min and the
of ∼400 ␮m radius was employed. Survey spectra were acquired catalyst temperature was measured by a thermocouple placed in
using a pass energy of 300 eV. High resolution, core level spec- contact with the catalyst bed. The signal from the thermocouple
tra were acquired using a pass energy of 50 eV. All spectra were was acquired using USB--6008 Multifunction I/O and NI-DAQmx
charge referenced against the C 1s peak at 285.0 eV to correct for (National Instruments) data acquisition board. The effluent gas was
charging effects during acquisition. Quantitative surface chemical fed into an inline multichannel infrared gas analyzer (IR200, Yoko-
analyses were calculated from the high resolution, core level spec- gawa) to analyze the exit gas reactants and products. The volume
tra following the removal of a non-linear (Shirley) background. The percent of CO, CO2 and O2 gases were determined simultaneously
manufacturer’s Avantage software was used which incorporates and logged along with the catalyst temperature during the course of
the appropriate sensitivity factors and corrects for the electron the conversion experiment using Labview-configured data acqui-
energy analyser transmission function. sition software. The long-term stability of selected catalysts was
studied under a continuous stream of feed gas for 20 h at 530 ◦ C.
The catalytic activity was expressed by the conversion of CO in the
2.2.10. Catalytic studies
effluent gas and indicated as a CO conversion percentage by the
The catalytic activity of the as-prepared catalysts for CO oxida- [CO]in −[CO]out
formula, XCO (%) = [CO]
x100 where X represents the per-
tion was measured in a continuous flow fixed-bed feed reactor. in

The feed gas mixture contained 4% CO, 20% O2 and balance He


46 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

a result of the direct interaction of the microwave radiation with


CeO2_Conv
Adsorbed Amount (cm3/g)
(a) CeO2_MW
the material inducing a different degree of supersaturation in the
solution and a different nucleation rate [48]. Microwave radiation
60 is non-ionizing electromagnetic radiation with high penetration
depth which leads to the formation of crystallites with higher
dimensional uniformity due to the absence of thermal gradients
during the synthesis [49]. Water is an ideal solvent for microwave
30 synthesis due to its high dipole moment. The dipole moment of the
solvent interacts with the electromagnetic radiation. The creation
of ‘hot spots’, evenly distribution over the precursor surface, for a
short period of time leads to enhanced mass transport of species
from the solution to the nuclei [49].
0 Fig. 2 shows plots of mass loss (Fig. 2a) and heat flow (Fig. 2b)
0.0 0.2 0.4 0.6 0.8 1.0
versus temperature for all the materials synthesized using both
Relative Pressure (P/Po) microwave and conventional (at 60 and 95 ◦ C) methods. All the MW
sol-gel materials (Fig. 2a) exhibited a greater thermal stability com-
Ce0.5Sm0.5O2_MW pared to the Conv sol-gel material synthesized at 60 ◦ C, whereas

40
(b) Ce0.5Sm0.5O2_Conv the Conv2 sol-gel material appears to show less mass loss than
Absorbed Amount (cm3/g)

the Conv sol-gel. This can be due to the temperature effect in the
solution, where Conv2 sol-gel material is closer to the MW tem-
perature conditions (100 vs. 130 ◦ C), though at slower ramp rate.
So, it is anticipated that an enhanced mass transport of species is
taking place from the solution to the first-formed nuclei at this high
20
temperature and nucleation of different precursor entities is being
favored which are more stable in the MW and Conv2 cases. In par-
ticular, the mixed oxide Conv sol-gel material lost 40% of its mass
at a temperature lower than 150 ◦ C, whereas all the MW sol-gel
materials preserved 75% of their mass even at temperatures higher
0
0.0 0.2 0.4 0.6 0.8 1.0
than 600 ◦ C. In the DSC profiles shown in Fig. 2b, the Conv sol-gel
Relative Pressure (P/Po) material exhibits a strong endothermic peak at 100–125 ◦ C due
to the loss of water of hydration. An exothermic peak due to the
Fig. 8. N2 adsorption isotherms at 77 K obtained over (a) CeO2 and (b) Ce0.5 Sm0.5 O2 solid–solid transition (from the precursor state to the crystalline
synthesized by Conv sol-gel and MW sol-gel methods. structure) appeared at 365 ◦ C. For the MW sol-gel catalysts, this
endothermic peak appeared at a lower temperature (ca. 80–100 ◦ C),
centage conversion and [CO] is the CO molar flow in the inlet and whereas two additional exothermic peaks appear at 270 and 500 ◦ C,
outlet feed gas. corresponding to the decomposition of different intermediates and
lattice formation, respectively.
2.2.11. Kinetic studies
For kinetic measurements, the catalyst loading for CO oxidation
experiments was varied as 15, 30, 50, 75 or 100 mg while keeping 3.2. Micro/nano-structure
the GHSV of the feed fixed (3000 cm3 h−1 ). The CO oxidation
 reac- 
tion rates were calculated using the formula rCO = XCO / W/FCO , Fig. 3a presents the XRD diffractograms of CeO2 for the MW and
where XCO is the fractional CO conversion, W is the mass of the cat- Conv sol-gel materials. For the MW sol-gel material, two different
alyst in g and FCO is the CO molar flow rate. The apparent activation set of conditions were adopted, namely 130 ◦ C, 800 W and 130 ◦ C,
energies of selected catalysts were determined from the Arrhenius 1500 W, for synthesis under mild and severe microwave condi-
plots using data from the linear region of <15% CO conversion. tions, respectively. The (111), (200), (220), (311), (222), (400), and
(331) peaks of the CeO2 fluorite structure are observed in both sam-
3. Results and discussion ples [50], whereas some peaks corresponding to the Ce(OH)x (CO3 )y
were traced too (JPDS No. 41-0013) [50]. The intensity of the lat-
3.1. Mixed metal oxide precursor ter peaks is becoming smaller in the case of MW sol-gel method.
The crystallite size of the Conv sol-gel material was found to be
Fig. 1 presents the FTIR absorption spectra obtained for 29 nm, compared to 27 and 21 nm for the MW (mild) MW (severe)
Ce0.50 Sm0.50 O2 samples synthesized by microwave and conven- sol-gel material respectively (Table 1). More severe microwave
tional methods (Conv/60 ◦ C), in the as-synthesized form (before heating leads to a higher vaporization rate in the solution and a
calcination). The peaks at 1420 and 1360 cm−1 can be observed dramatic increase in the reactant concentration in the liquid phase.
in both spectra, indicative of the formation of cerium formate in a Condensation reactions and crystallization processes continue in a
bidentate configuration [46]. The peak at 1036 cm−1 corresponds vapor-soaked slurry solid, which give rise to the observed smaller
to the C O single bond [47]. The peak at 1550 cm−1 was of much crystallite size for the MW (severe) sol-gel material. Fig. 3b presents
higher intensity for the MW sol-gel compared to the Conv sol-gel the effect of synthesis method (microwave vs. conventional at
synthesized nanoparticles. This peak corresponds to carboxylate 60 ◦ C) on the crystalline structure of the Ce0.5 Sm0.5 O2 mixed oxide.
salts [47] and it clearly demonstrates a change in the synthesis Microwave heating gives rise to a particle size of 14 nm compared
pathway, showing that different intermediate species and/or dif- to 28 nm for the conventional process (at 60 ◦ C). This difference is
ferent population of intermediate species with different structural caused by the much shorter microwave processing time and the
features are formed under microwave radiation/heating. This is in fundamental differences between MW and conventional heating
agreement with the literature where the reaction selectivity under (such as quenching, homogeneous distribution of temperature in
microwave radiation differs from the conventional route mainly as the MW conditions) [49].
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 47

Fig. 9. SEM microphotographs of (a) Conv sol-gel Ce0.5 Sm0.5 O2 , (b) MW sol-gel Ce0.8 Sm0.2 O2 , (c) MW sol-gel Sm2 O3 , (d) MW sol-gel CeO2 , (e) Conv sol-gel CeO2 and (f) Conv2
sol-gel Ce0.5 Sm0.5 O2.

Fig. 4 presents the XRD patterns obtained for the MW sol-gel (from the walls of the vessel to the core of the solution) in the case of
Cex Sm1-x O2 mixed oxides (x = 0, 0.2, 0.5, 0.8 and 1) along with the MW conditions [49]. No Raman band corresponding to Sm2 O3 was
CeO2 Conv2 and Ce0.5 Sm0.5 O2 Conv2 both synthesized at 100 ◦ C for found (cubic 375 cm−1 ) which is consistent with the incorporation
comparison purposes. Again, the CeO2 fluorite peaks are observed of Sm3+ ions into the ceria cubic lattice.
in all samples. No XRD peaks corresponding to Sm oxide phases The nanostructure and morphology of the materials was inves-
are identified for the Ce–Sm mixed oxides, showing the forma- tigated using TEM. A careful examination of the bright field TEM
tion of (Ce,Sm)O2 solid solutions in all cases. There is a progressive images for the MW sol-gel Ce0.8 Sm0.2 O2 and CeO2 materials (Fig. 6a
broadening of the XRD peaks as Sm is introduced and its concentra- and Fig. 7a respectively) shows that the Ce0.8 Sm0.2 O2 mixed oxide
tion increased in the CeO2 structure. The calculated crystallite sizes exhibits a smaller crystallite size than CeO2 , in agreement with
for CeO2 , Ce0.8 Sm0.2 O2 , Ce0.5 Sm0.5 O2 and Ce0.2 Sm0.8 O2 are 27, 19, the XRD data. Both materials are well-crystallised and octahedral
15 and 13 nm respectively (Table 1). These crystallite sizes are in in shape even though the terminating surfaces of the crystals are
good agreement with Sm-doped CeO2 mixed oxides prepared by not well-defined due to inherent limitations of the TEM. From the
mechanochemical synthesis (10–30 nm) [51] and in all cases (MW HRTEM images (Fig. 6b and Fig. 7b), the lattice fringes of the crystal-
and conventional methods) a Sm-induced drop in the crystallite lites and their growth in many different orientations can be seen. It
size was measured. The XRD peak maxima are also observed to has been reported that doping ceria oxide with lower valent cations,
progressively shift to lower 2 values with increased Sm doping such as La3+ , Sm3+ induces strain within the lattice, which results in
indicative of a lattice parameter expansion. This is expected consid- a better CO oxidation catalytic performance. In the present study,
ering the larger ionic radius of the Sm3+ cations (107 pm) compared Williamson-Hall plots were used in order to calculate the lattice
to the Ce4+ cations (97 pm). strain (see Table 1) [45]. The lattice strain was found to be compres-
Raman spectra of the microwave assisted synthesized CeO2 and sive in all cases, giving rise to a shorter Ce-O bond than expected.
Ce0.8 Sm0.2 O2 samples, are presented in Fig. 5. In agreement with In the case of Gd-doped Ceria, where Gd3+ ions have a similar ionic
the XRD data, the Raman spectra reflect the formation of the cubic radius to Sm3+ , it was found that a compressive strain of 0.3% results
fluorite lattice and confirm the (Ce,Sm)O2 solid solution forma- in a Ce-O bond contraction of 1% [53]. In the materials reported here
tion. The band at 460–470 cm−1 corresponds to the allowed Raman the compressive strain was found to be in the range of 0.45-1.9%
mode (F2 g) of the fluorite metal lattice. For pure CeO2 oxide, the depending on the composition (Table 1). The shorter Ce-O bond
Raman spectrum was found to be symmetric around the band at length results in Ce4+ ions being shifted away from the oxygen
465 cm−1 and the F2 g mode corresponded to the symmetric vibra- vacancies, making the vacancies more mobile and easier to migrate.
tion of oxygen ions around Ce4+ ions in the CeO6 octahedra [52]. In It is reasonable to expect that the distortion of the oxygen sub-
the samarium doped CeO2 material, the F2 g band becomes asym- lattice would be higher closer to the vacancies migration pathway
metrical and slightly shifted to lower wavenumber values, due to [54].
the cell expansion resulting from the substitution of Sm3+ ions in
the CeO2 lattice and the subsequent oxygen loss around the cation
3.3. Texture and morphology characteristics of the mixed metal
sites. In addition, a weak broad band in the range of 530–620 cm−1
oxides
was observed in samarium doped CeO2 , attributed to the oxygen
vacancies in the lattice [52]. It should be mentioned here that the
Figs. 8a and b present the low-pressure N2 adsorp-
high synthesis temperature during the MW preparation is believed
tion/desorption isothermal curves at 77.3 K for MW and C
that enhance the formation of oxygen vacancies. This is linked with
sol-gel synthesized CeO2 and Ce0.5 Sm0.5 O2 . All the synthesized
the enhanced mass transport (e.g. Sm3+ species incorporating into
materials demonstrated a type IV sorption isotherm based on
the CeO2 lattice) due to high temperature at short times. This mass
IUPAC [55], typical of mesoporous materials (pore width 2–50 nm).
transport of Sm3+ species is anticipated to happen homogeneously,
The hysteresis loop appearing above a relative pressure of 0.45
due to the undistrurbed temperature field, in the synthesis solution
is clearly attributable to capillary condensation of nitrogen gas
48 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

5.48
(a) Ce0.8Sm0.2O2_MW
Ce0.2Sm0.8O2_MW

5.47 284oC CeO2_MW


400oC 634oC CeO2_CONV
TCD Signal (a.u)

TCD Signal
5.46

CeO2_MW
Ce0.5Sm0.5O2_MW
5.45
CeO2_Conv
Ce0.2Sm0.8O2_MW
Ce0.8Sm0.2O2_MW
5.44
150 300 450 600 750
o
Temperature ( C) 150 300 450 600 750

Temperature (oC)
m/e 2
(b) m/e 18
Fig. 11. CO2 -TPD profiles obtained over the CeO2 single oxide and Cex Sm1-x O2
(x = 0.2, 0.5, 0.8, 1) mixed oxide catalysts, synthesized using Conv sol-gel and MW
MS Signal (a.u)

sol-gel methods, respectively.

Conv and Conv2 materials, Conv2 demonstrate a more profound


effect of Sm doping on the thermal stability. There was no clear
effect of the synthesis method on the porosity type.

3.4. Surface chemistry of the mixed metal oxides

Fig. 10a presents the H2 -TPR profiles obtained for the


150 300 450 600 750
Cex Sm1-x O2 materials. According to the literature, for a typical
Temperature (oC) reduction profile of CeO2 oxide, the peak at 247 ◦ C can be attributed
to the sub-surface reduction of CeO2 [57,58]. On the other hand, the
Fig. 10. (a) H2 -TPR profiles obtained over the CeO2 single oxide and Cex Sm1-x O2 peaks at approximately 467 ◦ C and 1000 ◦ C are assigned to the for-
(x = 0.2, 0.5, 0.8, 1) mixed oxide catalysts synthesized using Conv sol-gel and MW mation of non-stoichiometric Ce oxides and Ce respectively [57,58].
sol-gel, respectively, (b) On line mass spectroscopy signal for H2 (m/z = 18) and H2 O
(m/z = 18) recorded during the H2 -TPR experiment as presented in (a).
Concerning the pure CeO2 oxides, it appears that both the MW and
C sol-gel methods result in materials with very similar reduction
profiles. However, the introduction of Sm3+ ions significantly mod-
inside the mesopores, while its H3-type classification indicates the ifies the reduction patterns. In particular, it is observed that the
presence of pores with a slit-shape. Moreover, the characteristic Ce0.8 Sm0.2 O2 mixed oxide exhibits a reduction peak at 284 ◦ C, this
single-step observed next to the point of loop closure (P/P0 ≈ 0.5) being a lower reduction temperature compared to the other oxides,
strongly implies that desorption is achieved through cavitation. whereas the Ce0.5 Sm0.5 O2 mixed oxide exhibits a reduction peak at
The mechanism is associated with the evaporation of N2 from around 450 ◦ C, showing the presence of less reducible species (at
the mesopores through the spontaneous nucleation and growth low temperature regime), which is in agreement with the different
of gaseous bubbles inside the condensed nitrogen [56]. At higher CO oxidation performance of the microwave prepared materials
relative pressures (up to P/P0 ≈ 0.99) the adsorption curve does (Ce0.8 Sm0.2 O2 > Ce0.5 Sm0.5 O2 ). Also, in the case of Ce0.5 Sm0.5 O2 , the
not reach a saturation point, but instead extends indefinitely in high temperature (HT) peak for the reduction of the bulk mate-
the vertical direction. This behavior represents the unrestricted rial appears shifted to higher temperatures > 750 ◦ C compared to
multilayer adsorption of N2 due to the presence of a macroporous the Ce0.8 Sm0.2 O2 mixed oxide (around 630 ◦ C). The latter result is
surface (pore width >50 nm). This is in agreement with the mor- indicative of a higher structural integrity of the particular mixed
phology of these materials, as shown in the SEM images of Fig. 9. oxide under a reducing atmosphere. This is of particular technologi-
At lower relative pressures, (P/P0 < 0.05) the MW sol-gel materials cal interest for catalytic reactions where a reducing atmosphere and
demonstrate weak N2 physisorption most probably due to the lack thermal effects both impact on the material structure. The reduc-
of microporosity (pore width <2 nm). The BET surface area values tion of the catalytic oxides is accompanied by water production,
presented in this work are in good agreement with the values in as it can be seen in Fig. 10b by the mass spectrometer signal at
[51], where they studied the synthesis of 20 mol% samarium-doped m/z = 18.
cerium oxide nanoparticles from the mechanochemical reaction It is well known that the CO oxidation product, CO2 , can partic-
between samarium trichloride, sodium hydroxide and cerium ipate in competitive adsorption with CO (and H2 in the case of the
hydroxide, in which a surface area of 48.7 m2 /g was obtained. By CO-PROX reaction) having a negative effect on the overall reaction
comparing the porosity values of CeO2 Conv (35.1 m2 /g), CeO2 taking place on the catalyst surface. In this respect, the nature of the
Conv2 (6.7 m2 /g) and CeO2 MW (14.8 m2 /g), an effect of the support plays a crucial role in terms of stability in the presence of
synthesis method can be observed. This can be due to the higher CO2 in the feed. Acidic supports were reported to be more resistant
temperatures associated with the MW synthesis for a very short to deactivation than basic supports [59]. Ceria is a basic oxide and
period of time. Though, in MW CeO2 catalyst, the high temperature as such has the tendency to adsorb CO2 and form carbonates. TPD-
is homogeneously distributed leading to less dramatic collapse MS profiles of CO2 show broad peaks (Fig. 11) which correspond
of the surface area compared to Conv2 CeO2 . In the case of to several forms of adsorbed CO2 on the surfaces of the materials
Ce0.5 Sm0.5 O2 catalyst composition, the BET values for the MW, studied here. CO2 is being desorbed from the catalyst surface in the
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 49

Ce 3d v’’’ 20Ce
O 1s 20Ce
50Ce 50Ce
u 80Ce v
(a) u’’’ 100Ce (b) 80Ce
100Ce
u’
u’’ v’

CPS
CPS v’’

930 915 900 885 870 540 535 530 525 520
XT Binding Energy (eV)
Binding Energy (eV)
Text Legend
Sm 3d 20Ce
50Ce
80Ce
(c)
CPS

YL

XB

1130 1120 1110 1100 1090 1080 1070 1060


Binding Energy
XB (eV)
Binding Energy (eV)
Fig. 12. (a) Ce3d, (b) O 1s and (c) Sm 3d XPS core level spectra obtained over the Cex Sm1-x O2 (x = 0.2, 0.5, 0.8, 1) catalysts synthesized by MW sol-gel method.

Table 2
Reduction temperature, activation energy of CO2 desorption and T50 for the CeO2 and Cex Sm1-x O2 catalysts, prepared by Conv and MW sol-gel methods, in the CO oxidation
reaction.

Oxide System Reduction Temperatures (◦ C) CO2 Edes T50 (◦ C)

Ce0.2 Sm0.8 O2 -MW 312, 473 11.3 553


Ce0.5 Sm0.5 -MW 469, 737 453
Ce0.80 -Sm0.2 O2 -MW 284, 400, 645 8.7 406
CeO2 -MWa 370/667 – 390
CeO2 -MWb – – –
CeO2 CONV 328, 405, 633 24.7 524
Ce0.50 -Sm0.50 O2 CONV 19.6 512
c
:First order kinetics analysis.
a
Microwave conditions: 130 ◦ C–800 W.
b
Microwave conditions: 150 ◦ C–1200 W.

molecular form. The calculated values of Edes at corresponding tem- Raman data. In order to investigate in more detail the presence or
peratures of CO2 desorption are tabulated in Table 2. The values of not of Ce3+ states in the XPS spectra, it is useful to compare the data
Edes correlate very well with the catalyst activity in CO oxidation, given here for the pure CeO2 calcined catalysts with that of Mullins
namely: the lowest Edes corresponds to the most active catalyst et al. [61] for CeO2 bulk samples and Pfau and Schierbaum [62]
(due to competitive adsorption). This trend seems to be valid in and Henderson et al. [63] for CeO2 thin films. In comparison with
the case of the Cex Sm1-x O2 mixed metal oxides and aligned with the data of these other workers [61–63], there is a clear increase
their catalytic activity, where Ce0.8 Sm0.2 O2 presents the lowest Edes in photoelectron intensity between the v and v” Ce4+ peaks, at a
of CO2 desorption (8.7 kJ/mol) compared to CeO2 (11.2 kJ/mol) and binding energy position expected for the v’ Ce3+ peak. Both Pfau
Ce0.5 Sm0.5 O2 (19.6 kJ/mol). and Schierbaum [62] and Henderson et al. [63] have studied the
The XPS Ce 3d and O1 s core level spectra obtained for the emergence of the v’ peak, as a result of heating the CeO2 in the
Cex Sm1-x O2 catalytic materials are given in Fig. 12(a & b). In accor- UHV environment, leading to surface reduction of the CeO2 . From
dance with the nomenclature used in previous XPS work on cerium their results, both sets of authors have presented the changes in the
oxide, the peaks in Fig. 12a are labeled as v and u, corresponding Ce 3d spectral shape for the v, v’ and v” peak region. Furthermore,
to transitions associated with Ce 3d5/2 and Ce 3d3/2 , respectively Henderson et al. have estimated the CeO2-x stoichiometry corre-
[60–62]. The stronger peaks are v, v”, v”’, u, u”, u”’ corresponding to sponding to the different spectral shapes [63]. In comparing the
Ce4+ species, whereas the peaks labelled as vo , v’, u’, u0 correspond data recorded for our CeO2 calcined catalysts with the data given
to Ce3+ species. The Ce 3d spectra show that the cerium oxide is by Henderson et al. [63], the spectral shape for our data corresponds
predominantly a Ce4+ based oxide, in agreement with the XRD and to a stoichiometry of CeO1.95 . If the O2− vacancies associated with
50 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

the sub-stoichiometric CeO2 are located in the outer surface layer, Ce0.5Sm0.5O2-CONV2
100
then the O2− vacancy concentration in that layer would be 22% of CeO2-CONV2

% CO Conversion
Sm2O3-MW
the surface O2− sites [63]. This is an important finding given the 80 CeO2-MW
key role of oxygen vacancies in the CO oxidation reaction. The CO CeO2-CONV
oxidation reaction is based on the formation of oxygen vacant sites 60 Ce0.8Sm0.2O2-MW
which facilitate the reaction progress and they are restored by gas Ce0.5Sm0.5O2-MW

phase oxygen exposure. 40 Ce0.5Sm0.5O2-CONV


Ce0.2Sm0.8O2-MW
The Sm 3d5/2 peak for all of the Cex Sm1-x O2 mixed oxides are
shown in Fig. 12c. The commercial samaria powder exhibiting a 20
(a)
stoichiometry of Sm2 O2.9 was used as a reference and presented
0
only carbonaceous contaminant species, hence was considered to
100 200 300 400
o
500
be a good reference sample. The Sm 3d5/2 peak for the samaria Temperature ( C)
powder was observed at a binding energy of 1082.4 eV, in excellent CeO2-MW
agreement with the data of Elkins and Hagelin Weaver [64], and 100 Ce0.8Sm0.2O2-MW

% CO Conversion
the O 1s peak at a binding energy of 531.5 eV. The Sm 3d samaria Ce0.2Sm0.8O2-MW
80
peaks for the Cex Sm1-x O2 calcined catalysts show the same peak
shape as that of the samaria powder. Hence, in all of the Cex Sm1-x O2
60
samples, the Sm exists in a 3+ oxidation state. Interestingly, the Sm
3d5/2 peak progressively shifts to lower binding energies as the Sm 40
content in the Cex Sm1-x O2 catalysts is reduced. Thus, as the Sm (b)
content in Cex Sm1-x O2 is reduced, there is less charge transfer to 20
the neighboring O atoms.
The O 1s peak for the CeO2 and Cex Sm1-x O2 catalysts was com- 0
prised of two distinct peaks, one at ∼529 eV and the other at 100 200 300 400 500 600
o
∼531.5 eV. The lower binding energy peak corresponds to the lat- Temperature ( C)
tice oxygen [62,63,65]. The higher binding energy peak for CeO2
CO Conversion: Ce0.8Sm0.2O2-MW
has been ascribed to the presence of hydroxyl species, carbonate or 100 600
polarised O2− ions located adjacent to O vacancy sites [63,65,66].

Temperature ( C)
% CO Conversion 525
80

o
For all of the samples, there was a small C 1s peak at a binding Catalyst Temperature
450
energy of approximately 289.5 eV, indicative of the presence of a
surface carbonate. It has been reported previously that carbonate 60 (c) 375
species are expected at the surface of rare earth oxide materials 300
40
[64]. The intensity of the 531.5 eV peak (compared to the 529 eV lat- 225
tice oxide peak) for CeO1.95 sample by Henderson et al. [63] is very 20 150
small (due to the low O2− vacancy concentration), whereas for our
CeO1.95 oxide, the 531.5 eV peak has a substantially greater inten- 0 75
sity. The ceria films were deposited in-situ by Henderson et al. [63], 00 04 08 12 16 20
compared to the samples studied here, which have been exposed
Time (h)
to atmosphere. The O 1 s binding energy of metal hydroxides and
metal carbonates are typically around 531.0-531.5 eV [67]. There- Fig. 13. Catalytic performance (expressed as%CO conversion) as a function of tem-
fore, for the pure ceria catalysts reported here, the 531.5 eV peak perature, towards the CO oxidation reaction in the (a) absence (4% CO/20% O2 /He)
can be ascribed predominantly to the presence of surface hydrox- and (b) presence (1% CO/1% O2 /4% H2 /He) of H2 in the feed stream, obtained over
ide/carbonate species, with a small contribution from polarised O2− the Cex Sm1-x O2 catalysts, (c) Catalytic stability (expressed as % CO conversion) at
530 ◦ C obtained over the Ce0.8 Sm0.2 O2 catalyst for 20 h in stream.
ions located adjacent to O vacancy sites.
In our O 1 s spectra, the intensity of the 531.5 eV peak progres-
sively increases with Sm content. This is in agreement with our displays the catalytic activity under the feed gas containing 4% CO
observation of the O 1 s peak for samaria occurring at 531.5 eV. and 20% O2 with a He balance. By comparing the T50 values (T50 is
Furthermore, when a Sm3+ ion substitutes for a Ce4+ ion, one oxy- the temperature at which 50% CO conversion is achieved) for the
gen ion vacancy is introduced for every two Sm3+ cations [68]. As different oxides, it can be seen that the MW sol-gel CeO2 demon-
such, an increase in Sm3+ content will lead to a greater concen- strated higher catalytic activity, with a T50 = 392 ◦ C compared to a
tration of O2− vacancies in the Cex Sm1-x O2 catalysts and hence an T50 = 523 ◦ C for Conv sol-gel CeO2 . Though, performing the conven-
increase in the number of low co-ordinated and polarized O2− ions, tional sol-gel synthesis at much higher temperature (100 ◦ C) gives
located in the vicinity of O vacancy sites. In the O 1 s spectra for a CeO2 Conv2 catalyst with T50 = 372 ◦ C. The Conv2 CeO2 catalyst
the Cex Sm1-x O2 catalysts, the higher binding energy O 1 s peak is presented a high activity with a slower increase of XCO as a function
broader than the lower binding energy component, indicating the of temperature compared to the MW CeO2 . This is most likely due to
presence of multiple components. Hence, it appears that this peak the presence of more activated oxygen species and higher surface
centered at around 521.5 eV comprises of components associated area of the MW CeO2 . Similarly, the MW sol-gel Cex Sm1-x O2 mixed
with: (i) the substitution of Sm3+ for Ce4+ in Cex Sm1-x O2 ; (ii) surface oxides showed higher activities, demonstrated by the lower con-
hydroxide/carbonate species and (iii) the presence of O2− vacancies version temperature of T50 = 407 ◦ C and 454 ◦ C for Ce0.8 Sm0.2 O2 and
in CeO2 and Cex Sm1-x O2 materials. the Ce0.5 Sm0.5 O2 , respectively, compared to T50 = 513 ◦ C for Conv
sol-gel Ce0.5 Sm0.5 O2 (Fig. 13 & Table 3). However, the conventional
3.5. CO oxidation studies synthesis performed at 100 ◦ C led to a Ce0.5 Sm0.5 O2 Conv2 catalyst
with T50 = 427 ◦ C, a system with much better catalytic activity than
The CO oxidation catalytic performance of the different catalysts the Ce0.5 Sm0.5 O2 synthesized at 60 ◦ C (T50 = 513 ◦ C). Microwave
was compared using a continuous flow fixed-bed catalytic reactor prepared Sm2 O3 exhibited a conversion temperature of T50 = 401 ◦ C
in the absence and the presence of hydrogen in the feed gas. Fig. 13a but the activity deteriorates with the temperature. Overall, it is
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 51

Table 3
CO oxidation Catalytic Activity Results (% CO conversion) in the absence and presence of H2 , obtained over the CeO2 and Cex Sm1-x O2 catalysts, prepared by Conv and MW
sol-gel methods.

COOX1 COOX2 % CO @ 500 ◦ C

20% 50% 80% 100% 20% 50% 80% 100% COOX1 COOX2

CeO2 MW 341 392 440 518 330 382 432 496 98% 100%
CeO2 CONV 455 523 NA NA NA NA NA NA 38% NA
CeO2 CONV2 307 372 455 NA NA NA NA NA 89 NA
Ce0.8 Sm0.2 MW 356 407 457 533 353 403 452 516 94% 97%
Ce0.50 Sm0.50 O2 MW 400 454 509 NA 379 436 491 NA 76% 83%
Ce0.50 Sm0.50 O2 CONV 456 513 NA NA NA NA NA NA 42% NA
Ce0.50 Sm0.50 O2 CONV2 375 427 477 NA NA NA NA NA 89.3 NA
Ce0.20 Sm0.80 O2 MW 497 NA NA NA 513 NA NA NA 20% 18%

COOX1 : 4% CO/20% O2 /He.


COOX2 : 1% CO/1% O2 /4% H2 /He.

the nucleation rate and the smaller the critical sizes of the nuclei
that are needed to overcome the nucleation barrier. Furthermore,
it is known that MW leads to uniformity in the formed crystals
[49], whereas conventional heating, even when it is performed at
similar temperatures with the MW, leads to non uniformity due to
nucleation. MW can also give rise to Ostwald ripening phenomanon
while the growth process is taking place [70]. Also, in the case of the
MW prepared catalysts, efficient coupling of the microwave radi-
ation with the solution allows for volumetric heating and a more
homogeneous temperature field to be developed compared to both
of the traditional routes adopted herein [49]. The lower values of
porosity that have been presented earlier imply more dense MW
prepared catalysts, whereas the lower particle size implies sup-
pressed grain growth. It has been mentioned [71] that there is a
certain temperature range called ‘the kinetic window’where den-
sification is activated but grain growth is not activated. Another
important aspect of MW synthesis is the dielectric loss (␧)¨ of the
precursor material in the solution (metal citrates/nitrates in this
work) [49]. It seems that the low dielectric loss of the precursor
material suppresses any further effect of high temperature on the
grain growth of the material. In addition, the finer grain size is in
agreement with a higher surface/volume (S/V) ratio and a higher
number of active sites exposed on the surface, available for reac-
tion. In the case of conventional catalysts prepared at 60 ◦ C it seems
that low mass transport happens in the solution leading to limited
incorporation of Sm3+ into the CeO2 lattice and eventually giving
rise to a rather not so well-performed catalytic system. Also, it is
known that [72] the surface defects in CeO2 play the role of oxy-
gen adsorption sites. The MW quenching effect that is taking place
Fig. 14. XRD patterns obtained over the Cex Sm1-x O2 (x = 0.2, 0.5, 0.8, 1) catalysts
following the catalytic test described in Fig. 13a.
immediately after the end of the MW reaction (end of reaction
is linked with end of heating) facilitates the formation of a more
defective structure, thus leading to the creation of more oxygen
(reactant) adsorption sites.
observed that for CeO2 single oxide, in the case of MW and con-
Comparing the CO oxidation in the absence and presence of H2
ventional synthesis at 100 ◦ C an improvement of T50 is observed
(only for MW catalysts) in the feed gas, in terms of the T50 param-
by 26% and 29%, respectively, compared to the conventionally pre-
eter it can be stated that in the presence of H2 the reaction is being
pared CeO2 at 60 ◦ C. In the case of Ce0.5 Sm0.5 O2 composition when
facilitated by the drop in temperature. In the presence of H2 , oxi-
the conventional catalyst synthesized at 100 ◦ C outperfomes the
dation of H2 occurs, simultaneously with CO oxidation, leading to
MW and the conventional at 60 ◦ C by only 6% and 13%, respectively.
H2 O production. In the co-presence of H2 O, CO can react towards
Though the much faster MW reaction should kept in mind when a
the Water Gas Shift (WGS) and Reverse Water Gas Shift (RWGS)
meaningful comparison is soughtin terms of practical application.
reactions at T > 200 ◦ C. WGS can lead to an additional route in the
Consequently, there is a profound effect of the synthesis method
production of CO2 , which has been proved to exhibit an inhibiting
on the T50 value for both CeO2 and Cex Sm1-x O2 mixed oxides. This
effect on the CO oxidation reaction in the presence of H2 [73]. From
synthesis effect can be understood on the basis of parameters,
the CO2 TPD studies on these materials (Fig. 11) it is clear that CO2 is
such as solvent (water) used, temperature/temperature field in the
chemisorbed on the surface of CeO2 and Cex Sm1-x O2 mixed oxides
solution and time of synthesis. According to microwave chemistry
forming stable surface species. At this point, it is worthwhile con-
theory, the medium microwave radiation absorption capability is
sidering the oxide acidic character. It has been reported that acidic
described by the tan␦ parameter (0.123 for water) [69]. Hence,
oxides are more resistant to carbonate formation than basic oxides
water absorbs microwave radiation rather efficiently and rapid
[74]. In the case of Ce-containing materials, the acid-base proper-
heating occurs which leads to supersaturation of the metal precur-
ties vary with the cerium cation oxidation state, with Ce4+ being
sors’ solution. In general, the higher the supersaturation, the higher
52 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

Fig. 15. (a–c) the variation of the CO fractional conversion (XCO ) with W/FCO ratio, (d–f) Arrhenius plots for the MW CeO2 , Ce0.8 Sm0.2 O2 , Ce0.2 Sm0.8 O2 catalysts, respectively.
The apparent activation energies calculated from Arrhenius plots for the MW CeO2 , Ce0.8 Sm0.2 O2 , Ce0.2 Sm0.8 O2 are 35.74 ± 1.27 kJ mol−1 (R2 = 0.9937); 37.57 ± 1.54 kJ mol−1
(R2 = 0.99159) and 42.82 ± 4.29 kJ mol−1 (R2 = 0.95164), respectively.

more acidic than Ce3+ (XPS showed the presence of Ce3+ ions). Ce3+ as the population of mobile oxygen species is diluted by Sm dop-
as a hard acidic entity tends to form stable carbonates with a hard ing. It can also be speculated that CO induces a minor reduction of
CO3 2− base [75]. The catalytic activity of CeO2 can be understood on samaria and the reduced species of samaria decorate CeO2 -related
the basis of facile generation of oxygen vacant sites under reaction active sites for the reaction, thus lowering the activity; (iv) From the
conditions followed by the Ce4+ /Ce3+ redox cycle. A very impor- CO2 -TPD studies it can be seen that the amount of CO2 desorbed
tant finding from a technological viewpoint is the fact that for CO from the Sm-doped ceria oxides is lower than that of CO2 desorbed
oxidation the Ce0.8 Sm0.2 O2 catalyst is stable and very active (96% from the undoped ceria (conventional or microwave). This can be
conversion) in stream for 20 h with no deterioration of its catalytic due to an increase of the surface basicity of the Sm-modified oxides
activity (Fig. 13c). This can be attributed to the fact that this cata- [77] which present stronger adsorption sites for the CO2 (reaction
lyst chemisorbs a lower amount of CO2 as indicated by the CO2 -TPD product).
experiment (see Fig. 11). These catalysts, compared to other combinations of CeO2
Also it is worthwhile commenting on the lowering of the cat- dopants, exhibit a less favourable conversion temperature where
alytic activity that accompanies the Sm-doped materials which can the maximum catalytic activity is achieved; however, their perfor-
be explained as follows: (i) The ‘surface-bulk’ oxygen exchange pro- mance is still very promising, particularly as they can be produced
cess in Ceria begins at 650 K (337 ◦ C) according to Gorte et al. [21], rapidly and at low cost. Comparing the materials studied here with
meaning that at lower temperatures only the ceria surface partic- reported MOFs, e.g. ra-MOF where T50 and T100 (temperatures for
ipates, which limits the available pool of oxygen species for the 50% and 100% conversion) were 364 ◦ C and 500 ◦ C, respectively,
reaction at hand; (ii) Based on XPS studies (Fig. 12b) the lattice the superiority of the MW CeO2 and MW Cex Sm1-x O2 catalysts is
oxygen binding energy (eV) is shifted to lower values when Sm is notable given other benefits such as thermal stability and ease of
introduced. This would be expected to facilitate the oxygen avail- preparation and handling [78]. MOF catalysts reported by Jin Yeong
ability, but it seems that temperature is the determining factor Kim et al. [79] showed increased activity (T50 and T100 at 222 ◦ C and
for suppressing the above (as will be discussed in the next sec- 250 ◦ C, respectively) only after Pd nanoparticles were loaded onto
tion). Also, from the H2 -TPR studies it can be seen that Sm-doping them, a feature that elevates the overall cost of the material. The
increases the reducibility of the doped ceria in the low and mid- MW sol-gel CeO2 and Cex Sm1-x O2 catalysts also show an improved
temperature regime, whereas the undoped ceria presented a high activity compared to similar citrate sol-gel synthesized Ce0.6 Zr0.4 O2
reducibility at T > 500 ◦ C. This proves that indeed the temperature catalysts which exhibited a 10%CO conversion (T10 ) of 327 ◦ C [79].
is a key factor for the release of the potentially mobile oxygen Finally, it is important to note that the cubic fluorite structure of
species; (iii) Sm2 O3 does not belong to the easily reducible oxides the catalysts was preserved following the reaction as shown from
[76], and as such it is anticipated that the CO oxidation reaction, the XRD results given in Fig. 14. A shift of the peaks towards higher
that is basically happening through oxidation and reduction of the diffraction angles was noticed after the reaction, although, no peaks
contributing oxides (CeO2 and Sm2 O3 in the particular case) or by corresponding to carbonate species were identified. This peak shift
increasing the oxygen mobility, is being suppressed with Sm doping is indicative of reduction of the cubic ceria lattice [62], whereas the
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 53

absence of carbonate-relevant peaks shows the structural integrity ventionally synthesized material improves significantly when the
of the catalysts and their self-healing capability. preparation takes place at higher temperature (100 ◦ C). Microwave
assisted synthesis gives rise to a significant advantage over the
conventional sol-gel process by CO oxidation occurring at a signif-
3.6. Kinetic studies
icantly lower temperature, with T50 for the CeO2 and Ce0.5 Sm0.5 O2
materials being reduced by 131 ◦ C and 59 ◦ C, respectively, but
Kinetic studies were performed over selected catalytic systems
still maintaining a high CO conversion. Further benefits of the
so to enlighten the promising catalytic behavior of the MW pre-
microwave assisted synthesis are the enhanced presence of oxygen
pared materials.
vacant sites, giving rise to a lower temperature at which reducible
Effect of temperature. The plot of fractional CO convention
oxygen species can be provided and a refinement of the particle
vs. W/F for different reaction temperatures in the 180–280 ◦ C
size.
range is presented in Fig. 15(a)–(c) for CeO2 , Ce0.8 Sm0.2 O2 and
The affinity of the catalytic surface with CO2 (a product of both
Ce0.2 Sm0.8 O2 , respectively. A monotonic increase of CO conversion
the reactions under study) was studied using CO2 -TPD. A lower
can be observed as the temperature is increased (for fixed W/F val-
affinity of the surface with CO2 (as shown by a lower calculated
ues), indicating that CO oxidation over CeO2 and CeO2 -based mixed
Edes value) leads to higher activity of the catalyst. Of the microwave
oxides is being favored by temperature. This is in good agreement
assisted sol-gel synthesized materials, Ce0.8 Sm0.2 O2 shows the
with research already reported [80]. The reason might be that above
lowest Edes of CO2 desorption (8.7 kJ/mol) and a higher catalytic
a certain temperature (260–280 ◦ C), the oxygen that is chemically
activity compared to CeO2 (Edes = 11.2 kJ/mol) and Ce0.5 Sm0.5 O2
adsorbed on the catalyst surface is activated leading to species par-
(Edes = 19.6 kJ/mol). Kinetic studies for the MW catalysts have
ticipating in the CO oxidation and subsequently creating oxygen
demonstrated the monotic increase of CO conversion with tem-
vacancies and hence, improving the activity, since oxygen vacan-
perature and contact time. Similar values of Eact were obtained
cies are a crucial parameter of the oxidation reaction. This trend for
for CeO2 and Ce0.8 Sm0.2 O2 catalysts which shows that both sys-
the XCO increase, presented herein, coincides quite well with the
tems have reactive sites for the studied reaction. Though, the MW
Raman and H2 -TPR studies which show easily reducible oxygen
Ce0.8 Sm0.2 O2 catalyst brings the beneficient combination of high
species and oxygen vacant sites of these catalysts at low- and mid-
activity and thermal stability. Given the facile MW method, MW
temperature regimes (Ce0.8 Sm0.2 O2 ) and high temperature regime
Ce0.8 Sm0.2 O2 could be a scalable system of industrial interest.
(CeO2 ).
Effect of contact time (W/F). The effect of contact time (W/F,
W = catalyst weight (g); F = flow rate of the reactant (mL/min)) on Acknowledgements
the CO oxidation reaction is shown in Fig. 15(a)–(c) as well. A mono-
tonic increase in CO conversion was observed from W/F = 0.015 to KP would like to acknowledge the financial support from Abu
0.070 g s ␮mole. At a low contact time, the CO oxidation reaction Dhabi Educational Council (ADEC B3111) and Khalifa University
does not proceed sufficiently due to the high flow rate. It can be Internal Research Fund (L1 KUIRF-210103) for supporting this
stated that the catalyst active sites are covered with adsorbed research. Part of work presented here (XRD, SEM) is done at Khalifa
species and there is not enough time for them to be regenerated University Core Nanocharacterization Facilities. We acknowledge
and restore their activity for the CO oxidation reaction. As a result, KU and CNCF staff support. Work by AFZ and SYA was made possi-
only a low conversion of CO was observed. Increasing contact time ble by the grant number NPRP 6-351-1-072 from the Qatar National
or increasing catalyst loading, the regeneration and availability Research Fund (a member of Qatar Foundation).
of sufficient population of active sites can be achieved leading to
increase of the CO conversion. The apparent activation energies
calculated from Arrhenius plots (Fig. 15(d)–(f) for the CeO2 MW, References
Ce0.8 Sm0.2 O2 , Ce0.2 Sm0.8 O2 are 35.74 ± 1.27 kJ mol−1 (R2 = 0.9937);
37.57 ± 1.54 kJ mol−1 (R2 = 0.99159) and 42.82 ± 4.29 kJ mol−1 [1] J.L.G. Fierro, Metal Oxides: Chemistry and Applications, CRC Press, 2005.
[2] A. Trovarelli, P. Fornasiero, Catalytic science series Catalysis by Ceria and
(R2 = 0.95164), respectively. These results demonstrate as well the
Related Materials, Vol. 12, 2nd edition, Imperial College Press, 2013.
superiority of the CeO2 MW and Ce0.8 Sm0.2 O2 MW for the reaction [3] S.C. Singhal, K. Kendall, High-Temperature Solid Oxide Fuel Cells:
at hand. The proximity of Eact values for the doped and undoped Fundamentals Design and Applications, Elsevier, 2003.
ceria has also been reported [80] and has been attributed to the [4] S. Betelu, K. Polychronopoulou, C. Rebholz, I. Ignatiadis, Novel CeO2 -based
screen-printed potentiometric electrodes for pH monitoring, Talanta 87 (1)
oxygen vacant sites and grain growth among other parameters. The (2011) 126–135.
comparable Eact values for the CeO2 and Ce0.8 Sm0.2 O2 catalysts, [5] S.J. Schmieg, D.N. Belton, Effect of hydrothermal aging on oxygen
along with the enhanced thermal stability induced by Sm doping, storage/release and activity in a commercial automotive catalyst, Appl. Catal.
B Environ. 6 (1995) 127–144.
make the Ce0.8 Sm0.2 O2 catalyst a very promising system for CO [6] A. Bueno-Loı́pez, K. Krishna, M. Makkee, J.A. Moulijn, Enhanced soot oxidation
oxidation. by lattice oxygen via La3+ -doped CeO2 , J. Catal. 230 (2005) 237–248.
[7] Rui Si, Ya-Wen Zhang, Li-Min Wang, Shi-Jie Li, Bing-Xiong Lin, Wang-Sheng
Chu, Zi-Yu Wu, Chun-Hua Yan, Enhanced thermal stability and oxygen
4. Conclusions storage capacity for Cex Zr1-x O2 (x = 0.4–0.6) solid solutions by hydrothermally
homogenous doping of trivalent rare earths, J. Phys. Chem. C 111 (2) (2007)
787–794.
In this study, microwave assisted and conventional sol-gel CeO2 [8] W.Y. Hernanden, O.H. Laguna, M.A. Centeno, J.A. Odriozola, Structural and
and Cex Sm1-x O2 mixed oxides were synthesized and their activ- catalytic properties of lanthanide (La, Eu Gd) doped ceria, J. Solid State Chem.
184 (2011) 3014–3020.
ity towards CO catalytic oxidation in the absence and presence
[9] D. Jampaiah, K.M. Tur, S.J. Ippolito, Y.M. Sabri, J. Tardio, S.K. Bhargava, B.M.
of H2 was evaluated. FTIR studies showed that during the synthe- Reddy, Structural characterization and catalytic evaluation of transition and
sis intermediate species population is different for the microwave rare earth metal doped ceria-based solid solutions for elemental mercury
oxidation, RSC Adv. 3 (2013) 12963–12974.
assisted compared to the conventional sol-gel method, but there
[10] O.H. Laguna, M.A. Centeno, M.J. Boutonnet, A. Odriozola, Fe-doped ceria solids
is no significant effect of the synthesis method on the material synthesized by the microemulsion method for CO oxidation reactions, Appl.
porosity. However, the different population of intermediate species Catal. B. 106 (2011) 621–629.
that develop during microwave assisted synthesis leads to precur- [11] P. Fornasiero, R. Dimonte, G.R. Rao, J. Kaspar, S. Meriani, A. Trovarelli, M.
Graziani, Rh-Loaded CeO2 -ZrO2 solid-solutions as highly efficient oxygen
sor materials which are overall more thermally stable, preserving exchangers: dependence of the reduction behavior and the oxygen storage
75% of their mass up to 600 ◦ C. The thermal stability of the con- capacity on the structural-properties, J. Catal. 151 (1) (1995) 168–177.
54 K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55

[12] K.C. Petallidou, K. Polychronopoulou, J.L.G. Fierro, A.M. Efstathiou, [39] A. Mishra, R. Prasad, A review on preferential oxidation of carbon monoxide
Low-temperature water-gas shift on Pt/Ce0.8 La0.2 O2-␦ -CNT: the effect of in hydrogen rich gases, Bull. Chem. React. Eng. Catal. 6 (1) (2011) 1–14.
Ce0.8 La0.2 O2-␦ /CNT ratio, Appl. Catal A: Gen. 504 (2015) 585. [40] P.C. Hulteberg, J.G.M. Brandin, F.A. Silversand, M. Lundberg, Preferential
[13] K.C. Petallidou, K. Polychronopoulou, S. Boghosian, S. Garcia-Rodriguez, A.M. oxidation of carbon monoxide on mounted and unmounted noble-metal
Efstathiou, Water–gas shift reaction on Pt/Ce1–x Tix O2-␦ : the effect of Ce/Ti catalysts in hydrogen-rich streams, Int. J. Hydrogen Energy 30 (2005) 1235.
ratio, J. Phys. Chem. C 117 (48) (2013) 25467. [41] I. Lopez, T. Valdes-Solis, G. Marban, An attempt to rank copper-based catalysts
[14] K. Polychronopoulou, J.L.G. Fierro, A.M. Efstathiou, The phenol steam used in the CO-PROX reaction, Int. J. Hydrogen Energy 33 (2008) 197–205.
reforming reaction over mgO-based supported Rh catalysts, J. Catal. 228 (2) [42] Xiaowei Xie, Yong Li, Zhi-Quan Liu, Masatake Haruta, Wenjie Shen,
(2004) 417–432. Low-temperature oxidation of CO catalysed by Co3 O4 nanorods, Nat. Lett. 458
[15] K. Polychronopoulou, A. Bakandritsos, V. Tzitzios, J.L.G. Fierro, A.M. Efstathiou, (2009) 746–749.
Absorption enhanced reforming of phenol by steam over supported-Fe [43] Yang Lou, Jian Ma, Xiaoming Cao, Li Wang, Qiguang Dai, Zhenyang Zhao,
catalysts, J. Catal. 241 (1) (2006) 132–148. Yafeng Cai, Wangcheng Zhan, Yanglong Guo, P. Hu, Guanzhong Lu, Yun Guo,
[16] K. Polychronopoulou, A.M. Efstathiou, Spillover of labile OH, H, and O species Promoting effects of In2 O3 on Co3 O4 for CO oxidation: tuning O2 activation
in the H2 production by steam reforming of phenol over supported-Rh and CO adsorption strength simultaneously, ACS Catal. 4 (2014) 4143–4152.
catalysts, Catal. Today 116 (3) (2006) 341–347. [44] Yang Lou, Xiao-Ming Cao, Jinggang Lan, Li Wang, Qiguang Dai, A. Yun Guo,
[17] T. Baidya, A. Gupta, P.A. Deshpandey, G. Madras, M.S. Hegde, High oxygen Jian Ma, Zhenyang Zhao, Yanglong Guo, P. Huab, Guanzhong Lu,
storage capacity and high rates of CO oxidation and NO reduction catalytic Ultralow-temperature CO oxidation on an In2 O3 –Co3 O4 catalyst: a strategy to
properties of Ce1-x Snx O2 and Ce0.78 Sn0.2 Pd0.02 O2-␦ , J. Phys. Chem. C 113 (10) tune CO adsorption strength and oxygen activation simultaneously, Chem.
(2009) 4059–4068. Commnun. 50 (2014) 6835–6838.
[18] K. Kuntaiah, P. Sudarsanam, B.M. Reddy, A. Vinu, Nanocrystalline [45] K. Ramakanth, Basic of Diffraction and Its Application, I.K. International
Ce1-x Smx O2-␦ (x = 0.4) solid solutions: structural characterization versus CO Publishing House Pvt. Ltd., New Dehli, 2007.
oxidation, RSC Adv. 3 (2013) 7953–7962. [46] C. Li, K. Domen, K. Maruya, T. Onishi, Spectroscopic identification of adsorbed
[19] B.M. Reddy, Structural characterization and catalytic activity of nanosized species derived from adsorption and decomposition of formic acid, methanol,
Cex M1-x O2 (M = Zr and Hf) mixed oxides, J. Phys. Chem. C 112 (2008) and formaldehyde on cerium oxide, J. Catal 125 (1990) 445–455.
11729–11737. [47] A. Gondolini, E. Mercadelli, A. Sanson, S. Albonetti, L. Doubova, S. Boldrini,
[20] K. Eguchi, T. Setoguchi, T. Inoue, H. Arai, Electrical properties of ceria-based Effects of the microwave heating on the properties of gadolinium-doped
oxides and their application to solid oxide fuel cells, Solid State Ionics 52 cerium oxide prepared by polyol method, J. Eur. Ceram. Soc. 33 (2013) 67–77.
(1992) 165. [48] V.K. Saxena, Usha Chandra, 2016. Microwave Synthesis: a Physical Concept
[21] S. Zhao, R.J.A. Gorte, Comparison of ceria and Sm-doped ceria for hydrocarbon 10.5772/22888, 978-953-307-573-0.
oxidation reactions, Appl. Catal. A: Gen. 277 (1–2) (2004) 129–136. [49] Helen J. Kitchen, Simon R. Vallance, Jennifer L. Kennedy, Nuria Tapia-Ruiz,
[22] S. Phokha, S. Pinitsoontorn, P. Chirawatkul, Y. Poo-arporn, S. Maensiri, Lucia Carassiti, Andrew Harrison, A. Gavin Whittaker, Timothy D. Drysdale,
Synthesis, characterization, and magnetic properties of monodisperse CeO2 Samuel W. Kingman, Duncan H. Gregory, Modern microwave methods in
nanospheres prepared by PVP-assisted hydrothermal method, Nanoscale Res. solid-state inorganic materials chemistry: from fundamentals to
Lett. 7 (2012) 425. manufacturing, Chem. Rev. 114 (2014) 1170–1206.
[23] M. Alifanti, B. Baps, N. Blangenois, J. Naud, P. Grange, B. Delmon, [50] Jianmin Chen, Qilu Yao, Jia Zhu, Xiangshu Chen, Zhang-Hui Lu, Rh-Ni
Characterization of CeO2 -ZrO2 mixed oxides. comparison of the citrate and nanoparticles immobilized on Ce(OH)CO3 nanorods as highly efficient
sol-gel preparation methods, Chem. Mater. 15 (2) (2003) 395–403. catalysts for hydrogen generation from alkaline solution of hydrazine, Int. J.
[24] G. Qi, R.T. Yang, R. Chang, MnOx -CeO2 mixed oxides prepared by Hydrogen Energy 41 (2016) 3946–3954.
co-precipitation for selective catalytic reduction of NO with NH3 at low [51] James P. Hos, Paul G. McCormick, Mechanochemical synthesis and
temperatures, Appl. Catal. B: Environ. 51 (2004) 93–106. characterisation of nanoparticulate samarium-doped cerium oxide, Scr.
[25] H. Wang, Jun-Jie Zhu, Jian-Min Zhu, Xue-Hong Liao, Shu Xu, Tao Ding, Mater. 48 (2003) 85–90.
Hong-Yuan Chen, Preparation of nanocrystalline ceria particles by [52] P. Sudarsanam, B. Mallesham, D. Naga Durgasri, B.M. Reddy, Physicochemical
sonochemical and microwave assisted heating methods, Chem. Phys. Chem. characterization and catalytic CO oxidation performance of nanocrystalline
Chem. Phys. 4 (2002) 3794–3799. Ce–Fe mixed oxides, RSC Adv. 4 (2014) 11322.
[26] G. Avgouropoulos, T. Ioannides, Selective CO oxidation over CuO-CeO2 [53] A. Kossoy, A.I. Frenkel, Q. Wang, E. Wachtel, I. Lubomirsky, Local structure and
catalysts prepared via the urea-nitrate combustion method, Appl. Catal. A 244 strain-Induced distortion in Ce0.8 Gd0.2 O1.9 , Adv. Mater. 22 (14) (2010)
(2003) 155. 1659–1662.
[27] D. Gamarra, C. Belver, M. Fernadez-Garcia, A. Martinez-Arias, Selective CO [54] Dongwei Ma, Zhansheng Lu, Yanan Tang, Tingxian Li, Zhenjie Tang, Zongxian
oxidation in excess H2 over copper-ceria catalysts: identification of active Yang, Effect of lattice strain on the oxygen vacancy formation and hydrogen
entities/species, J. Am. Chem. Soc. 129 (2007) 12064–12065. adsorption at CeO2(111) surface, Phys. Lett. A 378 (2014) 2570–2575.
[28] R. Lo Nigro, R. Toro, G. Malandrino, I.L. Fragala, Hetero- epitaxial growth of [55] K.S.W. Sing, D.H. Everett, R.A.W. Hall, L. Moscou, R.A. Pierotti, J. Rouquerol,
nanostructured cerium dioxide thin films by MOCVD on a (001) TiO2 et al., Reporting physisorption data for gas/solid systems with special
substrate, Chem. Mater. 15 (2013) 1434–1440. reference to the determination of surface area and porosity, IUPAC 57 (4)
[29] M. Abi Jaoude, K. Polychronopoulou, S.J. Hinder, M.S. Katsiotis, M.A. Baker, Y.E. (1985) 603–619.
Greish, S.M. Alhassan, Synthesis and properties of 1D Sm-doped CeO2 [56] J. Rouquerol, F. Rouquerol, P. Llewellyn, G. Maurin, K.S. Sing, Adsorption by
composite nanofibers fabricated using a coupled electrospinning and sol–gel Powders and Porous Solids: Principles, Methodology and Applications,
methodology, Ceram. Int. 42 (9) (2016), http://dx.doi.org/10.1016/j.ceramint. Academic Press, 2013.
2016.03.197. [57] B.M. Reddy, L. Katta, G. Thrimurthulu, Novel nanocrystalline Ce1-x Lax O2-␦
[30] S. Gravani, K. Polychronopoulou, V. Stolojan, Q. Cui, P.N. Gibson, S.J. Hinder, Z. (x = 0.2) solid solutions: structural characteristics and catalytic performance,
Gu, C.C. Doumanidis, M.A. Baker, C. Rebholz, Growth and characterization of Chem. Mater. 22 (2010) 467–475.
ceria thin films and Ce-doped ␥-Al2O3 nanowires using sol-gel techniques, [58] M.A. Ebiad, D.R. AbdEl-Hafiz, R.A. Elsalamony, L.S. Mohamed, Ni supported
Nanotechnology 21 (46) (2010) 465606. high surface area CeO2 –ZrO2 catalysts for hydrogen production from ethanol
[31] Aurélien Vantomme, Zhong-Yong Yuan, Gaohui Du, Bao-Lian Su, steam reforming, RSC Adv. 2 (2012) 8145–8156.
Surfactant-assisted large-Scale preparation of crystalline CeO2 nanorods, [59] L. Pratti, A. Villa, Gold Catalysts: Preparation, Characterization and
Langmuir 21 (3) (2005) 1132–1135. Applications, CRC Press, 2016.
[32] A.G. Saskia, Microwave chemistry, Chem. Soc. Rev. 26 (1997) 233–238. [60] P. Burrows, A. Hamnett, A.F. Orchard, G. Thornton, J. Chem. Soc. Dalton Trans.
[33] Shangzhao Shi, Jiann-Yang Hwang, Microwave-assisted wet chemical 17 (1976) 1686.
synthesis: advantages, significance, and steps to industrialization, J. Miner. [61] D.R. Mullins, S.H. Overbury, D.R. Huntley, Electron spectroscopy of single
Mater. Charact. Eng. 2 (2) (2003) 101–110. crystal and polycrystalline cerium oxide surfaces, Surf. Sci. 409 (1998)
[34] Sridhar Komarneni, Nanophase materials by hydrothermal, microwave- 307–319.
hydrothermal and microwave-solvothermal methods, Curr. Sci. 85 (12) [62] A. Pfau, K.D. Schierbaum, The electronic structure of stoichiometric and
(2003) 1730. reduced CeO2 surfaces: an XPS UPS and HREELS study, Surf. Sci. 321 (1994)
[35] Sridhar Komarneni, Hiroaki Katsuki, Nanophase materials by a novel 71–80.
microwave- hydrothermal process, Pure Appl. Chem. 74 (9) (2002) [63] M.A. Henderson, C.L. Perkins, M.H. Engelhard, S. Thevuthasan, C.H.F. Peden,
1537–1543. Redox properties of water on the oxidized and reduced surfaces of CeO2 (111),
[36] C.S. Riccardi, R.C. Lima, M.L. dos Santos, P.R. Bueno, J.A. Varela, E. Longo, Surf. Sci. 526 (2003) 1–18.
Preparation of CeO2 by a simple microwave–hydrothermal method, Solid [64] H.E. Elkins, H.E. Hagelin-Weaver, Oxidative coupling of methane over
State Ionics 180 (2009) 288–291. unsupported and alumina-supported samaria catalysts, Appl. Catal., A 454
[37] M.L. Dos Santos, R.C. Lima, C.S. Riccardi, R.L. Tranquilin, P.R. Bueno, J.A. Varela, (2013) 100–114.
E. Longo, Preparation and characterization of ceria nanospheres by [65] A.E.C. Palmqvist, M. Wirde, U. Gelius, M. Muhammed, Surfaces of doped
microwave-hydrothermal method, Mater. Lett. 62 (2008) 4509–4511. nanophase cerium oxide catalysts, Nanostruct. Mater. 11 (1999) 995–1007.
[38] Li Xiazhang, Hu Zonglin, Zhao Xiaobing, Lu Xiaowang, [66] G.P. Holgado, J.P. Munuera, A.R. González-Elipe, XPS study of oxidation
Ce1–xSmxO2–-attapulgite nanocomposites: synthesis via simple microwave processes of CeOx defective layers, Appl. Surf. Sci. 158 (2000) 164–171.
approach and investigation of its catalytic activity, J. Rare Earths 31 (12) [67] NIST XPS database, http://srdata.nist.gov/xps/.
(2013) 1157.
K. Polychronopoulou et al. / Journal of Molecular Catalysis A: Chemical 428 (2017) 41–55 55

[68] S. Yamazaki, T. Matsui, T. Ohashi, Y. Arita, Defect structures in doped CeO2 [75] Ralph G. Pearson, Hard and soft acids and bases, J. Am. Chem. Soc. 85 (22)
studied by using XAFS spectrometry, Solid State Ionics 136–137 (2000) (1963) 3533–3539.
913–920. [76] Jenshi B. Wang, De-Hao Tsai, Ta-Jen Huang, Synergistic catalysis of carbon
[69] P. Lidstrom, J. Tierney, B. Wathey, J. Westmana, Microwave assisted organic monoxide oxidation over copper oxide supported on samaria-doped ceria, J.
synthesis—a review, Tetrahedron 57 (2001) 9225. Catal. 208 (2002) 370–380.
[70] M. Gharibeh, G.A. Tompsett, W.C. Conner, Microwave reaction enhancement: [77] Yusuke Niwa, Ken-ichi Aika, The effect of lanthanide oxides as a support for
the rapid synthesis of SAPO-11 molecular sieves, Top. Catal. 49 (2008) ruthenium catalysts in ammonia synthesis, J. Catal. 162 (1996) 138–142.
157–166. [78] Jin Yeong Kim, Mingshi Jin, Kyung Joo Lee, Jae Yeong Cheon, Sang Hoon Joo, Ji
[71] W. Chen, X.H. Wang, Sintering dense nanocrystalline ceramics without Man Kim, Hoi Ri Kim, In situ- generated metal oxide catalyst during CO
final-stage grain growth, Nature 404 (2000) 168–171. oxidation reaction transformed from redox-active metal-organic
[72] C. Li, K. Domen, K. Maruya, T. Onishi, Dioxygen adsorption on well-outgassed framework-supported palladium nanoparticles, Nanoscale Res. Lett. 7 (461)
and partially reduced cerium oxide studied by FT-IR, J. Am. Chem. Soc. 111 (2012).
(1989) 7683–7687. [79] E. Bekyarova, P. Fornasiero, J. KasÏpar, M. Grazian, CO oxidation on
[73] A. Di Benedetto, G. Landi, L. Lisi, G. Russo, Role of CO2 on CO preferential Pd/CeO2 –ZrO2 catalysts, Catal. Today 45 (1998) 179.
oxidation over CuO/CeO2 catalyst, Appl. Catal B. 142–143 (2013) 169–177. [80] Lj. Kundakovic, M. Flytzani-Stephanopoulos, Cu- and Ag-Modified cerium
[74] G.C. Bond, D.T. Thompson, Catalysis by gold, Catal. Rev. Sci. 41 (1999) 319. oxide catalysts for methane oxidation, J. Catal. 179 (1998) 203–221.

View publication stats

You might also like