You are on page 1of 17

Nano Research

DOI 10.1007/s12274-016-1296-2

Thermally stable Ir/Ce0.9La0.1O2 catalyst for high


temperature methane dry reforming reaction
Fagen Wang1,2 (), Leilei Xu3, Weidong Shi2, Jian Zhang4, Kai Wu5, Yu Zhao6, Hui Li6, He Xing Li6,
Guo Qin Xu1,4 (), and Wei Chen1,4,7 ()

1
Laboratory of Energy and Environment Interface Engineering, National University of Singapore Suzhou Research Institute, Suzhou
215123, China
2
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
3
School of Environmental Science and Engineering, Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution
Control, Collaborative Innovation Center of the Atmospheric Environment and Equipment Technology, Nanjing University of Information
Science & Technology, Nanjing 210044, China
4
Department of Chemistry, National University of Singapore, 3 Science Drive 3, Singapore 117543, Singapore
5
College of Chemistry and Molecular Engineering, Peking University, Beijing 100871, China
6
Department of Chemistry, Shanghai Normal University, Shanghai 200234, China
7
Singapore-Peking University Research Center for a Sustainable Low-Carbon Future, 1 CREATE Way, #15-01, CREATE Tower, Singapore
138602, Singapore

Received: 20 July 2016 ABSTRACT


Revised: 12 September 2016 In this study, the use of a thermally stable Ir/Ce0.9La0.1O2 catalyst was investigated
Accepted: 23 September 2016 for the dry reforming of methane. The doping of La2O3 into the CeO2 lattice
enhanced the chemical and physical properties of the Ir/Ce0.9La0.1O2 catalyst,
© Tsinghua University Press such as redox properties, Ir dispersion, oxygen storage capacity, and thermal
and Springer-Verlag Berlin stability, with respect to the Ir/CeO2 catalyst. Hence, the Ir/Ce0.9La0.1O2 catalyst
Heidelberg 2016 exhibits higher activity and stabler performance for the dry reforming of methane
than the Ir/CeO2 catalyst. This observation can be mainly attributed to the
KEYWORDS stronger interaction between the metal and support in the Ir/Ce0.9La0.1O2 catalyst
thermally stable catalyst, stabilizing the catalyst structure and improving the oxygen storage capacity,
Ir/Ce0.9La0.1O2, leading to negligible aggregation of Ir nanoparticles and the Ce0.9La0.1O2 support
metal–support interaction, at high temperatures, as well as the rapid removal of carbon deposits at the
methane dry reforming boundaries between the Ir metal and the Ce0.9La0.1O2 support.

1 Introduction decomposition of organic wastes, lead to continuous


increase in the concentrations of greenhouse gases in
Considerable emissions of CO2, NOx, and CH4, the atmosphere. The greenhouse gas effect results
attributed to the combustion of fossil fuels and the in the adverse consequence of climate change, which

Address correspondence to Fagen Wang, fagen.wang@gmail.com; Guoqin Xu, chmxugq@nus.edu.sg; Wei Chen, chmcw@nus.edu.sg
2 Nano Res.

in turn endangers humans. For sustainable human the large-scale industrialization of the dry reforming
development, it is imperative to reduce the emissions of methane. For commercialization, it is imperative
of or utilize these greenhouse gases [1]. to develop a rational design of catalysts exhibiting
CO2 is the one of the main components contributing characteristics of no sintering and no carbon deposition.
to the greenhouse gas effect; hence, CO2 reduction has To avoid sintering, a possible way is to isolate the
attracted considerable attention. In recent years, several nanoparticles in a confined space (similar to core–shell
carbon capture and sequestration (CCS) technologies catalysts) to avoid aggregation [20]. Another potential
and catalytic transformations have been developed way is to strengthen the interaction between the
for CO2 reduction [2–4]. Among catalytic processes, metal and support in supported catalysts to prevent
the dry reforming of methane is crucial, as CH4 and enlargement [21]. As compared to the sophisticated
CO2 produce syngas, which can be further used for methods required for synthesizing confined catalysts,
synthesizing energy chemicals, such as hydrocarbons the strengthening of the interaction between the metal
or olefins [5, 6]; this process not only decreases green- and support in supported catalysts is more suitable,
house gas emissions but also provides important raw attributed to simple preparation methods that can
feedstock for chemical industries, which are beneficial facilitate easy scale-up. To maximize the interaction
from aspects of environment and energy. between the metal and support in supported catalysts,
Among the catalysts investigated for the high- the size and shape of the active metal and support
temperature dry reforming of methane, Ni catalysts need to be finely tuned to obtain a stable catalyst
exhibit the best performance, attributed to their high structure [19, 21–24]. To avoid carbon deposition, the
activity for C–H bond cleavage [7, 8]. Unfortunately, gasification rate of the carbon precursors at the
methane rapidly decomposes on Ni nanoparticles, boundary between the metal and support should be
and the Ostwald ripening of the catalyst results more rapid as compared to the formation rate of the
in considerable amounts of deposited carbon and precursor at the metal sites, which can be achieved by
aggregates of Ni nanoparticles, leading to catalyst using oxides exhibiting oxygen storage capacity (OSC)
deactivation [9]. Hence, a majority of studies have [25]. OSC aids in the removal of carbon deposits by
been focused on the synthesis of small Ni nanoparticles exploiting high oxygen mobility, related to the surface
located in a confined space for decreasing carbon oxygen and partial bulk oxygen species [26]. In addition,
deposits and limiting Ni particle sintering, such as the OSC can lead to the significant enhancement of the
immobilization of Ni nanoparticles into mesoporous metal–support interaction, affording enhanced ability
materials or yolk–shell frameworks [5, 10–13] and the for removing carbon [27].
addition of other elements, such as Fe and Pt, for From the above discussion, catalysts exhibiting
enhancing the resistance of carbon in Ni catalysts properties of strong metal–support interaction and
[14, 15]. Noble metals catalysts, such as those of Rh, Pt, high OSC, such as ceria-based catalysts, are suggested
and Ir, also exhibit high activity for the dry reforming as excellent candidates for the high-temperature dry
of methane [16–18]. Although negligible carbon reforming of methane [28, 29]. However, catalyst per-
deposits are observed, the sintering of noble metal formance gradually decreases because of the thermal
nanoparticles is inevitable, owing to the difficulty instability of pure ceria at high temperature, leading to
in the stabilization of small metal nanoparticles at the decreased Brunauer–Emmett–Teller surface areas
high temperatures. Furthermore, depending on the and enlarged ceria nanoparticles [30, 31]. The sintering
interaction between the noble metal and support of the active metal and formation of carbon deposits
material, these small noble metal nanoparticles tend are observed because of the weakened metal–support
to either aggregate into large clusters or react with the interaction and unstable catalyst structure [32, 33].
support, resulting in decreased active sites, and hence For enhancing thermal stability, doped ceria, such as
reduced catalytic performance [19]. Ce–Zr–O, Ce– Pr–O, Ce–Gd–O, and Ce–La–O, has been
These detrimental issues of catalyst sintering and synthesized to improve the metal–support interaction
carbon deposition at high temperatures have prevented with the aim of limiting sintering of the active metal

| www.editorialmanager.com/nare/default.asp
Nano Res. 3

and reducing the deposition of carbon [21, 34–36]. for 4 h at a heating rate of 1 K/min. For comparison, an
Previously, our group has reported that the doping Ir/CeO2 catalyst was prepared by the same procedure,
of ceria results in the significant enhancement of but without La(NO3)3.
the metal–support interaction in catalysts, leading to
increased catalytic performance and stability for the 2.2 Catalyst characterization
dry reforming of methane [21, 37]. However, although Power X-ray diffraction (XRD) patterns were recorded
the sintering of the active metal is limited and deposition on a PANalytical Empyren DY708 diffractometer
of carbon is reduced, the sintering of the support is operated at 40 kV and 30 mA using Ni-filtered Cu Kα
still observed. Hence, for further limiting the sintering (0.15418 nm) radiation. Transmission electron micros-
of doped ceria, La2O3-doped ceria is synthesized as copy (TEM) images were recorded on a JEOL 2010
the support for the dry reforming of methane in this instrument operated at 200 kV. Specimens were
study. Results revealed that doping with La2O3 leads prepared by the ultrasonic dispersion of samples in
to the enhancement of the interaction between the ethanol. Droplets of the suspension were deposited
metal and support in the Ir/Ce0.9La0.1O2 catalyst, and on a thin carbon film supported on a standard copper
significant improvement in the performance and grid. X-ray photoelectron spectroscopy (XPS) mea-
stability of the Ir/Ce0.9La0.1O2 catalyst, as compared surements were carried out on an AMICUS spectrometer
to the Ir/CeO2 catalyst, for the dry reforming of using Mg Kα radiation. The charging effect was
methane is observed. Furthermore, the sintering of corrected by adjusting the binding energy of C 1s
the Ir nanoparticles and Ce0.9La0.1O2 support was not to 284.6 eV. Temperature-programmed desorption of
observed at a high temperature of 1,073 K for 1,000 h, oxygen (O2-TPD) from supports was performed in a
and negligible carbon deposits were observed on the U-shaped reactor connected to a mass spectrometer
used Ir/Ce0.9La0.1O2 catalyst, demonstrating immense (PrismaPlus). First, the supports were pretreated by
potential for the Ir/Ce0.9La0.1O2 catalyst as a candidate 20% O2/He at 673 K for 30 min, the samples were then
for facilitating the industrialization of the dry reforming purged with He, and the temperature was cooled to
of methane. room temperature. Next, the samples were heated
to 1,073 K at a rate of 5 K/min. The signal of oxygen
(m/e = 32) was detected by the mass spectrometer.
2 Experimental
Temperature-programmed reduction of hydrogen
2.1 Catalyst preparation (H2-TPR) from fresh catalysts and supports were con-
ducted in a U-shaped reactor connected to the mass
The Ir/Ce0.9La0.1O2 catalyst was prepared by deposition– spectrometer (PrismaPlus). Samples were loaded and
precipitation. First, a homogeneous mixture of pretreated under He (50 mL/min) at 573 K for 1 h.
CO(NH2)2, Ce(NO3)3, and La(NO3)3 in a molar ratio of After cooling to room temperature, a mixture of 10%
20:9:1 was heated to 363 K under stirring for 3 h. Next, H2/He (50 mL/min) was introduced over the samples
the mixture was filtered and washed with hot water, to attain a steady state. Then, the temperature was
and the obtained slurry was dried at 373 K for 12 h, increased to 1,023 K at a rate of 5 K/min and maintained
followed by calcination at 673 K for 4 h at a heating at that temperature for 30 min. The signal of H2 (m/e = 2)
rate of 1 K/min, affording the Ce0.9La0.1O2 support. The was detected by the mass spectrometer. The OSCs for
Ce0.9La0.1O2 support and H2IrCl6 solution (nominal Ir the supports and catalysts were measured after the
loading of 2 wt.%) were then homogeneously mixed samples were cooled to 573 K under He (50 mL/min).
in water at 338 K, followed by the slow addition of A stream of O2/He (10 vol.%) was periodically injected
Na2CO3(aq) with stirring until the pH of the mixture over the reduced sample until saturation. Temperature-
changed to pH 8–9. After the suspension was aged programmed desorption of CO 2 (CO 2-TPD) was
for 2 h, it was filtered and thoroughly washed several performed in the U-shaped reactor. First, the samples
times using hot water. Finally, the powder was dried were pretreated by He (20 mL/min) at 573 K for
overnight at 373 K, followed by calcination at 1,023 K 30 min. Next, 5% CO2/He (50 mL/min) was passed

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


4 Nano Res.

over the samples at room temperature for 30 min. velocity (GHSV) of 18,000 mL/(g·h). Catalytic perfor-
After purging with He (50 mL/min), the samples mance was tested from 873 to 1,023 K at intervals
were heated to 1,073 K at a rate of 5 K/min, and the of 50 K. To attain a steady state, measurements were
signal of CO2 (m/e = 44) was detected by the mass performed for at least 1 h at each designated tem-
spectrometer (PrismaPlus). Temperature-programmed perature. Stability tests were conducted at a GHSV of
surface reaction desorption of CH4 (CH4-TPSR) was 18,000 mL/(g·h) at 1,023 K for 200 h and at a GHSV of
performed in the U-shaped reactor. First, the samples 36,000 mL/(g·h) at 1,073 K for 1,000 h. The effluent
were reduced at 1,023 K for 30 min by 10% H2/He and from the reactor was mixed with an internal standard
then cooled to room temperature. Next, 5% CH4/He gas (N2, 5 mL/min), followed by injection into an
(50 mL/min) was introduced to attain a stable state. on-line GC system (Aglient 7890A). The mixture was
Finally, the temperature was increased to 1,023 K at a separated on a packed column (TDX-01) and analyzed
rate of 5 K/min and maintained at that temperature by a thermal conductivity detector (TCD). Conversions
for 30 min. The signals of CH4 (m/e = 16), CO2 (m/e = of CH4 and CO2 were calculated by the following
44), CO (m/e = 18), H2O (m/e = 18), and H2 (m/e = 2) were formula
monitored by the mass spectrometer (PrismaPlus).
FCH4 ,inlet  FCH4 ,outlet
Temperature-programmed surface reaction desorption X CH4   100%
FCH4 ,inlet
of CO2+CH4 (CO2+CH4-TPSR) was performed in the
U-shaped reactor. First, the samples were reduced at FCO2 ,inlet  FCO2 ,outlet
X CO2   100%
1,023 K for 30 min by 10% H2/He and then cooled to FCO2 ,inlet
room temperature, followed by adsorption with 25%
The molar ratio of H2 to CO was calculated as
CO2/He (50 mL/min) for 30 min. Next, 5% CH4/He
follows
(50 mL/min) was introduced to attain a stable state.
Finally, the temperature was increased to 1,023 K at a nH2 FH2 ,outlet

rate of 5 K/min and maintained at that temperature nCO FCO,outlet
for 30 min. The signals of CH4 (m/e = 16), CO2 (m/e = 44),
CO (m/e = 18), H2O (m/e = 18), and H2 (m/e = 2) were In all formulae, Fi represents the flow rate of the i
detected by the mass spectrometer (PrismaPlus). species.
Temperature-programmed oxidation (TPO) of the used Kinetic experiments were also conducted in the
catalysts was conducted in the U-shaped reactor. First, same fixed-bed reactor at atmosphere pressure. In
the used samples were heated to 573 K for 30 min each run, 50 mg of catalyst (80–100 mesh) mixed with
200 mg SiC was loaded for each experiment. The
under He to eliminate possible contaminants. Then,
height of the catalyst bed was approximately 12 mm.
the samples were heated from room temperature
Conversions of CH4 and CO2 were limited to less than
to 1,073 K under 5% O2/He (50 mL/min). The effluent
15% by adjusting the reaction temperature in the
composition of CO2 (m/e = 44) was detected by the
range of 823 to 873 K. For confirming mass and heat
mass spectrometer (PrismaPlus).
transfer limitations in kinetic experiments, the internal
2.3 Catalytic activity tests and kinetic parameter diffusion, external mass diffusion, and heat diffusion
measurements were calculated according to typical equations [38, 39].

The dry reforming of methane was performed in


a continuous-flow fixed-bed tubular quartz reactor 3 Results and discussion
(i.d. = 6 mm, L = 300 mm). The fresh catalyst powder
3.1 Physical and chemical properties of the fresh
was reduced under 20% H2/N2 (30 mL/min) at 1,023 K
catalysts and supports
for 30 min. After purging the reactor with N2, a mixture
of undiluted CH4 and CO2 (molar ratio CH4/CO2 = 1) Figure 1 shows the XRD patterns of the fresh Ir/CeO2
was introduced into the reactor at a gas hourly space and Ir/Ce0.9La0.1O2 catalysts. All samples exhibited

| www.editorialmanager.com/nare/default.asp
Nano Res. 5

defects that positively affect the dispersion of Ir


species [23].
Figure 2 shows the high-resolution TEM (HRTEM)
images of the reduced Ir/CeO2 and Ir/Ce0.9La0.1O2
catalysts. For all samples, randomly oriented polyhedral
nanoparticles were observed. The sizes of CeO2 and
Ce0.9La0.1O2 ranged from 10 to 20 nm. Larger Ir nano-
particles of approximately 6.0 nm were observed on
CeO2 surfaces (Fig. 2(b)), while smaller Ir nanoparticles
of approximately 2.5 nm were observed on Ce0.9La0.1O2
surfaces (Fig. 2(d)). Next, Ir dispersion was appro-
ximately estimated by the reciprocal of the Ir particle
size. As shown in Table 1, the dispersion of Ir for the
Ir/CeO2 catalyst was approximately 16.6%, slightly
less than that observed for the Ir/Ce0.9La0.1O2 catalyst
Figure 1 XRD patterns of the fresh and used Ir catalysts. (a) Fresh (approximately 40.0%). The above observations are in
Ir/CeO2, (b) fresh Ir/Ce0.9La0.1O2, (c) used Ir/CeO2-200, (d) used good agreement with the XRD results obtained earlier;
Ir/Ce0.9La0.1O2-200, and (e) used Ir/Ce0.9La0.1O2-1000. these result also reflect the effect of ceria- based solid
solutions on the stabilization of the active metal
diffraction peaks of CeO2 without La2O3 signals,
dispersion [36].
indicating that La3+ is doped into the CeO2 lattice
Figure 3 shows the H2-TPR profiles of the supports
in the Ir/Ce0.9La0.1O2 catalyst. The lattice parameter of
and Ir catalysts. With respect to the reduction behavior
Ce0.9La0.1O (5.4257 Å) was slightly greater than that
of the supports, main broad reduction peaks were
of CeO2 (5.3940 Å), indicating that the CeO2 lattice
observed at 650–700 and 890–910 K, attributed to the
expands with the introduction of La3+, attributed to
reduction of surface lattice oxygen in CeO2 [40]. Also
the fact that the ionic radius of La3+ (106 pm) is greater
than that of Ce4+ (92 pm). The average crystallite sizes
of Ce0.9La0.1O2 and CeO2 were 10 and 13 nm, respec-
tively. The smaller size of Ce0.9La0.1O2 suggested that
nanoparticle aggregation in Ce0.9La0.1O2 is effectively
hindered, better than in CeO2, contributing to the
higher thermal stability of the crystallite structure. The
higher thermal stability is attributed to the doping of
La3+ into the ceria lattice, which decreases the lattice
parameter of the crystallite as well as the thermal
expansion of the ceria lattice. With respect to the Ir
species, diffraction peaks of IrO2 were observed in the
fresh Ir/CeO2 catalyst but not detected for the fresh
Ir/Ce0.9La0.1O2 catalyst. Furthermore, the size of IrO2
nanoparticles in the fresh Ir/CeO2 catalyst was appro-
ximately 6 nm, while that of IrO2 nanoparticles in the
fresh Ir/Ce0.9La0.1O2 catalyst could not be measured
because no signals were detected. This result is
indicative of higher dispersion for the Ir species in
the Ir/Ce0.9La0.1O2 catalyst, attributed to the stronger Figure 2 HRTEM images of reduced Ir/CeO2 ((a) and (b)) and
metal–support interaction as well as increased oxygen Ir/Ce0.9La0.1O2 ((c) and (d)) catalysts.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


6 Nano Res.

Table 1 O2 desorption, H2 consumption, Ir dispersion, and OSC of samples


Samples O2 desorption (molO/g) H2 consumption (mol/g) OSC (molO/g) Ir dispersiona (%)
CeO2 648 614 634 —
Ce0.9La0.1O2 806 821 796 —
Ir/CeO2 — 910 924 16.6
Ir/Ce0.9La0.1O2 — 1,176 1,065 40.0
a 1
Dispersion of Ir was approximately calculated by D   100% , where dIr is the diameter of Ir nanoparticles; it is determined from
d Ir
HRTEM images of the fresh samples (Fig. 2).

addition, corresponding minor reduction peaks were


observed at 753 and 880 K, attributed to the further
removal of surface oxygen from the supports. The
above results indicated that the redox property and
hydrogen consumption (Table 1) for the Ir/Ce0.9La0.1O2
catalyst, as compared to the Ir/CeO2 catalyst, are
significantly improved because of the stronger metal–
support interaction, which facilitates the reduction of
Ir species and surface oxygen from the support [43].
The reduction of bulk oxygen from ceria always
requires a temperature of greater than 1,100 K [37];
Figure 3 H2-TPR profiles of the supports and Ir catalysts. thus, it is difficult to reduce bulk oxygen with the
current temperature employed.
shown to the left of Fig. 3 is the magnified image To demonstrate the higher oxygen mobility and
showing the minor reduction peaks observed at a higher oxygen vacancy formation in the Ir/Ce0.9La0.1O2
temperature of less than 600 K: Two larger reduction catalyst, as compared to the Ir/CeO2 catalyst, O2-TPD
peaks were observed at approximately 512 and 550 K from the support was carried out (Fig. 4), and the OSCs
for Ce0.1La0.1O2, while significantly smaller peaks were of the catalysts were determined (Table 1). From the
observed for CeO2, attributed to the surface-active O2-TPD profiles, O2 desorption was centered at 845 K
oxygen (superoxide species) of CeO2 [41, 42]. The for CeO2, while it peaked at 880 K for Ce0.9La0.1O2.
relatively higher amounts of superoxide species The oxygen removal amount for Ce0.9La0.1O2 was
implied that significantly more active oxygen species
are formed in Ce0.9La0.1O2 oxide, which in turn exhibits
enhanced chemical activity with hydrogen. Hydrogen
consumption values for CeO2 and Ce0.9La0.1O2 were
614 and 821 mol/g (Table 1), respectively. The higher
hydrogen consumption for Ce0.9La0.1O2 is attributed
to the structural modification of the ceria lattice,
attributed to the substitution of Ce4+ by La3+; this
substitution favors the diffusion of O2– to the Ce0.9La0.1O2
support surfaces. With respect to the reduction profiles
of the Ir catalysts, hydrogen was mainly consumed at
550–630 and 510 K for the Ir/CeO2 and Ir/Ce0.9La0.1O2
catalysts, respectively, attributed to the co-reduction
of IrO2 and the partial surface oxygen of CeO2 [37]. In Figure 4 O2-TPD profiles of the CeO2 and Ce0.9La0.1O2 supports.

| www.editorialmanager.com/nare/default.asp
Nano Res. 7

806 molO/gcat, slightly greater than that observed after subtracting the background for tentatively
for CeO2 (648 molO/gcat); this result is attributed to quantifying the relative percentage of Oβ, which is a
the replacement of high-valance Ce4+ site in the ceria crucial parameter for measuring oxygen vacancy. The
lattice by low-valence La3+, which favors the movement relative concentrations of Oβ were approximately 54%
of bulk oxygen and leads to a higher oxygen mobility and 59% for the reduced Ir/CeO2 and Ir/Ce0.9La0.1O2
for Ce0.9La0.1O2 [44]. The OSC for the Ir/Ce0.9La0.1O2 catalysts, respectively. The slightly higher concen-
catalyst was 1,065 molO/gcat, slightly greater than trations of Oβ demonstrated the positive effect of
that observed for the Ir/CeO2 catalyst (924 molO/gcat). the doping of La3+ on the enhancement of oxygen
This result indicated that higher amounts of oxygen vacancies [45]. With respect to the Ce 3d spectrum,
vacancies are filled by oxygen atoms in the Ir/Ce0.9La0.1O2 eight peaks corresponding to four pairs of spin-orbit
catalyst during oxygen treatment [36]. splitting doublets were observed. The u’ and v’ peaks
Figure 5 shows the XPS spectra of Ir 4f, O 2p, Ce 3d, were attributed to Ce3+ 3d3/2 and Ce3+ 3d5/2, respec-
and La 3d in the reduced Ir/Ce0.9La0.1O2 and Ir/CeO2 tively, the u, u’’, and u’’’ peaks were attributed to
catalysts. In the Ir 4f spectrum, peaks were observed Ce4+ 3d3/2, and those of v, v’’, and v’’’ were attributed
at binding energies of 60.8 and 63.6 eV, attributed to to Ce4+ 3d5/2 [46]. The co-existence of Ce3+ and Ce4+
Ir0 species; this result confirms the presence of Ir suggested the presence of oxygen defects in the sample,
on the surface of the support after pretreatment by which is in agreement with the detection of Oβ in the
hydrogen reduction. In the O 1s spectrum, broad O 1s profiles. The La 3d spectrum exhibited four peaks
peaks were observed at 531.8 and 528.9 eV, attributed at 833.3, 837.7, 850.4, and 854.8 eV, respectively. These
to the oxygen species corresponding to defect (Oβ) and peaks are different as compared to those observed
lattice oxygen (O), respectively. The O 1s spectra were for the binding energies of standard La2O3 (835.8
curve-fitted using a Gaussian–Lorentzian procedure and 852.6 eV), attributed to the different coordination

Figure 5 XPS spectra of (a) Ir 4f, (b) O 1s, (c) Ce 3d, and (d) La 3d of the Ir/Ce0.9La0.1O2 catalysts.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


8 Nano Res.

conditions of La3+ and the La–O distance in Ce0.9La0.1O2 that of Ce4+, leading to a slightly weaker interaction
as compared with that in pure La2O3 [47]. with CO2 [50]. However, CO2 desorption areas for
Ce0.9La0.1O2 and Ir/Ce0.9La0.1O2 were greater than those
3.2 Dynamic reaction analysis of the dry reforming observed for CeO2 and Ir/CeO2, indicating that more
of methane over the supports and catalysts CO2 can be adsorbed on the doped samples. Increased
3.2.1 Activation of CO2 on supports and catalysts adsorption should result in increased CO2 activation,
leading to enhanced catalytic performance for the dry
Figure 6 shows the CO2-TPD profiles of the supports reforming of methane.
and Ir catalysts. For all supports, the desorption of
CO2 was observed at three temperatures. The first 3.2.2 Interaction of CH4 with the supports and catalysts
(CH4-TPSR)
peak at 474 K was attributed to the interaction of
monodentate carbonate with CeO2. The second peak For investigating the interaction of CH4 with the
at 664 K was attributed to bidentate carbonate [48]. supports and catalysts, CH4-TPSR profiles were
The third peak at 945 K was attributed to the internal obtained (Fig. 7). Traces of H2 were detected from
polydentate carbonates in the bulk of oxides [49]. 980 and 800 K for CeO2 and Ce0.9La0.1O2, respectively.
After the loading of Ir, CO2 desorption was observed Negligible consumption of methane by the supports
at two temperature regions of 460–542 and 710–730 K, was attributed to weak interaction between methane
attributed to monodentate and bidentate carbonates, and the supports (Fig. 7(a) and 7(b)). By contrast, the
respectively. Polydentate carbonate was absent owing significant enhancement of the interaction between
to the decomposition of internal carbonates during methane and the Ir catalysts was observed (Fig. 7(c)
the high-temperature calcination, and because it could and 7(d)). First, minor products CO2 and H2O were
not be formed in the CO2 treatment at room tem- observed at temperatures of less than 800 K, attributed
perature. Notably, the CO2 desorption temperatures to the interaction between CHx intermediates and
were slightly shifted to lower temperatures for the the surface oxygen in ceria [21, 51]. Second, the main
doped oxides and catalysts, as compared to the products CO and H2 were formed at temperatures of
undoped oxides and catalysts, suggesting that the greater than 800 K, attributed to the surface reaction
doping of La3+ results in decreased basicity of the between CHx intermediates and accessible bulk oxygen
samples. This decreased basicity possibly originates in CeO2 [52]. The Ir/CeO2 catalyst exhibited loss of CO
from the lower oxidation state of La3+ as compared to and H2 at 600 and 700 K, respectively, slightly less
than that observed for the Ir/Ce0.9La0.1O2 catalyst (700
and 750 K, respectively). From the indirect dissociation
mechanism for the activation and transformation of
methane on noble catalysts, methane is proposed to
be activated on the active metal, affording adsorbed
H and CHx intermediates; these CHx intermediates
react with the bulk oxygen of oxide, generating CO
and H that are adsorbed on the metal site at the
boundary between the metal and support [3, 53, 54].
For the desorption of H2 and CO from the metal sites,
metal–H and metal–COad bonds should be broken.
As the energy required to break these bonds can
only be provided by the temperature in this case,
the slightly higher initial desorption temperature of
H2 and CO on the Ir/Ce0.9La0.1O2 catalyst indicated
stronger bond strength between CO and H2 and the
Figure 6 CO2-TPD profiles of the supports and the Ir catalysts. Ir sites as compared with those between CO and H2

| www.editorialmanager.com/nare/default.asp
Nano Res. 9

Figure 7 CH4-TPSR of the supports ((a) CeO2 and (b) Ce0.9La0.1O2) and catalysts ((c) Ir/CeO2 and (d) Ir/Ce0.9La0.1O2).

and the Ir/CeO2 catalyst. Methane consumption for can possibly be attributed to the different electronic
the Ir/Ce0.9La0.1O2 catalyst was observed at 920 K, surroundings of Ir nanoparticles after CO2 adsorption.
slightly less than that observed for the Ir/CeO2 catalysts Minor products H2 and CO2 were centered at 690 and
(940 K), indicative of the higher ability for methane 632–800 K, respectively, for the Ir/CeO2 catalyst, while
conversion on the Ir/Ce0.9La0.1O2 catalyst caused by these products were observed at 710 and 655–810 K,
the smaller size of Ir nanoparticles and the easier respectively, for the Ir/Ce0.9La0.1O2 catalyst. These
movement of partial bulk oxygen in the Ce0.9La0.1O2 peaks are attributed to the reaction between CHx
support. intermediates and the surface oxygen of supports. At
temperatures of greater than 800 K, main products H2
3.2.3 Temperature-programmed surface reaction of methane
and CO were observed for the catalysts, attributed
on the supports and catalysts after CO2 adsorption (CO2+
CH4-TPSR) to the interaction of CHx intermediates with the
accessible bulk oxygen of supports. The above results
Figure 8 gives the results of CO2+CH4-TPSR after CO2 demonstrate that both the surface and bulk oxygen
adsorption on the supports and catalysts. Methane of the support are involved in the transformation of
consumption was clearly not detected, and only CHx intermediates, which can be considered as a route
traces of H2 were observed for CeO2 and Ce0.9La0.1O2 for the elimination of carbon deposits during the dry
supports (Fig. 8(a) and 8(b)). In contrast, methane reforming of methane.
consumption peaks were clearly observed at 964 and
3.2.4 Kinetic parameters and turnover frequencies (TOFs)
940 K for Ir/CeO2 and Ir/Ce0.9La0.1O2 (Fig. 8(c) and 8(d)),
for the dry reforming of methane over the catalysts
respectively. As compared with the result observed
from Fig. 7, the higher CH4 consumption temperatures Before measuring the kinetic parameters and TOF, the

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


10 Nano Res.

Figure 8 CO2+CH4-TPSR of the supports ((a) CeO2 and (b) Ce0.9La0.1O2) and catalysts ((c) Ir/CeO2 and (d) Ir/Ce0.9La0.1O2).

criteria for external mass diffusion, internal diffusion, rather similar to previously reported values [55–57].
and the combined interphase and intra-particle Although the Ea observed for CH4 over the Ir/Ce0.9La0.1O2
heat and mass transport were calculated in Table 2. catalyst was slightly higher, the pre-exponential factor
The criterion values for methane and CO2 over the was significantly enhanced, suggesting that the collision
Ir/CeO2 and Ir/Ce0.9La0.1O2 catalysts were less than 0.15, frequency of CH4 is promoted at the catalyst surface,
indicative of no external mass-transport limitations. caused by the higher dispersion of Ir. With respect
Similarly, the internal criterion values of methane to the Ea values observed for CO2, the lower value
and CO2 over the catalysts were below 1, indicative for the Ir/Ce0.9La0.1O2 catalyst indicated easier CO2
of no internal mass diffusion limitations. The Mears activation, attributed to the improved basicity caused
criterion for the combined interphase and intraparticle by doping with La3+ (Fig. 6). With respect to intrinsic
heat and mass transport of methane and CO2 was rate measurement, the TOF values for CH4 and
far less than 3, further indicative of no interphase CO2 for the Ir/CeO2 catalyst were 0.09 and 0.12 s–1,
and intraparticle heat transfer or mass-transport respectively, slightly less than the corresponding values
limitations. These values revealed that the mass- and observed for the Ir/Ce0.9La0.1O2 catalyst (0.12 and
heat-transfer limitations can be neglected in kinetic 0.21 s–1). This result reflected the higher performance
measurements. for the Ir/Ce0.9La0.1O2 catalyst. As compared with the
The kinetic parameters and TOF values for the dry previously reported TOF values, those observed for
reforming of methane (Table 3) were compared for CH4 and CO2 on the Ir/CeO2 and Ir/Ce0.9La0.1O2 catalysts
the catalysts. In the Arrhenius plots (Fig. 9), the Ea were similar to the values observed for the Ir/Al2O3
values for CH4 were 60.7 and 70.1 kJ/mol, while those catalyst, albeit less than those observed for Ru/Al2O3,
for CO2 were 74.4 and 64.9 kJ/mol for the Ir/CeO2 and Pd/Al2O3, and Rh/Al2O3 catalysts [57]. This result is
Ir/Ce0.9La0.1O2 catalysts, respectively. The Ea values are attributed to the different abilities for the activation

| www.editorialmanager.com/nare/default.asp
Nano Res. 11

Table 2 Rates of methane and CO2, and the criterion values for internal diffusion, external diffusion, and combined interphase and
a
intra-particle heat, and mass transport
Mears criterion for combined
CH4 rate CO2 rate Mears criterion for Weisz–Prater criterion
Catalyst interphase and intra-particle
(kmol/(kmolcat·s)) (kmol/(kmolcat·s)) external diffusion for internal diffusion
heat and mass transport
 rA  b Rn  c R 2
 rA(obs)  rA R 2
kcCAb DeCAs CAb De

CH4 CO2 CH4 CO2 CH4 CO2


Ir/CeO2 1.96E–05 3.08E–05 0.003 0.002 0.23 0.34 3.3E–05 4.7E–05
Ir/Ce0.9La0.1O2 2.28E–05 3.84E–05 0.005 0.004 0.33 0.56 4.6E–05 7.8E–05
<0.15 <1 <3
a
Catalyst particle size R = 180 m; bulk density of catalyst b = 707 kg/m ; solid catalyst density c = 7,132 kg/m ; reaction orders of
3 3

CH4 and CO2 are estimated 1 and 0.5, respectively [20]; bulk gas concentration CAb and gas concentration on the catalyst surface CAs are
both 0.0006 kmol/m3; mass-transfer coefficient kc = 1.02 m/s, the value of which was adopted from Ref. [39]; effective gas-phase diffusivity
De = 2.6 × 105 m2/s, the value of which was adopted from Ref. [39].

Table 3 Comparison of the kinetic parameters of MDR in this study with those reported previously
CH4 CO2 NIr b
Catalysts Refs.
Ea (kJ/mol) TOF (s–1) Ea (kJ/mol) TOFa (s–1) (mol/gcat)

2%Ir/CeO2 60.7 0.09 74.4 0.12 17 This work


2%Ir/Ce0.9La0.1O2 71.8 0.12 64.9 0.21 40 This work
c
1%Ir/Al2O3 111.7 0.07 131.0 0.18 —  [57]
1%Ru/Al2O3 92.5 0.53 131.0 1.36 —  [57]
1%Pd/Al2O3 87.1 0.36 90.4 0.64 —  [57]
1%Rh/Al2O3 66.1 0.22 52.7 0.32 —  [57]
1%Pt/ZrO2 77.0 0.77 62.8 — —  [17]
1%Pt/5%ZrO2/Al2O3 54.8 1.8 68.7 — — [17]
0.5%Rh/Al2O3 83.7 — — — — [3]
2%Rh/La2O3-ZrO2 143 — 113 — — [55]
0.6%Rh/La2O3-SiO2 61.5 — 49.4 — — [56]
a b
TOF: turnover frequencies of CH4 and CO2 over the Ir/Ce0.9La0.1O2 and Ir/CeO2 catalysts at 823 K. Surface Ir sites were calculated by
D
N Ir,surface   106 μmol/g cat , where ω is the weight faction of Ir, D is the dispersion of Ir, and MIr is the relative atomic weight of Ir;
M Ir
1
dispersion of Ir was approximately calculated by D   100% , where dIr is the diameter of Ir nanoparticles, which is determined
d Ir
from the HRTEM images of the fresh samples (Fig. 2); ratios of pre-exponential factors for CH4 and CO2 rates over the Ir/Ce0.9La0.1O2
and Ir/CeO2 catalysts are 145 and 0.15, respectively, calculated from the intercept of the Arrhenius plots shown in Fig. 9. cNot available.

or transformation of CH4 molecules on different metal Ir/Ce 0.9La 0.1O2 catalysts for the dry reforming of
nanoparticles and CO2 molecules on the supports. methane. CH4 and CO2 conversions progressively
increased with increasing temperature, attributed
3.3 Performance and stability of the catalysts for
to the endothermic characteristics associated with the
the dry reforming of methane
dry reforming of methane: High temperature promotes
Figure 10 shows the performances of the Ir/CeO2 and high conversion of feedstock. At 823 K, methane and

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


12 Nano Res.

the formation of oxygen vacancies for activating CO2.


Notably, the above conversion values were less than the
equilibrium conversion, indicating that dry reforming
is an equilibrium-limited reaction. In addition, the
conversion of CO2 was higher than that of CH4,
attributed to the occurrence of the reverse water-gas
shift reaction at high temperatures.
Figure 11 shows the effect of different GHSVs on
the performance of the Ir/Ce0.9La0.1O2 catalyst for the
dry reforming of methane. The results indicated that
both CH4 and CO2 conversions gradually decrease with
increasing GHSV. At the lowest GHSV of 6,000 mL/(g·h),
the CH4 and CO2 conversions for the Ir/CeO2 catalyst
Figure 9 Arrhenius plots for CH4 and CO2 reaction rates over were 70% and 80%, respectively, slightly less than
the Ir/Ce0.9La0.1O2 and Ir/CeO2 catalysts.
the corresponding conversions of 79% and 88% over
the Ir/Ce0.9La0.1O2 catalyst. In contrast, at the highest
GHSV of 36,000 mL/(g·h), CH4 and CO2 conversions
for the Ir/CeO2 catalyst decreased to 31% and 47%,
respectively, and the corresponding values for the
Ir/Ce0.9La0.1O2 catalyst decreased to 47% and 61%. The
decreased conversions observed at high GHSVs are
speculated to be caused by the shortened contact time
of reactants. With increasing GHSV, the flow rate of
reactants increased, and more reactant molecules were
introduced, reducing the contact time of reactants and
the more insufficient active sites, which inevitably
decreases catalytic conversion [58].
Figure 12 shows the results obtained from the stability
tests over the two catalysts. At 1,023 K and a GHSV
Figure 10 Performances of Ir/CeO2 and Ir/Ce0.9La0.1O2 catalysts for
of 18,000 mL/(g·h), CH4 conversion decreased from
the dry reforming of methane (reaction conditions: mass of catalyst:
100 mg, molar ratio CH4/CO2 = 1, GHSV = 18,000 mL/(g·h)).
52% to 45% and CO2 conversion decreased from 71%
to 66% for the Ir/CeO2 catalyst after 200 h on stream.
CO2 conversions for the Ir/CeO2 and Ir/Ce0.9La0.1O2 By contrast, almost constant CH4 and CO2 conversions
catalysts were less than 10%. Increasingly, CH4 and of approximately 65% and 79%, respectively, were
CO2 conversions for the Ir/CeO2 catalyst improved to observed for the Ir/Ce0.9La0.1O2 catalyst (Fig. 12(a)).
55% and 71%, respectively; these conversions increased Even with the increase of GHSV to 36,000 mL/(g·h)
to 74% and 85% over the Ir/Ce0.9La0.1O2 catalyst at a and temperature to 1,073 K, CH4 and CO2 conversions
temperature of 1,073 K. Considering the structural were still retained at 60% and 70%, respectively, at
properties of the catalysts, higher methane conversion 1,000 h for the Ir/Ce0.9La0.1O2 catalyst, further confirming
for the Ir/Ce0.9La0.1O2 catalyst was attributed to the the high stability performance of the Ir/Ce0.9La0.1O2
smaller size of Ir nanoparticles, which provides more catalyst. Results obtained from the stability tests were
active sites for the activation and transformation of strongly indicative of the higher thermal stability of
methane. On the other hand, the higher CO2 conversion the Ir/Ce0.9La0.1O2 catalyst for the dry reforming of
for the Ir/Ce0.9La0.1O2 catalyst was attributed to the methane, attributed to the enhanced interaction between
modification of La2O3, which enhances the basicity of the metal and support for significantly limiting the
supports (CO2-TPD) for the adsorption of CO2 and size enlargement of Ir and the support (as discussed

| www.editorialmanager.com/nare/default.asp
Nano Res. 13

Figure 11 Effect of space velocity on the performance of the Ir/CeO2 and Ir/Ce0.9La0.1O2 catalysts for dry reforming of methane.

Figure 12 (a) Stability test for the dry reforming of methane over the Ir/Ce0.9La0.1O2 and Ir/CeO2 catalysts at 1,023 K and GHSV =
18,000 mL/(g·h) for 200 h, and (b) stability performance over the Ir/Ce0.9La0.1O2 catalyst at 1,073 K and GHSV = 36,000 mL/(g·h) for 100,0 h.

below), as well as maintaining the sites for molecules Ce0.9La0.1O2 or CeO2, and t represents reaction time
activation and the interfaces for intermediate tran- (200 or 1,000 h).
sformation. Besides, the enhanced OSC attributed to From the XRD patterns of the used samples (Fig. 1),
the doping of La3+ for the Ir/Ce0.9La0.1O2 catalyst also the Ir diffraction peak was not observed for the used
contributes to the higher ability for the gasification of Ir/Ce0.9La0.1O2-200 and Ir/Ce0.9La0.1O2-1000 catalysts,
carbon precursors, leading to significantly less carbon indicating that the high dispersion of Ir is possibly
deposits. maintained in the used Ir/Ce0.9La0.1O2 catalysts. In
contrast, Ir diffraction was observed for the used
3.4 Characterization of the used catalysts
Ir/CeO2-200 catalyst, and the size of the Ir nano-
For determining the relationship between the catalytic particles increased to 10 nm, indicative of the sintering
performance and catalyst structure, the used samples of Ir nanoparticles in the used Ir/CeO2 catalyst. With
were characterized by XRD, XPS, TEM, and TPO. respect to the supports, the size of the CeO2 nano-
In the following measurements, used catalysts will particles clearly increased from 13 to 18 nm, while that
be referred to as used Ir/A-t, where A represents of the Ce0.9La0.1O2 nanoparticles slightly increased

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


14 Nano Res.

from 10 to 13 nm only after 200 h on stream. Although


Ce0.9La0.1O2 reacted for 1,000 h at 1,073 K, its size was
only 15 nm for the used Ir/Ce0.9La0.1O2-1000 catalyst.
This size variation suggested that significantly limited
aggregation is observed for the Ce0.9La0.1O2 support
used in the Ir/Ce0.9La0.1O2 catalyst, indicating that its
structure is more stable than that of the Ir/CeO2
catalyst.
The XPS profiles of Ir 4f, O 1s, Ce 3d, and La 3d
of the used Ir/Ce0.9La0.1O2 catalysts shown in Fig. 5
also provide some insights into the stability of the
catalyst structure. Both the Ir 4f spectra on the used
Ir/Ce0.9La0.1O2-200 and used Ir/Ce0.9La0.1O2-1000 catalysts
exhibited two peaks at 60.8 and 63.6 eV, respectively,
attributed to metallic Ir0, indicating that Ir0 is retained
on the Ce0.9La0.1O2 support surface. O 1s profiles of the
used Ir/Ce0.9La0.1O2 catalysts demonstrated the presence
of oxygen defect and lattice oxygen species. The ratio
of the oxygen defects on the used Ir/Ce0.9La0.1O2-1000
catalyst surface was 9.5%, similar to a ratio of 11.0%
observed for the used Ir/Ce0.9La0.1O2-200 catalyst surface,
indicative of stable oxygen species in the catalyst. The
strength and shape of the Ce 3d and La 3d spectra for
the used catalysts were similar to those observed for
the fresh samples, indicative of the retained state of
Ce and La species in the used Ir/Ce0.9La0.1O2 catalysts. Figure 13 HRTEM images of used Ir/CeO2 ((a) and (b)) and
In the TEM images of the used Ir/CeO2-200 catalyst Ir/Ce0.9La0.1O2 ((c) and (d)) catalysts after 200 h on stream at
(Fig. 13(a) and 13(b)), approximately 9–10 nm Ir nano- 1,023 K and used Ir/Ce0.9La0.1O2 ((e) and (f)) catalysts after 1,000 h
particles observed on the ceria surface were covered on stream at 1,073 K.
by carbon deposits. In the former, the enlarged Ir
nanoparticles would decrease the boundaries between gasification of carbon precursors. Thus, significantly
the metal and support and weaken the metal–support more stable performance is observed for the Ir/
interaction. In the latter, the encapsulated Ir nano- Ce0.9La0.1O2 catalyst. In addition, although the size of
particles would block the activation of methane. Ir nanoparticles on the used Ir/Ce0.9La0.1O2-1000 catalyst
Both these effects led to decreased performance for slightly increased to 2–3 nm (Fig. 13(e) and 13(f)), the
the Ir/CeO2 catalyst in the stability test. In contrast, observed aggregation was too limited to affect catalyst
2–3 nm-sized Ir nanoparticles were still in intimate performance, attributed to the stronger metal–support
contact with the surface of the Ce0.9La0.1O2 support, interaction and the sufficient metal-support boundaries
and carbon deposits were not clearly observed for the to transform intermediates and gasify carbon pre-
used Ir/Ce0.9La0.1O2-200 catalyst (Fig. 13(c) and 13(d)). cursors at a high temperature of 1,073 K.
The constant size of the Ir nanoparticles was attributed For analyzing the carbon deposits observed on
to the stronger metal–support interaction of the Ir/ the used catalysts, TPO was performed, and Fig. 14
Ce0.9La0.1O2 catalyst in limiting the movement and shows the results obtained. CO2 was released at
aggregation of the Ir nanoparticles [23], which in turn 670 and 870–1,005 K, corresponding to amorphous
preserves the boundaries between the metal and carbon and graphitic carbon, respectively, for the
support for the transformation of intermediates and used Ir/CeO2-200 catalyst [59]. The ratio of the areas

| www.editorialmanager.com/nare/default.asp
Nano Res. 15

improvement in the interaction between the metal


and support, leading to enhanced oxygen storage
capacity and redox behavior, improved Ir dispersion,
and a stable catalyst structure. The thermally stable
Ir/Ce0.9La0.1O2 catalyst afforded higher performance
and stability for the dry reforming of methane,
attributed to stronger metal-interaction in limiting the
size enlargement of Ir nanoparticles and the Ce0.9La0.1O2
support, as well as the improved oxygen storage
capacity for decreasing carbon deposition.

Acknowledgements
Figure 14 TPO results of used Ir/CeO2 and Ir/Ce0.9La0.1O2
The authors acknowledge the financial supports from
catalysts.
National Natural Science Foundation of China (Nos.
indicated that mainly graphitic carbon is formed. 21503142 and 21503113), Singapore National Research
Instead, almost amorphous carbon (620 K) with marginal Foundation CREATE-SPURc program (No. R-143-001-
graphitic carbon (980 K) was observed for the used 205-592), Singapore MOE Tier II (No. R143-000-542-112),
Ir/Ce0.9La0.1O2-200 catalyst, and only graphitic carbon and Academia-Industry Collaborative Innovation
(980 K) without amorphous carbon was observed for Foundation from Jiangsu Science and Technology
the Ir/Ce0.9La0.1O2-1000 catalyst. This result indicated Department (No. BY2014139).
that the nature of the carbon deposits is significantly
affected by catalysts and reaction conditions. Amorphous Electronic Supplementary Material: Supplementary
carbon possibly undergoes the transformation to material (TG analysis in the used samples) is available
graphitic carbon at higher temperatures. Although in the online version of this article at http://dx.doi.org/
carbon deposits were observed in the TPO profiles, 10.1007/s12274-016-1296-2.
their weight estimated by thermogravimetry (TG) was
negligible (Fig. S1 in the Electronic Supplementary References
Material). Only 0.45 wt.% of carbon was deposited
on the used Ir/Ce0.9La0.1O2-200 catalysts, which is less [1] Song, C. S. Global challenges and strategies for control,
than that observed on the used Ir/CeO2-200 catalysts conversion and utilization of CO2 for sustainable development
(0.70 wt.%). This result is attributed to the enhanced involving energy, catalysis, adsorption and chemical processing.
interaction between Ir and the support, which in Catal. Today 2006, 115, 2–32.
turn leads to the improvement in the OSC of the [2] Satthawong, R.; Koizumi, N.; Song, C. S.; Prasassarakich, P.
Ir/Ce0.9La0.1O2 catalyst and facilitates the gasification Bimetallic Fe–Co catalysts for CO2 hydrogenation to higher
hydrocarbons. J. CO2 Util. 2013, 3–4, 102–106.
of carbon precursors by the surface and accessible
[3] Pakhare, D.; Spivey, J. A review of dry (CO2) reforming of
bulk oxygen of support; hence, low amounts of
methane over noble metal catalysts. Chem. Soc. Rev. 2014,
carbon deposits (0.6 wt.%) are observed on the used
43, 7813–7837.
Ir/Ce0.9La0.1O2-1000 catalyst.
[4] Olajire, A. A. Valorization of greenhouse carbon dioxide
emissions into value-added products by catalytic processes.
4 Conclusions J. CO2 Util. 2013, 3–4, 74–92.
[5] Xie, T.; Zhao, X. Y.; Zhang, J. P.; Shi, L. Y.; Zhang, D. S. Ni
In summary, Ir/Ce0.9La0.1O2 and Ir/CeO2 catalysts were nanoparticles immobilized Ce-modified mesoporous silica
used for the dry reforming of methane. As compared via a novel sublimation-deposition strategy for catalytic
to the Ir/CeO2 catalyst, the Ir/Ce0.9La0.1O2 catalyst reforming of methane with carbon dioxide. Int. J. Hydrogen
formed by the doping of La2O3 resulted in significant Energy 2015, 40, 9685–9695.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


16 Nano Res.

[6] Jiao, F.; Li, J. J.; Pan, X. L.; Xiao, J. P.; Li, H. B.; Ma, H.; [19] Li, W. Z.; Kovarik, L.; Mei, D. H.; Liu, J.; Wang, Y.; Peden,
Wei, M. M.; Pan, Y.; Zhou, Z. Y.; Li, M. R. et al. Selective C. H. F. Stable platinum nanoparticles on specific MgAl2O4
conversion of syngas to light olefins. Science 2016, 351, spinel facets at high temperatures in oxidizing atmospheres.
1065–1068. Nat. Commun. 2013, 4, 2481.
[7] Liu, C. J.; Ye, J. Y.; Jiang, J. J.; Pan, Y. X. Progresses in the [20] Adijanto, L.; Bennett, D. A.; Chen, C.; Yu, A. S.; Cargnello,
preparation of coke resistant Ni-based catalyst for steam and M.; Fornasiero, P.; Gorte, R. J.; Vohs, J. M. Exceptional
CO2 reforming of methane. ChemCatChem 2011, 3, 529–541. thermal stability of Pd@CeO2 core–shell catalyst nano-
[8] Du, X. J.; Zhang, D. S.; Gao, R. H.; Huang, L.; Shi, L. Y.; structures grafted onto an oxide surface. Nano Lett. 2013,
Zhang, J. P. Design of modular catalysts derived from 13, 2252–2257.
NiMgAl-LDH@m-SiO2 with dual confinement effects [21] Wang, F. G.; Xu, L. L.; Zhang, J.; Zhao, Y.; Li, H.; Li, H. X.;
for dry reforming of methane. Chem. Commun. 2013, 49, Wu, K.; Xu, G. Q.; Chen, W. Tuning the metal–support
6770–6772. interaction in catalysts for highly efficient methane dry
[9] Pechimuthu, N. A.; Pant, K. K.; Dhingra, S. C. Deactivation reforming reaction. Appl. Catal. B: Environ. 2016, 180,
studies over Ni−K/CeO2−Al2O3 catalyst for dry reforming 511–520.
of methane. Ind. Eng. Chem. Res. 2007, 46, 1731–1736. [22] Singh, S.; Zubenko, D.; Rosen, B. A. Influence of LaNiO3
[10] Yang, W. W.; Liu, H. M.; Li, Y. M.; Zhang, J.; Wu, H.; He, shape on its solid-phase crystallization into coke-free reforming
D. H. Properties of yolk–shell structured Ni@SiO2 catalysts. ACS Catal. 2016, 6, 4199–4205.
nanocatalyst and its catalytic performance in carbon dioxide [23] Farmer, J. A.; Campbell, C. T. Ceria maintains smaller metal
reforming of methane to syngas. Catal. Today 2016, 259, catalyst particles by strong metal–support bonding. Science
438–445. 2010, 329, 933–936.
[11] Xie, T.; Shi, L. Y.; Zhang, J. P.; Zhang, D. S. Immobilizing [24] Li, Y.; Shen, W. J. Morphology-dependent nanocatalysts:
Ni nanoparticles to mesoporous silica with size and location Rod-shaped oxides. Chem. Soc. Rev. 2014, 43, 1543–1574.
control via a polyol-assisted route for coking- and sintering- [25] Masias, K. L. S.; Peck, T. C.; Fanson, P. T. Thermally robust
resistant dry reforming of methane. Chem. Commun. 2014, core–shell material for automotive 3-way catalysis having
50, 7250–7253. oxygen storage capacity. RSC Adv. 2015, 5, 48851–48855.
[12] Zhao, X. Y.; Li, H. R.; Zhang, J. P.; Shi, L. Y.; Zhang, D. S. [26] Bedrane, S.; Descorme, C.; Duprez, D. Investigation of
Design and synthesis of NiCe@m-SiO2 yolk–shell framework the oxygen storage process on ceria- and ceria–zirconia-
catalysts with improved coke- and sintering-resistance in dry supported catalysts. Catal. Today 2002, 75, 401–405.
reforming of methane. Int. J. Hydrogen Energy 2016, 41, [27] Cai, W. J.; Wang, F. G.; Daniel, C.; van Veen, A. C.;
2447–2456. Schuurman, Y.; Descorme, C.; Provendier, H.; Shen, W. J.;
[13] Du, X. J.; Zhang, D. S.; Shi, L. Y.; Gao, R. H.; Zhang, J. P. Mirodatos, C. Oxidative steam reforming of ethanol over
Coke- and sintering-resistant monolithic catalysts derived Ir/CeO2 catalysts: A structure sensitivity analysis. J. Catal.
from in situ supported hydrotalcite-like films on Al wires for 2012, 286, 137–152.
dry reforming of methane. Nanoscale 2013, 5, 2659–2663. [28] Postole, G.; Nguyen, T.-S.; Aouine, M.; Gélin, P.; Cardenas,
[14] Theofanidis, S. A.; Galvita, V. V.; Poelman, H.; Marin, G. L.; Piccolo, L. Efficient hydrogen production from methane
B. Enhanced carbon-resistant dry reforming Fe-Ni catalyst: over iridium-doped ceria catalysts synthesized by solution
Role of Fe. ACS Catal. 2015, 5, 3028–3039. combustion. Appl. Catal. B: Environ. 2015, 166–167, 580–591.
[15] Bobrova, L. N.; Bobin, A. S.; Mezentseva, N. V.; Sadykov, [29] Matei-Rutkovska, F.; Postole, G.; Rotaru, C. G.; Florea, M.;
V. A.; Thybaut, J. W.; Marin, G. B. Kinetic assessment of Pârvulescu, V. I.; Gelin, P. Synthesis of ceria nanopowders
dry reforming of methane on Pt + Ni containing composite by microwave-assisted hydrothermal method for dry
of fluorite-like structure. Appl. Catal. B: Environ. 2016, 182, reforming of methane. Int. J. Hydrogen Energy 2016, 41,
513–524. 2512–2525.
[16] Mark, M. F.; Maier, W. F. CO2-reforming of methane on [30] Laosiripojana, N.; Assabumrungrat, S. Catalytic dry reforming
supported Rh and Ir catalysts. J. Catal. 1996, 164, 122–130. of methane over high surface area ceria. Appl. Catal. B:
[17] Souza, M. M. V. M.; Aranda, D. A. G.; Schmal, M. Reforming Environ. 2005, 60, 107–116.
of methane withe carbon dioxide over Pt/ZrO2/Al2O3 catalysts. [31] Wang, F. G.; Cai, W. J.; Tana; Provendier, H.; Schuurman,
J. Catal. 2001, 204, 498–511. Y.; Descorme, C.; Mirodatos, C.; Shen, W. J. Ageing analysis
[18] Ashcroft, A. T.; Cheetham, A. K.; Green, M. L. H.; Vernon, of a model Ir/CeO2 catalyst in ethanol steam reforming.
P. D. F. Partial oxidation of methane to synthesis gas using Appl. Catal. B: Environ. 2012, 125, 546–555.
carbon dioxide. Nature 1991, 352, 225–226. [32] Odedairo, T.; Chen, J. L.; Zhu, Z. H. Metal-support

| www.editorialmanager.com/nare/default.asp
Nano Res. 17

interface of a novel Ni-CeO2 catalyst for dry reforming of Energy 2011, 36, 13566–13574.
methane. Catal. Commun. 2013, 31, 25–31. [46] Larachi, F.; Pierre, J.; Adnot, A.; Bernis, A. Ce 3d XPS study
[33] Du, X. J.; Zhang, D. S.; Shi, L. Y.; Gao, R. H.; Zhang, J. P. of composite CexMn1−xO2−y wet oxidation catalysts. Appl.
Morphology dependence of catalytic properties of Ni/CeO2 Surf. Sci. 2002, 195, 236–250.
nanostructures for carbon dioxide reforming of methane. J. [47] Jia, T. K.; Wang, W. M.; Long, F.; Fu, Z. Y.; Wang, H.;
Phys. Chem. C 2012, 116, 10009–10016. Zhang, Q. J. Fabrication, characterization and photocatalytic
[34] Hou, T. F.; Yu, B.; Zhang, S. Y.; Zhang, J. H.; Wang, D. Z.; activity of La-doped ZnO nanowires. J. Alloy Compd. 2009,
Xu, T. K.; Cui, L.; Cai, W. J. Hydrogen production from 484, 410–415.
propane steam reforming over Ir/Ce0.75Zr0.25O2 catalyst. Appl. [48] Luo, M. F.; Zhong, Y. J.; Zhu, B.; Yuan, X. X.; Zheng, X. M.
Catal. B: Environ. 2015, 168–169, 524–530. Temperature-programmed desorption study of NO and CO2
[35] Wisniewski, M.; Boréave, A.; Gélin, P. Catalytic CO2 over CeO2 and ZrO2. Appl. Surf. Sci. 1997, 115, 185–189.
reforming of methane over Ir/Ce0.9Gd0.1O2−x. Catal. Commun. [49] Appel, L. G.; Eon, J. G.; Schmal, M. The CO2–CeO2
2005, 6, 596–600. interaction and its role in the CeO2 reactivity. Catal. Lett.
[36] Petallidou, K. C.; Efstathiou, A. M. Low-temperature water- 1998, 56, 199–202.
gas shift on Pt/Ce1−xLaxO2−δ: Effect of Ce/La ratio. Appl. [50] Aneggi, E.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A.
Catal. B: Environ. 2013, 140–141, 333–347. Promotion effect of surface lanthanum in soot oxidation over
[37] Wang, F. G.; Xu, L. L.; Yang, J.; Zhang, J.; Zhang, L. Z.; Li, ceria-based catalysts. Top Catal. 2007, 42, 319–322.
H.; Zhao, Y.; Li, H. X.; Wu, K.; Xu, G. Q. et al. Enhanced [51] Li, K. Z.; Wang, H.; Wei, Y. G. Syngas generation from
catalytic performance of Ir catalysts supported on ceria-based methane using a chemical-looping concept: A review of
solid solutions for methane dry reforming reaction. Catal. oxygen carriers. J. Chem. 2013, 2013, 294817.
Today, in press, DOI: 10.1016/j.cattod.2016.03.055. [52] Otsuka, K.; Wang, Y.; Sunada, E.; Yamanaka, I. Direct partial
[38] Mears, D. E. Tests for transport limitations in experimental oxidation of methane to synthesis gas by cerium oxide. J.
catalytic reactors. Ind. Eng. Chem. Process Des. Develop. Catal. 1998, 175, 152–160.
1971, 10, 541–547. [53] Wei, J. M.; Iglesia, E. Isotopic and kinetic assessment of
[39] Oyama, S. T.; Zhang, X. M.; Lu, J. Q.; Gu, Y. F.; Fujitani, the mechanism of methane reforming and decomposition
T. Epoxidation of propylene with H2 and O2 in the explosive reactions on supported iridium catalysts. Phys. Chem. Chem.
regime in a packed-bed catalytic membrane reactor. J. Catal. Phys. 2004, 6, 3754–3759.
2008, 257, 1–4. [54] Wei, J. M.; Iglesia, E. Mechanism and site requirements for
[40] He, H.; Dai, H. X.; Au, C. T. Defective structure, oxygen activation and chemical conversion of methane on supported
mobility, oxygen storage capacity, and redox properties of Pt clusters and turnover rate comparisons among noble
RE-based (RE = Ce, Pr) solid solutions. Catal. Today 2004, metals. J. Phys. Chem. B 2004, 108, 4094–4103.
90, 245–254. [55] Pakhare, D.; Schwartz, V.; Abdelsayed, V.; Haynes, D.;
[41] Huang, M.; Fabris, S. Role of surface peroxo and superoxo Shekhawat, D.; Poston, J.; Spivey, J. Kinetic and mechanistic
species in the low-temperature oxygen buffering of ceria: study of dry (CO2) reforming of methane over Rh-substituted
Density functional theory calculations. Phys. Rev. B 2007, La2Zr2O7 pyrochlores. J. Catal. 2014, 316, 78–92.
75, 081404. [56] Múnera, J. F.; Cornaglia, L. M.; Cesar, D. V.; Schmal, M.;
[42] Guzman, J.; Carrettin, S.; Corma, A. Spectroscopic evidence Lombardo, E. A. Kinetic studies of the dry reforming of
for the supply of reactive oxygen during CO oxidation methane over the Rh/La2O3−SiO2 catalyst. Ind. Eng. Chem.
catalyzed by gold supported on nanocrystalline CeO2. J. Am. Res. 2007, 46, 7543–7549.
Chem. Soc. 2005, 127, 3286–3287. [57] Solymosi, F.; Kutsán, G.; Erdöhelyi, A. Catalytic reaction
[43] Trovarelli, A. Catalytic properties of ceria and CeO2- of CH4 with CO2 over alumina-supported Pt metals. Catal.
containing materials. Catal. Rev. 1996, 38, 439–520. Lett. 1991, 11, 149–156.
[44] Reddy, B. M.; Thrimurthulu, G.; Katta, L.; Yamada, Y.; [58] Xu, L. L.; Zhang, J.; Wang, F. G.; Yuan, K. D.; Wang, L. J.;
Park, S. E. Structural characteristics and catalytic activity of Wu, K.; Xu, G. Q.; Chen, W. One-step synthesis of ordered
nanocrystalline ceria-praseodymia solid solutions. J. Phys. mesoporous CoAl2O4 spinel-based metal oxides for CO2
Chem. C 2009, 113, 15882–15890. reforming of CH4. RSC Adv. 2015, 5, 48256–48268.
[45] Wang, F. G.; Cai, W. J.; Provendier, H.; Schuurman, Y.; [59] Wang, R.; Liu, X. B.; Chen, Y. X.; Li, W. Z.; Xu, H. Y.
Descorme, C.; Mirodatos, C.; Shen, W. J. Hydrogen pro- Effect of metal-support interaction on coking resistance of
duction from ethanol steam reforming over Ir/CeO2 catalysts: Rh-based catalysts in CH4/CO2 reforming. Chin. J. Catal.
Enhanced stability by PrOx promotion. Int. J. Hydrogen 2007, 28, 865–869.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

You might also like