You are on page 1of 13

pubs.acs.

org/acscatalysis Research Article

Hydrogenation of CO2 to Methane over a Ru/RuTiO2 Surface: A DFT


Investigation into the Significant Role of the RuO2 Overlayer
Jie Yu, Yabing Zeng, Qirou Jin, Wei Lin,* and Xin Lu*
Cite This: ACS Catal. 2022, 12, 14654−14666 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Understanding metal−support interactions is crit-


ical to the rational design of a catalyst for CO2 methanation. In this
regard, the density functional theory is employed to shed light on
the factors that determine the difference between RuTiO2- and
Downloaded via KHON KAEN UNIV on March 10, 2024 at 10:15:42 (UTC).

TiO2-supported Ru nanoparticles. Structural observations and the


calculated ratio of cohesive energy and adsorption energy (Ecoh/
Eads) suggest that supported Ru6 could form as an epitaxial layer
along with RuTiO2 as well as display strong metal−support
interactions with TiO2 and thereby is used as a surface model to
simulate the nanoparticles employed in the experiment. Fur-
thermore, electronic analysis reveals that due to the existence of a
RuO2 overlayer, more electrons are transferred from the support to
the Ru cluster. Benefiting from this, CO2 is lower in adsorption
energy since electrons from Ru6/RuTiO2 are less likely to fill in the antibonding orbital of Ru−O interaction. Analysis of the
minimum-energy pathway indicates that the methanation of CO2 is led by C−O direct bond cleavage rather than the formate
pathway in the first place for both surfaces, which is consistent with the FTIR spectroscopy results. Besides, we noticed that different
reaction mechanisms control methane synthesis from the onset of CHO* formation. On the one hand, CHO* prefers an associative
mechanism on Ru6/RuTiO2 due to the lower d-band center of the support and facile formation of CH2O* species. On the other
hand, closer proximity of the d-band center to the Fermi level (EFermi) and preferable CH* formation promote CHO* on Ru6/TiO2
to undergo a dissociative pathway. Our comparative studies suggest that the RuO2 overlayer plays a key role in determining the
reaction mechanism of CO2 methanation for Ru/r-TiO2.
KEYWORDS: density functional theory, d-band center, RuO2 overlayer, methane formation, associative/dissociative mechanism

1. INTRODUCTION kJ/mol), multiple inactive elementary steps with high energy


Anthropogenic carbon dioxide (CO2) emission is one of the barriers limit its application. Thus, a higher temperature is
major culprits for the greenhouse effect, so much effort has commonly required to ensure a high conversion rate. However,
been devoted to searching for alternative low-carbon- this brings in another problem, where the reverse water−gas
containing fossil fuels.1,2 In addition to keeping the source of shift reaction would occur at this temperature range and
pollution within limits, postdisposal of already produced CO2 reduce CH4 selectivity.16 To meet the need for both methane
as the raw material has already emerged as a hot topic globally selectivity and conversion rate, a suitable catalyst should be
for the CO2-reduction research area. This is because CO2 is a designed. Among various choices, Ru-based catalysts have been
renewable and environment-friendly carbon feedstock that can regarded by far as the optimal ones for CO2 methanation at
be applied to the synthesis of chemicals such as methanol, low temperatures because they are much more stable and
methane, and so forth.3−8 CO2 methanation (CO2 + 4H2 = active over time than Ni-based catalysts which experience
CH4 + 2H2O), the so-called Sabatier reaction, has attracted severe deactivation resulting from coking and sintering and are
extensive attention for two primary reasons.9,10 On the one sensitive to oxidizing environments.17 For instance, Ru(0001)
hand, it takes advantage of the excessive electric energy has been verified computationally to exhibit higher activity for
originating from biomass which could serve to produce H2 gas
and thereby realizes the recyclable usage of CO2.11−13 On the Received: September 14, 2022
other hand, higher selectivity is achieved from much lower Revised: November 3, 2022
pressure compared with methanolization which is another Published: November 16, 2022
common CO2 hydrogenation product.14,15
Even though methane synthesis via CO2 hydrogenation is
feasible from the thermodynamic aspect (ΔG298.15K = −130.8

© 2022 American Chemical Society https://doi.org/10.1021/acscatal.2c04539


14654 ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 1. (a) Schematic representation of constructing RuTiO2 and TiO2. (b) Variation of the ratio of Ecoh/Eads as a function of the number of Ru
atoms in Run (n = 2−8) for both Run/RuTiO2 and Ru/TiO2. (c,d) AIMD plots of Ru6/RuTiO2 and Ru6/TiO2. (e) Most stable Ru6/RuTiO2 and
Ru6/TiO2, where Ru atoms are labeled accordingly, and the charge density difference of Ru6/TiO2 and Ru6/RuTiO2, where the isosurface value is
set to 0.0035 e/Å3. Yellow and blue isosurfaces denote the charge accumulation and depletion regions, respectively.

CO2 methanation than Ni(111).18 Moioli and Züttel found sequence, an equally high CH4 selectivity is achievable as the
that 0.5% Ru/Al2O3 exhibited much higher catalyst perform- traditional TiO2-supported Ru catalyst.29 In addition, air
ance for CO2 methanation compared with Al2O3-supported Ni annealing features Ru/r-TiO2 with a much higher conversion
of the same loading.19 By using the same catalysts, González- rate. It is argued that in these samples, Ru/r-TiO2 is only
Velasco et al. verified that CH4 formation of Ru/Al2O3 is about partially reduced and thereby preserves an overlayer of
10 times than that of Ni/Al2O3.20 ruthenium oxide (RuOx). The RuOx overlayer grants the Ru
For supported transition-metal catalysts, metal oxide cluster better dispersion, which has been verified by character-
supports exert significant influence on the intermediate ization techniques. Even though the SMSI is put forward as a
adsorption, morphology of the active phase, and catalytic plausible explanation, more information is needed correspond-
performance. For Ru-based catalysts, TiO2 is arguable to be the ing to how the modified surface changes the chemical
most suitable support.21−28 It is proposed that the high CH4 properties (i.e., charge density redistribution and adsorption
selectivity of Ru/TiO2 derives from strong metal−support of intermediate) of the catalyst. Besides, experimental results
interactions (SMSIs), which is determined by the difference in offer little insight into the detailed mechanisms for each
the crystalline phase. For instance, higher CH4 selectivity of elementary step of methane synthesis, which serves as the
Ru/r-TiO2 (rutile) compared with Ru/a-TiO2 (anatase) is foundation for a deeper understanding of the significant role of
attributed to better interfacial lattice compatibility between the RuO2 overlayer. Herein, we employed the density
RuO2 and r-TiO2.24,25,27,28 Regardless of high CH4 selectivity, functional theory (DFT) to comparatively study the CO2
improvement of the conversion rate for Ru/TiO2 is still methanation on Ru/r-TiO2. Based on our calculation results,
challenging. CO is considered to be an important intermediate key issues are dealt with including (1) the electronic properties
involved in CO2 methanation.29−31 Zhang et al. employed a of catalysts with or without RuO2, (2) various modes of CO2
coreduction method which allows for the generation of TiO2−x chemisorption on both surface models, (3) most preferred
overlayers at Ru/r-TiO2 interfaces. Based on their experimental pathways involved in CO2 methanation on both surface
results, the SMSI is responsible for readily breaking the C−O models, and (4) how RuOx affects the catalyst performance of
bonds of both CO2 and CO, thereby promoting methane Ru/r-TiO2.
selectivity.32,33 Behm et al. also observed the same promotion
effect by encapsulation Ru in r-TiO2−x (defective r-TiO2).34 2. COMPUTATIONAL METHODS
Even though high selectivity is achievable from their studies, All the spin-polarized DFT calculations with a periodic
the conversion rate remains very low. It is therefore suspected boundary were carried out using the Vienna ab initio
that interaction between Ru and TiO2 is not sufficient for the simulation package.35,36 The valence electrons and core
optimal conversion rate. interactions were described through the Perdew−Burke−
Our attention has been drawn to a recent study by Wang et Ernzerhof exchange−correlation functional37 and the projector
al. For Ru/r-TiO2 annealed through both air and H2 gas in augmented wave.38 The electron configuration of Ru is
14655 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

[Kr]4d75s1, where (4p6, 4d7, 5s1) are explicitly treated as The activation barrier (Ea) and reaction energy (Er) of each
valence electrons. The electron configuration of Ti is elementary reaction step for the CO2 methanation over Run/
[Ar]3d24s2, where (3s2, 3p6, 3d2, 4s2) are explicitly treated as RuTiO2 and Run/TiO2 are defined as
valence electrons. An energy cutoff of 400 eV and a Gaussian Ea = E TS E IS (4)
smearing width of 0.05 eV were applied to models containing
all surfaces and intermediates. The dispersion-corrected vdW- Er = E FS E IS (5)
D3 functional was considered to evaluate the van der Waals
effect.39 It is argued that an appropriate description of Ti 3d where EIS, ETS, and EFS represent the total energies of the IS,
and Ru 4d orbitals, DFT + U, should be considered.40 TS, and FS, respectively.
However, our trial calculation indicates that the Hubbert U The free energies of all ISs, FSs, and TSs at 500 K were
corrections have little effect on energetics and do not affect our estimated by harmonic approximation. In addition, the free
analysis of the density of states (DOS). Therefore, we have not energies of the gaseous species were calculated as
considered Hubbert U in our studies. A (2 × 2 × 1) k-point
grid using the Monkhorst−Pack scheme was found to give Gg(T ) = Eelec + E ZPE + Cp dT TS
converged results.41 All the surfaces with or without
intermediates were optimized, and the unconstrained atoms where Cp is the heat capacity of the gaseous species as a
stopped optimization when the forces on them were below function of T, which is obtained from the NIST website.47
0.05 eV/Å. The transition state (TS) was located using the
climbing image nudged elastic band method.42 To check 3. RESULTS AND DISCUSSION
whether the predicted TS connects the initial state (IS) and 3.1. Determination of Run/RuTiO2 and Run /TiO2
final state (FS), the vibrational frequency analysis was also Sharing Similar Properties (Electronic and Structural)
employed to find out if there is only one imaginary frequency as Nanoparticles Employed in the Experiment. In
and whether it corresponds to the first-order saddle point in general, it is hardly approachable to precisely simulate the
the minimum-energy path. A supercell was constructed from supported nanoparticles employed in the experiment due to
the optimized primitive unit cell of rutile TiO2 (r-TiO2) with the existence of various isomers and little understanding
lattice parameters of a = 4.64 Å, b = 4.64 Å, and c = 2.96 Å. concerned with their bonding environments. Therefore, an
The (110) facet was selected because it is the most stable one efficient approach in a theoretical study is to determine the
among all the exposed planes of r-TiO2 and has been supported nanoclusters which display similar properties (both
confirmed by experiments as the exposed phase.28,43 Then, electronic and structural) or serve as the reactive sites for
the r-TiO2 surface (simply denoted as TiO2 unless mentioned catalysis.48 On the one hand, it is reported that small-sized Ru
otherwise) with a nine-layer slab was modeled by a (4 × 2) nanoparticles form as an epitaxial layer along with
supercell (a = 12.15 Å and b = 12.94 Å). As presented in RuTiO2((0001)Ru//(110)RuTiO2).29 On the other hand, it is
d

Figure 1a, RuOx/TiO2 (denoted as RuTiO2) was built from a proposed that well-dispersed Ru nanoparticles supporting on
heterojunction between a six-layer slab of the TiO2(110) facet TiO2 exhibit SMSIs and optimal catalysis performance even
of the same size and a three-layer slab of the RuO2(110) though the epitaxial layer is not observed.26−29 To simulate the
facet.44 During the structure optimization, the bottom three nanoparticles employed in the experiment, we devoted our
layers were kept fixed, while the others including the surface efforts to find if the Ru nanoclusters form as an epitaxial layer
and intermediates were allowed to relax. A 20 Å vacuum slab on RuTiO2 or exhibit strong interaction with TiO2. In this
was applied to avoid interaction between each adjacent slab. context, numerous isomers of the Run (n ≤ 8) cluster
The basic units for the Ru cluster supported on both TiO2 supported on both RuTiO2 and TiO2 have been constructed,
and RuTiO2 were determined by employing several important optimized, and compared with each other depending on their
parameters. Among them, the adsorption energy (Eads) was relative adsorption energy (Eads), from which the most stable
used to describe the interaction between the Ru cluster and ones are determined [corresponding optimized structures on
support of the individual Ru atom, which is defined as45 the perfect Run/RuTiO2 and Run/TiO2 (n = 1−8), as shown in
Figures S1 and S2, and the defective Run/RuTiO2 and Run/
Eads = (E Run + Esupport E Ru n /support)/n (1) TiO2 (n = 1−8) are shown in Figures S3 and S4].
As displayed in Figure 1a, there are two types of oxygen
where n is the total number of adsorbed Ru. ERun, Esupport, and
d
atoms uncovered in the RuO2 and TiO2 overlayers, wherein
ERun/support are the total energy of n number of Ru clusters O2c composes the outmost layer and displays a bridged
d

configuration and O3c remains in the same plane with either Ru


retaining the shape from the support, bare support (RuTiO2 or or Ti atoms. For the most stable RuTiO2- and TiO2-supported
TiO2) surface, and supported Run. The cohesive energy (Ecoh) Ru monomer, Ru binds to two O2c, and their respective Eads
reflects the clustering ability of the Ru cluster on the bare are −4.85 and −5.20 eV. When n becomes 2 and 3, Ru prefers
support, which is defined as45 bonding with O3c, and their Eads increases apparently, which is
Ecoh = [nE Ru1 + Esupport E Ru n /support ]/n similar to that of Run/a-TiO2.45 For the most preferred
(2) configuration of the Ru4 isomer, the cluster adopts a
tetrahedron-like configuration on both supports and the
where ERu1 is the total energy of a single Ru atom in the gas
d

increasing amplitude of Eads gradually reduces. As a matter of


phase. Besides, the Ru−Ru interaction energy (ERu−Ru) defines fact, the supported Ru4 cluster generally prefers a 3D
the metal−metal interaction of individual Ru atoms inside the configuration and is regarded as the turning point, in which
cluster, which is defined as45,46 metal−metal interaction rather than metal−support interaction
gradually becomes the dominant driving force. As expected,
E Ru Ru = Ecoh Eads (3) ERu−Ru is smaller than Eads for Ru3/TiO2 but becomes larger
14656 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Table 1. Atomic Bader Charges (|e|) for Ru/RuTiO2, Ru/TiO2 Surfaces, and CO2 Adsorption on Both Surfacesa
C Oa Ob CO2 (total) Ru (total)
Ru/RuTiO2 plain surface +2.54
RT_CO2_MS* +1.05 −0.94 −1.02 −0.91 +3.42
RT_CO2_LS* +1.02 −0.84 −1.04 −0.86 +3.44
RT_CO2_SS* +1.35 −0.91 −1.03 −0.59 +3.03
Ru/TiO2 plain surface +1.70
T_CO2_MS* +1.13 −1.01 −0.97 −0.85 +2.79
T_CO2_LS* +1.06 −0.99 −0.98 −0.91 +2.75
T_CO2_SS* +1.15 −0.88 −1.06 −0.79 +2.41
Ru(0001) plain surface 0
Ru_CO2_MS* +1.19 −1.02 −1.00 −0.83 +0.39
a
Refer to Table S1 for a detailed analysis of the atomic Bader charges (|e|) of individual Ru atoms from the supported Ru clusters.

than ERu−Ru for Ru4/TiO2. Even though ERu−Ru in Ru4/RuTiO2 Similar to Run/RuTiO2, these 2D clusters are more reactive
is still less than Eads, the difference is much smaller than the than Run/TiO2 (n < 6) owing to more interfacial area provided
former three (Figure S5). When n is greater than 5 (n = 6−8), to activate the carbon dioxide.29,51 Compared with others
nanoclusters lie planar and parallel to both RuTiO2 and TiO2 [Run/TiO2 (n ≥ 8)], a stronger MSI is observed for Ru6/TiO2
surfaces. and Ru7/TiO2 owing to their relatively lower ratio of Ecoh/Eads.
Similar to the SMSI, epitaxial layer formation of the metal For comparison, Ru6/TiO2 is selected as our surface model
cluster is affected not only by the metal−support interaction29 (simply denoted as Ru/TiO2 in the following discussions
but also by the metal−metal interaction.48,49 To quantitatively unless mentioned otherwise).
describe their relative dominance, the ratio of cohesive energy To evaluate the dynamic stability of both Ru/RiTiO2 and
and adsorption energy is used as a descriptor, in which a Ru/TiO2, NVT ensemble simulations of the ab initio molecular
smaller Ecoh/Eads indicates that the cluster shape is primarily dynamics (AIMD) are applied under 500 K with the time step
determined by the Ru−support interaction.50 As shown in set as 1 fs. As shown in Figure 1c,d, both surface models
Figure 1b, compared with Run/TiO2, a much lower ratio of remain stable without obvious distortion in configuration over
Ecoh/Eads is observed for all investigated Run/RuTiO2. Besides, a time period of 5 ps.
the ratio does not experience much variation when n ≥ 6. We further characterized the structural properties of both
Taken together, we reasoned that both of these factors Ru/RuTiO2 and Ru/TiO2. As displayed in Figure 1e, the
contribute to the epitaxial layer formation of Run (n = 6−8) on optimized structures of both supported Ru clusters are similar
RuTiO2((0001)Ru//(110)RuTiO2), as shown in Figure S1.
d
in shape. However, epitaxial formation makes the cluster shape
Limited by the lattice parameters of the unit cell in our of Ru6 supporting on RuTiO2 resemble the (0001) facet of the
study, an extended supercell (6 × 2) is further employed to ruthenium surface. The average bond distance of Ru−O3c is
investigate if the trend continues for Run/RuTiO2 (n ≥ 8). As 2.34 Å with little variations, which indicates that the cluster
shown in Figures S6 and S7, the ratio of Ecoh/Eads and exhibits intimate contact with the RuO2 overlayer. On the
structural observation (Figure S7) further prove that the contrary, individual Ru−O3c varies greatly in distance for Ru/
epitaxial layer could exist even in larger-sized clusters, which is TiO2, wherein Rua and Ruf are much closer to the support than
consistent with the experimental results.29 Thus, Run/RuTiO2 the other four. Since the epitaxial layer is unable to form, the
(n = 6−8) are all qualified to simulate supported Ru cluster surface is more protruding than that of Ru6/RuTiO2
nanoparticles in the experiment. Given that similar CO2 and thereby exhibits less intimacy with the support. This is also
configurations are slightly more stabilized on Ru6/RuTiO2 in agreement with the above results drawn from the Ecoh/Eads
(see Figure S9a), it is determined as our surface model (simply ratio. To explore the electronic interaction that accounts for
denoted as Ru/RuTiO2 in the following discussions unless the configurational variations, the Bader charge of the Ru
mentioned otherwise). cluster was calculated to characterize the charge transfer at the
Furthermore, the ratio of Ecoh/Eads is applied to study the interface. As listed in Table 1, the Ru clusters lose a total
growth pattern of Ru clusters on TiO2. As displayed in Figure amount of 2.54 and 1.70 |e| when supported on RuTiO2 and
1b, the ratio of Ecoh/Eads is almost the same for Ru6/TiO2 and TiO2, respectively. In light of these findings, it is concluded
Ru7/TiO2 but starts to increase when the cluster size becomes that Ru/RuTiO2 exhibits a stronger interaction between the
8. By modeling with a (6 × 2) supercell, we noticed that the cluster and support, which coincides with the experiment’s
increase of Ecoh/Eads is not only from Ru7/TiO2 to Ru8/TiO2 results.29
but also from Ru10/TiO2 to Ru11/TiO2. Hence, it is expected Similar to CeO2, TiO2 has great potential to generate oxygen
that the cluster shape of TiO2-supported Run will become vacancies (Ov), which might influence the shape of the
controlled by Ru−Ru interaction and prefer 3D configurations supported Ru clusters. Hence, Ru clusters mounting on both
as long as the modeling clusters are large enough, which defective RuTiO2 and TiO2 were also examined with one of
complies well with the experiment results.29 It is also noticed the O2c being removed. For the most stable configuration of
that the Ecoh/Eads values for these supported Ru clusters are Run/RuTiO2−x (n = 1−8) shown in Figure S3, the preference
much lower than that of Run/a-TiO245 whereas higher than for 2D configuration is still fit for the cluster with a size as large
that of Run/RuTiO2. Based on these findings, it is concluded as 7. This also verifies the positive interaction between the
that Run exhibit SMSIs with r-TiO226−28 but disfavor the cluster and support. As the cluster keeps increasing, the 3D
formation of an epitaxial layer on it.29 As shown in Figure S8, configuration becomes dominant, suggesting that the cluster
the most stable configurations of Run (n ≥ 6) remain 2D. tends to agglomerate. Accordingly, a similar trend has been
14657 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 2. (a) Optimized structures of the most stable chemisorbed CO2 and their respective adsorption energy on the Ru/RuTiO2, Ru/TiO2, and
Ru(0001) surfaces. The white, red, green, and gray spheres represent titanium, oxygen, ruthenium, and carbon atoms, respectively. The yellow,
orange, and pink spheres represent oxygen from CO2 adsorbed on Ru/RuTiO2, Ru/TiO2, and Ru(0001), respectively. (b,c) −COHP curves of all
adsorbed bonds belonging to the most stable CO2 adsorption configuration on Ru/RuTiO2 (RT_CO2_MS*) and Ru/TiO2 (T_CO2_MS*),
where the Fermi level (EFermi) is set to zero.

observable for the TiO2−x supported Run (see Figure S4). It is participate in CO2 stabilization. The carbon atom is activated
therefore expected that Ru clusters prefer forming a planar by Rue and thereby stretches the bond length of C−Oa (1.38
layer (2D) on the perfect support, whereas a defective support Å) and C−Ob (1.28 Å) relative to inert CO2 (1.18 Å). The
promotes the 3D configuration of the Ru cluster in order to corresponding Eabs is −0.73 eV. The less stable one favors a
avoid surface tension. This is quite different from Aun/TiO2 bidentate structure, denoted as RT_CO2_LS*, in which C
where a 2D configuration is more preferred on the defective rather than Ob builds connectivity with Rub and thereby loses
surface according to the calculation results by Grunwaldt et one Ru−O bond. Despite being less competitive in adsorption
al.52 Both findings further prove the significant role of the energy, the C−Oa bond distance is prolonged apparently,
support in determining the shape of the metal cluster. indicating that it is more activated than the C−Ob bond. The
3.2. CO2 Adsorption on Ru/RuTiO2 and Ru/TiO2. To configuration being the second less stable is RT_CO2_SS*,
understand how the difference in the support affects the where only one Ru−O bond is formed between CO2 and the
catalysis performance, we started by comparatively examining Ru cluster. Owing to the lowest binding strength, bond lengths
various chemisorption modes of CO2 on Ru/RuTiO2 and Ru/ of both C−O from adsorbed CO2 decrease.
TiO2, the optimized configurations of which are displayed in Besides, we also examined the CO2 chemisorption modes on
Figures 2a and S9b,c. The structural parameters related to the Ru/TiO2. As expected, tridentate T_CO2_MS* is the most
C−Oa, C−Ob, and Oa−C−Ob angles are summarized in Table stable configuration adsorbed on Ru/TiO2 owing to the lowest
2. In addition, the CO2 adsorption modes on Ru(0001) are adsorption energy (−0.46 eV). Three atoms from the cluster,
also studied for crosswise comparison. including Rua, Rub, and Ruf, are involved in building
3.2.1. Structural Analysis. For Ru/RuTiO2, the most stable connections with the oxygen atoms from adsorbed CO2.
adsorption mode is denoted as RT_CO2_MS*, wherein three With respect to the less stable one T_CO2_LS*, also being in
Ru−O bonds including Ru a−Oa, Ruf−Oa, and Rub−Ob a tridentate pattern, three Ru−O bonds (including Rub−Ob,
Ruc−Oa, and Rud−Oa) are found. Despite having a similar
Table 2. Parameters Related to C−Oa and C−Ob Bond configuration to T_CO2_MS*, it is slightly higher in
Distances (Å) and the Oa−C−Ob Angle (deg) of Different adsorption energy due to the extra deformation energy
CO2 Chemisorption Modes on Ru/RuTiO2, Ru/TiO2, and required to release the lattice strain from the smaller grid
Ru(0001) Surfaces composed of Rub, Ruc, Rud, and Rue. Compared with the other
two adsorption modes, T_CO2_SS* has lower binding
adsorption mode C−Oa C−Ob ∠Oa−C−Ob strength. As for Ru(0001), tridentate Ru_CO2_MS* is found
Ru/RuTiO2 RT_CO2_MS* 1.38 1.28 121.0 as the most stable configuration, which is much higher in Eabs
RT_CO2_LS* 1.50 1.22 118.9 than its counterparts adsorbed on the other two surface models
RT_CO2_SS* 1.30 1.23 131.6 (Figure 2a).
Ru/TiO2 T_CO2_MS* 1.34 1.30 121.9 3.2.2. Electronic Structure Analysis. In addition to
T_CO2_LS* 1.39 1.29 120.0 structural analysis, the Bader charge analysis is employed to
T_CO2_SS* 1.46 1.21 120.0 elucidate the electronic factors that determine the mutual
Ru(0001) Ru_CO2_MS* 1.30 1.30 123.0 effects between the chemisorbed CO2 and three surface
14658 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 3. Potential energy diagram of the most favorable pathway for CO2 methanation on Ru/RuTiO2. Refer to Table S2 for detailed information
about Ea and Er of the specific elementary step and Figure S11 for the optimized structures of IS, TS, and FS involved in this pathway.

Figure 4. Potential energy diagram of the most favorable pathway for CO2 methanation on Ru/TiO2. Refer to Table S3 for detailed information
about Ea and Er of the specific elementary step, and Figure S14 for the optimized structures of IS, TS, and FS involved in this pathway.

models. As summarized in Table 1, the overall Bader charge of energy compared with that of Ru/TiO2. To elucidate the
all adsorbed carbon dioxide is negative, indicating that the reason behind this, electronic configurations of RT_CO2_MS*
direction of charge transfer is from the Ru cluster to CO2. For and T_CO2_MS* are further investigated via the crystal
Ru/RuTiO2 and Ru/TiO2, electron donations are more orbital Hamilton population (COHP), which describes the
localized around Ru atoms which are bound to CO2 (refer electronic contributions of the bonds taking place in the CO2−
to Table S1 for more information). For example, both Ru interaction (i.e., Oa−Rua, Oa−Ruf, Ob−Rub, and C−Rue).
tridentate RT_CO2_MS* and T_CO2_MS* coordinate to The relative bond strength is quantized through integrated
Rua, Rub, and Ruf, all of which contribute primarily to the COHP (ICOHP). As shown in Figure 2b,c, for both
overall electron loss. Therefore, it is speculated that the relative configurations, all Ru−C and Ru−O bonds contribute greatly
stability of the chemisorbed CO2 is affected by CO2- to stabilizing carbon dioxide. Compared with T_CO2_MS*,
coordinated Ru atoms, rather than the whole cluster. On the the values of ICOHP for RT_CO2_MS* are more negative,
contrary, electron donation from Ru(0001) is much more indicating stronger interaction between adsorbed CO2 and Ru/
delocalized throughout the entire surface plane, which is RuTiO2. In addition, small portions of contribution from
reflected by incompatible electron gain and loss between CO2- antibonding orbitals are observed right below the Fermi level
coordinated Ru atoms and CO2 (Table 1). This is also the for all Ru−O interactions, which negatively affects the O−Ru
cause of Ru_CO2*_MS* being the least stable among the binding strength. For a more stable CO2 chemisorption
three. configuration, the Ru cluster should be more positively charged
Furthermore, it is noticed that a similar configuration of to prevent electrons from filling in the antibonding orbitals of
CO2 adsorbed on Ru/RuTiO2 has relatively lower adsorption the Ru−O bond. Comparatively, a more stabilized config-
14659 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

uration RT_CO2_MS* is attributed to the higher Ru+/Ru OH* (S3′). Comparing both elementary steps suggests that
ratio of Ru/RuTiO2. the hydrogen atom bonding with CO* is the preferable one
3.3. Reaction Mechanisms of CO2 Methanation on because of its much more favorable barrier (0.57 eV) than that
Ru/RuTiO2 and Ru/TiO2. We continued to investigate the of OH* (1.34 eV). The other associative pathway with respect
influence of the RuO2 overlayer on the reaction mechanism of to COH* (S3″) generation is also invalidated owing to the
CO2 methanation. For an integrated CO2 methanation route reaction barrier of TS3″ being as high as 1.46 eV. As a result,
containing various elementary steps, each one is generally CHO* is the only possible associative product.
classified as follows: associative mechanism in which 3.3.1.3. CH2O* Formation via the Associative Mechanism.
adsorbates accept a hydrogen atom and dissociative mecha- Upon CHO* formation, further hydrogenation through either
nism where the C−O bonds of the intermediates are broken. carbon (S4, CH2O* formation) or oxygen atoms (S4″,
Before the investigation, it is worth noting that the Eads of the CHOH* formation) is plausible in theory. For the one
two atomic hydrogens are −1.63 eV for Ru/RuTiO2 and −1.72 involving CH2O* generation via O−H coordination, the C−H
eV for Ru/TiO2. Owing to the facile H2 dissociation, the distance is slightly reduced in TS4, which is proportional to a
isolated H atom is treated as the hydrogen source for all low activation barrier (0.22 eV). In the FS, the carbon atom
associative pathways. The logic of this section is presented as abandons the bridged configuration between Rud and Rue;
follows: as a start, we screened out the most preferred reaction instead, a Ruc−C linkage is established and it stabilizes the
mechanism from various ones on a stepwise basis, which is system only by −0.04 eV. On the contrary, Ob rather than C
determined by their respective activation energy barrier (Ea) couples with H* to produce CHOH*. However, with the
and reaction energy (Er) (summarized in Tables S2 and S3). reaction barrier and reaction energy substantially increasing to
Thereafter, the integrated route leading to methane synthesis is 1.25 and 0.89 eV, respectively, this pathway is neither
put forward for both Ru/RuTiO2 and Ru/TiO2 (the potential kinetically nor thermodynamically feasible. In addition to the
energy diagrams of the most favorable pathway for both surface associative pathways, formyl could also undergo a dissociative
models are displayed in Figures 3 and 4, while the other mechanism. CHO* first isomerizes to iso-CHO* (S4′-1) which
pathways are displayed in Figures S10 and S13). is stabilized by 0.50 eV. Thereafter, C−O bond cleavage (S4′-
3.3.1. Reaction Mechanisms for CO2 Methanation on Ru/ 2) is triggered with an intrinsic barrier of 0.75 eV.
RuTiO2. 3.3.1.1. First Elementary Step. Normally, the reaction Comparatively, the formation of CH2O* via the associative
mechanisms for the first elementary step of CO2 activation pathway is more preferred. However, the pathway involving
include dissociation [CO2* = CO* + O* (S1)] and association the C−O bond cleavage of iso-CHO* should also be
[CO2* + H* = HCOO* (S1′) and CO2* + H* = COOH* considered regardless of the higher barrier. Detailed discussion
(S1″)]. The dissociative mechanism starts from the chem- about their reaction mechanisms will be given in Section
isorbed RT_CO2_MS*, which breaks into CO* and O*. It is 3.3.1.5.
the most competitive pathway owing to the lowest barrier 3.3.1.4. Remaining Elementary Steps Leading to Methane
(0.21 eV) and the largest energy gap between the reactant and Formation. For the following elementary step, hydrogen
product (−1.06 eV). The associative pathway toward HCOO* donated by Rue causes the driving force to generate CH2OH*
occurs via donating the hydrogen atom from Rue to the carbon (S5). In the TS, hydrogenation features TS5 with reduced Ob−
of RT_CO2_MS*, where the barrier of TS1′ is 0.44 eV (see H distance (1.34 Å) and increased reaction barrier (0.84 eV).
Figure S12). The resulting bi-HCOO* stabilizes the system by For the oxygen-guided hydrogenation activity from either CO*2
only 0.16 eV and therefore shows less thermodynamic or CHO*, a relatively high barrier together with the
competitiveness. From the onset of formate species, it is endothermic nature of the reaction energy is required. We
found that the reverse reaction of HCOO* + H = H2COO* attributed this scenario to the passivation of the Ru−O bonds.
(S2′) is much lower in Ea than the forward one. Consequently, Furthermore, CH2OH* experiences bond cleavage between
HCOO* behaves as the “inspector”, which does not contribute C−O coordination and generates coexistent species [CH2 +
to methane formation. This finding complies well with the OH]* (S6). Despite the slightly destabilized final product, a
experimental results.29 Concerned with the other associative lower activation energy compared with the former one makes
mechanism, the coupling of CO2* with H comes along with this step kinetically favorable. The energy barrier of subsequent
exceedingly stretched Ob−Rub (2.32 Å), and our calculation OH* hydrogenation (S7) is 0.67 eV, which is followed by the
reveals that this leads to the unfavorable formation of trans- desorption of H2O* (S8). The desorption process is very
COOH due to the highest barrier among three (0.72 eV). The exothermic by 0.78 eV, which could be rationalized by CH*2
kinetic instability can be rationalized by substantial deforma- being transformed into a much more stable form of quadruple
tion energy produced to break the Ru−Ob bond. CO2 is more coordination when inert H2O is produced (refer to Table S4
conducive to undergo the dissociative pathway in the first for its Eads). Thereafter, quadruple coordination is readopted
elementary step which is the same as Ru(mono-layered)/γ- for CH*2 to accommodate another hydrogen atom which
Al2O3 and Ru-doped CeO2 and Ru/zeolite.53,54 elevates the system by 0.67 eV. The subsequent association is
3.3.1.2. CHO* Formation. Followed by CO2* dissociation, with respect to CH2* (S9) and CH3* (S10) both being lower in
either O* or CO* could be hydrogenated (S2 or S2″). A activation energy (0.42 eV for CH*2 and 0.23 eV for CH*3 ). For
comparison between both reveals that hydroxyl formation is the last elementary step, the second molecule of H2O* is
more favorable than the other one due to the benefit of lower produced (S11) and desorbed (S12) with a 0.59 eV reaction
activation and reaction energies. Furthermore, other pathways barrier and a 0.12 eV desorption energy after methane
including CO* dissociation (S2‴) and CO* protonation generation.
(S2‴′) are also reviewed, despite that neither one is kinetically 3.3.1.5. Methane Formation Led by the Dissociative
accessible. Therefore, it is concluded that coke formation is Mechanism. As shown in Figure S12, the dissociative
prevented on Ru/RuTiO2. To move on, we comparatively mechanism on Ru/RuTiO 2 starts with the H-assisted
studied the associative pathways involving either CO* (S3) or association of CH* to generate CH*2 (S5′) with the reaction
14660 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

barrier and energy being 0.29 and −0.84 eV, respectively. The As a result of facile OH* formation, the hydrogenation
increasing steric effect makes the subsequent H2O* generation product is reconsidered between CHO* (T3) and H2O*
(S6′) and desorption (S7′) very easy. For the next step, CH*2 (T3′). From one aspect, the pathway involving CO*
further accepts a hydrogen atom (S8′) and renders CH*3 hydrogenation has a reaction barrier of 0.90 eV. However,
binding to Rub. It is found that the reverse reaction of S8′ is the reaction energy is endothermic by 0.36 eV which promotes
barrierless even though the reaction energy for the forward one its reverse reaction. On the other hand, the TS for OH*
is only endothermic by 0.37 eV. Methane formation is realized hydrogenation is calculated to be slightly higher (0.92 eV).
through CH3* hydrogenation (S9′) with a reaction barrier of H2O* generation is followed by its desorption (T4″) with a
1.19 eV and a reaction energy of 0.55 eV. Owing to the higher 0.44 eV increase in reaction energy, so it is also feasible under
barrier and endothermic nature of S9′, it is concluded that the 500 K. Followed by T4″, CO* will continue to associate with
dissociative mechanism is less favorable. hydrogen like step T3. However, it turns out to be very
3.3.2. Reaction Mechanisms for CO2 Methanation on Ru/ unlikely from the aspect of both thermodynamics and kinetics.
TiO2. Next, we directed our attention to Ru/TiO2, from which Based on the above reasoning, the only possible pathway is T3.
all possible mechanisms involved in each elementary step for 3.3.2.3. Formation of [CH + O]* via the Dissociative
CO2 methanation are reviewed carefully and the most Mechanism. As a turning point, both associative and
preferred route is proposed. dissociative mechanisms are examined to find out which one
3.3.2.1. First Elementary Step. The first elementary step for is applicable to CHO* on Ru/TiO2. As for the former one
CO2 activation on Ru/TiO2 could lead to the generation of (T4′), the reaction energy is endothermic by 1.14 eV, so
[CO + O]*, HCOO*, and COOH*, which are classified as the CH2O* formation is prohibited. For this elementary step, TS4′
aforementioned mechanisms accordingly. To consider the cannot be located owing to the barrierless reverse reaction. We
dissociative mechanism (T1), we referred to the T_CO2_MS* also examined the dissociative pathway (T4), in which
which is the most stable among all investigated CO 2 preactivation involving isomerization into iso-CHO* is
chemisorption modes (Figure S9c). A nearly negligible barrier required, thereby reducing the adsorption energy by 0.48 eV.
and large energy gap between the initial and final step suggest Thereafter, dissociation into two separate fragment species
that this pathway is very competitive. A comparative study of (i.e., CH* and O*) occurs with only a 0.29 eV reaction barrier,
the isolated species between two surface models shows that the together with a −0.34 eV reaction energy. It is proposed that
adsorption energy of individual O* is higher on Ru/TiO2 than unstable CHO* is very active toward direct bond cleavage on
the other owing to the weaker Ru−O interaction. This causes a the flat Ru surface.55−58 Thus, dissociation is considered to be
slight increase in adsorption energy related to the FS (see the dominant reaction mechanism for formyl on Ru/TiO2,
which is much different from the other (Ru/RuTiO2).
Table S4 for details). On the other hand, HCOO* is produced
3.3.2.4. Remaining Elementary Steps Leading to Methane
as a result of the associative mechanism (T1′) with the
Formation. For the next elementary step, dissociative products
activation and reaction energy being 1.37 and 0.05 eV,
CH* and O* are compared with each other with respect to H-
respectively. As for the other associative product (T1″),
coupling. On the one hand, the CH* association (T5) leads to
COOH* is formed with the highest activation energy among
CH*2 with 0.55 eV corresponding to the reaction barrier of
the three. Taken together, both dissociative mechanisms are TS5. On the other hand, the hydrogenation of adsorbed O* is
prohibited kinetically. We also noticed that, compared with found to be unlikely because of the steric effect. Upon CH*2
Ru/RuTiO2, Ru/TiO2 has a higher association barrier but a formation, the cluster becomes more crowded, and therefore,
lower dissociation barrier. A detailed explanation of this H-coupling is only conducive to OH* (T6). Our results
interesting finding will be provided in Section 3.4. Never- suggest that it is a thermoneutral process which comes along
theless, a dissociative mechanism is more preferred to the with a reaction barrier of 1.09 eV. The subsequent H2O
associative one, which is the same as that for Ru/RuTiO2. desorption (T7) is endothermic by 0.40 eV. Further
3.3.2.2. CHO* Formation. When it comes to the next hydrogenation is related to both CH*2 (T8) and CH*3 (T9),
elementary step, either O* (T2) or CO* (T2′) could accept which are both much lower in barrier (0.31 eV for CH*2 and
hydrogen. On the one hand, O−H coupling requires slight 0.34 eV for CH3*). For the last two elementary steps, H-
shrinkage of their mutual distance in the TS2, which coupling is kinetically much more favorable for OH* (T11, Ea
corresponds to the reaction barrier of 0.67 eV. At the same = 0.91 eV) than O* (T10, Ea = 1.49 eV). The desorption of
time, with the FS being lower in adsorption energy by 0.62 eV, the second H2O (T12) is endothermic by 0.19 eV.
this pathway turns out to be much more exothermic than that 3.3.2.5. Methane Formation Led by the Associative
of Ru/RuTiO2 (S2). However, the higher adsorption energy of Mechanism. Despite unfavorable CH2O* formation, the H-
isolated OH* seems to fail to explain this (Table S4). Instead, assisted associative pathway originating from CHO* is also
we attributed it to the less electronic repulsion between CO* considered. As shown in Figure S15, CH2OH* is formed
and OH* owing to their remote distance. On the other hand, through the H-assisted association of CH2O* (T5′), and the
the activation energy for CO* hydrogenation is slightly lower. calculated reaction barrier is 1.47 eV, which is considerably
However, the thermoneutral nature of the reaction energy higher than S5 for Ru/RuTiO2. Thereafter, C−O bond
makes it less favorable at low temperatures. Therefore, the cleavage occurs and produces two fragments CH*2 and OH*
hydrogenation of OH* is more favorable, which is the same as (T6′) with TS6′ being only 0.44 eV. The examination of the
that for Ru/RuTiO2. In addition, direct CO* bond cleavage FS reveals that triple coordination is very unstable and thereby
(T2″) is energetically inaccessible, which contradicts the drives CH*2 to adopt a bridged configuration between Ruc and
proposed mechanism on the TiO2−x-encapsulated Ru cluster.33 Rud, and this accounts for the substantial decrease in reaction
This finding suggests that the difference in the metal−support energy. To move on, first, H2O* is generated from OH*
interaction (MSI) has a significant influence on determining hydrogenation (T7′) with the reaction barrier TS7′ being 0.73
the reaction mechanism for CO2 methanation. eV and the reaction energy being 0.33 eV. H2O desorption
14661 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 5. (a,b) Schematic representation of the most preferred pathway of CO2 methanation for Ru/RuTiO2 and Ru/TiO2. (c) Chart of energy
barriers for the RDS of Ru/RuTiO2 (associative and dissociative mechanisms) and Ru/TiO2 (associative and dissociative mechanisms).

Figure 6. (a) Comparisons of the reaction barrier (Ea) and reaction energy (Er) for Ru/RuTiO2 and Ru/TiO2 with respect to the dissociative
mechanism. (b) Charge density differences of CH3* on Ru/RuTiO2 and Ru/TiO2, where the isosurface value is set to 0.0009 e/Å3. The green and
brown isosurfaces denote the charge accumulation and depletion regions, respectively. (c) Comparisons of the reaction barrier (Ea) and reaction
energy (Er) for Ru/RuTiO2 and Ru/TiO2 with respect to the associative mechanism.

(T8′) comes along with a 0.80 eV energy released as a result of prohibited (T4′ and T5′), even though the energy barriers of
the steric effect. Methane is formed via the subsequent RDS for both reaction mechanisms are very close.
donation of a hydrogen atom to CH*2 (T9′) and CH*3 (T10′). 3.3.3.1. Comparison of the Dissociative Mechanism
The reaction barriers calculated for respective hydrogenation between Ru/RuTiO2 and Ru/TiO2. Furthermore, the dis-
are 0.13 and 0.70 eV. sociative mechanism starting from CHO* is comparatively
3.3.3. Comparison of the Reaction Mechanism between investigated on a stepwise basis between two surfaces, and the
Ru/RuTiO2 and Ru/TiO2. On the basis of the above analysis, corresponding energy barrier and reaction energy are
the reaction mechanism of CO2 methanation for two surface presented in Figure 6a. In summary, two factors limit the
models is schematically elucidated in Figure 5a,b. For Ru/ dissociative mechanism of Ru/RuTiO2. On the one hand,
RuTiO2, direct C−O cleavage is the dominant pathway in the compared with Ru/TiO2, CHO* dissociation on Ru/RuTiO2
first place, which is followed by consecutive hydrogenation of (S4′-2) is much less preferred. Isolated species analysis shown
both O* and CO*. After CHO* formation, H-assisted in Figure S16 indicates that CH* is too unstable to maintain
association to generate CH2O* is the most preferred pathway, the bridged configuration between Rub and Rue as the one
found in the FS of S4′-2 and thereby easily relaxes into a
which is followed by the formation of CH2OH*, CH*2 , CH*3 ,
different configuration. This leads to a considerable increase in
and methane. CH2O* hydrogenation is determined to be the
the adsorption energy of the FS [CH + O + OH]*. The kinetic
rate-determining step (RDS) with an activation barrier of 0.84 explanation of the unfavorable TS will be further given in
eV. Comparatively, the dissociative mechanism is less favorable Section 3.4. On the other hand, the hydrogenation of CH*3 to
with the RDS calculated to be 1.19 eV for S9 [CH3* + H* = generate methane on Ru/RuTiO2 is kinetically and thermo-
CH4(g)]. For Ru/TiO2, different reaction mechanisms control dynamically less favorable. For both surfaces, the stabilization
the most favorable pathway of CO2 methanation. To be of CH*3 mainly involves interaction between C and Rub.
specific, the dissociative mechanism is also dominant in the However, it is noted that one of the hydrogen atoms also
first elementary step, which is followed by the generation and interacts with Rua and contributes to the coordination of CH3*
isomerization of CHO*. From the onset of iso-CHO*, the with Ru/RuTiO2 (see Figure 6b). This could be further
dissociative mechanism is more favorable than the associative verified by Eads of the isolated CH*3 being more negative
one, which generates [CH + O]*. Thereafter, CH* accepts (Table S4). The H−Rua interaction greatly increases the
three hydrogen atoms to produce CH2, CH3*, and methane, stability of isolated CH3*, which makes it very retarded toward
respectively. On the contrary, the associative mechanism is less further hydrogenation. On the contrary, this phenomenon is
favorable because of two consecutive steps being kinetic not detected when it comes to Ru/TiO2 because of the uneven
14662 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Table 3. The d-Band Center (ε) of Ru Atoms on Ru/RuTiO2 and Ru/TiO2a


εa εb εc εd εe εf εavg
Ru/RuTiO2 without OH* −1.76 −1.72 −1.51 −1.72 −1.73 −1.48 −1.65
with OH* −1.78 −1.57 −1.98 −1.86 −1.79
Ru/TiO2 without OH* −1.48 −1.63 −1.35 −1.43 −1.52 −1.50 −1.48
with OH* −1.68 −1.47 −1.41 −1.36 −1.48
a
εa to εf represent the d-band centers of Rua to Ruf, respectively.

cluster surface (Rua−O3c is much shorter than Rub−O3c). As a study the dissociation or association reactivity of the
result, both the reaction barrier and reaction energy of S9′ are intermediate. For instance, closer proximity of the d-band
much higher than T9. Furthermore, we observed a similar center (ε) relative to the Fermi level (EFermi) forecasts a better
CH3* configuration on Run/RuTiO2 (n = 7 and 8), suggesting electron-donating nature of the substrate and thereby weakens
that the unfavorability of the dissociative mechanism is the bonds that are to be activated.61,62 The decomposition
universal (Figure S16). energy barrier therefore decreases. On the contrary, ε, which is
Even though the dissociative mechanism has been far from the Fermi level is an indicator of inferior ability to
determined to be the dominant pathway for Ru/TiO2, donate electrons and thereby favors bond formation such as
methane formation is greatly limited as a result of T10. Unlike C−H and O−H interactions.63−65 By far, this theory has been
S10′ for Ru/RuTiO2, the hydrogen atom is forced to move successfully employed to study CO2 dissociation on Al2O3-
away from O* and take a bridged configuration between Rub supported Ru clusters.66 Conceptually, the d-band center is
and Rue in order to avoid the steric effect. Despite a substantial expressed as
decrease in energy, the subsequent hydroxyl formation (T10)
+
initiated from such IS is both kinetically and thermodynami- d (E )E dE
cally prohibited. A similar configuration is also observed for = +
Ru8/TiO2. Hence, it is speculated to be one of the reasons d (E) dE
leading to a lower conversion rate of the Ru/r-TiO2 catalyst.29
3.3.3.2. Comparison of the Associative Mechanism where ρd represents the DOSs projected onto the d-orbital of
between Ru/RuTiO2 and Ru/TiO2. The associative mechanism individual Ru atoms (Rua to Ruf) from the supported cluster
from the onset of CHO* is also studied for both surfaces. As and E represents energy relative to EFermi. For the transition-
shown in Figure 6c, two unfavorable elementary steps are metal Ru consisting of both α and β spins, the overall d-band
proposed to account for the inhibition of this pathway on Ru/ center is the average of the d-band center for α and β spins,
TiO2. For the first one, both the barrier and reaction energy of respectively.
the H-assisted association of CHO* (T4′) are equally high, The calculated d-band centers for Rua to Ruf from both Ru/
which promotes the reverse reaction. Thermodynamically, we RuTiO2 and Ru/TiO2 are summarized in Table 3, along with
attributed this to the unstable isolated CH2O* (Table S4). For the corresponding projected DOS (pDOS) presented in
the second one, the calculated barrier for CH2O* hydro- Figures S17andS18. The average d-band center for the
genation (T5′) is as high as 1.47 eV, which further prevents the integrated Ru cluster is defined as εavg = (εa + εb + εc + εd +
associative mechanism from proceeding. A detailed explanation εe + εf)/6, wherein εa to εf refers to the d-band center of Rua to
of the kinetic reason for these two steps will be offered in the Ruf, respectively. As for Ru/RuTiO2, the d-band centers of Rua,
next section. Rub, Rud, and Rue are akin to each other, indicating that they
3.4. Electronic Analysis of the Difference in the are equivalent in the valence state. As for the other two, εc
Reaction Mechanisms for CO2 Methanation on Ru/ (−1.51) and εf (−1.48) are relatively closer to the Fermi level,
RuTiO2 and Ru/TiO2. As mentioned already, the reaction which is attributed to the low coordination of Ruc and Ruf. As
mechanism for CO2 methanation varies greatly from Ru/ a result, they slightly upshift the overall d-band center of the
RuTiO2 to Ru/TiO2. So far, explanations have been solely cluster (εavg) to −1.65. On the other hand, the d-band centers
given from the perspective of the intermediates (IS and FS) of the TiO2-supported cluster are quite different among Ru
and thereby lack insightful understanding concerned with the atoms with the lowest and highest one assigned for Rub and
TS. In light of this, further proof is required. Ruc, respectively. Consequently, the calculated εavg for the Ru
It is generally accepted that for an intermediate adsorbed on cluster is −1.48, which is much closer to EFermi than that of Ru/
a metal support, the d-orbital of the metal support with a RuTiO2. Considering that the Ru cluster is more positively
narrow bandwidth is very sensitive during adsorbate−metal charged on Ru/RuTiO2, it is expected to be the main reason
interaction via either accepting or donating electrons, which for the downshifted d-band center.
results in the splitting of adsorbate resonance into two parts Next, we paid attention to evaluate the influence introduced
(i.e., bonding orbital below the Fermi level and antibonding by the difference in the d-band center upon CO2 methanation
orbital above the Fermi level). Furthermore, strong interaction between two surface models. It is worth noting that the
also provides a driving force for the environmentally sensitive intermediate is very sensitive toward the reaction site, so
metal d-band to either downshift or upshift as a result of orbital consideration should be given to the average d-band center of
hybridization. As a parameter to characterize the ability to eject these areas rather than that of the integrated cluster.66 For the
electrons from the d-orbital of the metal to the antibonding initial elementary step of CO2 activation, it has been
orbital of the adsorbate, the d-band center proposed by mentioned that the dissociation barrier is higher on Ru/
Nørskov has been utilized to study the relative stability of RuTiO2, whereas the association barrier is higher on Ru/TiO2.
adsorbed molecules or atoms.59,60 In addition, they demon- For both surface models, Rua, Rub, Rue, and Ruf are the
strated that such a descriptor could be extensively applied to reactive sites, and the average d-band centers are −1.68 and
14663 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

−1.44 for Ru/RuTiO2 and Ru/TiO2, respectively. As a result, as an epitaxial layer along with RuTiO2 and exhibits SMSIs
more kinetic preference to carbonate and formate formation with TiO2 and therefore could be employed to simulate the
on Ru/RuTiO2 than to the other is ascribed to the downshift nanoparticles in the experiment. Compared with Ru/TiO2,
of the d-band center, which facilitates C−H and O−H bond more electrons are transferred from RuTiO2 to the Ru cluster,
formation. Nevertheless, we should keep in mind that which complies well with the surface characterization. This
dissociation is still the dominant mechanism for Ru/RuTiO2 makes individual Ru atoms in the cluster more positively
because of the reasons already mentioned above. charged and thereby facilitates CO 2 adsorption. The
For both surface models, the fourth elementary step is minimum-energy pathway analysis reveals that for both surface
considered as the turning point, for which either the associative models, CO2 undergoes direct C−O bond cleavage rather than
(CHO* + H* = CH2O*) or the dissociative pathway (CHO* the formate path in the first place, which is consistent with the
= CH* + O*) determines the fate of formyl. Previous results FTIR spectroscopy results. Furthermore, we also noticed the
have already elucidated that the relative stability of the final difference in the reaction mechanism from the onset of CHO*
product accounts for the difference in the reaction energy and generation, where Ru/RuTiO2 favors further hydrogenation
thereby determines the reaction mechanism. Otherwise, an toward CH2O* (an associative mechanism), whereas Ru/TiO2
electronic explanation should also be given with respect to the favors direct C−O bond cleavage after isomerization (a
reaction barrier. As listed in Table 3 (see Figures S19 and S20 dissociative mechanism). The preference of CHO* hydro-
in the Supporting Information for pDOS), coadsorbed OH* genation for Ru/RuTiO2 is attributed to much more stabilized
on Ru/RuTiO2 introduces little change toward εb and εc. CH2O* species, along with the downshift of the d-band center.
However, an apparent downshift of both εd and εe is detected On the contrary, CHO* dissociation is favored for Ru/TiO2
which pushes εavg (−1.79) to move away from EFermi. On the which is due to both the thermodynamic advantage of
other hand, Ru/TiO2 with the coadsorption of OH* has little generating [CH + O]* as well as the uplifted d-band center.
effect on εb and εd, whereas εc and εe either shift to a lower or Overall, our calculation results prove that RuO2 plays a very
higher region, respectively. The average d-band center for important role in determining the reaction mechanism of CO2
these four atomic Ru (−1.48) remains almost the same as εavg methanation for the Ru/r-TiO2 catalyst.
calculated from the integrated Ru cluster. As a result, the
downshifted d-band center could be put forward to explain the
low barrier of the associative mechanism on Ru/RuTiO2,

*
ASSOCIATED CONTENT
sı Supporting Information
whereas the upshifted d-band center could be used to explain The Supporting Information is available free of charge at
the low barrier of the dissociative mechanism on Ru/TiO2. https://pubs.acs.org/doi/10.1021/acscatal.2c04539.
Based on these results, it is obvious that the alternation of the
reaction mechanism from the onset of CHO* is not only Computational details of most stable structures of
influenced by the structural preference of the product but also perfect and defective Run/RuTiO2 and Run/TiO2,
driven by electronic factors. other less unfavorable reaction mechanisms for each
To continue, since OH* adsorption on Ru/RuTiO2 causes a elementary step, and the figures of pDOS (PDF)
downshift of the d-band center related to the reactive sites
(Rub, Ruc, Rud, and Rue), its influence on the reaction barrier
of CHO* hydrogenation is also studied. As shown in Figure
■ AUTHOR INFORMATION
Corresponding Authors
S21a, the presence of OH* substantially decreases both the Wei Lin − College of Chemistry, Fuzhou University, Fuzhou
activation energy and reaction energy of CHO* hydrogenation 350108, China; Fujian Provincial Key Laboratory of
compared with those without OH*. On the one hand, OH* Theoretical and Computational Chemistry, Xiamen
stabilizes CH2O* by releasing the surface tension of the cluster University, Xiamen 361005, China; orcid.org/0000-
and thereby decreasing the reaction energy. On the other hand, 0002-5046-4765; Email: wlin@fzu.edu.cn
the downshifted d-band center resulting from OH* adsorption Xin Lu − State Key Laboratory of Physical Chemistry of Solid
activates the formyl species and facilitates C−H bond Surfaces, College of Chemistry and Chemical Engineering and
formation. In comparison, the reaction mechanism in terms Fujian Provincial Key Laboratory of Theoretical and
of direct bond cleavage of CHO* on Ru/TiO2 with or without Computational Chemistry, Xiamen University, Xiamen
OH* is also reviewed. As shown in Figure S21b, the reaction 361005, China; orcid.org/0000-0003-4968-9462;
barriers for the two pathways are close to each other, Email: xinlu@xmu.edu.cn
suggesting that OH* adsorption has little effect on the
dissociative mechanism kinetically. This is further verified by Authors
the little fluctuation of the d-band center on Ru/TiO2 when Jie Yu − State Key Laboratory of Physical Chemistry of Solid
OH* is introduced. Furthermore, the d-band center could also Surfaces, College of Chemistry and Chemical Engineering,
be exploited to explain the way how CH*3 interacts with H* on Xiamen University, Xiamen 361005, China
Ru/RuTiO2 (displayed in Figure S22). As with Ruc (εc = Yabing Zeng − College of Chemistry, Fuzhou University,
−1.57), CH3* binding to Rud (εd = −1.98) has a much lower Fuzhou 350108, China
reaction barrier of hydrogenation, which results in facilitated Qirou Jin − State Key Laboratory of Physical Chemistry of
C−H bond formation. Solid Surfaces, College of Chemistry and Chemical
Engineering, Xiamen University, Xiamen 361005, China
4. CONCLUSIONS Complete contact information is available at:
In order to uncover the influence of the RuO2 overlayer on https://pubs.acs.org/10.1021/acscatal.2c04539
Ru/TiO2 with respect to CO2 methanation, the DFT is
employed to comparatively investigate the difference between Notes
Ru/RuTiO2 and Ru/TiO2. According to our results, Ru6 forms The authors declare no competing financial interest.
14664 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

■ ACKNOWLEDGMENTS
This work was financially supported by the National Natural
(18) Mohan, O.; Shambhawi, S.; Xu, R.; Lapkin, A. A.; Mushrif, S.
H. Investigating CO2 Methanation on Ni and Ru: DFT Assisted
Microkinetic Analysis. ChemCatChem 2021, 13, 2420−2433.
Science Foundation of China (92161117 and 21973014). (19) Moioli, E.; Züttel, A. A Model-Based Comparison of Ru and Ni
Catalysts for the Sabatier Reaction. Sustainable Energy Fuels 2020, 4,
■ REFERENCES
(1) Benson, E. E.; Kubiak, C. P.; Sathrum, A. J.; Smieja, J. M.
1396−1408.
(20) Quindimil, A.; De-La-Torre, U.; Pereda-Ayo, B.; Davó-
Quiñonero, A.; Bailón-García, E.; Lozano-Castelló, D.; González-
Electrocatalytic and Homogeneous Approaches to Conversion of CO2 Marcos, J. A.; Bueno-López, A.; González-Velasco, J. R. Effect of
to Liquid Fuels. Chem. Soc. Rev. 2009, 38, 89−99. Metal Loading on the CO2 Methanation: A Comparison between
(2) Brook, E. Windows on the Greenhouse. Nature 2008, 453, 291− Alumina Supported Ni and Ru Catalysts. Catal. Today 2020, 356,
292. 419−432.
(3) Rodriguez, J. A.; Evans, J.; Feria, L.; Vidal, A. B.; Liu, P.; (21) Aziz, M. A. A.; Jalil, A. A.; Triwahyono, S.; Ahmad, A. CO2
Nakamura, K.; Illas, F. CO2 Hydrogenation on Au/TiC, Cu/TiC, and Methanation over Heterogeneous Catalysts: Recent Progress and
Ni/TiC Catalysts: Production of CO, Methanol, and Methane. J. Future Prospects. Green Chem. 2015, 17, 2647−2663.
Catal. 2013, 307, 162−169. (22) You, Y. F.; Xu, C. H.; Xu, S. S.; Cao, S.; Wang, J. P.; Huang, Y.
(4) Ray, K.; Deo, G. A Potential Descriptor for the CO 2 B.; Shi, S. Q. Structural Characterization and Optical Property of
Hydrogenation to CH4 over Al2O3 Supported Ni and Ni-Based TiO2 Powders Prepared by the Sol-Gel Method. Ceram. Int. 2014, 40,
Alloy Catalysts. Appl. Catal., B 2017, 218, 525−537. 8659−8666.
(5) Wang, L.; Wang, L.; Zhang, J.; Liu, X.; Wang, H.; Zhang, W.; (23) Wang, F.; Zhang, S.; Li, C.; Liu, J.; He, S.; Zhao, Y.; Yan, H.;
Yang, Q.; Ma, J.; Dong, X.; Yoo, S. J. Selective Hydrogenation of CO2 Wei, M.; Evans, D. G.; Duan, X. Catalytic Behavior of Supported Ru
to Ethanol over Cobalt Catalysts. Angew. Chem., Int. Ed. 2018, 57, Nanoparticles on the (101) and (001) Facets of Anatase TiO2. RSC
6104−6108. Adv. 2014, 4, 10834−10840.
(6) He, Z.; Qian, Q.; Ma, J.; Meng, Q.; Zhou, H.; Song, J.; Liu, Z.; (24) Balaraju, M.; Rekha, V.; Devi, B. L. A. P.; Prasad, R. B. N.;
Han, B. Water-Enhanced Synthesis of Higher Alcohols from CO2 Prasad, P. S. S.; Lingaiah, N. Surface and Structural Properties of
Hydrogenation over a Pt/Co3O4 Catalyst under Milder Conditions. Titania-Supported Ru Catalysts for Hydrogenolysis of Glycerol. Appl.
Angew. Chem., Int. Ed. 2016, 55, 737−741. Catal., A 2010, 384, 107−114.
(7) Frusteri, F.; Migliori, M.; Cannilla, C.; Frusteri, L.; Catizzone, E.; (25) Abe, T.; Tanizawa, M.; Watanabe, K.; Taguchi, A. CO2
Aloise, A.; Giordano, G.; Bonura, G. Direct CO2-to-DME Hydro- Methanation Property of Ru Nanoparticle-Loaded TiO2 Prepared
genation Reaction: New Evidences of a Superior Behaviour of Fer- by a Polygonal Barrel-Sputtering Method. Energy Environ. Sci. 2009, 2,
Based Hybrid Systems to Obtain High DME Yield. J. CO2 Util. 2017, 315−321.
18, 353−361. (26) Lin, Q.; Liu, X. Y.; Jiang, Y.; Wang, Y.; Huang, Y.; Zhang, T.
(8) Dubey, A.; Nencini, L.; Fayzullin, R. R.; Nervi, C.; Crystal Phase Effects on the Structure and Performance of Ruthenium
Khusnutdinova, J. R. Bio-Inspired Mn(I) Complexes for the Nanoparticles for CO2 Hydrogenation. Catal. Sci. Technol. 2014, 4,
Hydrogenation of CO2 to Formate and Formamide. ACS Catal. 2058−2063.
2017, 7, 3864−3868. (27) Kim, A.; Debecker, D. P.; Devred, F.; Dubois, V.; Sanchez, C.;
(9) Sabatier, P.; Senderens, J. New Synthesis of Methane. C. R. Hebd. Sassoye, C. CO2 Methanation on Ru/TiO2 Catalysts: On the Effect of
Seances Acad. Sci. 1902, 134, 514−516. Mixing Anatase and Rutile TiO2 Supports. Appl. Catal., B 2018, 220,
(10) Sabatier, P.Catalysis in Organic Chemistry; D. Van Nostrand 615−625.
Company: New York, 1922. (28) Xu, J.; Su, X.; Duan, H.; Hou, B.; Lin, Q.; Liu, X.; Pan, X.; Pei,
(11) Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, G.; Geng, H.; Huang, Y. Influence of Pretreatment Temperature on
M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on Methana- Catalytic Performance of Rutile TiO2-Supported Ruthenium Catalyst
tion�from Fundamentals to Current Projects. Fuel 2016, 166, 276− in CO2 Methanation. J. Catal. 2016, 333, 227−237.
296. (29) Zhou, J.; Gao, Z.; Xiang, G.; Zhai, T.; Liu, Z.; Zhao, W.; Liang,
(12) Götz, M.; Lefebvre, J.; Mörs, F.; McDaniel Koch, A.; Graf, F.; X.; Wang, L. Interfacial Compatibility Critically Controls Ru/TiO2
Bajohr, S.; Reimert, R.; Kolb, T. Renewable Power-to-Gas: A Metal-Support Interaction Modes in CO2 Hydrogenation. Nat.
Technological and Economic Review. Renewable Energy 2016, 85, Commun. 2022, 13, 327.
1371−1390. (30) Loveless, B. T.; Buda, C.; Neurock, M.; Iglesia, E. CO
(13) Proaño, L.; Arellano-Treviño, M. A.; Farrauto, R. J.; Figueredo, Chemisorption and Dissociation at High Coverages During CO
M.; Jeong-Potter, C.; Cobo, M. Mechanistic Assessment of Dual Hydrogenation on Ru Catalysts. J. Am. Chem. Soc. 2013, 135, 6107−
Function Materials, Composed of Ru-Ni, Na2O/Al2O3 and Pt-Ni, 6121.
Na2O/Al2O3, for CO2 Capture and Methanation by in-Situ Drifts. (31) Li, X.; Lin, J.; Li, L.; Huang, Y.; Pan, X.; Collins, S. E.; Ren, Y.;
Appl. Surf. Sci. 2020, 533, 147469. Su, Y.; Kang, L.; Liu, X.; Zhou, Y.; Wang, H.; Wang, A.; Qiao, B.;
(14) Bonura, G.; Cordaro, M.; Cannilla, C.; Arena, F.; Frusteri, F. Wang, X.; Zhang, T. Controlling CO2 Hydrogenation Selectivity by
The Changing Nature of the Active Site of Cu-Zn-Zr Catalysts for the Metal-Supported Electron Transfer. Angew. Chem., Int. Ed. 2020, 59,
CO2 Hydrogenation Reaction to Methanol. Appl. Catal., B 2014, 152, 19983−19989.
152−161. (32) Zhang, Y.; Yan, W.; Qi, H.; Su, X.; Su, Y.; Liu, X.; Li, L.; Yang,
(15) Zhang, Q.; Zuo, Y. Z.; Han, M. H.; Wang, J. F.; Jin, Y.; Wei, F. X.; Huang, Y.; Zhang, T. Strong Metal-Support Interaction of Ru on
Long Carbon Nanotubes Intercrossed Cu/Zn/Al/Zr Catalyst for Co/ TiO2 Derived from the Co-reduction Mechanism of RuxTi1‑xO2
CO2 Hydrogenation to Methanol/Dimethyl Ether. Catal. Today Interphase. ACS Catal. 2022, 12, 1697−1705.
2010, 150, 55−60. (33) Zhang, Y.; Yang, X.; Yang, X.; Duan, H.; Qi, H.; Su, Y.; Liang,
(16) Younas, M.; Loong Kong, L.; Bashir, M. J. K.; Nadeem, H.; B.; Tao, H.; Liu, B.; Chen, D. Tuning Reactivity of Fischer−Tropsch
Shehzad, A.; Sethupathi, S. Recent Advancements, Fundamental Synthesis by Regulating TiOX Overlayer over Ru/ TiO2 Nano-
Challenges, and Opportunities in Catalytic Methanation of CO2. catalysts. Nat. Commun. 2020, 11, 3185.
Energy Fuels 2016, 30, 8815−8831. (34) Abdel-Mageed, A. M.; Wiese, K.; Parlinska-Wojtan, M.;
(17) Zhao, Z.; Jiang, Q.; Wang, Q.; Wang, M.; Zuo, J.; Chen, H.; Rabeah, J.; Brü ckner, A.; Behm, R. J. Encapsulation of Ru
Kuang, Q.; Xie, Z. Effect of Rutile Content on the Catalytic Nanoparticles: Modifying the Reactivity toward CO and CO2
Performance of Ru/TiO2 Catalyst for Low-Temperature CO2 Methanation on Highly Active Ru/TiO2 Catalysts. Appl. Catal., B
Methanation. ACS Sustainable Chem. Eng. 2021, 9, 14288−14296. 2020, 270, 118846.

14665 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666
ACS Catalysis pubs.acs.org/acscatalysis Research Article

(35) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab (54) Sharma, S.; Sravan Kumar, K. B.; Chandnani, Y. M.; Phani
Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Kumar, V. S.; Gangwar, B. P.; Singhal, A.; Deshpande, P. A.
Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169−11186. Mechanistic Insights into CO2 Methanation over Ru-Substituted
(36) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the CeO2. J. Phys. Chem. C 2016, 120, 14101−14112.
Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter (55) Shetty, S.; van Santen, R. A. CO Dissociation on Ru and Co
Mater. Phys. 1999, 59, 1758−1775. Surfaces: The Initial Step in the Fischer−Tropsch Synthesis. Catal.
(37) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Today 2011, 171, 168−173.
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (56) Ciobica, I. M.; van Santen, R. A. Carbon Monoxide
(38) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B: Dissociation on Planar and Stepped Ru(0001) Surfaces. J. Phys.
Condens. Matter Mater. Phys. 1994, 50, 17953−17979. Chem. B 2003, 107, 3808−3812.
(39) Lee, K.; Murray, É . D.; Kong, L.; Lundqvist, B. I.; Langreth, D. (57) Morgan, G. A.; Sorescu, D. C.; Zubkov, T.; Yates, J. T. The
C. Higher-Accuracy Van Der Waals Density Functional. Phys. Rev. B: Formation and Stability of Adsorbed Formyl as a Possible
Condens. Matter Mater. Phys. 2010, 82, 081101. Intermediate in Fischer-Tropsch Chemistry on Ruthenium. J. Phys.
(40) Fabris, S.; de Gironcoli, S.; Baroni, S.; Vicario, G.; Balducci, G. Chem. B 2004, 108, 3614−3624.
Reply to “Comment on ‘Taming Multiple Valency with Density (58) Inderwildi, O. R.; Jenkins, S. J.; King, D. A. Mechanistic Studies
Functionals: A Case Study of Defective Ceria’”. Phys. Rev. B: Condens. of Hydrocarbon Combustion and Synthesis on Noble Metals. Angew.
Matter Mater. Phys. 2005, 72, 237102. Chem., Int. Ed. 2008, 47, 5253−5255.
(41) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone (59) Medford, A. J.; Shi, C.; Hoffmann, M. J.; Lausche, A. C.;
Integrations. Phys. Rev. B: Solid State 1976, 13, 5188−5192. Fitzgibbon, S. R.; Bligaard, T.; Nørskov, J. K. CATMAP: A Software
(42) Mills, G.; Jónsson, H. Quantum and Thermal Effects in H2 Package for Descriptor-Based Microkinetic Mapping of Catalytic
Dissociative Adsorption: Evaluation of Free Energy Barriers in Trends. Catal. Lett. 2015, 145, 794−807.
Multidimensional Quantum Systems. Phys. Rev. Lett. 1994, 72, (60) Hammer, B.; Nørskov, J. K. Electronic Factors Determining the
1124−1127. Reactivity of Metal Surfaces. Surf. Sci. 1995, 343, 211−220.
(43) Perron, H.; Domain, C.; Roques, J.; Drot, R.; Simoni, E.; (61) Mavrikakis, M.; Hammer, B.; Nørskov, J. K. Effect of Strain on
Catalette, H. Optimisation of Accurate Rutile TiO2 (110), (100), the Reactivity of Metal Surfaces. Phys. Rev. Lett. 1998, 81, 2819−2822.
(101) and (001) Surface Models from Periodic DFT Calculations. (62) Pallassana, V.; Neurock, M. Electronic Factors Governing
Theor. Chem. Acc. 2007, 117, 565−574. Ethylene Hydrogenation and Dehydrogenation Activity of Pseudo-
(44) Yang, C.; Zhao, Z. Y. Investigation of Energy Band Alignments morphic PdML/Re(0001), PdML/Ru(0001), Pd(111), and PdML/
and Interfacial Properties of Rutile NMO2/TiO2 (Nm = Ru, Rh, Os, Au(111) Surfaces. J. Catal. 2000, 191, 301−317.
and Ir) by First-Principles Calculations. Phys. Chem. Chem. Phys. (63) Xing, B.; Wang, G. C. Insight into the General Rule for the
2017, 19, 29583−29593. Activation of the X−H Bonds (X = C, N, O, S) Induced by
(45) Zhang, S. T.; Li, C. M.; Yan, H.; Wei, M.; Evans, D. G.; Duan, Chemisorbed Oxygen Atoms. Phys. Chem. Chem. Phys. 2014, 16,
X. Density Functional Theory Study on the Metal-Support Interaction 2621−2629.
between Ru Cluster and Anatase TiO2(101) Surface. J. Phys. Chem. C (64) Zhang, J.; Jin, H.; Sullivan, M. B.; Lim, F. C. H.; Wu, P. Study
2014, 118, 3514−3522. of Pd−Au Bimetallic Catalysts for CO Oxidation Reaction by DFT
(46) Zhang, J.; Zhang, M.; Han, Y.; Li, W.; Meng, X.; Zong, B. Calculations. Phys. Chem. Chem. Phys. 2009, 11, 1441−1446.
Nucleation and Growth of Palladium Clusters on Anatase TiO2(101) (65) Huo, C. F.; Li, Y. W.; Wang, J.; Jiao, H. Insight into CH4
Surface: A First Principle Study. J. Phys. Chem. C 2008, 112, 19506− Formation in Iron-Catalyzed Fischer−Tropsch Synthesis. J. Am.
19515. Chem. Soc. 2009, 131, 14713−14721.
(47) Afeefy, H. Y.; Liebman, J. F.; Stein, S. E.Neutral (66) Yang, J.; Zhao, X.; Bu, S.; Fan, W. Theoretical Insights into the
Thermochemical Data. In NIST Chemsitry WebBook; Linstorom, P. Role of Metal−Support Interactions of an Al2O3-Supported Ru4
J., Mallard, W. G., Eds.; NIST Standard Reference Database Number Cluster in CO2 Dissociation. J. Phys. Chem. C 2018, 122, 17287−
69; National Institute of Standards and Technology: Gaithersburg, 17300.
MD. http://webbook.nist.giv (retrieved October 13, 2022).
(48) Tao, H.; Li, Y.; Cai, X.; Zhou, H.; Li, Y.; Lin, W.; Huang, S.;
Ding, K.; Chen, W.; Zhang, Y. What Is the Best Size of Subnanometer
Copper Clusters for CO2 Conversion to Methanol at Cu/TiO2
Interfaces? A Density Functional Theory Study. J. Phys. Chem. C
2019, 123, 24118−24132.
(49) Naya, S.; Akita, A.; Morita, Y.; Fujishima, M.; Tada, H.
Crystallographic Interface Control of the Plasmonic Photocatalyst
Consisting of Gold Nanoparticles and Titanium(IV) Oxide. Chem. Sci.
2022, 13, 12340−12347.
(50) Engel, J.; Francis, S.; Roldan, A. The influence of support
materials on the structural and electronic properties of gold
nanoparticles�a DFT study. Phys. Chem. Chem. Phys. 2019, 21,
19011−19025.
(51) Hongmanorom, P.; Ashok, J.; Chirawatkul, P.; Kawi, S.
Interfacial Synergistic Catalysis over Ni Nanoparticles Encapsulated in
Mesoporous Ceriaa for CO2 Methanation. Appl. Catal., B 2021, 297,
120454.
(52) Lopez, N.; Nørskov, J. K.; Janssens, T. V. W.; Carlsson, A.;
Puig-Molina, A.; Clausen, B. S.; Grunwaldt, J. -D. The Adhesion and
Shape of Nanosized Au Particles in a Au/TiO2 catalyst. J. Catal. 2004,
225, 86−94.
(53) Yan, Y.; Wang, Q.; Jiang, C.; Yao, Y.; Lu, D.; Zheng, J.; Dai, Y.;
Wang, H.; Yang, Y. Ru/Al2O3 Catalyzed CO2 Hydrogenation:
Oxygen-Exchange on Metal-Support Interfaces. J. Catal. 2018, 367,
194−205.

14666 https://doi.org/10.1021/acscatal.2c04539
ACS Catal. 2022, 12, 14654−14666

You might also like