Next Article in Journal
Progress and Recent Trends in the Application of Nanoparticles as Low Carbon Fuel Additives—A State of the Art Review
Next Article in Special Issue
A Novel Turn-On Fluorescent Sensor Based on Sulfur Quantum Dots and MnO2 Nanosheet Architectures for Detection of Hydrazine
Previous Article in Journal
Intracellular Trafficking and Distribution of Cd and InP Quantum Dots in HeLa and ML-1 Thyroid Cancer Cells
Previous Article in Special Issue
The Effect of Surface Hydroxyls on the Humidity-Sensitive Properties of LiCl-Doped ZnSn(OH)6 Sphere-Based Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the CH4 Conversion over Metal Dimers from First Principles

1
School of Physical Science and Technology, Inner Mongolia University, Hohhot 010021, China
2
Key Laboratory of Photonic and Electronic Bandgap Materials, College of Chemistry and Chemical Engineering, Ministry of Education, Harbin Normal University, Harbin 150025, China
3
Department of Chemistry, University of Puerto Rico, Rio Piedras Campus, San Juan, PR 00931, USA
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(9), 1518; https://doi.org/10.3390/nano12091518
Submission received: 26 March 2022 / Revised: 26 April 2022 / Accepted: 27 April 2022 / Published: 29 April 2022
(This article belongs to the Special Issue Nanostructures for Surfaces, Catalysis and Sensing)

Abstract

:
Inspired by the advantages of bi-atom catalysts and recent exciting progresses of nanozymes, by means of density functional theory (DFT) computations, we explored the potential of metal dimers embedded in phthalocyanine monolayers (M2-Pc), which mimics the binuclear centers of methane monooxygenase, as catalysts for methane conversion using H2O2 as an oxidant. In total, 26 transition metal (from group IB to VIIIB) and four main group metal (M = Al, Ga, Sn and Bi) dimers were considered, and two methane conversion routes, namely *O-assisted and *OH-assisted mechanisms were systematically studied. The results show that methane conversion proceeds via an *OH-assisted mechanism on the Ti2-Pc, Zr2-Pc and Ta2-Pc, a combination of *O- and *OH-assisted mechanism on the surface of Sc2-Pc, respectively. Our theoretical work may provide impetus to developing new catalysts for methane conversion and help stimulate further studies on metal dimer catalysts for other catalytic reactions.

1. Introduction

Global warming is gaining increasing concern worldwide. Greenhouse gases include carbon dioxide, methane, nitrous oxides, and other gases. According to EPA, carbon dioxide accounted for ca. 80% of all greenhouse gas emissions from human activities in the USA in 2019. Though methane has a lower emission (10%), it is also a major greenhouse gas since its greenhouse effect is 21 to 23 times that of carbon dioxide [1]. Therefore, the effective conversion of methane into value-added chemicals (instead of direct burning) is of both environmental and commercial importance [2,3,4].
The direct conversion of methane mainly associates with the high C−H bond strength (~434 kJ/mol) in the non-polar and highly symmetric methane [5,6,7,8]. In traditional industries, methane is first converted into syngas, then transferred to liquid hydrocarbons by Fischer–Tropsch process, which not only causes waste of resources, but also requires a high equipment maintenance cost [9,10,11,12,13,14]. Biological enzyme catalysis may be a good alternative since it has the advantages of high product selectivity and mild reaction conditions: compared with the current industrial process, the direct conversion of methane to methanol that occurs in the methane monooxygenase (MMO) from Methylococcus capsulatus is much more efficient [15,16,17,18]. Due to their high catalytic efficiency, biological enzymes can greatly increase the rate of chemical reactions, saving time and cost. Unfortunately, structural instability and sensitivity to the environment greatly limit their performance in industrial applications. A promising way to conquer such challenges is to mimic the MMO with the binuclear active sites [19,20,21]. Zeolites such as ZSM-5 are able to form stable binuclear centers (diiron or dicopper) in similar enzymes, and exhibit unprecedented high performance in methane conversion, but their catalytic mechanism and structure–property relationship remain unclear. The direct conversion of methane at room-temperature and atmospheric pressure is still an unsolved but high-rewarding challenge [22,23].
Nanozymes are a class of nanomaterials with unique enzyme-like properties, which have very similar active sites and catalytic mechanisms to biological enzymes. The first nanozymes were discovered in 2007, since then more than 300 nanomaterials have been found to have enzymatic activity [24,25,26]. Since “single-atom catalysis” was proposed in 2011, the concept of single-atom nanozymes (SAzymes) has also emerged as a research hotspot [27,28,29,30]. Compared with single-atom catalysts, bi-atom catalysts may possess improved catalytic performance [31,32]. For example, Yan et al. showed that Pt dimers embedded in graphene have 17-fold and 45-fold higher catalytic activity for the hydrolytic dehydrogenation of ammonia borane than its corresponding single-atom and nanoparticle counterparts [33]. Li et al. demonstrated that Cu dimers supported on C2N layers exhibited superior performance for CO oxidation compared to Cu1@C2N [34], and showed excellent performance with a small confinement potential of −0.23 V for electrochemical CO2 reduction [35]. However, to date, few studies have been reported on the catalytic performance of supported metal dimers for methane conversion. Inspired by the advantages of nanozymes and bi-atom catalysts, we designed a series of supported metal dimer catalysts for methane conversion by mimicking the binuclear centers in biological enzymes based on density functional theory computations.
The two-dimensional (2D) phthalocyanine-based (Pc) catalysts have a high surface area to volume ratio, abundant binding sites for anchoring metal atoms and the ability to prevent these metal atoms from aggregating into clusters. In 2011, Abel et al. [36] successfully prepared FePc complex and characterized the samples using scanning tunneling microscope (STM) at room temperature. Later on, Matsushita et al. [37] synthesized a rectangular phthalocyanine with two adjacent transition metal sites. Since Pc and transition metals are of low-cost, environmentally benign, more readily available than precious metals, the Pc-supported transition metal catalysts can be produced in a low-cost manner. Note that DFT calculations have been widely used to provide guidance in conversion/bonding/adsorption of molecules/clusters and reactions [38,39]. Here, first-principles calculations were conducted to explore the potential of all the 3d, 4d, and 5d non-toxic transition metals and the four main group metal (M = Al, Ga, Sn and Bi) dimers supported on the Pc (M2-Pc) for methane conversion.

2. Computational Methods

All the computations were carried out by spin-polarized density functional theory (DFT) calculations including van der Waals (vdW) corrections (DFT-D2) [40], as implemented in Vienna Ab initio Simulation Package (VASP) using the projector augmented wave (PAW) method [41]. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional was adopted [42]. The energy cutoff for the plane-wave basis set was chosen as 550 eV, the systemic energy tolerance and the remaining total force were set as 1 × 10−5 eV and 0.01 eV Å−1, respectively. The Brillouin zone was sampled with a 5 × 5 × 1 k-points grid of the Monkhorst–Pack scheme [43] for geometry optimization, and a denser k-mesh of 15 × 15 × 1 for electronic structure computations. To avoid interactions between periodic images, a vacuum space of 15 Å was used in the perpendicular direction of the 2D layer. The reaction energy barriers were estimated using the climbing-image nudged elastic band (CI-NEB) method [44,45], and the transition states were obtained by relaxing the force below 0.05 eV/Å. The entropic effects were not included in estimating reaction barriers. The binding energy ( E b ) of a metal atom was computed from the following equation.
E b = E t o t E s u b 2 E M / 2
where E t o t ,   E s u b , and E M represent the total energy of the complex of substrate and metal atoms, the energy of the substrate, and the energy of a free M atom, respectively. According to this definition, a more negative Eb value indicates a higher thermodynamic stability. The adsorption energy ( E a d s ) of an adsorbate was computed according to the following equation:
E a d s = E t o t E c a t E a d s o r b a t e
where the E t o t is the total energy of an adsorbate adsorbed on the catalyst, E c a t and E a d s o r b a t e represent the energies of the catalyst and a free adsorbate, respectively. The reaction energy ( E r e a ) and activity energy barrier ( E a c t ) were calculated using the following expressions:
E r e c = E F S E I S
E a c t = E T S E I S
in which E F S , E I S , and E T S denote the energies of the final, initial, and transition states, respectively. The reaction mechanism can also be effectively modeled by ab initio molecular dynamics (AIMD) simulations at specific temperatures [46].

3. Results and Discussion

This work describes our efforts to study the catalytic activity of M2-Pc on the conversion of methane. In this paper, we use H2O2 as the oxidant, because its reaction by-product is only water, which is a green catalyst [47]. Moreover, H2O2 has been widely used as an oxidant to study the conversion of methane [48,49,50,51,52]. The CH4 oxidation with H2O2 via both *OH- and *O-assisted mechanisms was investigated in detail [48].

3.1. Geometric Structure and Stability of M2-Pc

First, the geometric structure of the Pc monolayer was optimized, and the lattice parameters a and b in the Pc monolayer of 14.13 Å were used. As shown in Figure S1a, the unique cavity structure of Pc can provide ideal anchoring sites for the metal atoms to be connected, to four isoindole rings, preventing their migrating and aggregating. The computed energy band gap of Pc sheet is 0.94 eV (in Figure S1b,c). The geometric structures and related information of the optimized M2-Pc are shown in Figure S2 and Table S1. Obviously, due to the different radii of the metal atoms, the structures of M2-Pc are slightly different. The anchored metal atoms with smaller atomic radii form an in-plane configuration in the Pc cavity (Al, Mn, Fe, Co, Ni, Cu, Ga, Ru, Rh, Re, Os, and Ir), while others with larger atomic radii are pulled out of the Pc plane and lead to a buckled structure.
To confirm the stability of metal dimers embedded in the Pc sheet, the binding energy ( E b ) was calculated. Meanwhile, the corresponding metal bulk cohesive energy ( E b u l k ) were compared (Table S1), which are less negative than E b , indicating that the interaction between the metal atoms and Pc monolayer is very strong, i.e., the anchoring of metal dimers on the Pc has strong coupling and good stability. We also performed AIMD simulations of the Ag2-Pc monolayer, whose binding energy is the least favorable (−4.13 eV) among the considered systems (−13.47~−4.13 eV), and found that the monolayer structure was well kept during 5 ps’s annealing at 300, 800, 1000, 1300, and 1500 K, respectively, and bond breakage occurred at 1500 K (Figure S3). Therefore, all the models in our work have high thermal stability, and in the next section, we will study the catalytic mechanism for methane conversion on these M2-Pc catalysts.

3.2. Decomposition of H2O2 on M2-Pc

Since the adsorption and dissociation of oxidants are important in methane oxidation, we considered both side-on and end-on configurations of the H2O2 adsorption (Figure S4) on the examined 30 M2-Pc, including 26 transition metals from IB to group VIIIB and four main group metals (M = Al, Ga, Sn and Bi).
Upon adsorption, H2O2 will spontaneously decompose on the surface of the nine M2-Pc (M = Sc, Ti, V, Y, Zr, Nb, Hf, Ta, and W), all of which are highly exothermic (as shown in Table S2). As shown in Figure 1, H2O2 can be dissociated when the electronic state of the embedded atoms is a semi-occupied state. Among them, a H2O2 is spontaneously decompose into a H2O and an adsorbed oxygen ( H 2 O 2   O + H 2 O ) on the Nb2-Pc with the energy release of −6.08 eV, and into two adsorbed hydroxy groups ( H 2 O 2 2 O H ) on the five M2-Pc (M = Ti, V, Y, Hf, and Ta with the exothermic energy of −7.51, −6.70, −5.03, −8.71, and −9.04 eV, respectively), while into either a water molecule and an atomic oxygen (by releasing the heat of −3.75, −6.29, and −4.78 eV, respectively), or two OH species (by releasing the heat of −5.19, −8.16, and −7.41 eV, respectively) on the Sc2-Pc, Zr2-Pc, and W2-Pc. All the surface-adsorbed oxo species (*O) occupy the bridge position of the metal dimers. We also examined the magnetic moments and Bader charges on the metal atoms of these nine M2-Pc monolayers (M = Sc, Ti, V, Y, Zr, Nb, Hf, Ta, W) (Table S3), among which the Sc2-Pc, Ta2-Pc, Y2-Pc, and W2-Pc have spin states in singlet, the Nb2-Pc, Ti2-Pc, Zr2-Pc, and Hf2-Pc are triplet, while V2-Pc is in quintet spin state. According to the dissociation structure and energy, H2O2 dissociation on these nine catalysts (Figure S5 and Table S2) are much more exothermic than the previously reported dissociation reaction of H2O2 on Pd(111) and Au/Pd(111) surfaces (−1.76 and −1.58 eV for two adsorbed hydroxy groups, and −2.27 and −2.06 eV for absorbed oxygen) [53]. Therefore, it is believed that the decomposition of H2O2 on these nine catalysts is likely to occur under environmental conditions. Unexpectedly, Fe2 and Cu2 metal centers, very common active center in biological systems and some biomimetic compounds [54], cannot decompose H2O2 to form reactive intermediates on Pc sheet.

3.3. Catalytic Conversion of Methane on the M2-Pc

Previous theoretical studies by Yoo et al. [53] showed that when *OH and *O species exist on the surface, the activation energy barrier of the C−H bond on the precious metal will be decreased [48,55,56]. Based on the spontaneous decomposition of H2O2 on M2-Pc, two mechanisms of CH4 conversion will be examined: the *OH-assisted mechanism on Sc2-Pc, Ti2-Pc, V2-Pc, Y2-Pc, Zr2-Pc, Hf2-Pc, Ta2-Pc, and W2-Pc, the *O-assisted mechanism on Nb2-Pc, Sc2-Pc, Zr2-Pc, and W2-Pc.

3.3.1. OH-Assisted Methane Conversion

We first examined the methane conversion over Sc2-Pc, Ti2-Pc, V2-Pc, Y2-Pc, Zr2-Pc, Hf2-Pc, Ta2-Pc, and W2-Pc via *OH-assisted mechanism. The two quenching reactions of *OH, namely the disproportionation of two OH groups to H2O and O ( 2 O H O + H 2 O ) [57] and the self-reaction of H2O2  ( H 2 O 2 + 2 O H O 2 + 2 H 2 O ) [58], will result in a low reaction efficiency, thus the quenching reactions were investigated before the calculation.
The disproportionation of two *OH groups to H2O and *O on the Ti2-Pc, V2-Pc, Y2-Pc, Hf2-Pc, and Ta2-Pc were firstly evaluated. As shown in Table S4, the reaction on Y2-Pc, Hf2-Pc, and Ta2-Pc are difficult (Gibbs free energies are 1.83, 1.72, and 1.63 eV, respectively), in comparison, the free energy change on Ti2-Pc and V2-Pc are both less than 1 eV. However, the energy barriers on Ti2-Pc and V2-Pc are as high as 1.95 and 2.59 eV, respectively (Figure S6), indicating that the reaction is also difficult to proceed on these two catalysts. Though on Sc2-Pc, Zr2-Pc, and W2-Pc catalysts, the decomposition of H2O2 into a H2O and an adsorbed oxygen are more energy-efficient than the splitting into two *OH, considering that the splitting into two *OH on these three catalysts is spontaneous, we also calculated the methane conversion via *OH-assisted mechanism on these three catalyst surfaces.
The formation of O2 is also a competitive reaction, which means that H2O2 may become its own scavengers. The energy parameters for the self-reaction on the six M2-Pc (M = Ti, V, Zr, Sc, Y, and Hf) are given in Table S5, the rather high Gibbs free energies for the reaction on the Ti2-Pc, V2-Pc, and Zr2-Pc (2.61, 2.69, and 3.08 eV, respectively) indicate that the self-reaction on these three catalysts is not thermodynamically favorable. No matter what initial structure is built on the Ta2-Pc and W2-Pc, they will become two adsorbed hydroxy groups after relaxation, i.e., the reaction will not proceed on either Ta2-Pc or W2-Pc. The calculated reaction energies over the Sc2-Pc, Y2-Pc, and Hf2-Pc are 0.54, 0.65, and −0.18 eV, respectively; however, the activation energy barriers are 2.27, 1.66, and 1.36 eV, respectively (Figure S7), which means that the *OH is also difficult to quench on these three catalysts.
The above results showed that *OH will not be quenched on the eight catalysts examined in this section. Based on this finding, we investigated the methane conversion reaction via *OH-assisted mechanism.
First, CH4 is weakly adsorbed on the catalyst covered by OH, and the C−M bond lengths are 2.59, 2.42, 2.37, 2.81, 2.51, 2.45, 2.64, and 2.42 Å on the Sc2-Pc, Ti2-Pc, V2-Pc, Y2-Pc, Zr2-Pc, Ta2-Pc, W2-Pc, and Hf2-Pc, respectively. The adsorption energies are −0.15, −0.53, −0.52, −0.41, −0.49, −0.61, 0.28, and −0.22 eV, respectively. The reaction energies of surface *OH groups attracting H from CH4 are 0.76, 0.81, 0.59, 0.95, −0.14, 0.58, and 0.83 eV on Sc2-Pc, Ti2-Pc, V2-Pc, Y2-Pc, Zr2-Pc, Ta2-Pc, and Hf2-Pc, respectively; the corresponding activation energy barriers are 1.09, 0.85, 1.41, 1.35, 0.86, 0.62, and 1.22 eV. The reaction of extracting H from CH4 on the W2-Pc is not considered because its repulsiveness to CH4 (Figure S8). These analyses suggested that Ti2-Pc, Zr2-Pc, and Ta2-Pc can catalyze the C-H breakage due to favorable reaction energies and relatively mall activation barriers (0.85, 0.86 and 0.62 eV). Note that though the activation barrier on Sc2-Pc is slightly high (1.09 eV), considering that the O-assisted methane conversion on this catalyst benefits from the assistance of *OH (see Section 3.3.2), we also investigated the the *OH-assisted mechanism on Sc2-Pc. Thus, four catalysts which are feasible to break the C-H bonds, namely, Sc2-Pc, Ti2-Pc, Zr2-Pc, and Ta2-Pc, will be further investigated.
Then, we calculated the subsequent reactions over Sc2-Pc, Ti2-Pc, Zr2-Pc, and Ta2-Pc. Figure 2a–d summarizes the corresponding potential energy profile and the optimized geometries along the reaction path on these four M2-Pc. After the cleavage of the C−H bond, CH3 and H2O are adsorbed on the four catalyst surfaces. Next, the desorption of *CH3, the desorption of *H2O, and the reaction with OH in the solution to generate CH3OH were considered, respectively. The best path for all these four catalysts is to react with OH in the solution. Note that *CH3 will react with ∙OH in the solution to form CH3OH on Sc2-Pc and Ti2-Pc, the reaction proceeds spontaneously on Sc2-Pc, and the energy barrier on Ti2-Pc is only 0.08 eV. While on Zr2-Pc and Ta2-Pc, OH will combine with the H atom of *OH to generate H2O instead of reacting with *CH3 (this process occurs spontaneously on both catalysts). After desorbing H2O, *CH3, and *OH will remain on the surface. Unexpectedly, on both Zr2-Pc and Ta2-Pc catalysts, *CH3 will not combine with *OH on the surface to form CH3OH, but will combine with ∙OH in the solution to generate CH3OH (the energy barriers are 0.84 and 0.08 eV, respectively, as shown in Figure 2c,d).
To summarize, we identified that four catalysts, namely, Sc2-Pc, Ti2-Pc, Zr2-Pc, and Ta2-Pc, show high CH4 conversion activity with the assistance of *OH. The rate-limiting steps of Sc2-Pc, Ti2-Pc, and Zr2-Pc are the cleavage of the first C−H bond, with energy barriers of 1.09, 0.85, and 0.86 eV, respectively. The rate-limiting steps of Ta2-Pc is the extraction of ∙OH in the solution, with energy barriers of 1.11 eV.

3.3.2. O-Assisted Methane Conversion

Then we examined *O-assisted methane conversion on the Nb2-Pc, W2-Pc, Zr2- Pc, and Sc2-Pc catalysts. The adsorption energies of *H2O on these four M2-Pc are 0.27, 0.13, 0.21, and 0.31 eV, respectively, indicating the feasibility to desorb *H2O from these slabs to form O-adsorbed catalysts. According to previous studies, the C−H bond cleavage may occur via either a surface-stabilized ( O + C H 4   O H + C H 3 ) or a radical-like mechanism ( O + C H 4     O H + · C H 3 ) [59,60]. Figure S9 shows the energy diagram of methane conversion on these four *O-adsorbed catalysts following the surface-stabilized or radical-like mechanism, and the competition reaction pathway of the *CH3 dehydrogenation is also considered ( C H 3 + O H C H 2 + H 2 O ). In contrast to the quintet state of Fe(IV)(oxo) in Fe(IV)O/MOF-74 [61], the M-O-M moiety (M = Sc, Zr, Nb, W) is singlet in the ground state, since the magnetic moment on either M or O is zero. Accordingly, the oxidation state of O and Sc/Zr/Nb/W could be assigned as −2 and +3/+4/+4/+4, respectively, which is quantitatively in line with our Bader charge analysis (Table S6). Notably, the spin state of Nb2-Pc and Zr2-Pc switches from the triplet to the singlet when forming M-O-M moiety, agreeing well with our recent theoretical observations in metal dimer-related catalysis [62]. The partial density of states (PDOS) of CH4 adsorption on M-O-M moiety (Figure S10) shows that there is orbital hybridization between M-O-M and CH4.
Among these examined catalysts, *O-assisted conversion of methane is energetically more favorable on the Sc2-Pc monolayer through a surface-stabilized mechanism (Figure S9). This is different from the traditional single-site catalyst, which prefers the free radical mechanism [60,61,62,63,64,65]. The potential energy profile and the reaction path of *O-assisted methane C−H bond cleavage on the Sc2-Pc surface are illustrated in Figure 3, in which the C−H bond activation is found as the rate-determining step for the first methanol formation. The reaction begins with the adsorption of CH4, which is physically adsorbed on the O-preadsorbed Sc2-Pc through van der Waals interaction as the initial state with the adsorption energy of −0.09 eV, where the C−H bond length is 1.10 Å, slightly longer than that in the free CH4 molecule (1.07 Å), and the distance between O and H is 1.84 Å. In the transition state, the distance between C and H is elongated to 1.47 Å, and the distance between O and H is shortened to 1.21 Å, both of which are between the initial state and the final state. In the final state, *OH and *CH3 will form one Sc−C (bond distance of 2.25 Å) and two Sc−O bonds (bond lengths 1.97 and 2.37 Å) on surface. In other words, the *O-assisted methane activation on the Sc2-Pc catalyst follows the surface-stabilized mechanism, and the energy barrier for activation of the first C−H bond is 0.63 eV, which is much lower than the *O-assisted Au (111) surface (1.33 eV) [53].
After the C-H bond cleavage on Sc2-Pc, the position of *O is transferred from the bridge site of two Sc atoms to the top of one Sc atom, and *CH3 is adsorbed on the other Sc atom. Unlike single-site active center catalysts, bi-atom active centers increase the adsorption strength of the intermediates, thereby preventing the combination of *CH3 and *OH, which can be seen from the very high energy barrier (1.65 eV) in Figure 3.
We also considered that *CH3 reacts with ∙OH in the solution to form *CH3OH. This reaction proceeds spontaneously (the initial state structure and the final state structure are shown in Figure S11) by releasing energy of 0.54 eV (the blue line in Figure 3). The desorption of *CH3OH requires 0.13 eV of energy. After desorbing CH3OH, *OH remains being adsorbed on the surface occupying the bridge position of Sc dimers. Subsequently, we investigated two reaction paths: *OH moves to the top site of Ta atom, and *OH on the bridge position continues to activate CH4. As shown in Figure 3, the C−H bond cleavage assisted by the *OH on the bridge site is endothermic by 0.87 eV, while the migration of the *OH on the bridge site to the top site is slightly endothermic by 0.23 eV. Therefore, the *OH prefers moving to the top site, and the subsequent reaction is the same as discussed in Section 3.3.1. Noted that the active site motif for H2O2-converting methane is M-O-M (where M is a metal center), different from the biological MMOs containing binuclear Fe centers for the oxidation of CH4 to CH3OH by O2, where the reaction involves the formation of a pair of highly active iron(IV)oxo groups in a “diamond core” arrangement [61].
Note that the DFT self-interaction errors can have significant effects on the reactivity of high-valent Fe species during the oxidation of methane in metal-organic frameworks [66]. Thus, we tested the reaction of the first C–H bond cleavage using the HSE06 functional [67] and compared with the PBE result, Ti2-Pc was taken as a representative due to its best catalytic performance among the catalysts examined in this work. We found that reaction barriers (0.84 vs 0.85 eV) and reaction energies (0.78 vs 0.81 eV) are very close from the two methods (Figure S12). Thus, we conjecture that PBE results are reliable for our systems, and we adopted PBE functional through our calculations.

4. Conclusions

To sum up, we designed low-cost bi-atom (M2-Pc) catalysts for CH4 conversion using a two-dimensional material Pc to support the metal dimers. Two methane conversion routes, namely *O-assisted and *OH-assisted mechanisms, over M2-Pc were systematically studied by means of density functional theory computations. Our computations identified four high-performance catalysts for methane conversion: the Sc2-Pc surface following a combined *O-assisted and *OH-assisted mechanism with the C−H bond breaking energy barrier of 0.63 eV, and Ti2-Pc, Zr2-Pc, and Ta2-Pc following *OH-assisted mechanism with energy barriers of 0.85, 0.86, and 1.11 eV, respectively, all these activation barriers are lower than that on the Au(111) surface (1.33 eV) [53]. This work clearly demonstrated that the M2-Pc monolayers can serve as low-cost and efficient bi-atom catalysts for methane conversion, which not only enrich the family of bi-atom catalysts, but also provides new strategy to design effective bi-atom catalysts for methane conversion and related reactions.

Supplementary Materials

The Supporting Information is available free of charge at https://www.mdpi.com/article/10.3390/nano12091518/s1. Table S1. The lattice parameters (a/b, in Å) of M2-Pc catalysts, their corresponding bond lengths (the bond lengths of metal to metal, metal to nitrogen, dM−M, dM−N, dM−Nc, where Nc represents the N atom adjacent to the C atom, in Å), and the binding energy (Eb, in eV) of metal dimer anchoring at Pc monolayer, as well as the cohesive energy (Ebulk, in eV) of metal in bulk, Table S2. (a) The reaction energy for H2O2 dissociation into O + H 2 O on M2-Pc (M = Sc, Zr, Nb, W), bond lengths (dM1–O/dM2–O, in Å) (M1 = M2 = M), as well as the H2O binding energies (Eads (H2O), eV), (b) The reaction energy for H2O2 dissociation into O H + O H on M2-Pc (M = Sc, Ti, V, Y, Zr, Hf, Ta, W), as well as the bond lengths (dM1–O/dM2–O, dO1–O2, dO1–H/dO2–H, Å) (M1 = M2 = M), O1 and O2 represent the two O atoms bonded to M1 and M2, respectively, Table S3. The total magnetic moments and magnetic moment (in μB) of two metal atoms of M2-Pc (M = Sc, Ti, V, Y, Zr, Nb, Hf, Ta, W) (structures as shown in Figure S2), and the Bader charge (q, in |e|) of the two atoms. Table S4. The calculated zero-point energy, entropy, and free energy change of the reaction of the abstraction of a hydrogen from one of the *OH groups to the formation of water and oxo species ( O H + O H O + H 2 O ) on M2-Pc (M = Ti, V, Y, Hf, Ta), Table S5. The calculated zero-point energy, entropy, and free energy change of the self-reaction of H2O2 (2 O H + H 2 O 2 O 2 + 2 H 2 O ) on M2-Pc (M = Sc, Ti, V, Y, Zr, Hf) Table S6. The partial charges of O (O) and M (M1/M2) for M1-O-M2 (in q) (M1 = M2 = Sc, Zr, Nb, W). Values were obtained from Bader charge analysis, Figure S1. Top and side view of the structure of Pc in a 2 × 2 × 1 supercell (a), and its band structure (b) and projected density of state (DOS) (c). The Fermi lever is set to zero, Figure S2. Top and side views of the optimized M2-Pc monolayers, Figure S3. The energy evolution with time progress of the 5 ps MD simulation of the Ag2-Pc at 300 K, 800, 1000, 1300, and 1500 K, as well as the snapshot of structure at the end of 5 ps, respectively, Figure S4. Two adsorption configurations of H2O2 on the M2-Pc, Figure S5. The structures of spontaneously dissociated H2O2 on the M2-Pc monolayers, Figure S6. Transformation process of *OH + *OH → *O + H2O on Ti2-Pc (a) and V2-Pc (b) surfaces. The inset was the atomic structure model of each step, Figure S7. Transformation process of H2O2 + 2(*OH) → O2 + 2H2O on Sc2-Pc, Y2-Pc and Hf2-Pc surface. The inset was the atomic structure model of each step, Figure S8. The initial state structure (a) and the final state structure (b) of methane adsorption on W2-Pc surface, Figure S9. The corresponding energy profile of methane conversion via *O-assisted mechanism, Figure S10. Partial density of states (PDOS) of CH4 adsorption on M-O-M moiety. The Fermi level was set to zero, Figure S11. The initial state structure (a) and the final state structure (b) of the reaction between *CH3 and OH in solution on Sc2-Pc, Figure S12. Energy diagram of the C−H bond cleavage on Ti2-Pc surface calculated by PEB and HSE06 functional.

Author Contributions

The study was planned and designed by F.L., J.Z. and Z.C.; First-principles calculations were performed by H.M. and B.H; The manuscript was prepared by H.M., B.H., F.L and Z.C.; J.Z. discussed the results and commented on the manuscript together. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (11704203, 11964024), the “Grassland Talents” project of Inner Mongolia autonomous region (12000-12102613), and the Young Scientific Talent Development Project of Inner Mongolia University (10000-21221505), and in USA by NSF-CREST Center for Innovation, Research and Education in Environmental Nanotechnology (CIRE2N) (Grant Number HRD-1736093). The authors gratefully acknowledge the computational support from Beijing PARATEAR.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon reasonable request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Z.; Grace, J.R.; Lim, J.C.; Zhang, L. Combustion of Low-Concentration Coal Bed Methane in a Fluidized Bed. Energy Fuels 2011, 25, 975–980. [Google Scholar] [CrossRef]
  2. Sushkevich, V.L.; Palagin, D.; Ranocchiari, M.; Van Bokhoven, J.A. Selective Anaerobic Oxidation of Methane Enables Direct Synthesis of Methanol. Science 2017, 356, 523–527. [Google Scholar] [CrossRef] [PubMed]
  3. Yuan, J.; Zhang, W.; Li, X.; Yang, J. A High Performance Catalyst for Methane Conversion to Methanol: Graphene Supported Single Atom Co. Chem. Commun. 2018, 54, 2284–2287. [Google Scholar] [CrossRef] [PubMed]
  4. Tang, P.; Zhu, Q.; Wu, Z.; Ma, D. Methane Activation: The Past and Future. Energy Environ. Sci. 2014, 7, 2580–2591. [Google Scholar] [CrossRef]
  5. Narsimhan, K.; Iyoki, K.; Dinh, K.; Román-Leshkov, Y. Catalytic Oxidation of Methane into Methanol over Copper-Exchanged Zeolites with Oxygen at Low Temperature. ACS Cent. Sci. 2016, 2, 424–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Guo, X.; Fang, G.; Li, G.; Ma, H.; Fan, H.; Yu, L.; Ma, C.; Wu, X.; Deng, D.; Wei, M.; et al. Direct, Nonoxidative Conversion of Methane to Ethylene, Aromatics, and Hydrogen. Science 2014, 344, 616–619. [Google Scholar] [CrossRef]
  7. Kwon, Y.; Kim, T.Y.; Kwon, G.; Yi, J.; Lee, H. Selective Activation of Methane on Single-Atom Catalyst of Rhodium Dispersed on Zirconia for Direct Conversion. J. Am. Chem. Soc. 2017, 139, 17694–17699. [Google Scholar] [CrossRef] [PubMed]
  8. Holmen, A. Direct Conversion of Methane to Fuels and Chemicals. Catal. Today 2009, 142, 2–8. [Google Scholar] [CrossRef]
  9. Jin, D.; Zhu, B.; Hou, Z.; Fei, J.; Lou, H.; Zheng, X. Dimethyl Ether Synthesis via Methanol and Syngas over Rare Earth Metals Modified Zeolite Y and Dual Cu–Mn–Zn Catalysts. Fuel 2007, 86, 2707–2713. [Google Scholar] [CrossRef]
  10. Wang, B.; Albarracín-Suazo, S.; Pagán-Torres, Y.; Nikolla, E. Advances in Methane Conversion Processes. Catal. Today 2017, 285, 147–158. [Google Scholar] [CrossRef] [Green Version]
  11. Luo, J.Z.; Yu, Z.L.; Ng, C.F.; Au, C.T. CO2/CH4 Reforming over Ni–La2O3/5A: An Investigation on Carbon Deposition and Reaction Steps. J. Catal. 2000, 194, 198–210. [Google Scholar] [CrossRef]
  12. Tang, S.; Ji, L.; Lin, J.; Zeng, H.C.; Tan, K.L.; Li, K. CO2 Reforming of Methane to Synthesis Gas over Sol–Gel-made Ni/γ-Al2O3 Catalysts from Organometallic Precursors. J. Catal. 2000, 194, 424–430. [Google Scholar] [CrossRef]
  13. McFarland, E. Unconventional Chemistry for Unconventional Natural Gas. Science 2012, 338, 340–342. [Google Scholar] [CrossRef]
  14. Da Silva, M.J. Synthesis of Methanol from Methane: Challenges and Advances on the Multi-Step (Syngas) and One-Step Routes (DMTM). Fuel Process. Technol. 2016, 145, 42–61. [Google Scholar] [CrossRef]
  15. Schroder, D.; Fiedler, A.; Hrusak, J.; Schwarz, H. Experimental and Theoretical Studies toward a Characterization of Conceivable Intermediates Involved in the Gas-Phase Oxidation of Methane by Bare FeO+. Generation of Four Distinguishable [Fe,C,H4,O]+ Isomers. J. Am. Chem. Soc. 1992, 114, 1215–1222. [Google Scholar] [CrossRef]
  16. Chen, Y.-M.; Clemmer, D.E.; Armentrout, P.B. Conversion of CH4 to CH3OH: Reactions of CoO+ with CH4 and D2, Co+ with CH3OD and D2O, and Co+(CH3OD) with Xe. J. Am. Chem. Soc. 1994, 116, 7815–7826. [Google Scholar] [CrossRef]
  17. Lieberman, R.L.; Rosenzweig, A.C. Crystal Structure of a Membrane-Bound Metalloenzyme that Catalyses the Biological Oxidation of Methane. Nature 2005, 434, 177–182. [Google Scholar] [CrossRef]
  18. Hakemian, A.S.; Rosenzweig, A.C. The Biochemistry of Methane Oxidation. Annu. Rev. Biochem. 2007, 76, 223–241. [Google Scholar] [CrossRef]
  19. Rosenzweig, A.C.; Frederick, C.A.; Lippard, S.J.; Nordlund, P. Crystal Structure of A Bacterial Non-Haem Iron Hydroxylase that Catalyses the Biological Oxidation of Methane. Nature 1993, 366, 537–543. [Google Scholar] [CrossRef]
  20. Lipscomb, J.D. Biochemistry of the Soluble Methane Monooxygenase. Annu. Rev. Microbiol. 1994, 48, 371–399. [Google Scholar] [CrossRef]
  21. Colby, J.; Stirling, D.I.; Dalton, H. The Soluble Methane Mono-Oxygenase of Methylococcus Capsulatus (Bath). Its Ability to Oxygenate n-Alkanes, n-Alkenes, Ethers, and Alicyclic, Aromatic and Heterocyclic Compounds. Biochem. J. 1977, 165, 395–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Balasubramanian, R.; Smith, S.M.; Rawat, S.; Yatsunyk, L.A.; Stemmler, T.L.; Rosenzweig, A.C. Oxidation of Methane by a Biological Dicopper Centre. Nature 2010, 465, 115–119. [Google Scholar] [CrossRef] [Green Version]
  23. Narsimhan, K.; Michaelis, V.; Mathies, G.; Gunther, M.; Griffin, W.R.; Román-Leshkov, Y. Methane to Acetic Acid over Cu-Exchanged Zeolites: Mechanistic Insights from a Site-Specific Carbonylation Reaction. J. Am. Chem. Soc. 2015, 137, 1825–1832. [Google Scholar] [CrossRef] [Green Version]
  24. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic Peroxidase-like Activity of Ferromagnetic Nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, S.E.; Zhang, L.; Ma, K.; Riegman, M.; Chen, F.; Ingold, I.; Conrad, M.; Turker, M.Z.; Gao, M.; Jiang, X.; et al. Ultrasmall Nanoparticles Induce Ferroptosis in Nutrient-Deprived Cancer Cells and Suppress Tumour Growth. Nat. Nanotechnol. 2016, 11, 977–985. [Google Scholar] [CrossRef] [Green Version]
  26. Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with Enzyme-like Characteristics (Nanozymes): Next-Generation Artificial Enzymes (II). Chem. Soc. Rev. 2019, 48, 1004–1076. [Google Scholar] [CrossRef] [PubMed]
  27. Qiao, B.; Wang, A.; Yang, X.; Allard, L.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-Atom Catalysis of CO Oxidation Using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef] [PubMed]
  28. Gawande, M.B.; Ariga, K.; Yamauchi, Y. Single-Atom Catalysts. Small 2021, 17, 2101584. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, A.; Li, J.; Zhang, T. Heterogeneous Single-Atom Catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  30. Deng, T.; Zheng, W.; Zhang, W. Increasing the Range of Non-Noble-Metal Single-Atom Catalysts. Chin. J. Catal. 2017, 38, 1489–1497. [Google Scholar] [CrossRef]
  31. Chen, Z.; Yan, J.-M.; Jiang, Q. Single or Double: Which Is the Altar of Atomic Catalysts for Nitrogen Reduction Reaction? Small Methods 2018, 3, 1800291. [Google Scholar] [CrossRef]
  32. Li, F.; Liu, X.; Chen, Z. 1 + 1′ > 2: Heteronuclear Bi-atom Catalyst Outperforms Its Homonuclear Counterparts for CO Oxidation. Small Methods 2019, 3, 1800480. [Google Scholar] [CrossRef]
  33. Yan, H.; Lin, Y.; Wu, H.; Zhang, W.; Sun, Z.; Cheng, H.; Liu, W.; Wang, C.; Li, J.; Huang, X.; et al. Bottom-up Precise Synthesis of Stable Platinum Dimers on Graphene. Nat. Commun. 2017, 8, 1070. [Google Scholar] [CrossRef] [PubMed]
  34. Li, F.; Li, Y.; Zeng, X.C.; Chen, Z. Exploration of High-Performance Single-Atom Catalysts on Support M1/FeOx for CO Oxidation via Computational Study. ACS Catal. 2015, 5, 544–552. [Google Scholar] [CrossRef]
  35. Zhao, J.; Zhao, J.; Li, F.; Chen, Z. Copper Dimer Supported on C2N-Layer as an Efficient Electrocatalyst for CO2 Reduction Reaction: A Computational Study. J. Phys. Chem. C 2018, 122, 19712–19721. [Google Scholar] [CrossRef]
  36. Abel, M.; Clair, S.; Ourdjini, O.; Mossoyan, M.; Porte, L. Single Layer of Polymeric Fe-Phthalocyanine: An Organometallic Sheet on Metal and Thin Insulating Film. J. Am. Chem. Soc. 2011, 133, 1203–1205. [Google Scholar] [CrossRef]
  37. Matsushita, O.; Derkacheva, V.M.; Muranaka, A.; Shimizu, S.; Uchiyama, M.; Luk’yanets, E.A.; Kobayashi, N. Rectangular-Shaped Expanded Phthalocyanines with Two Central Metal Atoms. J. Am. Chem. Soc. 2012, 134, 3411–3418. [Google Scholar] [CrossRef]
  38. Gueorguiev, G.K.; Pacheco, J.M.; Stafström, S.; Hultman, L. Silicon–metal Clusters: Nano-templates for Cluster Assembled Materials. Thin Solid Films 2006, 515, 1192–1196. [Google Scholar] [CrossRef]
  39. Dos Santos, P.B.; Rivelino, R.; de Brito Mota, F.; Gueorguiev, G.K.; Kakanakova-Georgieva, A. Dopant Species with Al-Si and N-Si Bonding in The MOCVD of AlN Implementing Trimethylaluminum, Ammonia and Silane. J. Phys. D Appl. Phys. 2015, 48, 295104. [Google Scholar]
  40. Bučko, T.; Hafner, J.; Lebègue, S.; Àngyán, J.G. Improved Description of the Structure of Molecular and Layered Crystals: Ab Initio DFT Calculations with van der Waals Corrections. J. Phys. Chem. A 2010, 114, 11814–11824. [Google Scholar] [CrossRef]
  41. Blöch, P. Projector Augmented-Wave Method. Phy. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  43. Monkhorst, H.; Pack, J. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  44. Henkelman, G.; Uberuaga, B.P.; Jónsson, H. A Climbing Image Nudged Elastic Band Method for Finding Saddle Points and Minimum Energy Paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef] [Green Version]
  45. Henkelman, G.; Jónsson, H. Improved Tangent Estimate in the Nudged Elastic Band Method for Finding Minimum Energy Paths and Saddle Points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef] [Green Version]
  46. Sangiovanni, D.G.; Gueorguiev, G.K.; Kakanakova-Georgieva, A. Ab initio Molecular Dynamics of Atomic-Scale Surface Rreactions: Insights into Metal Organic Chemical Vapor Deposition of AlN on Graphene. Phys. Chem. Chem. Phys. 2018, 20, 17751–17761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Campos-Martin, J.M.; Blanco-Brieva, G.; Fierro, J.L.G. Hydrogen Peroxide Synthesis: An Outlook beyond the Anthraquinone Process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef]
  48. Sajith, P.K.; Staykov, A.; Yoshida, M.; Shiota, Y.; Yoshizawa, K. Theoretical Study of the Direct Conversion of Methane to Methanol Using H2O2 as an Oxidant on Pd and Au/Pd Surfaces. J. Phys. Chem. C 2020, 124, 13231–13239. [Google Scholar] [CrossRef]
  49. Hammond, C.; Forde, M.M.; Ab Rahim, M.H.; Thetford, A.; He, Q.; Jenkins, R.L.; Dimitratos, N.; Lopez-Sanchez, J.A.; Dummer, N.F.; Murphy, D.M.; et al. Direct Catalytic Conversion of Methane to Methanol in an Aqueous Medium by Using Copper-Promoted Fe-ZSM-5. Angew. Chem. Int. Ed. 2012, 51, 5129–5133. [Google Scholar] [CrossRef]
  50. Ab Rahim, M.H.; Forde, M.M.; Jenkins, R.; Hammond, C.; He, Q.; Dimitratos, N.; Lopez-Sanchez, J.A.; Carley, A.F.; Taylor, S.H.; Willock, D.J.; et al. Oxidation of Methane to Methanol with Hydrogen Peroxide Using Supported Gold-Palladium Alloy Nanoparticles. Angew. Chem. Int. Ed. 2013, 52, 1280–1284. [Google Scholar] [CrossRef]
  51. Osadchii, D.Y.; Olivos-Suarez, A.L.; Szecsenyi, Á.; Li, G.; Nasalevich, M.A.; Dugulan, L.A.; Crespo, P.S.; Hensen, E.J.M.; Veber, S.L.; Fedin, M.V.; et al. Isolated Fe Sites in Metal Organic Frameworks Catalyze the Direct Conversion of Methane to Methanol. ACS Catal. 2018, 8, 5542–5548. [Google Scholar] [CrossRef] [Green Version]
  52. Xiao, P.; Wang, Y.; Nishitoba, T.; Kondo, J.N.; Yoko, T. Selective Oxidation of Methane to Methanol with H2O2 over An Fe-MFI Zeolite Catalyst Using Sulfolane Solvent. Chem. Commun. 2019, 55, 2896–2899. [Google Scholar] [CrossRef] [PubMed]
  53. Yoo, J.S.; Schumann, J.; Studt, F.; Abild-Pedersen, F.; Nørskov, J.K. Theoretical Investigation of Methane Oxidation on Pd(111) and Other Metallic Surfaces. J. Phys. Chem. C 2018, 122, 16023–16032. [Google Scholar] [CrossRef]
  54. Dinh, K.T.; Sullivan, M.M.; Serna, P.; Meyer, R.J.; Dincă, M.; Román-Leshkov, Y. Viewpoint on the Partial Oxidation of Methane to Methanol Using Cu- and Fe-Exchanged Zeolites. ACS Catal. 2018, 8, 8306–8313. [Google Scholar] [CrossRef] [Green Version]
  55. Tsai, C.; Latimer, A.; Yoo, J.S.; Studt, F.; Abild-Pedersen, F. Predicting Promoter-Induced Bond Activation on Solid Catalysts Using Elementary Bond Orders. J. Phys. Chem. Lett. 2015, 6, 3670–3674. [Google Scholar] [CrossRef]
  56. Yoo, J.S.; Khan, T.S.; Abild-Pedersen, F.; Nørskov, J.K.; Studt, F. On the Role of the Surface Oxygen Species during A–H (A = C, N, O) Bond Activation: A Density Functional Theory Study. Chem. Commun. 2015, 51, 2621–2624. [Google Scholar] [CrossRef]
  57. Li, J.; Staykov, A.; Ishihara, T.; Yoshizawa, K. Theoretical Study of the Decomposition and Hydrogenation of H2O2 on Pd and Au@Pd Surfaces: Understanding toward High Selectivity of H2O2 Synthesis. J. Phys. Chem. C 2011, 115, 7392–7398. [Google Scholar] [CrossRef]
  58. Serra-Maia, R.; Michel, F.M.; Kang, Y.; Stach, E.A. Decomposition of Hydrogen Peroxide Catalyzed by AuPd Nanocatalysts during Methane Oxidation to Methanol. ACS Catal. 2020, 10, 5115–5123. [Google Scholar] [CrossRef]
  59. Latimer, A.A.; Aljama, H.; Kakekhani, A.; Yoo, J.S.; Kulkarni, A.; Tsai, C.; Garcia-Melchor, M.; Abild-Pedersen, F.; Nørskov, J.K. Mechanistic Insights into Heterogeneous Methane Activation. Phys. Chem. Chem. Phys. 2017, 19, 3575–3581. [Google Scholar] [CrossRef]
  60. Latimer, A.A.; Kulkarni, A.R.; Aljama, H.; Montoya, J.H.; Yoo, J.S.; Tsai, C.; Abild-Pedersen, F.; Studt, F.; Nørskov, J.K. Understanding Trends in C–H Bond Activation in Heterogeneous Catalysis. Nat. Mater. 2017, 16, 225–229. [Google Scholar] [CrossRef]
  61. Fernan, S.; Leonardo, B. Density-Functional Theory Models of Fe(IV)O Reactivity in metal-organic Frameworks: SelfInteraction Error, Spin Delocalisation and The Role of Hybrid Exchange. Phys. Chem. Chem. Phys. 2020, 22, 12821–12830. [Google Scholar]
  62. Yu, L.; Li, F.; Zhao, J.; Chen, Z. Revisiting Catalytic Performance of Supported Metal Dimers for Oxygen Reduction Reaction via Magnetic Coupling from First Principles. Adv. Powder Mater. 2022, 1, 100031. [Google Scholar] [CrossRef]
  63. Latimer, A.A.; Kakekhani, A.; Kulkarni, A.; Nørskov, J.K. Direct Methane to Methanol: The Selectivity-Conversion Limit and Design Strategies. ACS Catal. 2018, 8, 6894–6907. [Google Scholar] [CrossRef]
  64. Da Silva, J.C.S.; Pennifold, R.C.R.; Harvey, J.N.; Rocha, W.R.A. Radical Rebound Mechanism for the Methane Oxidation Reaction Promoted by the Dicopper Center of A pMMO Enzyme: A Computational Perspective. Dalton Trans. 2016, 45, 2492–2504. [Google Scholar] [CrossRef] [Green Version]
  65. Chin, Y.H.; Buda, C.; Neurock, M.; Iglesia, E. Consequences of Metal-Oxide Interconversion for C–H Bond Activation during CH4 Reactions on Pd Catalysts. J. Am. Chem. Soc. 2013, 135, 15425–15442. [Google Scholar] [CrossRef]
  66. Bernd, E.; Francesco, B.; Michiel, C.M.G.; Evert, J.B. Methane-to-Methanol Oxidation by the Hydrated Iron(IV) Oxo Species in Aqueous Solution: A Combined DFT and Car-Parrinello Molecular Dynamics Study. J. Am. Chem. Soc. 2004, 126, 4355–4365. [Google Scholar]
  67. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Phys. Chem. C 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Top view of M2-Pc sheet. (b) The metals examined in this work. The poisonous metals were indicated in gray; the metals assist H2O2 decomposition into 2*OH or *O and H2O are highlighted in yellow and orange, respectively; the metals could decompose H2O2 into either 2*OH or *O and H2O are represented in blue; metals in white cannot spontaneously decompose H2O2.
Figure 1. (a) Top view of M2-Pc sheet. (b) The metals examined in this work. The poisonous metals were indicated in gray; the metals assist H2O2 decomposition into 2*OH or *O and H2O are highlighted in yellow and orange, respectively; the metals could decompose H2O2 into either 2*OH or *O and H2O are represented in blue; metals in white cannot spontaneously decompose H2O2.
Nanomaterials 12 01518 g001
Figure 2. Reaction pathway of *OH-assisted CH4 decomposition on the Sc2-Pc (a), Ti2-Pc (b), Zr2-Pc (c), and Ta2-Pc (d) monolayers. Blue, black, and red lines represent the reaction paths for *CH3 desorption, *H2O desorption, and the reaction with OH in the solution, respectively.
Figure 2. Reaction pathway of *OH-assisted CH4 decomposition on the Sc2-Pc (a), Ti2-Pc (b), Zr2-Pc (c), and Ta2-Pc (d) monolayers. Blue, black, and red lines represent the reaction paths for *CH3 desorption, *H2O desorption, and the reaction with OH in the solution, respectively.
Nanomaterials 12 01518 g002
Figure 3. Reaction pathway of *O-assisted CH4 decomposition on the Sc2-Pc monolayer. Blue, black, and red lines represent the three paths of *CH3 reacting with *OH, *CH3 reacting with ∙OH in solution, and the migration of *OH from bridge to top site, respectively.
Figure 3. Reaction pathway of *O-assisted CH4 decomposition on the Sc2-Pc monolayer. Blue, black, and red lines represent the three paths of *CH3 reacting with *OH, *CH3 reacting with ∙OH in solution, and the migration of *OH from bridge to top site, respectively.
Nanomaterials 12 01518 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Meng, H.; Han, B.; Li, F.; Zhao, J.; Chen, Z. Understanding the CH4 Conversion over Metal Dimers from First Principles. Nanomaterials 2022, 12, 1518. https://doi.org/10.3390/nano12091518

AMA Style

Meng H, Han B, Li F, Zhao J, Chen Z. Understanding the CH4 Conversion over Metal Dimers from First Principles. Nanomaterials. 2022; 12(9):1518. https://doi.org/10.3390/nano12091518

Chicago/Turabian Style

Meng, Haihong, Bing Han, Fengyu Li, Jingxiang Zhao, and Zhongfang Chen. 2022. "Understanding the CH4 Conversion over Metal Dimers from First Principles" Nanomaterials 12, no. 9: 1518. https://doi.org/10.3390/nano12091518

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop