ArticlePDF Available

Effect of Ce/Zr Composition on Structure and Properties of Ce1−xZrxO2 Oxides and Related Ni/Ce1−xZrxO2 Catalysts for CO2 Methanation

Authors:

Abstract and Figures

Ce1−xZrxO2 oxides (x = 0.1, 0.25, 0.5) prepared via the Pechini route were investigated using XRD analysis, N2 physisorption, TEM, and TPR in combination with density functional theory calculations. The Ni/Ce1−xZrxO2 catalysts were characterized via XRD analysis, SEM-EDX, TEM-EDX, and CO chemisorption and tested in carbon dioxide methanation. The obtained Ce1−xZrxO2 materials were single-phase solid solutions. The increase in Zr content intensified crystal structure strains and favored the reducibility of the Ce1−xZrxO2 oxides but strongly affected their microstructure. The catalytic activity of the Ni/Ce1−xZrxO2 catalysts was found to depend on the composition of the Ce1−xZrxO2 supports. The detected negative effect of Zr content on the catalytic activity was attributed to the decrease in the dispersion of the Ni0 nanoparticles and the length of metal–support contacts due to the worsening microstructure of Ce1−xZrxO2 oxides. The improvement of the redox properties of the Ce1−xZrxO2 oxide supports through cation modification can be negated by changes in their microstructure and textural characteristics.
The H 2 -TPR profiles of the Ce 1−x Zr x O 2 samples. 3.2.2. Ni/Ce 1−x Zr x O 2 Catalysts Figure 4 shows XRD patterns of the Ni/Ce 1−x Zr x O 2 catalysts. The broad peaks from the oxide NiO phase (PDF#00-047-1049) are detected in the XRD patterns of the as-prepared catalysts, while the narrow peaks from the metallic Ni 0 phase (PDF #00-004-0085) are observed in the XRD patterns of the catalysts aged under reductive conditions of CO 2 methanation. Table 4 summarizes the average size characteristics of nickel species in the as-prepared and used Ni/Ce 1−x Zr x O 2 catalysts according to the XRD and CO chemisorption data. The Ni/Ce 0.9 Zr 0.1 O 2 catalyst is characterized by a higher dispersion of initial NiO crystallites as well as Ni 0 crystallites formed upon reduction. Reaction conditions provoke the sintering of nickel species. The average size of Ni 0 crystallites in the aged catalysts is larger than the size of NiO crystallites in the as-prepared catalysts. The Ni/Ce 0.9 Zr 0.1 O 2 catalyst is characterized by a higher resistance of Ni 0 crystallites to sintering. The average size of Ni 0 crystallites in the Ni/Ce 0.9 Zr 0.1 O 2 catalyst is about two times smaller than in other catalysts. CO chemisorption results (Table 4) confirmed that the Ni/Ce 0.9 Zr 0.1 O 2 catalyst contains Ni 0 particles of the highest dispersion. The determined sizes of Ni 0 nanoparticles in the aged catalysts are in the increasing order of Ni/Ce 0.9 Zr 0.1 O 2 < Ni/Ce 0.75 Zr 0.25 O 2 < Ni/Ce 0.5 Zr 0.5 O 2 . The chemisorption method is considerably more sensitive to ultrafine particles compared to XRD analysis. The observed differences in sizes measured via the chemisorption and XRD techniques suggest that highly dispersed particles of metallic Ni 0 , undetectable via XRD, are present in the catalysts. The comparison of values of Ni 0 content determined from the Rietveld refinement of XRD data and XRF analysis (Supplementary material, Table S1) confirmed that a part of the loaded nickel in the catalysts is not detected via XRD analysis. The fraction of XRDundetectable nickel species in the catalysts decreases in the sequence: Ni/Ce 0.9 Zr 0.1 O 2 > Ni/Ce 0.75 Zr 0.25 O 2 > Ni/Ce 0.5 Zr 0.5 O 2 .
… 
Content may be subject to copyright.
Citation: Pakharukova, V.P.;
Potemkin, D.I.; Rogozhnikov, V.N.;
Stonkus, O.A.; Gorlova, A.M.;
Nikitina, N.A.; Suprun, E.A.; Brayko,
A.S.; Rogov, V.A.; Snytnikov, P.V.
Effect of Ce/Zr Composition on
Structure and Properties of
Ce1xZrxO2Oxides and Related
Ni/Ce1xZrxO2Catalysts for CO2
Methanation. Nanomaterials 2022,12,
3207. https://doi.org/10.3390/
nano12183207
Academic Editor: Giorgio Vilardi
Received: 25 August 2022
Accepted: 10 September 2022
Published: 15 September 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
nanomaterials
Article
Effect of Ce/Zr Composition on Structure and Properties of
Ce1xZrxO2Oxides and Related Ni/Ce1xZrxO2Catalysts for
CO2Methanation
Vera P. Pakharukova 1,* , Dmitriy I. Potemkin 1, Vladimir N. Rogozhnikov 1, Olga A. Stonkus 1,
Anna M. Gorlova 1,2 , Nadezhda A. Nikitina 1,3, Evgeniy A. Suprun 1, Andrey S. Brayko 1, Vladimir A. Rogov 1
and Pavel V. Snytnikov 1
1Boreskov Institute of Catalysis SB RAS, Pr. Lavrentieva 5, 630090 Novosibirsk, Russia
2
Department of Natural Sciences, Novosibirsk State University, Pirogova Street 2, 630090 Novosibirsk, Russia
3Department of Chemistry, Moscow State University, Leninskie Gory St., 1, 119991 Moscow, Russia
*Correspondence: verapakh@catalysis.ru; Tel.: +7-383-326-9597; Fax: +7-383-330-8056
Abstract:
Ce
1x
Zr
x
O
2
oxides (x = 0.1, 0.25, 0.5) prepared via the Pechini route were investigated
using XRD analysis, N
2
physisorption, TEM, and TPR in combination with density functional theory
calculations. The Ni/Ce
1x
Zr
x
O
2
catalysts were characterized via XRD analysis, SEM-EDX, TEM-
EDX, and CO chemisorption and tested in carbon dioxide methanation. The obtained Ce
1x
Zr
x
O
2
materials were single-phase solid solutions. The increase in Zr content intensified crystal structure
strains and favored the reducibility of the Ce
1x
Zr
x
O
2
oxides but strongly affected their microstruc-
ture. The catalytic activity of the Ni/Ce
1x
Zr
x
O
2
catalysts was found to depend on the composition
of the Ce
1x
Zr
x
O
2
supports. The detected negative effect of Zr content on the catalytic activity was
attributed to the decrease in the dispersion of the Ni
0
nanoparticles and the length of metal–support
contacts due to the worsening microstructure of Ce
1x
Zr
x
O
2
oxides. The improvement of the redox
properties of the Ce
1x
Zr
x
O
2
oxide supports through cation modification can be negated by changes
in their microstructure and textural characteristics.
Keywords:
Ni/Ce
1x
Zr
x
O
2
catalysts; mixed oxides; structure; microstructure; methanation;
carbon dioxide
1. Introduction
The CO
2
methanation process has received great interest during the last few years as
a promising method of the production of synthetic natural gas as a hydrogen storage ap-
proach [
1
3
]. Nickel-based catalysts exhibit high activity and selectivity in the methanation
of carbon oxides and are less expensive than systems containing noble metals [
3
5
]. Because
of this, nickel catalysts are widely investigated. The impact of a support on the dispersion
of nickel particles and catalytic performance of nickel catalysts is significant [
6
8
]. CeO
2
was shown to be one of the most effective support materials [
6
16
]. Ceria works as both
a support for nickel particles and a reaction promoter. The promoting effect is related to
the easiness and reversibility of Ce4+–Ce3+ transition associated with appearing or healing
oxygen vacancies [
17
]. Oxygen vacancies on the support surface were reported to participate
in the activation of CO2molecules [14,1822].
The modification of ceria by doping with foreign cations is a well-known strategy to
improve its redox properties [
17
]. Nickel catalysts supported on mixed Ce
1x
Zr
x
O
2
oxides
are promising systems in terms of catalytic activity and stability [
12
,
23
26
]. Zirconium is
recognized as a promoter of the reducibility and oxygen storage capacity of ceria [
27
29
].
Our previous study [
26
] showed that the formation of oxygen vacancies on the support
surface is essential for activity of Ni/Ce
1x
Zr
x
O
2
catalysts in the methanation of carbon
oxides. The reducibility of Ce
1x
Zr
x
O
2
oxides was experimentally shown to increase with
Nanomaterials 2022,12, 3207. https://doi.org/10.3390/nano12183207 https://www.mdpi.com/journal/nanomaterials
Nanomaterials 2022,12, 3207 2 of 16
the Zr content [
29
,
30
]. The regulation of the support redox properties through the variation
of Zr content is of great interest. However, there have been few studies on the effect of the
Ce/Zr ratio on the performance of Ni/Ce
1x
Zr
x
O
2
catalysts in CO
2
methanation. Ocampo
et al. [
23
] studied CO
2
methanation over 5 wt.% Ni/Ce
1x
Zr
x
O
2
catalysts differing in
Zr content (x = 0.28, 0.5, 0.86). The 5 wt.% Ni/Ce
0.5
Zr
0.5
O
2
catalyst showed the highest
activity. Nie et al. [
31
] studied NiO-CeO
2
-ZrO
2
mixed oxides containing 40 wt.% Ni.
The catalyst with a Ce/Zr molar ratio of 9:1 (Ce
0.9
Zr
0.1
O
2
) exhibited the best catalytic
properties. Atzori et al. [
32
] tested NiO-CeO
2
-ZrO
2
catalysts (30 wt.% Ni—Ce
1x
Zr
x
O
2
) in
CO
2
and CO co-methanation. The activity was the same for the catalysts with Zr content in
x = 0–0.5 and decreased at a higher content. To summarize, a correlation between the
activity of Ni/Ce
1x
Zr
x
O
2
catalysts and the Ce/Zr ratio has not been fully understood.
All the reported studies indicated that nickel catalysts based on highly doped Ce
1x
Zr
x
O
2
oxides (x > 0.5) are less effective. The data concerning systems based on Ce
1x
Zr
x
O
2
oxides
with a lower Zr content (x 0.5) are somewhat controversial.
The present work aims to provide further insight into the relation between the com-
position of the Ce
1x
Zr
x
O
2
support materials and the performance of Ni/Ce
1x
Zr
x
O
2
catalysts in CO
2
methanation. A series of Ce
1x
Zr
x
O
2
supports with different Zr con-
tents (x = 0.1, 0.25, 0.5) were prepared via the Pechini method. The supported catalysts
Ni/Ce
1x
Zr
x
O
2
containing 10 wt.% Ni were synthesized via the impregnation method.
DFT calculations and TPR studies were employed to evaluate the impact of Zr content on
the reducibility of the Ce
1x
Zr
x
O
2
oxides. A wide range of physical methods was used
to reveal the structure features of the Ce
1x
Zr
x
O
2
support materials and Ni/Ce
1x
Zr
x
O
2
catalysts. The catalytic properties of the Ni/Ce
1x
Zr
x
O
2
samples were analyzed with
regard to the structural features of both the support materials and catalysts.
2. Experimental Section
2.1. Computational Methods and Details
The DFT+U calculations were performed using the Vienna Ab initio Simulation Pack-
age (VASP) program [
33
]. The generalized gradient approximation (GGA) PBE96 functional
was employed [
34
]. The core electrons were described by projector augmented-wave (PAW)
potentials [
35
,
36
], and the valence electrons were described by a plane-wave basis set. The
DFT+U method (U
eff
= 5 eV) [
37
] was used to correct the strong Coulomb repulsion of
cerium and DFT-D3 [
38
] to take into account dispersion corrections. The cutoff energy was
400 eV. A 2
×
2
×
1 Monkhorst–Pack k-point set was used for the Brillouin-zone integration.
The face-centered cubic unit cell of a fluorite-type structure (space group: Fm
3
m) was
used as the initial geometry in the calculations [
39
,
40
]. Ce
1x
Zr
x
O
2
supercells (x = 0.25, 0.50,
0.75) were built by replacing Ce atoms with Zr ones (Supplementary Materials, Figure S1).
Ce
1x
Zr
x
O
2
(100) and (111) surfaces were modeled as p (2
×
2) and p (3
×
3) slab cells.
The (100) and (111) Ce
1x
Zr
x
O
2
supercells contained four and three layers, respectively. A
vacuum space of 15
Å was set between neighboring slabs to keep the spurious interaction.
The energy of oxygen vacancy formation (Ef) was computed as:
Ef= E(Ce1xZrxO2δ) + 1/2E(O2)E(Ce1xZrxO2),
where E(Ce
1x
Zr
x
O
2δ
) and E(Ce
1x
Zr
x
O
2
) are the energies for surfaces with and without
oxygen vacancy, and E(O2) is the energy of gas-phase O2.
2.2. Samples Preparation
2.2.1. Synthesis of Ce1xZrxO2Mixed Oxides
A series of Ce
1x
Zr
x
O
2
(x = 0.1, 0.25, 0.5) mixed oxides was prepared via the Pechini
method [
41
]. The Ce(NO
3
)
3
*6H
2
O (99.4%) and ZrO(NO
3
)
2
*8H
2
O (99.98%) salts were used
as precursors. Aqueous solutions of the salts with required Ce:Zr molar ratios (9:1; 3:1;
1:1) were prepared. Citric acid (CA) was added to the aqueous solutions at 80
C and
vigorously stirred for 30 min. The CA:metal molar ratio was 1:1. Subsequently, ethylene
Nanomaterials 2022,12, 3207 3 of 16
glycol C
2
H
4
(OH)
2
(EG) was added. The CA:EG molar ratio was 3:2. Next, the solution was
heated at 100
C to promote the polyesterification reaction and water evaporation with the
formation of a polymeric resin. The obtained solids were mechanically milled to powders
and calcined at 450 C for 8 h.
2.2.2. Synthesis of Ni/Ce1xZrxO2Catalysts
The catalysts containing 10 wt.% Ni were prepared by the impregnation of the obtained
Ce
1x
Zr
x
O
2
oxide materials. Metal precursor Ni(CH
3
COO)
2·
4H
2
O (99.0%) and ethylene
glycol (99.5%) were dissolved in the distilled water under stirring at 70
C for 20 min. Then,
the support material was added into the solution. The EG:Ni molar ratio was set to 5.3. The
suspension of the support and impregnating solution was stirred for 2 h at 70
C and dried
at 120
C for 12 h in air. The dried samples were heated at a rate of 2
C/min and calcined
in air at 400
C for 2 h. The catalysts were designated as Ni/Ce
1x
Zr
x
O
2
, with x indicating
Zr content.
2.3. Samples Characterization
2.3.1. XRF Analysis
Elemental compositions of the catalysts were determined via X-ray fluorescent spec-
troscopy (XRF) using an ARL-Advant’x device (Thermo Fisher, Vienna, Austria). Mea-
surements were carried out in a helium atmosphere using the Rh X-Ray tube. UniQuant
software was used to calculate the element percentages.
2.3.2. BET Surface Area Analysis
The BET specific surface areas (S
BET
, m
2
/g) of the Ce
1x
Zr
x
O
2
oxides were determined
by N
2
-physisorption at
196
C. The experiments were carried out using an ASAP 2400
instrument (Micrometrics, Norcross, GA, USA).
2.3.3. XRD Analysis
X-ray diffraction (XRD) measurements were carried out using a D8 Advance diffrac-
tometer (Bruker, Germany) equipped with a Lynxeye linear detector. The measurements
were carried out using the Cu K
α
radiation (
λ
= 1.5418 Å) in the 2
θ
range of 10–83
with
a step of 0.05
. XRD phase analysis was performed using the ICDD PDF-4+ database.
The evaluation of substructure parameters of the Ce1xZrxO2oxides was performed. The
separation of the crystallite size (D
XRD
) and microstrain (
d/d) effects to the line broad-
ening was performed by means of Williamson–Hall plots [
42
]. From the D
XRD
values,
the crystallite surface area (S
XRD
) values were calculated assuming the crystallites were
quasi-spherical:
SXRD =
6000
ρDXRD
where ρis the theoretical density of the material (g/cm3).
To estimate the degree of agglomeration of the Ce
1x
Zr
x
O
2
crystallites, an agglomera-
tion coefficient (ξ) was calculated on the basis of the ratio of SBET and SXRD values:
ξ=1SBET
SXRD
The average crystallite size of nickel-containing phases was estimated by the line
broadening analysis according to the Scherrer equation [
43
]. The Rietveld refinement was
carried out using the software package Topas v.4.2 (Bruker-AXS, Karlsruhe, Germany).
2.3.4. CO Pulse Chemisorption
The average sizes of metallic Ni
0
nanoparticles in catalysts after catalytic experiments
were determined using the pulse chemisorption of CO on the assumption that each sur-
face Ni atom adsorbed one CO molecule. The measurements were carried out using a
Nanomaterials 2022,12, 3207 4 of 16
Chemosorb analyzer (Modern Laboratory Equipment, Novosibirsk, Russia). An amount of
50 mg of each sample was placed inside a U-shape quartz reactor and reduced at 350
C in
H
2
flow (100 mL/min) for 30 min. The treated sample was subsequently cooled down to
room temperature, followed by Ar purge. After that, pulses of CO were fed to the reactor
(100
µ
L) until the amount of CO in the outlet stopped changing according to the thermal
conductivity detector. The amount of chemisorbed CO was estimated. CO adsorption over
the pure supports was negligible.
2.3.5. TEM-EDX
Transmission electron microscopy (TEM) studies were carried out using a JEM-2200FS
(JEOL, Tokyo, Japan) and a Themis Z electron microscope (Thermo Fisher Scientific, Eind-
hoven, The Netherlands) operated at 200 kV. Images in Scanning-TEM (STEM) mode were
acquired using a high-angle annular dark field (HAADF) detector. The local elemental
composition of the samples was studied using a Thermo Fisher Scientific Super-X EDX
spectrometer. The samples were ground, suspended in ethanol, and placed on a copper
grid coated with a holey carbon film.
2.3.6. SEM-EDX
Scanning electron microscopy–energy-dispersive X-ray analysis (SEM-EDX) studies
were carried out using a dual-beam scanning electron microscope, Tescan Solaris FE/SEM
(Tescan, Brno, Czech Republic). The experiments were performed in secondary electron
mode at an accelerating voltage of 20 kV. The microscope is equipped with an AztecLive
EDX spectrometer (Oxford Instruments, High Wycombe, UK) with a Silicon Drift Detector
and energy resolution of 128 eV. The cross-sections of catalyst granules in epoxy resin were
prepared for the examination. The cross-sections with a given flatness of 0.25
µ
m were
covered with a conductive carbon layer of 10–20 nm thickness.
2.3.7. H2-TPR
The temperature-programmed reduction (TPR) via hydrogen was performed with
40–60 mg of sample in a quartz reactor using a flow setup with a thermal conductivity
detector. The gas mixture containing 10 vol.% of H
2
in Ar was fed at 40 mL/min. The
rate of heating from 25 to 800
C was 10
C/min. The TPR curves were normalized per
sample mass.
2.4. Tests of CO2Methanation Activity
The catalytic tests were performed in a U-shaped tubular continuous-flow reactor
(i.d. = 3 mm) at ambient pressure and the temperature range from 200 to 400
C. The
temperature was controlled using a K-type thermocouple placed in the middle of the
catalyst bed. The feed gas contained 4 vol.% CO
2
, 16 vol.% H
2
, and Ar as balance. The
catalyst load was 150 mg, 0.2–0.5 mm fraction, and the flow rate—75 mL/min. Prior to
the experiment, each catalyst was reduced in 10 vol.% H
2
in Ar flow (50 mL/min) at
400
C for 1 h. The compositions of the inlet and outlet gas mixtures were determined
using a gas chromatograph, KHROMOS-1000 (Khromos, Dzerzhinsk, Russia), equipped
with a thermal conductivity detector (CaA molecular sieves column) and flame ionization
detector (Porapak Q column) with a methanator characterized by sensitivity to CO, CH
4
,
and CO
2
of ~1 ppm. The separation of CO, CH
4
, and CO
2
on the column followed by the
methanation of carbon oxides allowed the flame ionization detector to be used to analyze
their concentration. The equilibrium compositions were calculated using equilibrium
software HSC 7.0. It was assumed that equilibrium mixtures only contained gaseous
substances (CH
4
, CO, CO
2
, H
2
, and H
2
O), i.e., no carbon deposition processes were taken
into account. The catalytic properties of Ni/Ce
1x
Zr
x
O
2
catalysts were also compared with
those of industrial catalyst NIAP-07-05 for the methanation of carbon oxides, which contains
38 wt.% NiO, 12 wt.% Cr
2
O
3
, and 50 wt.% Al
2
O
3
(further denoted as 38Ni
Cr
Al). The
catalyst has a Ni0-specific surface area of 3.3 m2/g in a reduced state [26].
Nanomaterials 2022,12, 3207 5 of 16
3. Results and Discussion
3.1. Calculation Results
According to DFT+U calculations, the parameter of the Ce
1x
Zr
x
O
2
lattice decreases
with an increase in Zr content (Table 1) due to a smaller radius of the Zr
4+
cation (0.84 Å)
compared to the Ce
4+
cation (0.97 Å). Doping CeO
2
with Zr leads to a slight decrease in
Ce–O distances. The calculated Ce–O and Zr–O distances are in the range of 2.314–2.375 Å
and 2.205–2.258 Å, respectively.
Table 1.
Calculated geometric characteristics, energies of oxygen vacancy formation (E
f
), and surface
energies (Esurf) for Ce1xZrxO2.
Ce1xZrxO2
Composition Lattice Parameter (Å) rCe-O (Å) rZr-O (Å) Esurf (J/m2)Ef(eV)
(100) (111) (100) (111)
CeO25.46 2.375 - 1.76 0.72 2.03 2.71
Ce0.75Zr0.25 O25.38 2.347 2.258 1.79 0.74 1.45 2.23
Ce0.5Zr0.5 O25.26 2.326 2.242 1.81 0.75 1.22 1.95
Ce0.25Zr0.75 O25.17 2.314 2.220 1.85 0.79 1.97 2.54
ZrO25.08 - 2.205 1.94 0.89 4.26 5.08
The calculated values of the surface energy are presented in Table 1. The Ce
1x
Zr
x
O
2
(111) surface is more stable than the (100) surface. The CeO
2
(111) surface is known to
be the most stable surface among the low-index surfaces (100), (110), and (111) [
44
,
45
].
Zirconium incorporation slightly increases the surface energy at 0 < x < 0.5, while a sharp
increase in the energy is observed at x > 0.5.
The calculated energies of oxygen vacancy formation in Ce
1x
Zr
x
O
2
are also listed
in Table 1. The E
f
values decrease considerably in the range of compositions 0 < x < 0.5,
while at x > 0.5, the E
f
values increase. The Ce
0.5
Zr
0.5
O
2
surfaces are characterized by the
lowest energies of oxygen formation. It was found that Zr incorporation facilitates the
generation of oxygen vacancy. The obtained result agrees with data of interatomic potential
simulations reported by G. Balducci et al. [
46
,
47
], which showed that Ce
4+
/Ce
3+
reduction
energy is reduced even by small amounts of zirconium incorporated into ceria. Based
on DFT+U calculations, Yang et al. [
48
,
49
] also found that zirconium addition leads to
lowering energy of the oxygen vacancy formation. The structure distortion induced by Zr
cations can be responsible for the decrease in the reduction energy. It was assumed that the
smaller Zr
4+
cations counterbalance steric strains arising at formation Ce
3+
cations, which
are larger than Ce4+ cations [46,49].
3.2. Experimental Results
3.2.1. Ce1xZrxO2Support Materials
The powder XRD patterns of the Ce
1x
Zr
x
O
2
samples are shown in Figure 1. Only
Bragg peaks related to oxide with a fluorite-type cubic crystal structure (S.G. Fm
3
m)
are observed.
Table 2lists data on the structural parameters obtained by Rietveld refinement. The lattice
parameters of all the oxides are lower than the value characteristic of CeO
2
(a = 5.411 Å, PDF#
00-028-0753). The lattice shrinkage indicates the formation of substitutional solid solution
Ce
1x
Zr
x
O
2
. The gradual decrease in the lattice parameter with the increase in Zr content
is observed. The composition of the Ce
1x
Zr
x
O
2
solid solutions was evaluated with use of
Vegard’s rule. The linear dependence of the lattice parameter on the Zr content is provided
elsewhere [
26
]. Estimated values of zirconium content coincide with those set at the synthesis
(Table 2). This implies that obtained materials are single-phase Ce
1x
Zr
x
O
2
solid solutions. As
can be seen from Table 2, the isotropic temperature factors of atoms increase with Zr content.
This suggests the increase in crystal lattice distortion resulted from Zr incorporation.
Nanomaterials 2022,12, 3207 6 of 16
Figure 1.
XRD patterns fitted using the Rietveld refinement method for Ce
0.9
Zr
0.1
O
2
(
a
), Ce
0.75
Zr
0.25
O
2
(b), and Ce0.5Zr0.5O2(c) mixed oxides.
Table 2. Structural parameters of Ce1xZrxO2oxides determined from XRD data.
Sample
Ce1xZrxO2Lattice Parameter (Å) Estimated Zirconium
Content xBMe 1) * BO1) * Rwp ** χ2**
Ce0.9Zr0.1 O25.388 (1) 0.10 0.06 0.17 3.62 0.97
Ce0.75Zr0.25 O25.353 (1) 0.25 0.13 0.29 4.07 1.22
Ce0.5Zr0.5 O25.290 (1) 0.51 0.45 1.20 3.92 1.58
* Isotropic temperature factors were refined on an assumption of Ce
1x
Zr
x
O
2
composition set at the synthesis.
The factors of Ce and Zr atoms were constrained to be equal (BMe). ** Rietveld analysis agreement indices.
Nanomaterials 2022,12, 3207 7 of 16
Microstrain analysis also indicates the intensification of structure deformation with
an increasing Zr concentration in Ce
1x
Zr
x
O
2
oxides (Table 3). The microstrain value
for the Ce
0.5
Zr
0.5
O
2
sample is roughly two times higher than that for the Ce
0.9
Zr
0.1
O
2
sample. A difference in the radii of Ce
4+
and Zr
4+
cations is the main reason for the crystal
lattice distortion. As mentioned above, the structure strains induced by Zr doping can be
responsible for the improved reducibility of the Ce1xZrxO2mixed oxides.
Table 3.
Average crystallite sizes and microstrain values according to XRD data, specific surface
areas calculated from XRD-derived crystallite sizes and determined by BET method, agglomeration
coefficients, and average crystallite sizes according to HRTEM data.
Sample d/d DXRD (nm) SXRD (m2/g) SBET (m2/g) Agglomeration
Coefficient ξdHRTEM (nm)
Ce0.9Zr0.1 O2
(2.6
±
0.2)
×
10
37.0 120 71 0.41 6.4
Ce0.75Zr0.25 O2
(4.5
±
0.2)
×
10
35.5 157 83 0.47 5.5
Ce0.5Zr0.5 O2
(6.9
±
0.2)
×
10
35.0 180 53 0.70 4.1
All the oxides are highly dispersed; the D
XRD
are in the range of 5–7 nm. However,
quite low S
BET
values were obtained. Calculated S
XRD
values significantly exceed S
BET
ones (Table 3). Such a difference between S
XRD
and S
BET
values can be explained by the
agglomeration of primary crystallites. The analysis of estimated agglomeration coefficients
(Table 3) showed that an increase in Zr content in the Ce
1x
Zr
x
O
2
oxides is accompanied
with an increase in the agglomeration of crystallites. TEM data confirmed these results. The
TEM images shown in Figure 2a–c demonstrate the loose agglomeration of crystallites in the
low-doped Ce
0.9
Zr
0.1
O
2
sample and the significantly denser agglomeration of crystallites
in the high-doped Ce
0.5
Zr
0.5
O
2
sample. In the HRTEM images (Figure 2d–f), the individual
Ce
1x
Zr
x
O
2
crystallites are distinguishable. The mean size of Ce
1x
Zr
x
O
2
crystallites
(Table 3, d
HRTEM
) was determined from particle size distribution histograms shown in
insets in Figure 2d–f. There is a clear tendency for the Ce
1x
Zr
x
O
2
crystallite size to
decrease as the Zr content increases in accordance with the XRD results.
Figure 2.
TEM data for Ce
0.9
Zr
0.1
O
2
(
a
,
d
), Ce
0.75
Zr
0.25
O
2
(
b
,
e
), and Ce
0.5
Zr
0.5
O
2
(
c
,
f
) catalysts:
(
a
c
) TEM images and SAED patterns in insets with indication of fluorite phase rings; (
d
f
) HRTEM
images and particle size distribution histograms.
Nanomaterials 2022,12, 3207 8 of 16
The reduction behavior of the Ce
1x
Zr
x
O
2
oxides was studied. The H
2
-TPR profiles
are reported in Figure 3. The H
2
-TPR profile of the low-doped Ce
0.9
Zr
0.1
O
2
oxide exhibits
two broad peaks at 500
C and 820
C. Such a two-peak pattern is characteristic of CeO
2
oxide and reflects the stepwise reduction in the surface and bulk [
50
,
51
]. The TPR profiles
of the Ce
0.75
Zr
0.25
O
2
and Ce
0.5
Zr
0.5
O
2
oxides exhibit asymmetric peaks with significantly
increased intensities. The second high-temperature reduction peak is not observed. These data
indicate that surface and bulk reduction processes occur almost simultaneously. The observed
intensive asymmetric peak reflects the co-reduction in surface and bulk. This suggests the
improvement of the reducibility of the mixed Ce
1x
Zr
x
O
2
oxides with an increase in Zr
content in full agreement with previous TPR studies of Ce1xZrxO2oxides [2830].
Figure 3. The H2-TPR profiles of the Ce1xZrxO2samples.
3.2.2. Ni/Ce1xZrxO2Catalysts
Figure 4shows XRD patterns of the Ni/Ce
1x
Zr
x
O
2
catalysts. The broad peaks
from the oxide NiO phase (PDF#00-047-1049) are detected in the XRD patterns of the
as-prepared catalysts, while the narrow peaks from the metallic Ni
0
phase (PDF #00-004-
0085) are observed in the XRD patterns of the catalysts aged under reductive conditions of
CO2methanation.
Table 4summarizes the average size characteristics of nickel species in the as-prepared
and used Ni/Ce
1x
Zr
x
O
2
catalysts according to the XRD and CO chemisorption data. The
Ni/Ce
0.9
Zr
0.1
O
2
catalyst is characterized by a higher dispersion of initial NiO crystallites as
well as Ni0crystallites formed upon reduction. Reaction conditions provoke the sintering
of nickel species. The average size of Ni
0
crystallites in the aged catalysts is larger than
the size of NiO crystallites in the as-prepared catalysts. The Ni/Ce
0.9
Zr
0.1
O
2
catalyst is
characterized by a higher resistance of Ni
0
crystallites to sintering. The average size of
Ni
0
crystallites in the Ni/Ce
0.9
Zr
0.1
O
2
catalyst is about two times smaller than in other
catalysts. CO chemisorption results (Table 4) confirmed that the Ni/Ce
0.9
Zr
0.1
O
2
catalyst
contains Ni
0
particles of the highest dispersion. The determined sizes of Ni
0
nanoparticles
in the aged catalysts are in the increasing order of Ni/Ce
0.9
Zr
0.1
O
2
< Ni/Ce
0.75
Zr
0.25
O
2
<
Ni/Ce
0.5
Zr
0.5
O
2
. The chemisorption method is considerably more sensitive to ultrafine
particles compared to XRD analysis. The observed differences in sizes measured via the
chemisorption and XRD techniques suggest that highly dispersed particles of metallic Ni
0
,
undetectable via XRD, are present in the catalysts.
The comparison of values of Ni
0
content determined from the Rietveld refinement of
XRD data and XRF analysis (Supplementary material, Table S1) confirmed that a part of
the loaded nickel in the catalysts is not detected via XRD analysis. The fraction of XRD-
undetectable nickel species in the catalysts decreases in the sequence: Ni/Ce
0.9
Zr
0.1
O
2
>
Ni/Ce0.75Zr0.25O2> Ni/Ce0.5Zr0.5O2.
Nanomaterials 2022,12, 3207 9 of 16
Figure 4.
Fragments of XRD patterns fitted using the Rietveld refinement method for Ni/Ce
0.9
Zr
0.1
O
2
(
a
,
d
), Ni/Ce
0.75
Zr
0.25
O
2
(
b
,
e
), and Ni/Ce
0.5
Zr
0.5
O
2
(
c
,
f
) catalysts before (
a
c
) and after (
d
f
) the
catalytic reaction.
Table 4.
Average size characteristics of nickel species in the as-prepared and used Ni/Ce
1x
Zr
x
O
2
catalysts according to the XRD and CO chemisorption data.
Sample As-Prepared Catalysts Aged Catalysts
dNiOXRD (nm) dNi XRD (nm) dNichem (nm) SNichem (m2/gcat )
Ni/Ce0.9Zr0.1 O28.5(5) 20.0(5) 11.2 6
Ni/Ce0.75Zr0.25 O212.0(5) 54.0(5) 23.2 2.9
Ni/Ce0.5Zr0.5 O211.0(5) 53.0(5) 33.7 2
The HAADF STEM images and EDX mapping patterns are presented in Figure 5. The
analysis of EDX data revealed that the spatial distribution of Ce and Zr in the catalysts is
homogeneous. No Ce-rich or Zr-rich areas are observed. This confirms that Ce
1x
Zr
x
O
2
oxides used as supports are single-phase substitutional solid solutions. The analysis of
Ni distribution via EDX allows us to see Ni-rich nanoparticles of 5–10 nm in size in the
catalysts. These particles in all the studied samples were shown to be NiO particles
(Supplementary material Figure S2). Individual NiO nanoparticles in contact with support
particles and ones assembled into large aggregates are observed. The Ni/Ce
1x
Zr
x
O
2
catalysts were found to differ in the amount of non-agglomerated NiO nanoparticles
and the possibility of their fixation as single particles. Thus, a large number of single
NiO particles being in contact with support particles are observed in the images of the
Ni/Ce
0.9
Zr
0.1
O
2
catalyst (Figure 5b). More developed contacts between NiO particles and
Ce
0.9
Zr
0.1
O
2
support are likely responsible for the higher resistance of nickel particles
to sintering under conditions of CO
2
methanation. In the case of the Ni/Ce
0.5
Zr
0.5
O
2
Nanomaterials 2022,12, 3207 10 of 16
catalyst, agglomerated NiO nanoparticles with slight contact with the support are observed
(Figure 5h). The Ni particles undergo reduction and sintering during the reaction, as
revealed by XRD analysis. However, the highly dispersed particles fixed on the support
are retained (Supplementary material Figure S3).
Figure 5.
TEM data for Ni/Ce
0.9
Zr
0.1
O
2
(
a
,
b
), Ni/Ce
0.75
Zr
0.25
O
2
(
c
e
), and Ni/Ce
0.5
Zr
0.5
O
2
(
f
h
)
catalysts: TEM (
c
) and HAADF STEM (
a
,
d
,
f
,
g
) images and corresponding EDX mapping patterns
(
b
,
e
,
h
) showing distribution of Ni (red), Zr (blue), and Ce (green). The maps are presented in
background-corrected intensities.
The observed differences in the dispersion of nickel species in the catalysts seem to
be related to effect of the microstructure of the Ce
1x
Zr
x
O
2
oxides on the formation of
supported nanoparticles. As noted above, the Ce
1x
Zr
x
O
2
support materials differed in
particle organization and the agglomeration of crystallites. The microstructure features
of the catalysts were additionally studied via SEM coupled with EDX analysis. SEM anal-
ysis showed quite different internal organization and porous structure in grains of the
Ni/Ce
0.9
Zr
0.1
O
2
and Ni/Ce
0.5
Zr
0.5
O
2
catalysts. In the Ni/Ce
0.5
Zr
0.5
O
2
catalyst, coarse elon-
gated support particles are aggregated with the formation of large, unevenly distributed
voids (Figure 6a,b). Large NiO particles are visualized in the big voids, while the main
part of the support material is weakly covered with nickel compounds according to EDX
mapping (Figure 6c,f).
Nanomaterials 2022,12, 3207 11 of 16
Figure 6.
SEM images (
a
,
b
) and EDX mapping patterns (
c
f
) for nickel (red), zirconium (orange), and
cerium (blue) of cross-section area of the Ni/Ce0.5Zr0.5O2catalyst.
In the Ni/Ce
0.9
Zr
0.1
O
2
catalyst, aggregates of smaller and thinner support particles
have a more uniform distribution of narrow interparticle voids (Figure 7a,b). EDX map-
ping showed the more homogeneous distribution of nickel over the Ce
0.9
Zr
0.1
O
2
support
(Figure 7c,f) than over the Ce
0.5
Zr
0.5
O
2
support (Figure 6c,f). It appears that a network of
narrow channels provided a more uniform distribution of the nickel compounds over the
support surface in the impregnation step. As a result, the support is effectively covered
with nickel compounds and there is no pronounced gradient in the size of formed NiO
particles. Thus, the SEM-EDX results allowed one to explain the observed differences in
Ni/Ce
1x
Zr
x
O
2
catalysts in the dispersion of nickel species. The microstructure features of
the Ce
1x
Zr
x
O
2
oxides affected the dispersion of nickel compounds as well as their spatial
distribution over the supports. The microstructure of the Ce
0.9
Zr
0.1
O
2
support is more
beneficial for the uniform distribution of the nickel compounds without their segregation.
The formation of larger particles in the Ce
0.5
Zr
0.5
O
2
oxide seems to be caused by the
considerable aggregation of the constituent crystallites (Table 3, Figure 2). On the one
hand, greater aggregation is likely related to the higher dispersion of crystallites. Ultrafine
crystallites assemble to lower the surface energy [
52
,
53
]. On the other hand, differences in
the microstructure can be related with specifics of the formation of Ce
1x
Zr
x
O
2
oxides at
their preparation via the Pechini route. Thus, the intensity of combustion processes when
burning the organic polymer matrix depends on the relative contents of Ce and Zr atoms.
During the preparation of Ce
0.9
Zr
0.1
O
2
oxide, a higher content of easily oxidized Ce cations
intensifies combustion processes. The explosive formation of gaseous products favors the
formation of loose particles with the low agglomeration of crystallites.
Nanomaterials 2022,12, 3207 12 of 16
Figure 7.
SEM images (
a
,
b
) and EDX mapping patterns (
c
f
) for nickel (red), zirconium (green), and
cerium (blue) of cross-section area of the Ni/Ce0.9Zr0.1O2catalyst.
The catalytic performance of the Ni/Ce
1x
Zr
x
O
2
catalysts in CO
2
methanation was
studied. Figure 8shows light-off curves for the CO
2
methanation. For all the catalysts, the
CH
4
concentration increases with temperature, reaches a maximum, and then decreases
coinciding with the equilibrium values. All the Ni/Ce
1x
Zr
x
O
2
catalysts had comparable
or higher activity in comparison with the industrial 38Ni-Cr-Al catalyst containing a
significantly higher amount of nickel (38 wt.% vs. 10 wt.%). In a previous work, we
reported that the synergism of redox properties of the Ni/Ce
1x
Zr
x
O
2
system enhances
the catalytic performance in the methanation of carbon oxides [26].
Figure 8.
Temperature dependences of the outlet concentrations of CH
4
(
a
) and CO (
b
) during CO
2
methanation over the Ni/Ce
1x
Zr
x
O
2
catalysts and 38Ni-Cr-Al catalyst. The dashed lines correspond
to the calculated equilibrium concentrations.
Nanomaterials 2022,12, 3207 13 of 16
The Ni/Ce
0.9
Zr
0.1
O
2
catalyst exhibited the highest catalytic activity. The CO
2
half-
conversion temperature (T
50
) was 244, 266, and 280
C for Ni/Ce
0.9
Zr
0.1
O
2
, Ni/Ce
0.75
Zr
0.25
O
2
,
and Ni/Ce
0.5
Zr
0.5
O
2
, respectively. The observed decrease in activity in the sequence
Ni/Ce
0.9
Zr
0.1
O
2
>> Ni/Ce
0.75
Zr
0.25
O
2
> Ni/Ce
0.5
Zr
0.5
O
2
correlates with a decrease in Ni
0
-
specific surface area: 6 >> 2.9 > 2 m
Ni2
/g
cat
. The revealed higher dispersion of Ni
0
particles
and more developed contacts between them and support surface can explain the higher
catalytic activity of the Ni/Ce
0.9
Zr
0.1
O
2
catalyst. As was shown above, the Ni/Ce
0.9
Zr
0.1
O
2
catalyst contained Ni
0
nanoparticles with d
Nichem
sizes of 11.2 nm, and the least active
Ni/Ce
0.5
Zr
0.5
O
2
catalyst contained Ni
0
nanoparticles with d
Nichem
sizes of 33.7 nm (Table 4).
It was also revealed that the most active Ni/Ce
0.9
Zr
0.1
O
2
catalyst is characterized by more
developed contacts between nickel particles and the support (Figure 5b). It is well accepted
that high Ni
0
dispersion and a developed metal–support interface area are highly important
for the catalytic activity for CO2methanation [6,8,23,54,55].
It was recently suggested that zirconium addition impacts the dispersion of nickel
nanoparticles in Ni/Ce
1x
Zr
x
O
2
catalysts as well as the degree of metal–support interac-
tion [
23
,
31
]. The results of this study imply that the cation composition of the Ce
1x
Zr
x
O
2
oxides can affect their microstructure and texture features and, consequently, the dispersion
of supported nickel nanoparticles. This influence is likely related to the synthesis technique
used. The Ce
1x
Zr
x
O
2
oxides under study were prepared via the Pechini method. A similar
effect was observed by Iglesias et al. [
24
] in the study of Ni/Ce
1x
Zr
x
O
2
catalysts with
supports prepared using the coprecipitation method. It was shown that an increase in Zr
content diminishes the specific surface area of the Ce
1x
Zr
x
O
2
supports with a decrease in
the dispersion of supported Ni0nanoparticles. It was also reported that the increase in Zr
content reduces the specific surface area of the Ce
1x
Zr
x
O
2
oxides prepared via the sol–gel
technique [56] and the citrate complexation route [57].
4. Conclusions
In this study, Ce
1x
Zr
x
O
2
mixed oxides with different compositions were prepared
using the Pechini method and used as the supports for Ni/Ce
1x
Zr
x
O
2
catalysts. It was
demonstrated that an increase in Zr content enhances the distortion of the crystal structure
of Ce
1x
Zr
x
O
2
oxides and leads to improvements in their redox properties. It was also
found that microstructure features of Ce
1x
Zr
x
O
2
oxides change with cation composition.
The effect of the Ce
1x
Zr
x
O
2
composition on the structure and catalytic activity of
the Ni/Ce
1x
Zr
x
O
2
catalysts in CO
2
methanation was investigated. The activity of the
Ni/Ce
1x
Zr
x
O
2
catalysts decreased in the order: Ni/Ce
0.9
Zr
0.1
O
2
>> Ni/Ce
0.75
Zr
0.25
O
2
>
Ni/Ce
0.5
Zr
0.5
O
2
. The drop in the activity correlated with the decrease in the dispersion
of metallic Ni
0
nanoparticles. It was revealed that differences in the microstructural char-
acteristics of the Ce
1x
Zr
x
O
2
supports are responsible for differences in the dispersion of
supported Ni0nanoparticles and the length of the metal–support interface.
It was shown that improving the redox properties of Ce
1x
Zr
x
O
2
oxides, which are
important for catalysis, through cation modification can be counterbalanced by wors-
ening their microstructure characteristics, which determine the dispersion of supported
Ni0nanoparticles.
Supplementary Materials:
The following supporting information can be downloaded at: https://
www.mdpi.com/article/10.3390/nano12183207/s1, Figure S1: The models of fluorite Ce
1x
Zr
x
O
2
unit
cells: (a) x = 0, (b) x = 0.25, (c) x = 0.5, (d) x = 0.75, (e) x = 1; Figure S2: TEM image of as-prepared
Ni/Ce
0.9
Zr
0.1
O
2
catalyst (a), electron diffraction pattern (b); Figure S3: HAADF-STEM images and
corresponding EDX-mapping patterns for aged Ni/Ce
0.9
Zr
0.1
O
2
(a,b) Ni/Ce
0.5
Zr
0.5
O
2
(c,d) catalysts.
Table S1: Quantities of nickel compounds in the used Ni/Ce
1x
Zr
x
O
2
catalysts according to XRD phase
analysis and XRF analysis.
Author Contributions:
V.P.P.: investigation, writing—original draft, project administration, writing—
review and editing, funding acquisition. D.I.P.: investigation, visualization, writing—review and
editing. V.N.R.: methodology, investigation. O.A.S.: investigation, visualization. A.M.G.: investiga-
Nanomaterials 2022,12, 3207 14 of 16
tion. N.A.N.: methodology, investigation. E.A.S.: investigation. A.S.B.: methodology, investigation.
V.A.R.: investigation. P.V.S.: investigation, writing—review and editing. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by the Russian Science Foundation (project 21-73-20075).
Data Availability Statement: Not applicable.
Acknowledgments:
The XRF, HRTEM, and SEM studies were carried out using facilities of the shared
research center “National center of investigation of catalysts” at Boreskov Institute of Catalysis. The
authors also acknowledge resource center “VTAN” (Novosibirsk State University) for the access to
TEM equipment. The quantum–chemical calculations were carried out using computing resources of
the federal collective usage center Complex for Simulation and Data Processing for Mega-science
Facilities at NRC “Kurchatov Institute”.
Conflicts of Interest:
The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in this paper.
References
1.
Zhao, A.; Ying, W.; Zhang, H.; Hongfang, M.; Fang, D. Ni/Al
2
O
3
catalysts for syngas methanation: Effect of Mn promoter. J. Nat.
Gas Chem. 2012,21, 170–177. [CrossRef]
2.
Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation—
From fundamentals to current projects. Fuel 2016,166, 276–296. [CrossRef]
3.
Vannice, M.A. The Catalytic Synthesis of Hydrocarbons from Carbon Monoxide and Hydrogen. Catal. Rev.
1976
,14, 153–191.
[CrossRef]
4.
Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. Bimetallic Ni-Based Catalysts for CO
2
Methanation: A Review.
Nanomaterials 2020,11, 28. [CrossRef] [PubMed]
5. Mills, G.A.; Steffgen, F.W. Catalytic Methanation. Catal. Rev. 1974,8, 159–210. [CrossRef]
6.
Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO
2
catalysts with high CO
2
methanation activity and high
CH4 selectivity at low temperatures. Int. J. Hydrogen Energy 2012,37, 5527–5531. [CrossRef]
7.
Martin, N.M.; Velin, P.; Skoglundh, M.; Bauer, M.; Carlsson, P.-A. Catalytic hydrogenation of CO
2
to methane over supported Pd,
Rh and Ni catalysts. Catal. Sci. Technol. 2017,7, 1086–1094. [CrossRef]
8.
Le, T.A.; Kim, M.S.; Lee, S.H.; Kim, T.W.; Park, E.D. CO and CO
2
methanation over supported Ni catalysts. Catal. Today
2017
,
293–294, 89–96. [CrossRef]
9.
Nematollahi, B.; Rezaei, M.; Lay, E.N. Preparation of highly active and stable NiO–CeO
2
nanocatalysts for CO selective
methanation. Int. J. Hydrogen Energy 2015,40, 8539–8547. [CrossRef]
10.
Zhou, G.; Liu, H.; Cui, K.; Jia, A.; Hu, G.; Jiao, Z.; Liu, Y.; Zhang, X. Role of surface Ni and Ce species of Ni/CeO
2
catalyst in CO
2
methanation. Appl. Surf. Sci. 2016,383, 248–252. [CrossRef]
11.
Konishcheva, M.V.; Potemkin, D.I.; Snytnikov, P.V.; Zyryanova, M.M.; Pakharukova, V.P.; Simonov, P.A.; Sobyanin, V.A. Selective
CO methanation in H
2
-rich stream over Ni–, Co– and Fe/CeO
2
: Effect of metal and precursor nature. Int. J. Hydrogen Energy
2015
,
40, 14058–14063. [CrossRef]
12.
Pan, Q.; Peng, J.; Sun, T.; Gao, D.; Wang, S.; Wang, S. CO
2
methanation on Ni/Ce
0.5
Zr
0.5
O
2
catalysts for the production of
synthetic natural gas. Fuel Process. Technol. 2014,123, 166–171. [CrossRef]
13.
Nematollahi, B.; Rezaei, M.; Lay, E.N. Selective methanation of carbon monoxide in hydrogen rich stream over Ni/CeO
2
nanocatalysts. J. Rare Earths 2015,33, 619–628. [CrossRef]
14.
Konishcheva, M.V.; Potemkin, D.I.; Badmaev, S.D.; Snytnikov, P.V.; Paukshtis, E.A.; Sobyanin, V.A.; Parmon, V.N. On the
Mechanism of CO and CO2Methanation Over Ni/CeO2Catalysts. Top. Catal. 2016,59, 1424–1430. [CrossRef]
15.
Konishcheva, M.V.; Potemkin, D.I.; Snytnikov, P.V.; Sobyanin, V.A. The influence of CO, CO
2
and H
2
O on selective CO methanation
over Ni(Cl)/CeO
2
catalyst: On the way to formic acid derived CO-free hydrogen. Int. J. Hydrogen Energy
2019
,44, 9978–9986.
[CrossRef]
16.
Bendieb Aberkane, A.; Yeste, M.P.; Djazi, F.; Cauqui, M.Á. CO Methanation over NiO-CeO
2
Mixed-Oxide Catalysts Prepared by
a Modified Co-Precipitation Method: Effect of the Preparation pH on the Catalytic Performance. Nanomaterials
2022
,12, 2627.
[CrossRef]
17. Trovarelli, A. Catalysis by Ceria and Related Materials; Hutchings, G.J., Ed.; Imperial College Press: London, UK, 2002.
18.
Shanmugam, V.; Neuberg, S.; Zapf, R.; Pennemann, H.; Kolb, G. Effect of Support and Chelating Ligand on the Synthesis of Ni
Catalysts with High Activity and Stability for CO2Methanation. Catalysts 2020,10, 493. [CrossRef]
19.
Martin, N.M.; Hemmingsson, F.; Schaefer, A.; Ek, M.; Merte, L.R.; Hejral, U.; Gustafson, J.; Skoglundh, M.; Dippel, A.-C.;
Gutowski, O.; et al. Structure–function relationship for CO
2
methanation over ceria supported Rh and Ni catalysts under
atmospheric pressure conditions. Catal. Sci. Technol. 2019,9, 1644–1653. [CrossRef]
20.
Jin, T.; Okuhara, T.; Mains, G.J.; White, J.M. Temperature-programmed desorption of carbon monoxide and carbon dioxide from
platinum/ceria: An important role for lattice oxygen in carbon monoxide oxidation. J. Phys. Chem.
1987
,91, 3310–3315. [CrossRef]
Nanomaterials 2022,12, 3207 15 of 16
21.
Konishcheva, M.V.; Potemkin, D.I.; Snytnikov, P.V.; Stonkus, O.A.; Belyaev, V.D.; Sobyanin, V.A. The insights into chlorine doping
effect on performance of ceria supported nickel catalysts for selective CO methanation. Appl. Catal. B Environ.
2018
,221, 413–421.
[CrossRef]
22.
Znak, L.; Stołecki, K.; Zieli´nski, J. The effect of cerium, lanthanum and zirconium on nickel/alumina catalysts for the hydrogena-
tion of carbon oxides. Catal. Today 2005,101, 65–71. [CrossRef]
23.
Ocampo, F.; Louis, B.; Kiwi-Minsker, L.; Roger, A.-C. Effect of Ce/Zr composition and noble metal promotion on nickel based
CexZr1xO2catalysts for carbon dioxide methanation. Appl. Catal. A Gen. 2011,392, 36–44. [CrossRef]
24.
Iglesias, I.; Quindimil, A.; Mariño, F.; De-La-Torre, U.; González-Velasco, J.R. Zr promotion effect in CO
2
methanation over ceria
supported nickel catalysts. Int. J. Hydrogen Energy 2019,44, 1710–1719. [CrossRef]
25.
Ashok, J.; Ang, M.L.; Kawi, S. Enhanced activity of CO
2
methanation over Ni/CeO
2
-ZrO
2
catalysts: Influence of preparation
methods. Catal. Today 2017,281, 304–311. [CrossRef]
26.
Pakharukova, V.P.; Potemkin, D.I.; Stonkus, O.A.; Kharchenko, N.A.; Saraev, A.A.; Gorlova, A.M. Investigation of the Structure
and Interface Features of Ni/Ce
1x
Zr
x
O
2
Catalysts for CO and CO
2
Methanation. J. Phys. Chem. C
2021
,125, 20538–20550.
[CrossRef]
27.
Fornasiero, P.; Dimonte, R.; Rao, G.R.; Kaspar, J.; Meriani, S.; Trovarelli, A.; Graziani, M. Rh-Loaded CeO
2
-ZrO
2
Solid-Solutions
as Highly Efficient Oxygen Exchangers: Dependence of the Reduction Behavior and the Oxygen Storage Capacity on the
Structural-Properties. J. Catal. 1995,151, 168–177. [CrossRef]
28.
Hori, C. Thermal stability of oxygen storage properties in a mixed CeO
2
-ZrO
2
system. Appl. Catal. B Environ.
1998
,16, 105–117.
[CrossRef]
29.
Murota, T.; Hasegawa, T.; Aozasa, S.; Matsui, H.; Motoyama, M. Production method of cerium oxide with high storage capacity
of oxygen and its mechanism. J. Alloys Compd. 1993,193, 298–299. [CrossRef]
30.
Daturi, M.; Finocchio, E.; Binet, C.; Lavalley, J.-C.; Fally, F.; Perrichon, V.; Vidal, H.; Hickey, N.; Kašpar, J. Reduction of High
Surface Area CeO2–ZrO2Mixed Oxides. J. Phys. Chem. B 2000,104, 9186–9194. [CrossRef]
31.
Nie, W.; Zou, X.; Chen, C.; Wang, X.; Ding, W.; Lu, X. Methanation of Carbon Dioxide over Ni–Ce–Zr Oxides Prepared by One-Pot
Hydrolysis of Metal Nitrates with Ammonium Carbonate. Catalysts 2017,7, 104. [CrossRef]
32.
Atzori, L.; Rombi, E.; Meloni, D.; Sini, M.F.; Monaci, R.; Cutrufello, M.G. CO and CO
2
Co-Methanation on Ni/CeO
2
-ZrO
2
Soft-Templated Catalysts. Catalysts 2019,9, 415. [CrossRef]
33. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993,47, 558–561. [CrossRef] [PubMed]
34.
Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett.
1996
,77, 3865–3868.
[CrossRef] [PubMed]
35. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994,50, 17953–17979. [CrossRef]
36.
Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B
1999
,59, 1758–1775.
[CrossRef]
37.
Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability
of nickel oxide: An LSDA+U study. Phys. Rev. B 1998,57, 1505–1509. [CrossRef]
38.
Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,132, 154104. [CrossRef]
39.
Gerward, L.; Staun Olsen, J.; Petit, L.; Vaitheeswaran, G.; Kanchana, V.; Svane, A. Bulk modulus of CeO
2
and PrO
2
—An
experimental and theoretical study. J. Alloys Compd. 2005,400, 56–61. [CrossRef]
40.
French, R.H.; Glass, S.J.; Ohuchi, F.S.; Xu, Y.-N.; Ching, W.Y. Experimental and theoretical determination of the electronic structure
and optical properties of three phases of ZrO2.Phys. Rev. B 1994,49, 5133–5142. [CrossRef]
41.
Pechini, M.P. Method of Preparing Lead and Alkaline Earth Titanates and Niobates and Coating Method Using the Same to form
a Capacitor. U.S. Patent 3330697A, 11 July 1967.
42. Williamson, G.; Hall, W. X-ray line broadening from filed aluminium and wolfram. Acta Metall. 1953,1, 22–31. [CrossRef]
43.
Scherrer, P. Bestimmung der Grosse und der Inneren Struktur von Kolloidteilchen Mittels Rontgenstrahlen, Nachrichten von der
Gesellschaft der Wissenschaften, Gottingen. Math.-Phys. Kl. 1918,2, 98–100.
44.
Nolan, M.; Grigoleit, S.; Sayle, D.C.; Parker, S.C.; Watson, G.W. Density functional theory studies of the structure and electronic
structure of pure and defective low index surfaces of ceria. Surf. Sci. 2005,576, 217–229. [CrossRef]
45.
Yang, Z.; Woo, T.K.; Baudin, M.; Hermansson, K. Atomic and electronic structure of unreduced and reduced CeO
2
surfaces: A
first-principles study. J. Chem. Phys. 2004,120, 7741–7749. [CrossRef] [PubMed]
46.
Balducci, G.; Kašpar, J.; Fornasiero, P.; Graziani, M.; Islam, M.S.; Gale, J.D. Computer Simulation Studies of Bulk Reduction and
Oxygen Migration in CeO2–ZrO2Solid Solutions. J. Phys. Chem. B 1997,101, 1750–1753. [CrossRef]
47.
Balducci, G.; Kašpar, J.; Fornasiero, P.; Graziani, M.; Islam, M.S. Surface and Reduction Energetics of the CeO
2
–ZrO
2
Catalysts.
J. Phys. Chem. B 1998,102, 557–561. [CrossRef]
48.
Yang, Z.; Wei, Y.; Fu, Z.; Lu, Z.; Hermansson, K. Facilitated vacancy formation at Zr-doped ceria(111) surfaces. Surf. Sci.
2008
,
602, 1199–1206. [CrossRef]
49.
Yang, Z.; Fu, Z.; Wei, Y.; Hermansson, K. The electronic and reduction properties of Ce
0.75
Zr
0.25
O
2
(110). Chem. Phys. Lett.
2008
,
450, 286–291. [CrossRef]
Nanomaterials 2022,12, 3207 16 of 16
50.
Giordano, F.; Trovarelli, A.; de Leitenburg, C.; Giona, M. A Model for the Temperature-Programmed Reduction of Low and High
Surface Area Ceria. J. Catal. 2000,193, 273–282. [CrossRef]
51.
Boaro, M.; Vicario, M.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. The use of temperature-programmed and dynamic/transient
methods in catalysis: Characterization of ceria-based, model three-way catalysts. Catal. Today 2003,77, 407–417. [CrossRef]
52.
Kamiya, H.; Gotoh, K.; Shimada, M.; Uchikoshi, T.; Otani, Y.; Fuji, M.; Matsusaka, S.; Matsuyama, T.; Tatami, J.; Higashitani, K.; et al.
Characteristics and behavior of nanoparticles and its dispersion systems. In Nanoparticle Technology Handbook; Elsevier: Amsterdam,
The Netherlands, 2008; pp. 113–176.
53.
Shrestha, S.; Wang, B.; Dutta, P. Nanoparticle processing: Understanding and controlling aggregation. Adv. Colloid Interface Sci.
2020,279, 102162. [CrossRef]
54.
Ocampo, F.; Louis, B.; Roger, A.-C. Methanation of carbon dioxide over nickel-based Ce
0.72
Zr
0.28
O
2
mixed oxide catalysts
prepared by sol–gel method. Appl. Catal. A Gen. 2009,369, 90–96. [CrossRef]
55.
Hao, Z.; Shen, J.; Lin, S.; Han, X.; Chang, X.; Liu, J.; Li, M.; Ma, X. Decoupling the effect of Ni particle size and surface oxygen
deficiencies in CO2methanation over ceria supported Ni. Appl. Catal. B Environ. 2021,286, 119922. [CrossRef]
56.
Dobrosz-Gómez, I.; Kocemba, I.; Rynkowski, J.M. Au/Ce
1x
Zr
x
O
2
as effective catalysts for low-temperature CO oxidation. Appl.
Catal. B Environ. 2008,83, 240–255. [CrossRef]
57.
Kaspar, J.; Fornasiero, P.; Balducci, G.; Di Monte, R.; Hickey, N.; Sergo, V. Effect of ZrO
2
content on textural and structural
properties of CeO
2
–ZrO
2
solid solutions made by citrate complexation route. Inorganica Chim. Acta
2003
,349, 217–226. [CrossRef]
... Ni-based catalysts can be classified into several types: (1) supported Ni-based catalysts on alumina, titanium, etc. [2][3][4][5][6][10][11][12][13][14]; (2) mixed oxides such as perovskites [15], aluminates [16], and Ce-Zr oxide based on the fluorite structure [17]; and (3) high Ni-loaded catalysts [18][19][20]. The latter ones have been extensively studied as catalysts for bio-oil upgrading [9,10,21,22] and methane decomposition [20]. ...
Article
Full-text available
In this paper, structural features of the NiO-SiO2 nanocrystalline catalyst synthesized by the sol-gel method were studied by X-ray diffraction (XRD), high-resolution transmission electron microscopy (TEM), and differential dissolution (DD). The XRD pattern of NiO-SiO2 significantly differs from the “ideal” NiO pattern: the peaks of the NiO-like phase are asymmetric, especially the 111 diffraction peak. The NiO-SiO2 nanocrystalline catalyst was investigated by means of XRD simulations based on two approaches: conventional Rietveld analysis and statistical models of 1D disordered crystals. Through a direct simulation of XRD profiles, structural information is extracted from both the Bragg and diffuse scattering. XRD simulations showed that the asymmetry of all the diffraction peaks is due to the presence of two NiO-like oxides with different lattice constants and different average sizes: ~90 wt% of mixed Ni-Si oxide (Ni:Si = 0.14:0.86) with average crystallite sizes (D ~ 27.5Å) and ~10 wt% of pure NiO (D ~ 50 Å). The high asymmetry of the 111 diffraction peak is due to the appearance of diffuse scattering caused by the inclusion of tetrahedral SiO2 layers between octahedral NiO layers. Such methods as TEM and DD were applied as independent criteria to prove the structural model, and the results obtained confirm the formation of mixed Ni-Si oxide.
Article
Full-text available
In this study, a series of NiO-CeO2 mixed-oxide catalysts have been prepared by a modified co-precipitation method similar to the one used for the synthesis of hydrotalcites. The syntheses were carried out at different pH values (8, 9 and 10), in order to determine the influence of this synthetic variable on the properties of the obtained materials. These materials were characterized by using different techniques, such as TGA, XRD, ICP, N2 adsorption-desorption isotherms, H2 temperature-programmed reduction (H2-TPR), and electron microscopy, including high-angle annular dark-field transmission electron microscopy (HAADF-TEM) and EDS. The characterization results revealed the influence of the preparation method, in general, and of the pH value, in particular, on the textural properties of the oxides, as well as on the dispersion of the Ni species. The catalyst prepared at a higher pH value (pH = 10) was the one that exhibited better behavior in the CO methanation reaction (almost 100% CO conversion at 235 °C), which is attributed to the achievement, under these synthetic conditions, of a combination of properties (metal dispersion, specific surface area, porosity) more suitable for the reaction.
Article
Full-text available
CO2 methanation has recently emerged as a process that targets the reduction in anthropogenic CO2 emissions, via the conversion of CO2 captured from point and mobile sources, as well as H2 produced from renewables into CH4. Ni, among the early transition metals, as well as Ru and Rh, among the noble metals, have been known to be among the most active methanation catalysts, with Ni being favoured due to its low cost and high natural abundance. However, insufficient low-temperature activity, low dispersion and reducibility, as well as nanoparticle sintering are some of the main drawbacks when using Ni-based catalysts. Such problems can be partly overcome via the introduction of a second transition metal (e.g., Fe, Co) or a noble metal (e.g., Ru, Rh, Pt, Pd and Re) in Ni-based catalysts. Through Ni-M alloy formation, or the intricate synergy between two adjacent metallic phases, new high-performing and low-cost methanation catalysts can be obtained. This review summarizes and critically discusses recent progress made in the field of bimetallic Ni-M (M = Fe, Co, Cu, Ru, Rh, Pt, Pd, Re)-based catalyst development for the CO2 methanation reaction.
Article
Full-text available
Carbon dioxide methanation was carried out over Ni-based catalysts on different supports and chelating ligands in microreactors. To investigate the influence of chelating ligands and supports, the Ni catalysts were prepared using different support such as CeO2, Al2O3, SiO2, and SBA-15 by a citric acid (CA)-assisted impregnation method. The properties of the developed catalysts were studied by X-ray diffraction (XRD), Transmission electron microscope (TEM), and X-ray photoelectron spectroscopy (XPS) measurement, and the results show that the addition of CA in the impregnation solution improved the dispersion, refines the particle size, and enhanced the interaction of nickel species. The catalytic performance of the developed Ni catalysts were evaluated by CO2 methanation in microreactors in the temperature range of 275 °C–375 °C under 12.5 bar pressure. All the catalysts exhibit high CO2 conversion and extremely high selectivity to methane. However, the catalysts prepared via CA-assisted method exhibited excellent activity and stability, compared with Ni catalysts prepared by a conventional impregnation method, which could be attributed to highly dispersed nickel particles with strong metal–support interaction. The activity of CO2 methanation followed the order of Ni/CeO2-CA > Ni/SBA-15-CA > Ni/Al2O3-CA > Ni/SiO2-CA > Ni/CeO2. The Ni/CeO2 catalysts have also been prepared using different chelating ligands such as ethylene glycol (EG), sucrose (S), oxalic acid (OA) and ethylene diamine tetra acidic acid (EDTA). Among the tested catalysts prepared with different support and chelating ligands, the Ni/CeO2 catalyst prepared via CA-assisted method gave superior catalytic performance and it could attain 98.6% of CO2 conversion and 99.7% methane selectivity at 325 °C. The partial reduction of the CeO2 support generates more surface oxygen vacancies and results in a high CO2 conversion and methane selectivity compared with other catalysts. The addition of CA as promoter favored the synergistic effect of Ni and support, which led to high dispersion, controls the size, and stabilizes the Ni nanoparticles. Furthermore, the Ni/CeO2-CA catalyst yields high CO2 conversion in a time-on-stream study due to the ability of preventing the carbon deposition and sintering of Ni particles under the applied reaction conditions. However, the Ni/Al2O3-CA and Ni/SBA-15-CA catalysts showed stable performance for 100 h of time on stream.
Article
Full-text available
Supported nickel catalysts were synthesized, characterized, and employed in the carbon oxides co-methanation process. Five NiO/CeO2-ZrO2 mixed oxides, with the same Ni content and different Ce/Zr molar ratios, were prepared by the soft-template method. They were characterized through ICP-AES, N2 adsorption, XRD, and TPR. Reduced Ni/CeO2-ZrO2 catalysts were obtained by submitting the oxide systems to reduction treatment in H2 at 400 °C. They were characterized by XRD, H2-TPD, and CO2 adsorption microcalorimetry and their catalytic performances in the carbon oxides co-methanation were investigated. Catalytic tests were performed in a fixed-bed continuous-flow microreactor at atmospheric pressure. The effect of experimental conditions (reaction temperature, space velocity, reactants molar ratio) was also studied. Almost complete CO conversion was obtained on any catalyst, whereas CO2 conversion was much lower and increased with Ce content, at least up to Ce/Zr = 1. The beneficial effect of the Ce content could be related to the increased NiO reducibility and to the higher ability to adsorb and activate CO2. However, at high Ce/Zr ratios, it is probably counterbalanced by an interplay of reactions involving CO and CO2.
Article
Full-text available
In situ structural and chemical state characterization of Rh/CeO2 and Ni/CeO2 catalysts during at- mospheric pressure CO2 methanation have been performed by a combined array of time-resolved analytical tools including ambient-pressure X-ray photoelectron spectroscopy, high-energy X-ray diffraction and diffuse reflectance infrared Fourier transform spectroscopy. The ceria phase is partially reduced during the CO2 methanation and in particular Ce3+ species seem to facilitate activation of the CO2 molecule. The activated CO2 molecule then reacts with atomic hydrogen provided from H2 dissociation on Rh and Ni sites to form formate species. For the most active catalyst (Rh/CeO2), transmission electron microscopy measurements show that the Rh nanopar- ticles are small (average 4 nm, but with a long tail towards smaller particles) due to a strong interaction between Rh particles and the ceria phase. In contrast, larger nanoparticles were ob- served for the Ni/CeO2 catalyst (average 6 nm, with no crystallites below 5 nm found), suggesting a weaker interaction with the ceria phase. The higher selectivity towards methane of Rh/CeO2 is proposed to be due to the stronger metal-support interaction.
Article
Full-text available
Carbon dioxide methanation is an interesting way to reduce greenhouse effect gases emission and, simultaneously, provide a renewable energy source of methane. Ceria and 15 at.% Zr-doped ceria supported nickel catalysts were characterized by means of various techniques (BET, XRD, Raman, H2-TPR, CO2-TPD, O2-TPO, OSC and H2-chemisorption) and evaluated in carbon dioxide methanation. Zr incorporation into catalyst formulation reduced catalyst's basicity but favored its reducibility, nickel availability and oxygen storage capacity. These characteristics gave rise to an improved catalytic performance both in terms of activity and stability: temperature required to achieve 50% conversion was reduced in 20 °C and low temperature (250 °C) stability was improved in around 8%. Initial rates approach was employed to determine reaction rates and apparent activation energies for CO2 methanation, which resulted in 113 and 121 kJ mol⁻¹for Ni/CeO2 and Ni/Ce0.85Zr0.15O2, respectively.
Article
The study describes decoupling the effect of surface oxygen vacancies and Ni particle size in the CO2 methanation over CeO2 supported Ni nanoparticles. The availability of surface oxygen vacancies with the interaction between Ni and CeO2 served to appreciably enhance the catalytic capacity for activation of CO2, generating higher conversion rates, larger selectivity to CH4 and lower activation energies relative to Ni/SiO2. A structure- sensitivity in the reaction over Ni/CeO2 was established without the interference of different concentrations of surface oxygen vacancies, where the turnover frequencies of CO2 decreased with increasing the Ni nanoparticle sizes (8-21 nm). The TPSR and Operando FT-IR analysis demonstrated that formate was the critical intermediate for CH4 formation. The CO formation with smaller nickel size resulted from the carboxyl species. In addition, the long-term evaluations and TGA measurements revealed small Ni nanoparticles suffered a temporary loss of activity due to the carbon deposition.
Article
Nanoparticles (NPs) are commonly defined as particles with size <100 nm and are currently of considerable technological and academic interest, since they are often the starting materials for nanotechnology. Novel properties develop as a bulk material is reduced to nanodimensions and is reflected in new chemistry, physics and biology. With reduction in size, a greater function of the atoms is at the surface, and promote different interaction with its environment, as compared to the bulk material. In addition, the reduction in size alters the electronic structure of the material, resulting in novel quantum effects. Size also influences mobility, primarily controlled by Brownian motion for NPs, and relevant in biological and environmental processes. However, the small size also leads to high surface energy, and NPs tend to aggregate, thereby lowering the surface energy. In all applications, the uncontrolled aggregation of NPs can have negative effects and needs to be avoided. There are however examples of controlled aggregation of NPs which give rise to novel effects. This review article is focused on the NP features that influences aggregation. Common strategies for synthesis of NPs from the gas and liquid phases are discussed with emphasis on aggregation during and after synthesis. The theory involving Van der Waals attractive force and electrical repulsive force as the controlling features of the stability of NPs is discussed, followed by examples of how repulsive and attractive forces can be manipulated experimentally to control NP aggregation. In some applications, NPs prepared by liquid methods need to be isolated for further applications. The process of solvent removal introduces new forces such as capillary forces that promote aggregation, in many cases, irreversibly. Strategies for controlling aggregation upon drying are discussed. There are also many methods for redispersing aggregated NPs, which involve mechanical forces, as well as manipulating capillary forces and surface characteristics. We conclude this review with a discussion of aggregation relevant real-world applications of NPs. This review should be relevant for scientists and technologists interested in NPs, since emphasis has been on the practical aspects of NP-based technology, and especially, strategies relevant to controlling NP aggregation.
Article
For the first time the influence of CO, CO2 and H2O content on the performance of chlorinated NiCeO2 catalyst in selective or preferential CO methanation was studied systematically. It was shown that the rate of CO methanation over Ni(Cl)/CeO2 increases with the increasing H2 concentration, is independent of CO2 concentration and decreases with increasing CO and H2O concentrations; the rate of CO2 methanation is weakly sensitive to H2 and CO2 concentrations and decreases with increasing CO and H2O concentrations. High catalyst selectivity was attributed to Ni surface blockage by strongly adsorbed CO molecules and ceria surface blockage by Cl, which both inhibit CO2 hydrogenation. For the first time, selective CO methanation over Ni(Cl)/CeO2 was studied for deep CO removal from formic acid derived hydrogen-rich gases characterized by high CO2 (40–50 vol%), low CO (30–1000 ppm) content and trace amounts of water. Composite Ni(Cl)/CeO2-η-Al2O3/FeCrAl wire mesh catalyst was demonstrated to be effective for this process at temperatures of 180–220°С, selectivity 30–70%, WHSV up to 200 L (STP)/(g∙h). The catalyst provides high process productivity, low pressure drop, uniform temperature distribution, and appears highly promising for the development of a compact CO cleanup reactor. Selective CO methanation was concluded to be a convenient way to CO-free hydrogen produced by formic acid decomposition.