You are on page 1of 11

Article

Cite This: J. Phys. Chem. C 2019, 123, 9788−9798 pubs.acs.org/JPCC

Methane Activation at the Metal−Support Interface of Ni4−


CeO2(111) Catalyst: A Theoretical Study
Rajib Kumar Singha, Yuta Tsuji, Muhammad Haris Mahyuddin, and Kazunari Yoshizawa*
Institute for Materials Chemistry and Engineering and IRCCS, Kyushu University, Fukuoka 819-0395, Japan
*
S Supporting Information

ABSTRACT: Methane activation is usually assumed to take place on top of metal surfaces
or metal clusters. It can also occur at the metal−support interface in metal-supported
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

catalysts with reducible oxides, such as CeO2. In the present work, we exploit density
functional theory with an additional Hubbard-like parameter (DFT + U) to calculate the
Downloaded via UNIV OF NEW ENGLAND on October 23, 2019 at 16:35:46 (UTC).

activation of methane at an O site interfacing a Ni4 metal cluster on a support, CeO2(111)


surface. Two reaction routes, namely, radical and nonradical routes, are taken into account.
We show that the nonradical route is favored with an apparent activation energy of 18.1
kcal/mol, which is lower than that for the radical route by 15.0 kcal/mol. In the nonradical
route, the formation of a four-centered transition-state structure is observed while a C−H
bond of methane is being cleaved to form an OH moiety and a CH3 fragment that is being
bound to the interfacial Ni atom. It is also found that the interfacial O atoms are out of the
CeO2 surface plane with Ce−O bond distances being much longer than those in the crystalline bulk CeO2, which allows them
to be easily reduced, and hence, the interfacial O atoms become more reactive toward methane, as compared to the surface O
atoms. The interactions between Ni4 cluster and the CeO2(111) surface result in the reduction of two Ce4+ ions to Ce3+,
improving the reducibility of the interfacial O atoms. This should be an important key to the facile methane activation.

1. INTRODUCTION Al2O3, MgO, and TiO2 supports, it is not the case for reducible
Even after decades of research, the activation and conversion of oxide supports such as CeO2, ZrO2, and CexZr1−xO2 solid
methane to value-added products in mild conditions still solutions because the interface between the metal and the
remain challenging, mainly due to the high symmetric support has been experimentally suggested to be more
geometry and the strong C−H bonds of methane.1,2 At active.21,39 Theoretically, density functional theory (DFT)
present, methane is industrially used in the steam reforming studies have shown a preference of bare metal or bare metal-
process to produce synthesis gas (syngas), which is followed by oxide surfaces as catalyst models,40−44 but these cannot
the Fischer−Tropsch synthesis for the conversion of syngas to represent the real catalytic systems because in the commercial
a number of chemicals.3,4 In contrast to such an expensive, scale the reaction is not performed over bare surfaces. In other
energy-demanding two-step process, the direct and low-cost DFT studies, metal oxides doped by single metal atom are
conversion of methane to methanol, formaldehyde, ethylene, considered as catalyst models,31,41,45−48 but again they are
and ethane, are not yet achieved on the commercial scale,5−7 industrially impractical due to the consideration of an ideal
although progressive efforts in this direction with the use of metal dispersion, that is 100%. Of course, there would be a
metal-exchanged zeolite catalysts have been extensively single atomic species on particular sites but that would be very
reported.8−13 rare. Therefore, DFT calculations that are carried out in more
Recently, metal-supported catalysts have attracted increasing realistic catalyst models are required to find the actual active
interests due to the mutual interactions between a metal and a site for methane activation on metal supported by reducible
support that lead to lower C−H activation energies of metal-oxide catalysts. Only then, the observed results can be
methane.14−23 Researchers have investigated various supports compared with the actual experimental observations and a
such as Al2O3, TiO2, MgO, ZrO2, CeO2, and so forth,24−30 better understanding of the reaction route for methane
finding that the redox property of CeO2 is very helpful in the activation can be obtained for the development of catalysts
activation of methane.28−30 More specifically, the reduction of that are more superior than the existing catalysts.
Ce4+ to Ce3+ improves the reducibility of the surface oxygen of To date, DFT studies on the elucidation of methane
CeO2,31,32 and thus, it helps to activate methane under lower activation over metal-support catalysts have been extensively
temperatures.29,30,33−36 Owing to this fact, CeO2 has been reported. For example, Guo and Wang37 calculated and
extensively studied by both experimental and theoretical reported partial oxidation of methane (POM) over Rh4−TiO2
research groups for methane conversion reactions.14,17,18,20,33
In most catalytic systems of metal supported by metal oxide, Received: December 13, 2018
the metal site is often suggested as the active site for methane Revised: March 8, 2019
activation.16−20,37,38 While this might be true for example Published: March 27, 2019

© 2019 American Chemical Society 9788 DOI: 10.1021/acs.jpcc.8b11973


J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

catalyst, showing that the top-site of the Rh4 cluster is the electron exchange−correlation was treated by the generalized
methane activation site. Similarly, Lustemberg et al.18 also gradient approximation with a Hubbard-like parameter56
suggested that the top-site of the metal cluster is the active site (GGA + U) based on the Perdew−Burke−Ernzerhof func-
for methane activation over Ni4 cluster supported on partially tional.57 DFT has difficulty in representing the nature of 4f
reduced and fully reduced CeO2 catalysts. Liu et al.17 reported electrons of Ce and d-states of metal oxides.48,58−60 U value is
methane activation over Cu4 cluster supported on partially a key factor to obtain theoretical results.61 The U values we
reduced CeO2 catalyst, where the top-site of the Cu4 cluster have used for this manuscript are well documented by Jiang et
was also reported as the methane activation site. Zhang et al.,38 al.62 As suggested by Calderon et al.63 and Jiang et al.,62 the
on the other hand, reported methane activation over Ni4− Ueff values are set to 6.3 and 6.4 eV, respectively, for Ce 4f and
Al2O3 catalyst showing that methane is adsorbed preferably on Ni 3d states. The calculations were performed using plane
the Al2O3 surface, but the activation occurs at the surface of wave basis sets with a cutoff energy of 400 eV. Brillouin zone
Ni4 cluster. Foppa et al.49 also showed that the activation sampling was done by using Monkhorst−Pack scheme with the
energy of methane over a Ni cluster supported by Al2O3 k-points being set to 2 × 2 × 1. The geometrical optimizations
surface is lower than that over the interfacial O site. The four- of the structures were done by using the conjugate gradient
atom Ni cluster is the smallest atom cluster which has a three- method. For the determination of TS, quasi-Newton algorithm
dimensional geometry, and it has the interface and non- was applied along with the climbing-image nudged elastic band
interface (top site) atoms at the same time when deposited on (CI-NEB) method.64,65 The initial reaction path for the TS
the support (CeO2) surface. So, we can compare the determination during methane activation was generated by the
differences in properties of the interfacial Ni-atoms and the image-dependent pair potentials method66 implemented in
non-interfacial Ni atom easily. This small cluster also reduces Virtual NanoLab 2016.67 The convergence of the structural
the computational cost. With a planar geometry, deposition of optimization was determined when the maximum forces on
Ni4 on the CeO2(111) surface will lead to the adsorption of unconstrained atoms were less than 0.05 eV/Å. Bader analysis
every Ni atom on the surface oxygen site. So, the Ni4− algorithm was applied for the atomic charge and spin density
CeO2(111) system will have only the interfacial Ni-atoms. analyses.68,69 The optimized structures were visualized by
Very recently, Lustemberg et al.19 calculated the reaction of VESTA.70
methane-to-methanol conversion over Ni4−CeO2−x catalysts, The Ni4 cluster was randomly placed above the CeO2(111)
where the activation of methane takes place at a flat surface of surface and then optimized. But there is no guarantee that the
the Ni4 cluster. However, such a particular methane activation structure obtained from the initial optimization correspond to
on flat Ni4 cluster surface can lead to some controversy, since the most stable structure. However, it is impossible to test all
the most stable geometry of Ni4 cluster is tetrahedral, not flat/ the possible adsorption geometries due to limited computa-
planar.50 tional resource. Fortunately, we are blessed with many
Here, we performed DFT + U calculations for methane optimization algorithms; and what is the so-called simulated
activation on a CeO2(111)-supported tetrahedral Ni4 cluster, annealing method is thought to be a powerful tool to find the
herein after referred to as Ni4−CeO2(111), where the global minimum structure of complicated systems with high
interfacial O atom is considered as the active site. Between probability.71 In the present study, in order to find the
two known reaction mechanisms, that is radical and nonradical energetically most stable adsorption structure of the Ni4
mechanisms, we found that the latter one is more preferred cluster, we employed the annealing method combined with
with a C−H activation energy being in good agreement with ab initio Born−Oppenheimer molecular dynamics imple-
the reported experimental value. This study encompasses a mented in VASP. Starting at 300 K, the temperature was
comprehensive discussion of possible factors contributing to gradually decreased down to 0 K. By scaling the velocities of
such a highly reactive interfacial O atom toward H-atom the atoms, a continuous decrease of the kinetic energy is
abstraction of methane. The role of interfacial oxygen had been achieved. The interval between velocity scaling events is
long proposed as the active site for C−H bond activation, but simulated in a microcanonical ensemble. For each annealing
to the best of our knowledge there is no theoretical report on simulation, a total of 5000 molecular dynamics steps with a
the activation of methane at the metal−support interfacial O- time step of 1 fs were conducted. We carried out further static
atom. There is no report about the transition state (TS), geometry optimizations at 0 K for the final annealed structure.
including its structure and the energy required for the Such a method has been well tested in a variety of systems.72,73
activation of methane at this site, either. Therefore, our In this work, the adsorption energy of Ni4 cluster on the
study holds a significant novelty and importance for the support CeO2(111) surface and that of methane on the Ni4−
readers from the field of catalysis for methane activation. CeO2(111) surface was, respectively, calculated according to
eqs 1 and 2, as follows
2. METHODOLOGY ads
E Ni = E Ni4 /CeO2 − (ECeO2 + E Ni4) (1)
2.1. Computational Methods. Spin-polarized DFT 4

calculations were carried out by using Vienna ab initio and


simulation package (VASP).51,52 The Grimme’s D3 method53
was chosen to include van der Waals correction because the ads
ECH 4
= ECH4−Ni4 /CeO2 − (E Ni4 /CeO2 + ECH4) (2)
D3 correction is known to accurately reproduce the binding
energies of n-alkanes, including methane, not only on a pure where ENi4/CeO2 is the total energy of Ni4 cluster supported on
metal surface but also on a metal oxide surface, and predict the
CeO2(111) surface, ECeO2 is the total energy of CeO2(111)
activation barriers for alkane C−H bond cleavage on the
surfaces, as reported by Antony et al.54,55 The interactions surface, ENi4 is the total energy of Ni4 cluster, and ECH4 is the
between the ion cores and electrons were described by total energy of methane in the gas phase. A large negative value
employing the projector augmented wave method. The of Eads indicates a strong adsorption energy, and vice versa.
9789 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

Activation energy (Ea) and apparent activation energy (Eaa) of at the interface are within the range of 1.81−1.84 Å, while the
methane were calculated according to the following Ni−Ni bond distance was measured to be about 2.4 Å.
equations.74,75 Apparent activation is the true activation energy Experimental reports suggest that small changes in the CeO2
reduced by adsorption energy of the reactant.74,75 surface drastically change its surface property.23 We have also
ads found that after the deposition of Ni4 cluster on the
Ea = E TS − ECH 4 (3) CeO2(111) surface, the oxidation state of surface Ce atoms
and changes. The Ce atoms indicated with blue and greenish
yellow colors in Figure 2 represent Ce3+ and Ce4+ oxidation
Eaa = E TS − (E Ni4 /CeO2 + ECH4) (4)
where ETS is the total energy of the TS.
2.2. Model for the Ni4−CeO2(111) Catalyst System. A
periodic slab model of (4 × 4) stoichiometric CeO2(111)
surface (i.e., no oxygen defects) was used. The model consists
of three layers of Ce atoms and six layers of O atoms, giving a
total of nine atomic layers with 144 atoms. A unit cell size of
15.57 Å × 15.57 Å × 24.54 Å with a vacuum space of 15 Å was
used to minimize the interactions between the periodically
repeated crystal faces. A nickel cluster consisting of four Ni
atoms was placed on the CeO2(111) surface (see Figure 1).

Figure 2. Change in oxidation state of surface Ce atoms or the surface


property of CeO2(111) after deposition of Ni4 cluster. The Ce atoms
indicated with blue and greenish yellow colors represent Ce3+ and
Ce4+ atoms, respectively, whereas the Ce atoms with pink color
represent the Ce atoms with oxidation state in between +3 and +4.

states, respectively, whereas the Ce atoms with pink color


represent Ce atoms with oxidation state in between +3 and +4.
Figure 1. Side and top views of the structure of Ni4−CeO2(111) It was observed that the Bader charge of Ce atoms of pure bulk
catalyst model. Ce, O, and Ni atoms are indicated by the greenish CeO2 ranges from 2.41 to 2.45 e, which correspond to Ce4+. In
yellow, red, and green balls, respectively. our model structure of Ni4−CeO2(111), we have found that
two of the surface Ce atoms, namely, Ce4 and Ce8 (just below
During calculations, only the molecule, the Ni4 cluster, and the Ni3 in Figure 2), have Bader charges of 2.11 and 2.03 e,
six topmost layers of the slab surface were allowed to fully respectively. Following the description reported by Tang et
relaxed while the remaining three layers were kept fixed to al.78 and Senftle et al.,33 we designated a Ce atom as Ce4+,
their bulk positions. Cen+, or Ce3+ when the Bader charge is 2.41 e or larger, in
Owing to their higher surface-to-volume ratio, metal clusters between 2.11 and 2.41, and 2.11 e or smaller, respectively,
are considered better than bulk surfaces.76 In particular, the where n is in between 3 and 4. One should notice that the
nanoscale and sub-nanoscale metal clusters are more value of n is close to 4. The Bader charge of the surface Ce
interesting than the bulk materials because of their distinct atoms is shown in Table S1 (see the Supporting Information to
optical, magnetic, and catalytic properties.77 The Ni4 cluster is this paper). Experimental reports also stated that the presence
the smallest atomic cluster that can have both interface and of Ce3+ atoms in a catalyst highly increases its methane
non-interface (top site) atoms when it is deposited on the activation property by increasing the surface oxygen reduc-
CeO2 surface. Moreover, it enables us to compare more easily ibility.78−81 Singha et al.82 reported that a Pd/CeO2 catalyst
the differences between interfacial and non-interfacial Ni atom containing about 33% Ce3+ atoms activates methane at 350 °C
properties. This small cluster also reduces the computational during the dry reforming reaction of methane and even
cost. produced syngas at that temperature, which is very low as
compared to the reported temperature for dry reforming of
3. RESULTS AND DISCUSSION methane (∼600 °C). Thus, the change in oxidation state of
Ce4+ to Ce3+ is beneficial for the purpose of methane
3.1. Changes in the Surface Property of CeO2(111)
activation. This redox property (Ce4+/Ce3+) is highly
after Deposition of Ni4 Cluster. After the deposition of a
referenced in various articles for different reactions.32,82−86
highly symmetric, most stable, tetrahedral Ni4 cluster50 on the
Such a redox property was experimentally suggested to enable
CeO2(111) surface, it is found that the center of the Ni4 cluster
methane activation on bare CeO2 surface.41,85 For simplicity in
resides on top of a surface Ce atom, as it can be seen form the
presentation of figures, we do not use the color code of the Ce
top view of the Ni4−CeO2 system (see Figure 1). The
ads
atoms in the subsequent figures in this manuscript.
calculated adsorption energy of Ni4 cluster on the surface (E Ni 4 In addition to the redox property of CeO2, the addition of
= −9.1 kcal/mol) is found to be quite large, showing that the Ni4 cluster on the surface of CeO2 leads to longer Ce−O bond
interaction between the Ni4 cluster and CeO2(111) surface is lengths at the metal−support interface, as compared to that in
strong. The Ni4 atomic cluster consists of three Ni atoms the bulk crystal (see Figure 3A). This figure also shows that the
attached to the CeO2(111) surface and one Ni atom remains Ce43+−O and Ce83+−O lengths are 2.61 and 2.69 Å,
unattached to the CeO2(111) surface. The Ni−O bond lengths respectively, suggesting that the Ce−O bond length is longer
9790 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

Figure 3. (A) Difference in the Ce−O bond length. (B) Charge density difference upon the adsorption of the Ni4 cluster. Violet and orange regions
indicate the increase and decrease in electron charge in the respective regions. (C) Atomic charges of Ni-cluster atoms.

Figure 4. (A) Possible methane adsorption sites over Ni4−CeO2(111) catalyst surface. (B) Initial state (IS) where methane is adsorbed at the
interfacial O site of Ni4−CeO2(111) catalyst.

when the Ce atom is in a lower oxidation state. The Ce9−O sintering or the formation of larger active particles. Here, Ni
bond (2.42 Å) is also found to be slightly longer than the bulk atoms of the Ni4 cluster are strongly bound to the CeO2
Ce−O bond length (2.39 Å). Such increases in the Ce−O surface by the strong Ce−O−Ni bonds, which are also true for
bond length indicate decreases in the bond strength. According larger Ni-clusters and/or nanoparticles. This explains why the
to Tang et al.,78 a weak Ce−O bond is highly reactive toward sintering of active metal particles supported on the CeO2
methane because it can form a very strong OH bond that leads surface is significantly suppressed.14,92 The Ce−O−Ni bonds
to a lower C−H activation energy of methane. Moreover, the act as anchors during catalysis and prevent sintering of Ni
O atom at the metal−support interface becomes out of plane particles during high temperature reactions.
due to the deposition of Ni nanocluster, allowing the interfacial The top Ni site (Ni3 atom), on the other hand, represents a
O atom to be more easily reducible as compared to the surface metallic Ni atom that has a completely different electronic
O atom of CeO2. property from the interfacial Ni atoms. As shown in Figure 3C,
Changes also occur at the interfacial Ni sites (Ni1, Ni2, and the top-site Ni atom of the Ni4 cluster is almost charge neutral
Ni4 atoms), which are in contact with the surface O atoms. and has an electronic structure similar to the metallic Ni.
According to Jung et al.,87 the interfacial Ni atoms are Despite this, metal nanoclusters were reported to show a
polarized by the electron transfer from the surface, and thus, significantly different property from the bulk metal.93,94
their electronic property is different from the metallic Ni-atom Reactivity of metallic nanoclusters supported on metal oxides
nanoclusters and from the bare Ni surface. The Bader charge is superior to the reactivity of bulk active metal. The bond
analysis of our catalyst model also revealed that the interfacial distance in bulk crystalline nickel is 2.49 Å,95 but the gas-phase
Ni atoms are positively charged (see Table S1). It is clear from Ni4 cluster has a Ni−Ni bond length of 2.31 Å. After
the calculation of the charge density difference, as indicated by adsorption on the CeO2(111) surface, the Ni−Ni bond lengths
the violet-colored region in Figure 3B, that some electron of the Ni4 cluster increase and fall in a range from 2.40 to 2.49
density is shared between the interfacial Ni (electron donor) Å, indicating that the Ni−Ni bond strength is weakened. The
and O (electron acceptor) atoms, which leads to a hybrid- reason for this could be that electrons are transferred from the
ization between the Ni-3d orbitals and the O-2p orbitals and Ni4 cluster to the CeO2(111) surface, as also supported by the
eventually results in the formation of a strong covalent-like positively charged Ni atoms on the CeO2(111) surface.
bond between them. This is because such a covalent-like bond 3.2. Adsorption of Methane over Ni4−CeO2(111). Our
formation has also been found in a similar system consisting of catalyst model shows that the Ni4−CeO2(111) has four
a Pt4 cluster adsorbed on the CeO2(111) surface as described possible methane adsorption sites: (i) surface O site, (ii)
by Nguyen et al.88 It was experimentally found that methane interfacial O site, (iii) interfacial Ni site, and (iv) top site of Ni4
activation over NiOx, where Ni atom is positively charged, cluster (see Figure 4A). For the (i) site, we can actually define
occurs at higher temperatures, suggesting that the activation three different sites, namely top, bridge, and hollow site of
energy of methane over positively charged Ni is higher.89−91 surface O (see Figures S1 and S2 in the Supporting
Therefore, the positive charge of the interfacial Ni atoms Information). However, the optimizations of the bridge and
should effect on the activation barrier of methane. hollow sites show that the H atom of the methane molecule
In experiments, it is very commonly found that active shifts toward the top site of the surface O atom. Therefore,
particles after a catalytic reaction are larger than they were in only four sites are possible for methane adsorption over the
the freshly prepared catalyst because during the reaction they Ni4−CeO2(111) catalyst surface.
can move over the catalyst support surface if the metal− Considering all the four adsorption sites, we found that the
support interaction is weak. However, a strong metal−support adsorption energies of methane on the (i) top site of surface O,
interaction prevents such a movement and hinders the (ii) interfacial O site, (iii) interfacial Ni site, and (iv) top site of
9791 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

the Ni4 cluster atom are calculated to be −5.2, −5.6, −4.5, and
−3.8 kcal/mol, respectively. These values suggest that the
methane adsorption is energetically most stable at the
interfacial O site (see Figure 4B), consistent with the work
of Foppa et al.,49 suggesting that an interface modifies energy
profiles of a catalytic process and provides adsorption sites
where intermediates are more stable, which modifies the
overall reaction rate. According to Su et al.,45 the most active O
atom in a metal-doped CeO2(111) surface should be the
unstable O atom adjacent to the dopant with longer Ce−O
Figure 5. Final state (FS-R) of the radical route after methane
bond lengths, whereas the least active O atoms are the surface activation at the interfacial O site of the Ni4−CeO2(111) catalyst.
O atoms of CeO2(111). From the longer Ce−O bond lengths
of the interfacial O atom, it is evident that adsorption of
methane should occur at the interfacial O site of Ni4−
CeO2(111) catalyst system, not at the surface O site.87 Spin
density of the surface and interfacial O-atoms are provided in
the Supporting Information to this manuscript (see Table S2
in the Supporting Information).
As described above, there are two possible interfacial sites,
that is, interfacial O and Ni sites. However, it is known that the
interaction between an O atom and an H atom is much
stronger than the one between a metal atom and an H atom.
Moreover, Liu et al.96 reported that the oxidation of CO
requires very little amount of energy at the interface O site of
Au−TiO2 catalyst and the oxygen adsorption on the interface
is the rate determining step of the reaction. Thus, after
considering all the possibilities based on experimental,14,21,39
theoretical45,96 reports and the surface property of our model
Ni4−CeO2(111) catalyst, it is really possible that the interfacial
O sites could be the active site for methane activation during Figure 6. Energy diagram for methane activation at the interfacial O
different methane activation processes over the Ni−CeO2 site via the radical route (denoted as R).
catalyst. At the most stable adsorption site of methane
(interfacial O site, see Figure 4B), herein after referred to as The CI-NEB calculations for the transition-state determi-
the IS of methane activation process, the C···O and O···H nation show that at the TS (see TS-R in Figure 6), the CH3
distances are 3.49 and 2.50 Å, respectively. fragment forms a near-planar geometry, where the H−C−H
3.3. Possible Mechanisms of Methane Activation at angles range from 117.7° to 118.7°. The activation energy (Ea)
the Interfacial O Site. Here, we consider two known reaction and apparent activation energy (Eaa) of methane are calculated
routes, namely radical and nonradical routes, for the activation to be 38.7 and 33.1 kcal/mol, respectively. The dissociating
of methane, which are broadly reported. There are C−H bond of TS-R is found to be 1.90 Å, which is only by
experimental97−100 and theoretical11,41,101 evidences suggesting 0.11 Å shorter than that of FS-R. Such similar TS-R and FS-R
that the methane activation can occur via the radical route over structures indicate that TS-R is a late TS that leads to a high
different catalysts producing a •CH3 radical. To gain better activation energy.
understanding on the radical and nonradical route and to 3.3.2. Nonradical Route: CH4 + O (Interfacial O) → CH3* +
check which route is feasible and favorable, it is necessary to OH*. Similar to the radical route discussed above, here we also
carry out calculations of both routes and compare the consider the interfacial O site as the adsorption and activation
activation energies and the TS structures. In the radical site of methane, but the resultant CH3 fragment gets adsorbed
route, the C−H bond cleavage of methane leads to the on the catalyst surface37 (see Figures 7 and 8). The “*” is used
formation of a •CH3 radical, whereas in the nonradical route it to denote the species adsorbed and/or bonded to the catalyst
leads to the formation of a coordinated CH*3 ligand.13 surface.
3.3.1. Radical Route: CH4 + O (Interfacial O) → •CH3 + To discuss the nonradical route of methane activation, we
OH*. In this reaction route, methane is adsorbed preferably on first investigate the binding site of the CH3 fragment. The CH3
the interfacial O site (IS), and then one H atom of methane is fragment can be adsorbed on the Ni4 cluster, CeO2(111)
abstracted via a radical-like TS (TS-R) to form a final state of surface, and interfacial O sites. Let us first discuss CH3
methyl radical (FS-R) (see Figures 5 and 6). In this case, the adsorption on the Ni4 cluster. As shown in Figure 7, we
H-atom abstraction of methane takes place at the active consider the following four adsorption sites: (i) on-top site of
interfacial O atom, forming an O−H bond and a planar •CH3 the interfacial Ni atom, (ii) bridging site between the
radical that remains in the gas phase.38 In the FS-R shown in interfacial Ni atom and the top site (Ni3) of the cluster, (iii)
Figure 5, the separated C···H and C···O distances as well as the top-site (Ni3) of the cluster, and (iv) hollow site of the three
formed O−H bond length are found to be 2.01, 2.97, and 0.99 Ni atoms of the Ni4 cluster.
Å, respectively. The FS-R is by 27.6 kcal/mol less stable than The geometrical optimization of the case (i) results in the
the IS, consistent with the report by Mayernick and Janik,102 same geometry as that for the case (ii), suggesting that the on-
showing that the energy of •CH3 + H* over the Ni4− top site of the interfacial Ni atoms is not suitable for the CH3
CeO2(111) catalyst is higher than that of CH4(g) + catalyst. binding. Figure 7A−C shows that there are three possible
9792 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

Table 1. Possible CH3 Adsorption Sites on the Ni4 Cluster,


Surface O Site, and Interfacial O Site
Ni−C C−O relative
CH3 adsorption site bond bond energyb
a
geometry and bonding mode length (Å) length (Å) (kcal/mol)
Figure 7A Ni4 cluster, bridged 2.097, 5.09
between Ni1 and 1.995
Ni3
Figure 7B Ni4 cluster, bridged 2.048, 2.23
between Ni2 and 2.056
Ni3
Figure 7C Ni4 cluster, bridged 2.098, 1.64
between Ni4 and 2.031
Ni3
Figure 7D Ni4 cluster, on-top 1.938 11.19
site (Ni3)
Figure 8A surface O site 1.424 −26.82
Figure 8B interface O site 1.448 −7.52
attached to Ni2
Figure 8C interface O site 1.454 −6.72
attached to Ni4
a
See Figures 7 and 8 for the structures. bEnergy relative to the energy
of IS (the initial state represented in Figure 4B).

adsorption site of methane, the top of Ni1 atom should be the


first CH3 binding site, although we found that the optimization
of such a binding leads to a structure with the CH3 fragment
bridged between the Ni1 and Ni3 atoms. Thus, we think that
the CH3 binding on the Ni1−Ni3 bridge site is mechanistically
better than that on the Ni2−Ni3 and Ni4−Ni3 bridging sites.
Figure 7. Possible final-state geometries for the nonradical route, In the case of (iii), where the CH3 group is adsorbed on the
where CH3 is adsorbed over the Ni4 cluster. on-top site (Ni3) of the Ni4 cluster (see Figure 7D), our
calculations show that the structure is much less stable than the
structures shown in Figure 7A−C. This agrees with the report
by Roy and Chattopadhyay,76 showing that the most stable
adsorption site for the CH3 fragment is the bridging site of the
Ni4 cluster and is also in good agreement with that for Ni4−
Al2O3(0001) catalyst.38 From the top view of Figure 7D, it can
be seen that the CH3* is not exactly bonded to the top of the
Ni4 cluster but it is slightly tilted toward the Ni1 atom of the
Ni4 cluster, close to the methane adsorption site. In the case of
(iv), there are three possible hollow sites for the CH3 binding
on the Ni4 cluster (see Figure S3). However, geometry
optimizations of these binding structures cannot reach the
required accuracy and eventually lead to the same geometry as
the case (ii), suggesting that the CH3 binding on the three-fold
hollow site of the Ni4 cluster is not possible in this case.
The CH3 group can also be adsorbed on the surface and
interface O sites. As shown in Figure 8A, the adsorption of
CH3 on the support surface makes the O atom attached to the
CH3 group out of the surface plane and thus increases the
reducibility of the surface O site. In the second case (see
Figure 8B,C), the CH3 fragment is adsorbed on two interfacial
Figure 8. Possible geometries after the adsorption of CH3 on the O sites. Table 1 shows that the structure shown in Figure 8A is
surface and interfacial O sites. more stable than the structure shown in Figure 8B,C by 19.3
and 20.1 kcal/mol, respectively. Also, the structures shown in
Figure 8B,C are by 12.61 and 11.81 kcal/mol, respectively,
bridging sites representing the structure of the case (ii), which more stable than the structure shown in Figure 7A, where the
are Ni1−Ni3, Ni2−Ni3, and Ni4−Ni3 bridging sites. Table 1 CH3 fragment is bonded to the Ni4 cluster with a bridging
shows that CH3 binding on the Ni2−Ni3 and Ni4−Ni3 geometry. This suggests that the CH3 adsorption energetically
bridging sites (see Figure 7B,C, respectively) is more stable prefers these O sites rather than the Ni4-cluster site, in contrast
than that on the Ni1−Ni3 bridging site (see Figure 7A). to the experimental fact showing that the deposition of carbon
However, considering the first point of contact between the is found mainly on the active metal surface, not on the catalyst
CH3 fragment and the Ni4 cluster as well as the most stable support surface.14,103−106
9793 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

For the activation of methane, we consider only the scenario formation of H−Ni−CH3 species during the activation of
where the resultant CH3 is bonded to the Ni1−Ni3 bridge site, methane when the active site is Ni atom, not the interfacial O
while the other two scenarios, where the resultant CH3 is atom. However, as shown in Figure 10b, the interfacial O
adsorbed on the CeO2(111) surface and interfacial O sites, active site leads to the formation of H−O−Ni−CH3 like
respectively, are excluded because they require reaction species in the TS. Such a geometry of TS-NR is similar to the
coordinates that pass through the high-barrier radical route. four-centered TS structure reported by Shiota and Yoshizawa
As we will discuss below, the C−H activation energy of (see Figure 10b in bracket).108 One more important
methane via the nonradical route is much lower than that via observation is that the Ce−O bond distances of the active
the radical route. interfacial O site (O15) are further increased to 3.41, 2.76,
As shown in Figure 9, after methane is adsorbed preferably 2.80 from 2.70, 2.42, 2.61 Å at the IS, after methane activation,
on the interfacial O site (IS), the H-atom abstraction of and following the formation of the O−H bond. The active
interfacial O atom is already out of plane and the Ce−O bond
length is longer than the Ce−O bond length in the bulk CeO2.
Now, after methane activation at the interfacial O site, the
removal of the active interfacial O atom leading to an
interfacial O vacancy should be easier than what it was at the
initial stage. This is what we will pursue in the future. We have
also calculated other possible routes for methane activation
over other active sites, which are presented in Supporting
Information to the manuscript (see Figures S4, S5, and Table
S3).

4. CONCLUSIONS
DFT + U calculations have been carried out for the activation
of methane at the interface of a Ni4 cluster supported on the
Figure 9. Energy diagram for methane activation at the interfacial O CeO2(111) surface. It has been found that the deposition of
site through the nonradical route (denoted as NR), where the the Ni4 cluster over the CeO2(111) surface changes the surface
resultant CH3 fragment is adsorbed the Ni1−Ni3 bridge site. properties of CeO2 dramatically. Our model structures show
that two of the surface Ce4+ atoms get reduced to the Ce3+
methane occurs via a TS for the nonradical route (TS-NR) to atoms, which are close to the deposited Ni4 cluster. The
form a CH3 bonded to the Ni1−Ni3 bridge site and an addition of the Ni4 cluster draws the surface O atoms of CeO2
interfacial OH moiety (FS-NR). The Ea and Eaa are calculated outward and out of the surface plane, which increases the
to be 23.7 and 18.1 kcal/mol, respectively, the latter which is in distances of the bonds between the surface Ce atom and the
good agreement with the experimental value (17.4 kcal/mol) interfacial O atom, making such O atoms very reactive for
reported by Singha et al.14 for POM over Ni nanoparticles methane activation.
supported on a CeO2 catalyst. These values are by 15.0 kcal/ The adsorption energy of methane on the interfacial O site
mol lower than those for TS-R, suggesting that methane of the Ni4−CeO2(111) catalyst is −5.6 kcal/mol, which is
activation via the nonradical path is much more facile than that larger in magnitude than that on the top site of the deposited
via the radical route. The FS-NR is found to be less stable by Ni4 cluster. The significant point of the observation is that the
only 5.1 kcal/mol than the IS, but much more stable than the interfacial O site could be a better activation site for methane.
FS-R (see Figure 6). This is consistent with the report by Our DFT + U calculations reveal that the methane activation
Mayernick and Janik,102 suggesting that the formation energy takes place via the nonradical route with an activation energy
of CH3* + H* on CeO2 surfaces is lower than that for •CH3 + of 23.7 kcal/mol, which can be transformed into an apparent
H* on the same surfaces. activation energy of 18.1 kcal/mol. The latter agrees well with
From the structure of TS-NR, it can be seen that the CH3 an experimentally observed value (17.4 kcal/mol).14 These
fragment has a tetrahedral-like geometry and is bonded to the calculated activation energies are lower than those for the
Ni1 atom with a separated C···H distance of 1.51 Å, while the radical route by 15.0 kcal/mol, indicating that the nonradical
abstracted H atom is approaching the interfacial O atom with a route is more favorable for methane activation at the interfacial
O···H distance of 1.22 Å, a structure that is in contrast to the O site of the Ni4−CeO2(111) catalyst.
TS-R structure (see Figure 10a). Yang et al.107 reported the Our study also has revealed that the most stable adsorption
geometry of the CH3 on the surface is found in a bridging site
of the Ni4 cluster, neither on the top of the interfacial Ni atom,
at the three-fold hollow site, nor at the top site of the Ni4
cluster. A low activation barrier at the interfacial O site
indicates that methane activation over the Ni4−CeO2(111)
catalyst can occur under low temperatures, making the catalyst
very potent for methane reforming reactions.

Figure 10. TS structures for methane activation over the Ni4−



*
ASSOCIATED CONTENT
S Supporting Information

CeO2(111) catalyst via the (a) radical and (b) nonradical routes. The The Supporting Information is available free of charge on the
distances between atoms and/or the bond lengths are in Å unit. ACS Publications website at DOI: 10.1021/acs.jpcc.8b11973.
9794 DOI: 10.1021/acs.jpcc.8b11973
J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

Initial states with different methane adsorption sites; (9) Snyder, B. E. R.; Vanelderen, P.; Bols, M. L.; Hallaert, S. D.;
change in oxidation state of surface Ce atoms; spin Böttger, L. H.; Ungur, L.; Pierloot, K.; Schoonheydt, R. A.; Sels, B. F.;
density of surface and interface O atoms; other Solomon, E. I. The active site of low-temperature methane
postulated geometries for FS of the nonradial route; hydroxylation in iron-containing zeolites. Nature 2016, 536, 317−321.
(10) Mahyuddin, M. H.; Tanaka, T.; Shiota, Y.; Staykov, A.;
activation energies and routes for methane activation at
Yoshizawa, K. Methane partial oxidation over [Cu2(μ-O)]2+ and
other plausible activation sites; and atomic coordinates [Cu3(μ-O)3]2+ active species in large-pore zeolites. ACS Catal. 2018,
of the optimized structures in the nonradical route 8, 1500−1509.
(PDF) (11) Mahyuddin, M. H.; Shiota, Y.; Staykov, A.; Yoshizawa, K.


Theoretical investigation of methane hydroxylation over isoelectronic
AUTHOR INFORMATION [FeO]2+- and [MnO]+-exchanged zeolites activated by N2O. Inorg.
Chem. 2017, 56, 10370−10380.
Corresponding Author (12) Mahyuddin, M. H.; Tanaka, T.; Staykov, A.; Shiota, Y.;
*E-mail: kazunari@ms.ifoc.kyushu-u.ac.jp. Phone: +81-92-802- Yoshizawa, K. Dioxygen activation on Cu-MOR zeolite: theoretical
2529. Fax: +81-92-802-2528. insights into the formation of Cu2O and Cu3O3 active species. Inorg.
ORCID Chem. 2018, 57, 10146−10152.
(13) Mahyuddin, M. H.; Shiota, Y.; Staykov, A.; Yoshizawa, K.
Yuta Tsuji: 0000-0003-4224-4532
Theoretical Overview of Methane Hydroxylation by Copper-Oxygen
Muhammad Haris Mahyuddin: 0000-0002-8017-7847 Species in Enzymatic and Zeolitic Catalysts. Acc. Chem. Res. 2018, 51,
Kazunari Yoshizawa: 0000-0002-6279-9722 2382−2390.
Notes (14) Singha, R. K.; Shukla, A.; Yadav, A.; Sivakumar Konathala, L.
The authors declare no competing financial interest. N.; Bal, R. Effect of metal-support interaction on activity and stability


of Ni-CeO2 catalyst for partial oxidation of methane. Appl. Catal., B
ACKNOWLEDGMENTS 2017, 202, 473−488.
(15) Mei, D.; Glezakou, V.-A.; Lebarbier, V.; Kovarik, L.; Wan, H.;
This work was supported by KAKENHI grant numbers Albrecht, K. O.; Gerber, M.; Rousseau, R.; Dagle, R. A. Highly active
JP24109014 and JP15K13710 from Japan Society for the and stable MgAl2O4-supported Rh and Ir catalysts for methane steam
Promotion of Science (JSPS) and the Ministry of Education, reforming: A combined experimental and theoretical study. J. Catal.
Culture, Sports, Science and Technology of Japan (MEXT), 2014, 316, 11−23.
the MEXT Projects of “Integrated Research Consortium on (16) Li, J.; Croiset, E.; Ricardez-Sandoval, L. Effect of metal−
Chemical Sciences”, “Elements Strategy Initiative to Form support interface during CH4 and H2 dissociation on Ni/γ-Al2O3: a
Core Research Center”, the Cooperative Research Program of density functional theory study. J. Phys. Chem. C 2013, 117, 16907−
“Network Joint Research Center for Materials and Devices”, 16920.
and the JST-CREST JPMJCR15P5. The computation was (17) Liu, Z.; Lustemberg, P.; Gutiérrez, R. A.; Carey, J. J.; Palomino,
mainly carried out using the computer facilities at Research R. M.; Vorokhta, M.; Grinter, D. C.; Ramírez, P. J.; Matolín, V.;
Institute for Information Technology, Kyushu University. Y.T. Nolan, M.; et al. In Situ Investigation of Methane Dry Reforming on
Metal/Ceria(111) Surfaces: Metal-Support Interactions and C−H
thanks JSPS KAKENHI grant numbers JP17K14440 and
Bond Activation at Low Temperature. Angew. Chem., Int. Ed. 2017,
JP18H04488.


56, 13041−13046.
(18) Lustemberg, P. G.; Ramírez, P. J.; Liu, Z.; Gutiérrez, R. A.;
REFERENCES Grinter, D. G.; Carrasco, J.; Senanayake, S. D.; Rodriguez, J. A.;
(1) Kwon, Y.; Kim, T. Y.; Kwon, G.; Yi, J.; Lee, H. Selective Ganduglia-Pirovano, M. V. Room-temperature activation of methane
activation of methane on single-atom catalyst of rhodium dispersed on and dry reforming with CO2 on Ni-CeO2(111) surfaces: effect of Ce3+
zirconia for direct conversion. J. Am. Chem. Soc. 2017, 139, 17694− sites and metal−support interactions on C−H bond cleavage. ACS
17699. Catal. 2016, 6, 8184−8191.
(2) Holmen, A. Direct conversion of methane to fuels and chemicals. (19) Lustemberg, P. G.; Palomino, R. M.; Gutiérrez, R. A.; Grinter,
Catal. Today 2009, 142, 2−8. D. C.; Vorokhta, M.; Liu, Z.; Ramírez, P. J.; Matolín, V.; Ganduglia-
(3) Luo, J. Z.; Yu, Z. L.; Ng, C. F.; Au, C. T. CO2/CH4 reforming Pirovano, M. V.; Senanayake, S. D.; et al. Direct Conversion of
over Ni−La2O3/5A: an investigation on carbon deposition and Methane to Methanol on Ni-Ceria Surfaces: Metal-Support
reaction steps. J. Catal. 2000, 194, 198−210. Interactions and Water-Enabled Catalytic Conversion by Site
(4) Tang, S.; Ji, L.; Lin, J.; Zeng, H. C.; Tan, K. L.; Li, K. CO2 Blocking. J. Am. Chem. Soc. 2018, 140, 7681−7687.
reforming of methane to synthesis gas over sol−gel-made Ni/γ-Al2O3 (20) Singha, R. K.; Shukla, A.; Yadav, A.; Sasaki, T.; Sandupatla, A.;
catalysts from organometallic precursors. J. Catal. 2000, 194, 424− Deo, G.; Bal, R. Pt−CeO2 nanoporous spheres − an excellent catalyst
430.
for partial oxidation of methane: effect of the bimodal pore structure.
(5) Wang, Z.-C.; Dietl, N.; Kretschmer, R.; Ma, J.-B.; Weiske, T.;
Schlangen, M.; Schwarz, H. Direct conversion of methane into Catal. Sci. Technol. 2017, 7, 4720−4735.
formaldehyde mediated by [Al2O3]+ at room temperature. Angew. (21) Scarabello, A.; Dalle Nogare, D.; Canu, P.; Lanza, R. Partial
Chem., Int. Ed. 2012, 51, 3703−3707. oxidation of methane on Rh/ZrO2 and Rh/Ce−ZrO2 on monoliths:
(6) Zhang, Q.; He, D.; Zhu, Q. Recent progress in direct partial catalyst restructuring at reaction conditions. Appl. Catal., B 2015,
oxidation of methane to methanol. J. Nat. Gas Chem. 2003, 12, 81− 174−175, 308−322.
89. (22) Kondratenko, V. A.; Berger-Karin, C.; Kondratenko, E. V.
(7) Palkovits, R.; Antonietti, M.; Kuhn, P.; Thomas, A.; Schüth, F. Partial oxidation of methane to syngas over γ-Al2O3-supported rh
Solid Catalysts for the Selective Low-Temperature Oxidation of nanoparticles: kinetic and mechanistic origins of size effect on
Methane to Methanol. Angew. Chem., Int. Ed. 2009, 48, 6909−6912. selectivity and activity. ACS Catal. 2014, 4, 3136−3144.
(8) Woertink, J. S.; Smeets, P. J.; Groothaert, M. H.; Vance, M. A.; (23) Pantaleo, G.; La Parola, V.; Deganello, F.; Singha, R. K.; Bal, R.;
Sels, B. F.; Schoonheydt, R. A.; Solomon, E. I. A [Cu2O]2+ core in Cu- Venezia, A. M. Ni/CeO2 catalysts for methane partial oxidation:
ZSM-5, the active site in the oxidation of methane to methanol. Proc. synthesis driven structural and catalytic effects. Appl. Catal., B 2016,
Natl. Acad. Sci. U.S.A. 2009, 106, 18908−18913. 189, 233−241.

9795 DOI: 10.1021/acs.jpcc.8b11973


J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

(24) Choudhary, T. V.; Choudhary, V. R. Energy-efficient syngas (45) Su, Y.-Q.; Filot, I. A. W.; Liu, J.-X.; Hensen, E. J. M. Stable Pd-
production through catalytic oxy-methane reforming reactions. Angew. doped ceria structures for CH4 activation and CO oxidation. ACS
Chem., Int. Ed. 2008, 47, 1828−1847. Catal. 2018, 8, 75−80.
(25) Paksoy, A. I.; Caglayan, B. S.; Aksoylu, A. E. A study on (46) Li, Y.-K.; Yuan, Z.; Zhao, Y.-X.; Zhao, C.; Liu, Q.-Y.; Chen, H.;
characterization and methane dry reforming performance of Co−Ce/ He, S.-G. Thermal methane conversion to syngas mediated by Rh1-
ZrO2 catalyst. Appl. Catal., B 2015, 168−169, 164−174. doped aluminum oxide cluster cations RhAl3O4+. J. Am. Chem. Soc.
(26) Elsayed, N. H.; Roberts, N. R. M.; Joseph, B.; Kuhn, J. N. Low 2016, 138, 12854−12860.
temperature dry reforming of methane over Pt-Ni-Mg/ceria-zirconia (47) Tian, D.; Li, K.; Wei, Y.; Zhu, X.; Zeng, C.; Cheng, X.; Zheng,
catalysts. Appl. Catal., B 2015, 179, 213−219. Y.; Wang, H. DFT insights into oxygen vacancy formation and CH4
(27) Berger-Karin, C.; Radnik, J.; Kondratenko, E. V. Mechanistic activation over CeO2 surfaces modified by transition metals (Fe, Co
origins of the promoting effect of tiny amounts of Rh on the and Ni). Phys. Chem. Chem. Phys. 2018, 20, 11912−11929.
performance of NiOx/Al2O3 in partial oxidation of methane. J. Catal. (48) Riley, C.; Zhou, S.; Kunwar, D.; De La Riva, A.; Peterson, E.;
2011, 280, 116−124. Payne, R.; Gao, L.; Lin, S.; Guo, H.; Datye, A. Design of effective
(28) Yao, H.; Yao, Y. F. Y. Ceria in automotive exhaust catalysts I. catalysts for selective alkyne hydrogenation by doping of ceria with a
Oxygen storage. J. Catal. 1984, 86, 254−265. single-atom promotor. J. Am. Chem. Soc. 2018, 140, 12964−12973.
(29) Trovarelli, A.; Dolcetti, G.; de Leitenburg, C.; Kaspar, J. CO2 (49) Foppa, L.; Margossian, T.; Kim, S. M.; Müller, C.; Copéret, C.;
hydrogenation over platinum group metals supported on CeO2: Larmier, K.; Comas-Vives, A. Contrasting the role of Ni/Al2O3
evidence for a transient metal-support interaction. Stud. Surf. Sci. interfaces in water−gas shift and dry reforming of methane. J. Am.
Catal. 1993, 75, 2781−2784. Chem. Soc. 2017, 139, 17128−17139.
(30) Haneda, M.; Mizushima, T.; Kakuta, N. Synergistic effect (50) Petkov, P. S.; Vayssilov, G. N.; Krüger, S.; Rösch, N. Structure,
between Pd and nonstoichiometric cerium oxide for oxygen activation stability, electronic and magnetic properties of Ni4 clusters containing
in methane oxidation. J. Phys. Chem. B 1998, 102, 6579−6587. impurity atoms. Phys. Chem. Chem. Phys. 2006, 8, 1282−1291.
(31) Feng, Y.; Wan, Q.; Xiong, H.; Zhou, S.; Chen, X.; Pereira (51) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy
Hernandez, X. I.; Wang, Y.; Lin, S.; Datye, A. K.; Guo, H. Correlating calculations for metals and semiconductors using a plane-wave basis
DFT calculations with CO oxidation reactivity on Ga-doped Pt/CeO2 set. J. Comput. Mater. Sci. 1996, 6, 15−50.
single-atom catalysts. J. Phys. Chem. C 2018, 122, 22460−22468. (52) Kresse, G.; Furthmüller, J. Efficient iterative schemes forab
(32) Paier, J.; Penschke, C.; Sauer, J. Oxygen defects and surface initiototal-energy calculations using a plane-wave basis set. Phys. Rev.
chemistry of ceria: quantum chemical studies compared to experi- B: Condens. Matter Mater. Phys. 1996, 54, 11169−11186.
ment. Chem. Rev. 2013, 113, 3949−3985. (53) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and
(33) Senftle, T. P.; van Duin, A. C. T.; Janik, M. J. Methane
accurate ab initio parametrization of density functional dispersion
activation at the Pd/CeO2 interface. ACS Catal. 2017, 7, 327−332.
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,
(34) Farmer, J. A.; Campbell, C. T. Ceria maintains smaller metal
132, 154104.
catalyst particles by strong metal-support bonding. Science 2010, 329,
(54) Antony, A.; Hakanoglu, C.; Asthagiri, A.; Weaver, J. F.
933−936.
Dispersion-corrected density functional theory calculations of the
(35) Senanayake, S. D.; Zhou, J.; Baddorf, A. P.; Mullins, D. R. The
molecular binding of n-alkanes on Pd(111) and PdO(101). J. Chem.
reaction of carbon monoxide with palladium supported on cerium
oxide thin films. Surf. Sci. 2007, 601, 3215−3223. Phys. 2012, 136, 054702.
(36) Cargnello, M.; Jaen, J. J. D.; Garrido, J. C. H.; Bakhmutsky, K.; (55) Antony, A.; Asthagiri, A.; Weaver, J. F. Pathways and kinetics of
Montini, T.; Gamez, J. J. C.; Gorte, R. J.; Fornasiero, P. Exceptional methane and ethane C-H bond cleavage on PdO(101). J. Chem. Phys.
activity for methane combustion over modular Pd@CeO2 subunits on 2013, 139, 104702.
functionalized Al2O3. Science 2012, 337, 713−717. (56) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.;
(37) Guo, D.; Wang, G.-C. Partial oxidation of methane on anatase Sutton, A. P. Electron-energy-loss spectra and the structural stability
and rutile defective TiO2 supported Rh4 cluster: a density functional of nickel oxide: An LSDA+U study. Phys. Rev. B: Condens. Matter
theory study. J. Phys. Chem. C 2017, 121, 26308−26320. Mater. Phys. 1998, 57, 1505−1509.
(38) Zhang, M.; Yu, Y.; Zhang, Y. DFT research of methane (57) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient
preliminary dissociation on aluminum catalyst. Appl. Surf. Sci. 2013, approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
280, 15−24. (58) Krcha, M. D.; Mayernick, A. D.; Janik, M. J. Periodic trends of
(39) Du, X.; Zhang, D.; Shi, L.; Gao, R.; Zhang, J. Morphology oxygen vacancy formation and C−H bond activation over transition
dependence of catalytic properties of Ni/CeO2 nanostructures for metal-doped CeO2(111) surfaces. J. Catal. 2012, 293, 103−115.
carbon dioxide reforming of methane. J. Phys. Chem. C 2012, 116, (59) Wan, Q.; Wei, F.; Wang, Y.; Wang, F.; Zhou, L.; Lin, S.; Xie,
10009−10016. D.; Guo, H. Single atom detachment from Cu clusters, and diffusion
(40) Dianat, A.; Seriani, N.; Ciacchi, L. C.; Bobeth, M.; Cuniberti, and trapping on CeO2(111): implications in ostwald ripening and
G. DFT study of reaction processes of methane combustion on atomic redispersion. Nanoscale 2018, 10, 17893−17901.
PdO(100). Chem. Phys. 2014, 443, 53−60. (60) Carey, J. J.; Nolan, M. Cation doping size effect for methane
(41) Mayernick, A. D.; Janik, M. J. Methane oxidation on Pd−Ceria: activation on alkaline earth metal doping of the CeO2(111) surface.
A DFT study of the mechanism over PdxCe1−xO2, Pd, and PdO. J. Catal. Sci. Technol. 2016, 6, 3544−3558.
Catal. 2011, 278, 16−25. (61) Cheng, Z.; Qin, L.; Guo, M.; Xu, M.; Fan, J. A.; Fan, L.-S.
(42) Hu, W.; Lan, J.; Guo, Y.; Cao, X.-M.; Hu, P. Origin of efficient Oxygen vacancy promoted methane partial oxidation over iron oxide
catalytic combustion of methane over Co3O4(110): active low- oxygen carriers in the chemical looping process. Phys. Chem. Chem.
coordination lattice oxygen and cooperation of multiple active sites. Phys. 2016, 18, 32418−32428.
ACS Catal. 2016, 6, 5508−5519. (62) Jiang, Y.; Adams, J. B.; van Schilfgaarde, M. Density-functional
(43) An, W.; Zeng, X. C.; Turner, C. H. First-principles study of calculation of CeO2 surfaces and prediction of effects of oxygen partial
methane dehydrogenation on a bimetallic Cu/Ni(111) surface. J. pressure and temperature on stabilities. J. Chem. Phys. 2005, 123,
Chem. Phys. 2009, 131, 174702. 064701.
(44) Dianat, A.; Seriani, N.; Ciacchi, L. C.; Pompe, W.; Cuniberti, (63) Calderon, C. E.; Plata, J. J.; Toher, C.; Oses, C.; Levy, O.;
G.; Bobeth, M. Dissociative Adsorption of Methane on Surface Oxide Fornari, M.; Natan, A.; Mehl, M. J.; Hart, G.; Buongiorno Nardelli,
Structures of Pd−Pt Alloys. J. Phys. Chem. C 2009, 113, 21097− M.; et al. The AFLOW standard for high-throughput materials science
21105. calculations. Comput. Mater. Sci. 2015, 108, 233−238.

9796 DOI: 10.1021/acs.jpcc.8b11973


J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

(64) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image (86) Sugiura, K.; Ogo, S.; Iwasaki, K.; Yabe, T.; Sekine, Y. Low-
nudged elastic band method for finding saddle points and minimum temperature catalytic oxidative coupling of methane in an electric field
energy paths. J. Chem. Phys. 2000, 113, 9901−9904. over a Ce−W−O catalyst system. Sci. Rep. 2016, 6, 25154.
(65) Henkelman, G.; Jónsson, H. Improved tangent estimate in the (87) Jung, C.; Tsuboi, H.; Koyama, M.; Kubo, M.; Broclawik, E.;
nudged elastic band method for finding minimum energy paths and Miyamoto, A. Different support effect of M/ZrO2 and M/CeO2
saddle points. J. Chem. Phys. 2000, 113, 9978−9985. (M=Pd and Pt) catalysts on CO adsorption: A periodic density
(66) Smidstrup, S.; Pedersen, A.; Stokbro, K.; Jónsson, H. Improved functional study. Catal. Today 2006, 111, 322−327.
initial guess for minimum energy path calculations. J. Chem. Phys. (88) Nguyen, T. Q.; Escaño, M. C. S.; Nakanishi, H.; Kasai, H.;
2014, 140, 214106. Maekawa, H.; Osumi, K.; Sato, K. DFT+U study on the oxygen
(67) Virtual NanoLab 2016. https://quantumwise.com (accessed adsorption and dissociation on CeO2-supported platinum cluster.
Jan 25, 2018). Appl. Surf. Sci. 2014, 288, 244−250.
(68) Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust (89) Branco, J. B.; Ferreira, A. C.; Gasche, T. A.; Pimenta, G.; Leal, J.
algorithm for bader decomposition of charge density. Comput. Mater. P. Low Temperature Partial Oxidation of Methane over Bimetallic
Sci. 2006, 36, 354−360. Nickel-fBlock Element Oxide Nanocatalysts. Adv. Synth. Catal. 2014,
(69) Sanville, E.; Kenny, S. D.; Smith, R.; Henkelman, G. Improved 356, 3048−3058.
grid-based algorithm for bader charge allocation. Comput. Chem. (90) Hickman, D. A.; Schmidt, L. D. Production of syngas by direct
2007, 28, 899−908. catalytic oxidation of methane. Science 1993, 259, 343−346.
(70) Momma, K.; Izumi, F. VESTA 3for three-dimensional (91) Mallens, E. P. J.; Hoebink, J. H. B. J.; Marin, G. B. The reaction
visualization of crystal, volumetric and morphology data. J. Appl. mechanism of the partial oxidation of methane to synthesis gas: A
Crystallogr. 2011, 44, 1272−1276. transient kinetic study over rhodium and a comparison with platinum.
(71) Prodhomme, P.-Y.; Raybaud, P.; Toulhoat, H. Free-energy J. Catal. 1997, 167, 43−56.
profiles along reduction pathways of MoS2 M-Edge and S-Edge by (92) Pantu, P.; Gavalas, G. R. Methane partial oxidation on Pt/CeO2
dihydrogen: A first-principles study. J. Catal. 2011, 280, 178−195. and Pt/Al2O3 catalysts. Appl. Catal., A 2002, 223, 253−260.
(72) Bedolla, P. O.; Feldbauer, G.; Wolloch, M.; Gruber, C.; Eder, S. (93) Wilcoxon, J. P.; Abrams, B. L. Synthesis, structure and
J.; Dörr, N.; Mohn, P.; Redinger, J.; Vernes, A. Density functional properties of metal nanoclusters. Chem. Soc. Rev. 2006, 35, 1162−
investigation of the adsorption of isooctane, ethanol, and acetic acid 1194.
on a water-covered Fe(100) surface. J. Phys. Chem. C 2014, 118, (94) Carrasco, J.; Barrio, L.; Liu, P.; Rodriguez, J. A.; Ganduglia-
Pirovano, M. V. Theoretical studies of the adsorption of CO and C on
21428−21437.
(73) Tsuji, Y.; Kitamura, Y.; Someya, M.; Takano, T.; Yaginuma, M.; Ni(111) and Ni/CeO2(111): Evidence of a strong metal−support
interaction. J. Phys. Chem. C 2013, 117, 8241−8250.
Nakanishi, K.; Yoshizawa, K. Adhesion of epoxy resin with hexagonal
(95) Lathiotakis, N. N.; Andriotis, A. N.; Menon, M.; Connolly, J.
boron nitride and graphite. ACS Omega 2019, 4, 4491−4504.
Tight binding molecular dynamics study of Ni clusters. J. Chem. Phys.
(74) Crosby, S. A.; Millward, G. E.; Butler, E. I.; Turner, D. R.;
1996, 104, 992−1003.
Whitfield, M. Kinetics of phosphate adsorption by iron oxyhydroxides
(96) Liu, Z.-P.; Gong, X.-Q.; Kohanoff, J.; Sanchez, C.; Hu, P.
in aqueous systems. Estuarine, Coastal Shelf Sci. 1984, 19, 257−270.
Catalytic role of metal oxides in gold-based catalysts: A first principles
(75) van Santen, R. A. Molecular Catalytic Kinetics Concepts. In
study of CO oxidation on TiO2 supported Au. Phys. Rev. Lett. 2003,
Novel Concepts in Catalysis and Chemical Reactors; Wiley-VCH Verlag:
91, 266102.
Weinheim, Germany, 2010; pp 1−30. (97) Luo, L.; Tang, X.; Wang, W.; Wang, Y.; Sun, S.; Qi, F.; Huang,
(76) Roy, G.; Chattopadhyay, A. P. Dissociation of methane on Ni 4
W. Methyl radicals in oxidative coupling of methane directly
cluster-A DFT study. Comput. Theor. Chem. 2017, 1106, 7−14. confirmed by synchrotron VUV photoionization mass spectroscopy.
(77) Johnston, R. L. Atomic and Molecular Clusters; CRC Press:
Sci. Rep. 2013, 3, 1625.
London, U.K., 2002. (98) Hayek, N. S.; Lucas, N. S.; Warwar Damouny, C.; Gazit, O. M.
(78) Tang, W.; Hu, Z.; Wang, M.; Stucky, G. D.; Metiu, H.; Critical surface parameters for the oxidative coupling of methane over
McFarland, E. W. Methane complete and partial oxidation catalyzed the Mn−Na−W/SiO2 catalyst. ACS Appl. Mater. Interfaces 2017, 9,
by Pt-doped CeO2. J. Catal. 2010, 273, 125−137. 40404−40411.
(79) Pakhare, D.; Spivey, J. A review of dry (CO2) reforming of (99) Kumar, G.; Lau, S. L. J.; Krcha, M. D.; Janik, M. J. Correlation
methane over noble metal catalysts. Chem. Soc. Rev. 2014, 43, 7813− of methane activation and oxide catalyst reducibility and its
7837. implications for oxidative coupling. ACS Catal. 2016, 6, 1812−1821.
(80) Das, S.; Thakur, S.; Bag, A.; Gupta, M. S.; Mondal, P.; Bordoloi, (100) Davydov, A. A.; Budneva, A. A.; Aliev, S. M.; Sokolovskii, V.
A. Support interaction of Ni nanocluster based catalysts applied in D. IR spectra of methane adsorbed on MgO. React. Kinet. Catal. Lett.
CO2 reforming. J. Catal. 2015, 330, 46−60. 1988, 36, 491−495.
(81) Yu, M.; Zhu, Y.-A.; Lu, Y.; Tong, G.; Zhu, K.; Zhou, X. The (101) Li, G.; Vassilev, P.; Sanchez-Sanchez, M.; Lercher, J. A.;
promoting role of Ag in Ni-CeO2 catalyzed CH4-CO2 dry reforming Hensen, E. J. M.; Pidko, E. A. Stability and reactivity of copper oxo-
reaction. Appl. Catal., B 2015, 165, 43−56. clusters in ZSM-5 zeolite for selective methane oxidation to methanol.
(82) Singha, R. K.; Yadav, A.; Shukla, A.; Kumar, M.; Bal, R. Low J. Catal. 2016, 338, 305−312.
temperature dry reforming of methane over Pd-CeO2 nanocatalyst. (102) Mayernick, A. D.; Janik, M. J. Methane activation and oxygen
Catal. Commun. 2017, 92, 19−22. vacancy formation over CeO2 and Zr, Pd substituted CeO2 surfaces. J.
(83) Stagg-Williams, S. M.; Noronha, F. B.; Fendley, G.; Resasco, D. Phys. Chem. C 2008, 112, 14955−14964.
E. CO2 reforming of CH4 over Pt/ZrO2 catalysts promoted with La (103) Singha, R. K.; Shukla, A.; Sandupatla, A.; Deo, G.; Bal, R.
and Ce oxides. J. Catal. 2000, 194, 240−249. Synthesis and catalytic activity of a Pd doped Ni-MgO catalyst for dry
(84) Scarabello, A.; Dalle Nogare, D.; Canu, P.; Lanza, R. Partial reforming of methane. J. Mater. Chem. A 2017, 5, 15688−15699.
oxidation of methane on Rh/ZrO2 and Rh/Ce−ZrO2 on monoliths: (104) Wang, N.; Shen, K.; Huang, L.; Yu, X.; Qian, W.; Chu, W.
Catalyst restructuring at reaction conditions. Appl. Catal., B 2015, Facile Route for Synthesizing Ordered Mesoporous Ni-Ce-Al Oxide
174−175, 308−322. Materials and Their Catalytic Performance for Methane Dry
(85) Pantaleo, G.; La Parola, V.; Deganello, F.; Calatozzo, P.; Bal, R.; Reforming to Hydrogen and Syngas. ACS Catal. 2013, 3, 1638−1651.
Venezia, A. M. Synthesis and support composition effects on CH4 (105) Gili, A.; Schlicker, L.; Bekheet, M. F.; Görke, O.; Penner, S.;
partial oxidation over Ni−CeLa oxides. Appl. Catal., B 2015, 164, Grünbacher, M.; Götsch, T.; Littlewood, P.; Marks, T. J.; Stair, P. C.;
135−143. et al. Surface carbon as a reactive intermediate in dry reforming of

9797 DOI: 10.1021/acs.jpcc.8b11973


J. Phys. Chem. C 2019, 123, 9788−9798
The Journal of Physical Chemistry C Article

methane to syngas on a 5% Ni/MnO catalyst. ACS Catal. 2018, 8,


8739−8750.
(106) Li, M.; van Veen, A. C. Tuning the catalytic performance of
Ni-catalysed dry reforming of methane and carbon deposition via Ni-
CeO2‑x interaction. Appl. Catal., B 2018, 237, 641−648.
(107) Yang, H.-Q.; Chen, Y.-Q.; Hu, C.-W.; Hu, H.-R.; Gong, M.-
C.; Tian, A.-M.; Wong, N.-B. C−H bond activation: Ni-
(d10 1S)+CH4→NiCH2+H2. A DFT study. J. Mol. Struct.: THEO-
CHEM 2001, 574, 57−74.
(108) Shiota, Y.; Yoshizawa, K. Methane-to-methanol conversion by
first-row transition-metal oxide ions: ScO+, TiO+, VO+, CrO+, MnO+,
FeO+, CoO+, NiO+, and CuO+. J. Am. Chem. Soc. 2000, 122, 12317−
12326.

9798 DOI: 10.1021/acs.jpcc.8b11973


J. Phys. Chem. C 2019, 123, 9788−9798

You might also like