Next Article in Journal
The Bioactive Properties of Carotenoids from Lipophilic Sea buckthorn Extract (Hippophae rhamnoides L.) in Breast Cancer Cell Lines
Next Article in Special Issue
Boron-Based Inverse Sandwich V2B7 Cluster: Double π/σ Aromaticity, Metal–Metal Bonding, and Chemical Analogy to Planar Hypercoordinate Molecular Wheels
Previous Article in Journal
Change in Lipofectamine Carrier as a Tool to Fine-Tune Immunostimulation of Nucleic Acid Nanoparticles
Previous Article in Special Issue
Sc@B28, Ti@B28, V@B28+, and V@B292−: Spherically Aromatic Endohedral Seashell-like Metallo-Borospherenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanisms in the Catalytic Reduction of N2O by CO over the M13@Cu42 Clusters of Aromatic-like Inorganic and Metal Compounds

1
College of Sciences, Xinjiang Production and Construction Corps Key Laboratory of Advanced Energy Storage Materials and Technology, Shihezi University, Shihezi 832000, China
2
Key Laboratory of Materials Modification by Laser, Ministry of Education, Ion and Electron Beams, Dalian University of Technology, Dalian 116024, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(11), 4485; https://doi.org/10.3390/molecules28114485
Submission received: 2 May 2023 / Revised: 23 May 2023 / Accepted: 27 May 2023 / Published: 1 June 2023
(This article belongs to the Special Issue Aromatic Inorganic and Metallic Compounds)

Abstract

:
Metal aromatic substances play a unique and important role in both experimental and theoretical aspects, and they have made tremendous progress in the past few decades. The new aromaticity system has posed a significant challenge and expansion to the concept of aromaticity. From this perspective, based on spin-polarized density functional theory (DFT) calculations, we systematically investigated the doping effects on the reduction reactions of N2O catalyzed by CO for M13@Cu42 (M = Cu, Co, Ni, Zn, Ru, Rh, Pd, Pt) core–shell clusters from aromatic-like inorganic and metal compounds. It was found that compared with the pure Cu55 cluster, the strong M–Cu bonds provide more structural stability for M13@Cu42 clusters. Electrons that transferred from the M13@Cu42 to N2O promoted the activation and dissociation of the N–O bond. Two possible reaction modes of co-adsorption (L-H) and stepwise adsorption (E-R) mechanisms over M13@Cu42 clusters were thoroughly discovered. The results showed that the exothermic phenomenon was accompanied with the decomposition process of N2O via L-H mechanisms for all of the considered M13@Cu42 clusters and via E-R mechanisms for most of the M13@Cu42 clusters. Furthermore, the rate-limiting step of the whole reactions for the M13@Cu42 clusters were examined as the CO oxidation process. Our numerical calculations suggested that the Ni13@Cu42 cluster and Co13@Cu42 clusters exhibited superior potential in the reduction reactions of N2O by CO; especially, Ni13@Cu42 clusters are highly active, with very low free energy barriers of 9.68 kcal/mol under the L-H mechanism. This work demonstrates that the transition metal core encapsulated M13@Cu42 clusters can present superior catalytic activities towards N2O reduction by CO.

Graphical Abstract

1. Introduction

Nowadays, air pollution has gradually become a serious environmental problem. Fuel combustion, the emissions of exhaust gases, and excessive fertilization in agriculture have led to a huge production of large amounts of toxic and harmful gases in the atmosphere. Among them, nitrous oxide (N2O) and carbon monoxide (CO) have been recognized as two common harmful gases among the emissions of exhaust gases. Specially, N2O has been recognized as the main gas that causes the greenhouse effect because it has a global warming potential that is 300 times larger than that of CO2 [1,2]. Moreover, it is also a stratospheric ozone depleter [3]. Meanwhile, CO is not only a potential greenhouse gas but can also cause serious harm to human health. In recent years, significant efforts have been devoted to the development and application of the reduction of the two harmful gases [4,5,6,7,8,9,10,11]. One of the most promising methods is to convert them into less harmful N2 and CO2 gases (N2O + CO→N2 + CO2). This process involves two steps: firstly, a N2O molecule reacts to form N2 and an adsorbed O atom (N2O→N2 + O*), then the O* atom reacts with a CO molecule to form CO2 (O* + CO→CO2) [12]. Notably, the direct reaction of N2O with CO molecules shows a high kinetic barrier (about 47.77 kcal/mol) [13], which seriously hinders the reaction’s ability to take place at room temperature. Therefore, it is necessary to develop catalysts that have the properties of being stable, economical, and highly active. Until now, many kinds of catalysts, including noble metal [14], metal zeolites/porphyrins [15], metal alloys, and perovskite-like catalysts [16,17], have been proven to achieve efficient conversions of N2O and CO to N2 and CO2. However, the high cost and difficulties in the mining process (environment, working conditions) have led to a deep search for metal catalysts with similar or improved properties to replace them.
Aromaticity has attracted the most attention in recent years. Li et al. prepared a series of all metal clusters MAl4 (M = Li, Na, Cu) and investigated the valence molecular orbitals of Al42− using photoelectron spectroscopy experiments combined with quantum chemical theory calculations [18]. Two π electrons occupied completely non-deterministic HOMO orbitals. This study identified MAl4 as an aromatic system, which has been confirmed in the scientific research. Zhu et al. reported the first synthesis of metallapentalyne having a transition metal-centered d orbital involved in conjugation in osmium metallapentalyne, thus converting the Hückel anti-aromatic nature of metallapentalyne into the Möbius aromatic nature of metallapentalyne [19]. These studies have generated great scientific importance in the extended concept of the aromaticity, which has extended the scope from organic chemistry to metal cluster systems and has opened up a new scientific frontier. In this context, Cu-based aromatic-like metal compound alloys have attracted considerable attentions because of copper’s low cost and excellent properties in various aspects, such as having high stability and long persistence in high-temperature conditions [20,21,22]. In recent years, metallic Cu nanoclusters or Cu-based alloy clusters have been widely applied in various catalytic processes, e.g., hydrogen evolution reaction (HER), NOx reduction, dry reforming of methane (DRM), etc. [23,24,25,26,27]. As for the catalytic reduction of N2O by CO reaction, Barabás et al. systematically investigated the performance of Cun (n = 4–15) cluster catalysts and found that Cu12 and Cu14 clusters were the best catalysts according to the thermodynamic analysis, even at ambient temperatures [28]. Lian et al. systematically studied M@Cu12 (M = Cu, Pt, Ru, Pd, Rh) clusters and confirmed that Ru@Cu12 and Pt@Cu12 clusters exhibited superior catalytic activity via a co-adsorption mechanism [29].
As a medium-sized nanocluster (the diameter is about 1 nm) with a typical core–shell structure, the icosahedral Cu55 cluster of aromatic-like inorganic and metal compounds is considered to be the global minimum of Cu55 clusters [30]. Extensive studies have been carried out to reveal the physical and chemical properties and potentials in the catalysis of this nanocluster material [31,32,33,34,35,36]. Mao et al. reported a DFT-based high-throughput screening method to successfully screen Cu55-nMn (M = Co, Ni, Ru, and Rh) core–shell alloy clusters and identified Cu-Ni alloy clusters as a superior electrocatalyst for HER [31]. Cao et al. systematically studied a single Pd-doped Cu55 nanoparticle towards propane dehydrogenation and demonstrated that this nanoparticle exhibited superior catalytic activity toward C–H bond activation and significantly reduced side reactions such as deep dehydrogenation [32]. Liu et al. identified a crown jewel-structured Pt12Cu43 cluster as a promising catalyst candidate for highly efficient and low lost oxygen reduction reaction (ORR) processes [33]. So far, pure or alloyed Cu55 nanoclusters have been proven to own great potentials in HER [31], ORR [33], H2 dissociation [35], propane dehydrogenation [32], and acetylene selective hydrogenation reactions [36]; however, to the best of our knowledge, no theoretical work has been carried out to investigate their activity towards the selective catalytic reduction of N2O via CO (i.e., CO oxidation by N2O). We aim to fill up that void with the current work.
In this paper, by means of DFT calculations, we systematically investigated the structural stability and catalytic performance of pure Cu55 core–shell clusters and substituted M13@Cu42 (M = Co, Ni, Zn, Ru, Rh, Pd, Pt) clusters of aromatic-like inorganic and metal compounds by encapsulating the transition metal atoms as a core in a CO oxidation by N2O reaction. The results confirm that the doping of the core metal atoms can greatly affect the structure and electronic and catalytic properties of Cu55 clusters. By carefully examining the N2O decomposition and CO oxidation processes of these core–shell clusters, we found that Ni13@Cu42 and Co13@Cu42 clusters can serve as promising candidates in CO oxidation by N2O. The results also reveal that the different reaction mechanisms and the metal modification doping of M13@Cu42 clusters play a key role in CO oxidation. Our calculations reveal endoplasmically doped, medium-size Cu55 clusters that can serve as stable, low-cost, and highly effective catalysts in the selective catalytic reduction of N2O by CO.

2. Computational Details

All spin-polarized DFT calculations in this paper were performed using the generalized gradient approximation (GGA) of the Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional while adopting the Vienna ab initio simulation package (VASP 5.4.4) software package [37,38,39]. The electron–ion exchange correlation interactions were calculated by using the projector augmented wave (PAW) method. Grimme’s semiempirical DFT-D3 scheme of dispersion correction was implemented to elaborate the van der Waals (vdW) interaction [40]. A plane wave basis was accompanied by a kinetic energy cut-off of 500 eV. The convergence criteria for the structure optimizations were set to 10−5 eV, and the Hellmann–Feynman force was less than 0.02 eV Å−1. Cluster structures with a vacuum space of 25 Å were applied to ensure negligible interactions within neighboring unit clusters. For the optimized geometries, the K-points was set to be 3 × 3 × 1; they were set to 9 × 9 × 1 for the density of states (DOSs) calculations. The minimum reaction paths for each step of the reaction were considered to be using the climbing image advancement elastic band (CI–NEB) method [41].
In order to characterize the stability of the transition metal (TM) core atoms-substituted M13@Cu42 clusters, the average binding energy (Eb) was adopted to evaluate the structural stability of core–shell bimetallic clusters, which was calculated according to the following equation:
E b = 13 E M + 42 E C u E M 13 @ C u 42 / 55
where E M 13 @ C u 42 , ECu, and EM are the total energy of M13@Cu42 cluster, energy of the Cu atom, and doped metal atom, respectively.
The adsorption energies (Eads) of the adsorption species on these clusters were defined by the following Equation (2):
E a d s = E t o t a l E s p e c i e s E c l u s t e r ,
where Etotal, Especies, and Ecluster are the total energies of the total adsorbed systems, isolated adsorption species, and clusters, respectively. A Bader population analysis was adopted to quantify the charge population in each atom in our calculation.
To support the choice of the functional combinations and the computational detail of the basis set described, we provided benchmark calculations of the geometrical parameters for N2O, CO, CO2, and N2. The Cu–Cu average bond length of the Cu55 cluster was calculated to be 2.32 Å, which is fairly consistent with the experimental result of 2.37 Å [42].

3. Results and Discussion

3.1. Structure

The structure of the aromatic-like inorganic and metal compounds Cu55 cluster is presented in Figure 1, which had icosahedral (Ih) symmetry and a diameter size of approximately 9.77 Å. The Cu55 cluster possessed a multishell structure, in which the outer layer was composed of 42 Cu atoms and the Ih core was formed by 13 copper central atoms. As shown in Figure 1, the surface of Cu55 consisted of 20 equivalent triangular fcc (111) facets. The fcc (111) facet in Cu55 clusters and aromatic structures are similar. Each facet contained two types of nonequivalent atoms, namely three Cu1 atoms at the intersection of five fcc (111) facets (vertex, T) and three Cu2 atoms in two contiguous fcc (111) facets (bridge site, B). The different Cu atoms induced different Cu–Cu bond lengths, such as 2.51 Å of dCu1–Cu2 and 2.59 Å of dCu2–Cu2. The distance of the Cu1 atoms to the inner layer atom was 4.77 Å with a coordination number of six, while that of the Cu2 atoms to the inner layer atom was 4.18 Å with a coordination number of eight. Compared with bulk copper, the different bond lengths in the Cu55 cluster endowed its atoms and bonds with higher activity, showing potential in certain catalysis reactions. After the substitution of the core atom, all of the M13@Cu42 (M = Co, Ni, Zn, Ru, Rh, Pd, Pt) clusters exhibited no obvious structural deformation.
Furthermore, the structural stability of the M13@Cu42 clusters was examined from the binding energy point (Figure 2). The binding energy of the pure Cu55 cluster was calculated to be 70.10 kcal/mol, which agrees well with a previous report [43]. The average binding energies of the subsequent clusters were even larger than those of intrinsic Cu55 except for Zn13@Cu42, indicating the strong M–Cu bonds and structural stability of these core-doped clusters, as listed in Table 1. This is reasonable from the viewpoint of the melting points of these metals, e.g., [44]. The energy gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of these clusters ranged from 0.01 eV to 0.59 eV. The bond lengths between the M dopants and out-layer Cu atoms ranged from 2.42 Å to 2.65 Å, which was accompanied by an average charge transfer of about 0.01 |e|~0.28 |e| from the TM atoms to the adjacent Cu atoms, as shown in Table 1. Such electron transfers greatly affected the clusters’ catalytic performances, as discussed in the following sections.

3.2. Catalytic Properties of M13@Cu42 Cluster

Before analyzing the CO oxidation by N2O, the adsorption of N2O on the catalysts was explored. For the neutral state, the N2O molecules presented a linear geometry with a N–O bond length of 1.20 Å. Generally, there were two types of geometries for N2O adsorption on the catalyst, namely N terminal and O terminal geometries. After geometric optimization, the adsorption energies of N2O binding to clusters through the N terminus were smaller than those through the O terminus (about 43.81 kcal/mol). The charge transfer from the system with O terminals to the N2O molecule exceeded −0.9|e|; meanwhile, there was no significant charge transfer in the N terminal adsorbed systems. The related parameters and structure of N2O adsorption on the Cu55 cluster are shown in Table S1 and Figure S1.
Six different adsorption sites on the surface of the pure Cu55 cluster were investigated to determine the optimal adsorption configuration for the reaction, including the hollow of the fcc (H1, H2) facets, top of the Cu1 atom (T1), top of the Cu2 atom (T2), bridge of the bond between Cu1 and Cu2 (B1), and bridge of the bond between Cu2 and Cu2 (B2). After geometric optimization with the O terminal, the N2O was decomposed to the N2 molecule, and the remaining O atom was adsorbed on the Cu55 cluster. The average distance of Cu–O was shortened from 2.20 Å to 1.90 Å, while the bond length of N–O was lengthened from 1.20 to 3.19 Å. Clearly, the N–O bond was broken, and the adsorbed O atom prepared for the subsequent oxidation reaction of CO on the cluster. The adsorption energy between the Cu55 cluster and N2O was obtained as ranging from −48.43 to −53.27 kcal/mol, as shown in Table S1. The adsorption energies of the H1 and H2 adsorption sites were comparable on the Cu55 cluster, but the H2 configuration was more easily activated than the H1 configuration. The optimization results indicated that the optimized structure of N2O adsorption at the B1, T2, and B2 sites was consistent with that located at the H2 sites. The subsequent CI–NEB calculation also showed that the remaining O atom was biased to be adsorbed at the H2 site after the N2 dissociation for the Cu55 cluster. The Bader charge analysis proved that approximately 0.96|e| to 0.97|e| was transferred from the clusters to the N2O molecule (H2 approximately 0.97|e|). Therefore, the H2 configuration served as the preferred adsorption site of N2O on the Cu55 cluster, which is consistent with the literature [45]. For further calculations, the H2 site adsorbed to the clusters by the O terminal was selected as the most stable adsorption geometry.
Similarly, the configurations of N2O at the H2 active site on the doped M13@Cu42 clusters were studied. The geometry of the N2O adsorbs was greatly changed. The atomic average distance of Cu–O was 1.92 Å, while the length of the N–O bond was elongated, ranging from 3.17 Å to 3.58 Å for the M13@Cu42 clusters. These changes indicated that the N2O absorbed by the M13@Cu42 clusters decomposed to N2 and an adsorbed O atom on the M13@Cu42 cluster for the follow-up reaction. All adsorption energies were negative (−46.12 to −70.33 kcal/mol), suggesting that the adsorption processes for all catalysts were favorable in terms of thermodynamic stability. The Bader charge analysis showed that the M13@Cu42 clusters transferred 0.97|e| to 1.11|e| charges to N2O molecule, which caused its reduction. In other words, the M13@Cu42 clusters significantly assisted in withdrawing charges from the N2O molecule, as indicated in Table 2.
After the analysis of N2O adsorption, the entire CO oxidation mechanisms as a result of the N2O reaction were investigated by combining the elementary steps associated with all eight kinds of potential catalysts. There were a total of two possible reaction mechanisms, including the E-R and L-H pathways. For the above eight potential catalysts, the most favorable potential energy curves of the CO oxidation as a result of N2O processes for all of the possible reaction pathways are presented in Figures S2–S7. The corresponding structures of the reaction intermediates were also explored; the important information is summarized in Table 2.
Under the L-H mechanism, N2O and CO were first co-adsorbed (*N2O–*CO) on the catalyst surface, bonding to the underlying two adjacent Cu atoms. Subsequently, the N2 was desorbed from the active sites to generate a gaseous N2 molecule, and the remaining O atoms were adsorbed on the Cu55 cluster. The average distance between *C and Cu was 1.84 Å, and the N–O bond length was 3.04 Å. The adsorption energies of the co-adsorbed CO and N2O molecules for the eight types of catalysts were calculated to range from −72.87 to −97.77 kcal/mol. The remaining O atom approached an adsorbed CO molecule and reacted to form CO2, which was also exothermic. The adsorption energies of the CO and absorbed O atom ranged from −0.92 to −7.61 kcal/mol. The barrier of the transition state (TS2+) for the formation of *O*CO ranged from 9.68 kcal/mol to 19.58 kcal/mol. The transition state involved *O and *CO adsorbates having a distance of 1.93–2.27 Å between them for the M13@Cu42 clusters. The CO2 molecule was directly dissociated from the M13@Cu42 clusters, indicating the accomplishment of the reaction. By examining the overall CO oxidation as a result of N2O processes via the L-H mechanism, the CO oxidation step was considered as the rate-limiting step in the eight catalysts (M13@Cu42 clusters), which was consistent with the results of the Cun (n = 4–15) clusters [28]. Before the rate-limiting step, the barriers for the N2O reduction steps on these eight catalysts were relatively exothermic and barrierless. The released energy ranged from −72.87 to −97.77 kcal/mol for the exothermic reaction of N2O reduction. As shown in Figure 3 and Figure 4, the Cu55 and Ni13@Cu42 clusters of catalysts successfully produced N2 and CO2 with barrier energies of 10.58 and 9.68 kcal/mol, respectively. Notably, the Ni13@Cu42 cluster possessed the lowest barrier for catalyzing the CO oxidation by the N2O reaction.
CO oxidation on the M13@Cu42 clusters by the E-R mechanism followed similar procedures as those occurring by the L-H mechanism; the difference is whether the CO reactant was physisorbed on the catalyst or not. In the transition state, only one O atom of the reaction intermediates was bound with Cu atoms. As a result, the corresponding barriers were 10.97 kcal/mol and 22.37 kcal/mol, which was higher than those of the L-H mechanism, as shown in Table 2.
Figure 5 summarizes the energy barriers of the rate-limiting step to CO oxidation by the Cu55, Co13@Cu42, Ni13@Cu42, and Ru13@Cu42 clusters in this work along with some other reported atomically dispersed catalysts. The barriers of the rate-limiting step to CO oxidation via the L-H mechanism for these four kinds of nanoclusters were obviously lower than that of Cu13 (16.60 kcal/mol) [29], Cu-graphene (CuG:19.14 kcal/mol), [46] and Fe-graphene (FeG:19.37 kcal/mol) [47] systems and were comparable with that of a Ru@Cu12 (11.66 kcal/mol) [29] catalyst. In addition, the barriers of the rate-limiting step to CO oxidation via the E-R mechanism for these four catalysts were also lower than that of Cu12 (18.45 kcal/mol) [28] and SiN4G (16.60 kcal/mol) [48], comparable with that of Pt@Cu12 (14.81 kcal/mol) [29] and V@Au12 (15.68 kcal/mol) [49], and higher than that of Cr@Au12 (4.15 kcal/mol or 6.23 kcal/mol) [49], showing great potential in the reaction.
To date, computations refer to hypothetically ideal conditions, where finite temperature and constant pressure are generally desired in experiments. Thus, the Gibbs free energy (ΔG) of N2O reduction by the CO reaction at 1 atm pressure was calculated and served as a function of a temperature of 298.15 K for each basic procedure; it was computed as:
Δ G = Δ E + Δ E Z P E T Δ S ,
where ΔE and ΔEZPE are the calculated DFT total energy, calculated zero-point energy, and entropic corrections (T Δ S) at T = 298.15 K. For gas phase molecules, the values of ZPE and S came from the NIST database [50]. The ΔG, ZPE, and S of the reactant by the thermodynamics systems were calculated (Tables S2 and S3). The reaction occurred readily under ambient conditions in the case of the M13@Cu42 cluster clusters (except for Pd13@Cu42), as shown in Figure 6.
The results clearly indicated that Cu55 and doped M13@Cu42 clusters (M = Co, Ni, Zn, Ru, Rh, Pt) enable the reaction process to properly proceed under ambient conditions. Therefore, Cu55 and doped M13@Cu42 clusters are promising catalysts that benefited from the fact that the nitrogen-oxygen bonds were broken without an energy barrier and that the nitrogen molecule readily separated from the cluster. At the same time, these catalysts were able to perform at low temperatures.
The CO oxidation by N2O reaction evaluation results showed that doping with different elements of the M13@Cu42 clusters and different reaction mechanisms had an important effect on the reaction activity. Extensive previous theoretical and experimental works have demonstrated that the rate-limiting step of CO oxidation by the N2O reaction depended on the properties of the catalysts. For example, N2O reduction possessed a higher energy barrier than CO oxidation on the Ag7Au6 catalyst [51]. Notably, the potential energy curve analysis indicated that the catalytic activity was tuned by controlling the different active mechanisms. Compared with the E-R mechanism, the bimetallic cluster catalyst under the L-H mechanism reduced the energy barrier of the CO oxidation process. According to the charge transfer in the two mechanisms, more electrons were transferred (approximately 0.20|e|) in the presence of CO, which was favorable compared with the absence of CO. The co-adsorbed CO promoted the conduction of the reaction to some extent. Therefore, the structural robustness and chemical tunability are the prominent advantages of the doped metal cage clusters, making them become a promising family of nanometer catalysts with practical application prospects.

3.3. Electronic Structure Analysis

The electronic structure of these bimetallic clusters was deeply analyzed to further understand the catalytic activity of the M13@Cu42 cluster. The surfaces of all of these clusters exhibited significant charge densities, which corresponded to the electronic states near the Fermi energy level. The electron configuration of N2O was 7 σ 2 2 π 4 3 π 0 , and the frontier molecular orbitals (FMOs) consisted of π –orbitals, where the 2 π –HOMO orbital was the bonding orbital of the N–N bond and the antibonding orbital of the N–O bond. 3 π –LUMO was the strong antibonding orbital between all atoms [52].
There was a small charge transfer from the N2O molecule to the metallic clusters when the N2O molecule was attached to the surface of clusters through its O end. When N2O was adsorbed, a sizeable electron density rearrangement appeared on the shell atoms; the core atoms were not significantly affected. In the alloy metal clusters, the electrons on the shell Cu atoms were transferred to the core dopant atom. Due to the reduction of electrons in the shell Cu atoms, the electrophilicity of the Cu atoms was enhanced, which makes it easier for alloy metal clusters to adsorb N2O compared with pure Cu55 clusters. The adsorbability of the M13Cu42 cluster was better than that of the Cu55 cluster in the presence of greater electron transfer, which was consistent with the adsorption energy analysis. CO oxidation served as the rate-limiting step during the reaction, and a cluster complex with residual O was produced in both mechanisms; therefore, the interaction between O and CO played an important role. The CO molecule was attached to the metallic clusters through the C atom, and the electron of the CO molecule was transferred to the metallic clusters. For example, the number of charge transfers from CO to the bimetallic clusters was larger than that from CO to the pure Cu55 cluster, which led to a higher catalytic performance for CO oxidation on the alloy clusters.
All of these clusters show prominent charge densities on the Cu cage surface, which correspond to the electronic states near the Fermi level and are responsible for the chemical reactivity. As shown in Figure 7a, the M13@Cu42 cluster with a lower d orbital center provided a stronger adsorption strength with the CO molecule. Such a trend of activity was also observed in previous studies [49,53]. This interaction makes it easier for N2O to obtain electrons, meaning that the N–O bond of N2O is more easy to break. Intuitively, less charge transfer between M–Cu indicates a weaker bonding between M and Cu atoms and an enhanced unsaturation of the Cu outer cage, resulting in a higher reactivity of the surface Cu atoms to CO. The activity of the M13@Cu42 clusters can be further associated with the d orbital center of the cluster, defined as [54]:
ε d = 0 E D E d E 0 D E d E ,
where D€ is the local density of states (LDOSs) of the d orbitals of the cluster at a given energy E; the integral is taken from all occupied states, and the highest occupied molecular orbital (HOMO) is set to zero. The LDOSs of the M13@Cu42 clusters are shown in Figure 7b and Figure S8.
According to the picture of the extended Hückel theory [55], a deeper d orbital level of the catalyst leads to a lower hopping matrix element and stronger binding strength with the adsorbate. Therefore, the binding ability and activity of the M13@Cu42 clusters are related to the d orbital centers of the clusters, and the catalytic performance can be optimized by selecting suitable doping elements and even by designing ideal catalysts for various reactions.

4. Conclusions

In conclusion, M13@Cu42 (M = Cu, Co, Ni, Zn, Ru, Rh, Pd, Pt) core–shell clusters of aromatic-like inorganic and metal compounds, where transition metal atoms acted as the core for the selective catalytic reduction of N2O via CO, were systematically investigated using periodic spin-polarized first-principles calculations. The results show that the stability of these M13@Cu42 clusters are significantly higher than that of the intrinsic Cu55 cluster. Meanwhile, the total charge transfers from the shell to the central doped atoms were shown to increase. The doping of the central metal atom affected the catalytic and electronic properties of the clusters. A portion of the M13@Cu42 clusters that had suitable binding capacities comprised a low potential barrier. Especially, the kinetic barrier of Ni13@Cu42 was 9.68 kcal/mol for CO oxidation under the L-H mechanism. The L-H mechanism, which stemmed from gas molecule co-adsorption on the M13@Cu42 clusters and transition metal modification doping, played a key role in the adsorption of CO oxidation. Our calculations revealed the use of endoplasmically doped copper clusters as a novel stable subnanocatalyst, which is a promising material for high-performance catalytic media.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28114485/s1.

Author Contributions

Z.L.: Investigation, data analysis, writing of the manuscript; H.W.: methodology, software, data analysis; Y.G.: supervision, funding, resources; J.Z.: conceptualization, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China under Grant No. 11864033 and 12264043.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Trogler, W.C. Physical properties and mechanisms of formation of nitrous oxide. Coord. Chem. Rev. 1999, 187, 303–327. [Google Scholar] [CrossRef]
  2. Ravishankara, A.; Daniel, J.S.; Portmann, R.W. Nitrous oxide (N2O): The dominant ozone-depleting substance emitted in the 21st century. Science 2009, 326, 123–125. [Google Scholar] [CrossRef] [PubMed]
  3. Dameris, M. Depletion of the ozone layer in the 21st century. Angew. Chem. Int. Ed. 2010, 49, 489–491. [Google Scholar] [CrossRef] [PubMed]
  4. Esrafili, M.D.; Janebi, H.; Mousavian, P. Single Al atom anchored on defective MoS2: An efficient catalytic site for reduction of greenhouse N2O gas by CO or C2H4 molecules. Appl. Surf. Sci. 2021, 569, 151001. [Google Scholar] [CrossRef]
  5. Shao, L.; Chen, J.; Wang, K.; Mei, J.; Tan, T.; Wang, G.; Liu, K.; Gao, X. Highly precise measurement of atmospheric N2O and CO using improved White cell and RF current perturbation. Sens. Actuators B Chem. 2022, 352, 130995. [Google Scholar] [CrossRef]
  6. Esrafili, M.D.; Nejadebrahimi, B. N2O reduction over a porous Si-decorated carbon nitride fullerene: A DFT study. Chem. Phys. Lett. 2019, 716, 11–16. [Google Scholar] [CrossRef]
  7. Sittiwong, J.; Jaturajamrenchai, T.; Wongkampuan, P.; Somwatcharajit, N.; Impeng, S.; Maihom, T.; Probst, M.; Limtrakul, J. Modulating the catalytic activity of metal-organic frameworks for CO oxidation with N2O through an oriented external electric field. Mol. Catal. 2021, 516, 111970. [Google Scholar] [CrossRef]
  8. Amaya-Roncancio, S.; Reinaudi, L.; Gimenez, M.C. Adsorption and dissociation of CO on metal clusters. Mater. Today Commun. 2020, 24, 101158. [Google Scholar] [CrossRef]
  9. Chen, P.; Gu, M.; Chen, G.; Liu, F.; Lin, Y. DFT study on the reaction mechanism of N2O reduction with CO catalyzed by char. Fuel 2019, 254, 115666. [Google Scholar] [CrossRef]
  10. Khan, A.A.; Ahmad, R.; Ahmad, I. Removal of nitrous and carbon mono oxide from flue gases by Si-coordinated nitrogen doped C60-fullerene: A DFT approach. Mol. Catal. 2021, 509, 111674. [Google Scholar] [CrossRef]
  11. You, Y.; Chen, S.; Li, J.; Zeng, J.; Chang, H.; Ma, L.; Li, J. Low-temperature selective catalytic reduction of N2O by CO over Fe-ZSM-5 catalysts in the presence of O2. J. Hazard. Mater. 2020, 383, 121117. [Google Scholar] [CrossRef] [PubMed]
  12. Boehme, D.K.; Schwarz, H. Gas-phase catalysis by atomic and cluster metal ions: The ultimate single-site catalysts. Angew. Chem. Int. Ed. 2005, 44, 2336–2354. [Google Scholar] [CrossRef]
  13. Delabie, A.; Vinckier, C.; Flock, M.; Pierloot, K. Evaluating the activation barriers for transition metal N2O reactions. J. Phys. Chem. A 2001, 105, 5479–5485. [Google Scholar] [CrossRef]
  14. Campa, M.C.; Doyle, A.M.; Fierro, G.; Pietrogiacomi, D. Simultaneous abatement of NO and N2O with CH4 over modified Al2O3 supported Pt, Pd, Rh. Catal. Today 2022, 384, 76–87. [Google Scholar] [CrossRef]
  15. Okemoto, A.; Harada, M.R.; Ishizaka, T.; Hiyoshi, N.; Sato, K. Catalytic performance of MoO3/FAU zeolite catalysts modified by Cu for reverse water gas shift reaction. Appl. Catal. A Gen. 2020, 592, 117415. [Google Scholar] [CrossRef]
  16. Kim, K.; Baek, S.; Kim, J.J.; Han, J.W. Catalytic decomposition of N2O on PdxCuy alloy catalysts: A density functional theory study. Appl. Surf. Sci. 2020, 510, 145349. [Google Scholar] [CrossRef]
  17. Richards, N.; Carter, J.H.; Parker, L.A.; Pattisson, S.; Hewes, D.G.; Morgan, D.J.; Davies, T.E.; Dummer, N.F.; Golunski, S.; Hutchings, G.J. Lowering the operating temperature of perovskite catalysts for N2O decomposition through control of preparation methods. ACS Catal. 2020, 10, 5430–5442. [Google Scholar] [CrossRef]
  18. Li, X.; Kuznetsov, A.E.; Zhang, H.-F.; Boldyrev, A.I.; Wang, L.-S. Observation of all-metal aromatic molecules. Science 2001, 291, 859–861. [Google Scholar] [CrossRef]
  19. Zhu, C.; Li, S.; Luo, M.; Zhou, X.; Niu, Y.; Lin, M.; Zhu, J.; Cao, Z.; Lu, X.; Wen, T. Stabilization of anti-aromatic and strained five-membered rings with a transition metal. Nat. Chem. 2013, 5, 698–703. [Google Scholar] [CrossRef]
  20. Niu, T.; Liu, G.; Chen, Y.; Yang, J.; Wu, J.; Cao, Y.; Liu, Y. Hydrothermal synthesis of graphene-LaFeO3 composite supported with Cu-Co nanocatalyst for higher alcohol synthesis from syngas. Appl. Surf. Sci. 2016, 364, 388–399. [Google Scholar] [CrossRef]
  21. Wang, X.; Qiu, S.; Feng, J.; Tong, Y.; Zhou, F.; Li, Q.; Song, L.; Chen, S.; Wu, K.H.; Su, P. Confined Fe–Cu clusters as sub-nanometer reactors for efficiently regulating the electrochemical nitrogen reduction reaction. Adv. Mater. 2020, 32, 2004382. [Google Scholar] [CrossRef]
  22. Nicholas, K.M.; Lander, C.; Shao, Y. Computational Evaluation of Potential Molecular Catalysts for Nitrous Oxide Decomposition. Inorg. Chem. 2022, 61, 14591–14605. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, Y.; Huang, L.; Zhu, X.; Fang, Y.; Dong, S. Coupling Cu with Au for enhanced electrocatalytic activity of nitrogen reduction reaction. Nanoscale 2020, 12, 1811–1816. [Google Scholar] [CrossRef]
  24. Kim, C.; Kim, J. Comparative evaluation of artificial neural networks for the performance prediction of Pt-based catalysts in water gas shift reaction. Int. J. Energy Res. 2022, 46, 9602–9620. [Google Scholar] [CrossRef]
  25. Ko, B.H.; Hasa, B.; Shin, H.; Jeng, E.; Overa, S.; Chen, W.; Jiao, F. The impact of nitrogen oxides on electrochemical carbon dioxide reduction. Nat. Commun. 2020, 11, 5856. [Google Scholar] [CrossRef] [PubMed]
  26. Pan, K.L.; Young, C.W.; Pan, G.T.; Chang, M.B. Catalytic reduction of NO by CO with Cu-based and Mn-based catalysts. Catal. Today 2020, 348, 15–25. [Google Scholar] [CrossRef]
  27. Andana, T.; Rappé, K.G.; Nelson, N.C.; Gao, F.; Wang, Y. Selective catalytic reduction of NOx with NH3 over Ce-Mn oxide and Cu-SSZ-13 composite catalysts–Low temperature enhancement. Appl. Catal. B Environ. 2022, 316, 121522. [Google Scholar] [CrossRef]
  28. Barabás, J.L.; Höltzl, T. Reaction of N2O and CO catalyzed with small copper clusters: Mechanism and design. J. Phys. Chem. A 2016, 120, 8862–8870. [Google Scholar] [CrossRef]
  29. Lian, X.; Guo, W.; He, B.; Yu, B.; Chen, S.; Qin, D.; Chen, F. Insights of the mechanisms for CO oxidation by N2O over M@Cu12 (M = Cu, Pt, Ru, Pd, Rh) core-shell clusters. Mol. Catal. 2020, 494, 111126. [Google Scholar] [CrossRef]
  30. Jiang, T.; Mowbray, D.; Dobrin, S.; Falsig, H.; Hvolbæk, B.; Bligaard, T.; Nørskov, J.K. Trends in CO oxidation rates for metal nanoparticles and close-packed, stepped, and kinked surfaces. J. Phys. Chem. C 2009, 113, 10548–10553. [Google Scholar] [CrossRef]
  31. Mao, X.; Wang, L.; Xu, Y.; Wang, P.; Li, Y.; Zhao, J. Computational high-throughput screening of alloy nanoclusters for electrocatalytic hydrogen evolution. npj Comput. Mater. 2021, 7, 46. [Google Scholar] [CrossRef]
  32. Cao, X.; Ji, Y.; Luo, Y. Dehydrogenation of propane to propylene by a Pd/Cu single-atom catalyst: Insight from first-principles calculations. J. Phys. Chem. C 2015, 119, 1016–1023. [Google Scholar] [CrossRef]
  33. Liu, Q.; Wang, X.; Li, L.; Song, K.; Qian, P.; Feng, Y.P. Design of platinum single-atom doped metal nanoclusters as efficient oxygen reduction electrocatalysts by coupling electronic descriptor. Nano Res. 2022, 15, 7016–7025. [Google Scholar] [CrossRef]
  34. Liu, C.; Dong, H.; Ji, Y.; Rujisamphan, N.; Li, Y. High-performance hydrogen evolution reaction catalysis achieved by small core-shell copper nanoparticles. J. Colloid Interface Sci. 2019, 551, 130–137. [Google Scholar] [CrossRef]
  35. Zhang, R.; Wang, Y.; Wang, B.; Ling, L. Probing into the effects of cluster size and Pd ensemble as active center on the activity of H2 dissociation over the noble metal Pd-doped Cu bimetallic clusters. Mol. Catal. 2019, 475, 110457. [Google Scholar] [CrossRef]
  36. Zhao, B.; Zhang, R.; Huang, Z.; Wang, B. Effect of the size of Cu clusters on selectivity and activity of acetylene selective hydrogenation. Appl. Catal. A Gen. 2017, 546, 111–121. [Google Scholar] [CrossRef]
  37. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef] [PubMed]
  38. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  39. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  40. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  41. 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]
  42. Tang, D.; Chen, Z.; Hu, J.; Sun, G.; Lu, S.; Hu, C. CO oxidation catalyzed by silver nanoclusters: Mechanism and effects of charge. Phys. Chem. Chem. Phys. 2012, 14, 12829–12837. [Google Scholar] [CrossRef]
  43. Austin, N.; Butina, B.; Mpourmpakis, G. CO2 activation on bimetallic CuNi nanoparticles. Prog. Nat. Sci. Mater. Int. 2016, 26, 487–492. [Google Scholar] [CrossRef]
  44. McEuen, P.; Kittel, C. Introduction to Solid State Physics; John Wiley & Sons: Hoboken, NJ, USA, 2005. [Google Scholar]
  45. Tang, D.; Zhang, J. Theoretical investigation on CO oxidation catalyzed by a copper nanocluster. RSC Adv. 2013, 3, 15225–15236. [Google Scholar] [CrossRef]
  46. Akça, A.; Karaman, O.; Karaman, C. Mechanistic insights into catalytic reduction of N2O by CO over Cu-embedded graphene: A density functional theory perspective. ECS J. Solid State Sci. Technol. 2021, 10, 041003. [Google Scholar] [CrossRef]
  47. Wannakao, S.; Nongnual, T.; Khongpracha, P.; Maihom, T.; Limtrakul, J. Reaction mechanisms for CO catalytic oxidation by N2O on Fe-embedded graphene. J. Phys. Chem. C 2012, 116, 16992–16998. [Google Scholar] [CrossRef]
  48. Junkaew, A.; Namuangruk, S.; Maitarad, P.; Ehara, M. Silicon-coordinated nitrogen-doped graphene as a promising metal-free catalyst for N2O reduction by CO: A theoretical study. RSC Adv. 2018, 8, 22322–22330. [Google Scholar] [CrossRef]
  49. Zhou, S.; Pei, W.; Du, Q.; Zhao, J. Foreign atom encapsulated Au 12 golden cages for catalysis of CO oxidation. Phys. Chem. Chem. Phys. 2019, 21, 10587–10593. [Google Scholar] [CrossRef]
  50. Chase, M.W. NIST–JANAF thermochemical tables for the bromine oxides. J. Phys. Chem. Ref. Data 1996, 25, 1069–1111. [Google Scholar] [CrossRef]
  51. Wongnongwa, Y.; Namuangruk, S.; Kungwan, N.; Jungsuttiwong, S. Mechanistic study of CO oxidation by N2O over Ag7Au6 cluster investigated by DFT methods. Appl. Catal. A Gen. 2017, 538, 99–106. [Google Scholar] [CrossRef]
  52. Piskorz, W.; Zasada, F.; Stelmachowski, P.; Kotarba, A.; Sojka, Z. DFT modeling of reaction mechanism and ab initio microkinetics of catalytic N2O decomposition over alkaline earth oxides: From molecular orbital picture account to simulation of transient and stationary rate profiles. J. Phys. Chem. C 2013, 117, 18488–18501. [Google Scholar] [CrossRef]
  53. Pei, W.; Zhou, S.; Bai, Y.; Zhao, J. N-doped graphitic carbon materials hybridized with transition metals (compounds) for hydrogen evolution reaction: Understanding the synergistic effect from atomistic level. Carbon 2018, 133, 260–266. [Google Scholar] [CrossRef]
  54. Zhou, S.; Yang, X.; Pei, W.; Liu, N.; Zhao, J. Heterostructures of MXenes and N-doped graphene as highly active bifunctional electrocatalysts. Nanoscale 2018, 10, 10876–10883. [Google Scholar] [CrossRef] [PubMed]
  55. Xin, H.; Linic, S. Communications: Exceptions to the d-band model of chemisorption on metal surfaces: The dominant role of repulsion between adsorbate states and metal d-states. J. Chem. Phys. 2010, 132, 221101. [Google Scholar] [CrossRef]
Figure 1. (a) Structure of the pristine Cu55 cluster; (b) atomic arrangement of one of the 20 equivalent facets in the Cu55 cluster; (c) TM-substituted M13@Cu42 clusters (TM = Co, Ni, Zn, Ru, Rh, Pd, Pt). The TM and Cu atoms are shown in blue and red, respectively.
Figure 1. (a) Structure of the pristine Cu55 cluster; (b) atomic arrangement of one of the 20 equivalent facets in the Cu55 cluster; (c) TM-substituted M13@Cu42 clusters (TM = Co, Ni, Zn, Ru, Rh, Pd, Pt). The TM and Cu atoms are shown in blue and red, respectively.
Molecules 28 04485 g001
Figure 2. Calculated binding energies of the intrinsic Cu55 cluster and M13@Cu42 (M = Cu, Co, Ni, Zn, Ru, Rh, Pd, Pt)-substituted clusters (absolute value).
Figure 2. Calculated binding energies of the intrinsic Cu55 cluster and M13@Cu42 (M = Cu, Co, Ni, Zn, Ru, Rh, Pd, Pt)-substituted clusters (absolute value).
Molecules 28 04485 g002
Figure 3. (a) N2O decomposition to N2 and O* and (b) CO oxidation by the remaining O* on the Cu55 cluster. (c) Corresponding optimized intermediates and transition states involved in N2O decomposition and CO oxidation on the Cu55 cluster (black numbers represent the bond length).
Figure 3. (a) N2O decomposition to N2 and O* and (b) CO oxidation by the remaining O* on the Cu55 cluster. (c) Corresponding optimized intermediates and transition states involved in N2O decomposition and CO oxidation on the Cu55 cluster (black numbers represent the bond length).
Molecules 28 04485 g003
Figure 4. (a) N2O decomposition to N2 and O* and (b) CO oxidation by the remaining O* on the Ni13@Cu42 cluster. (c) Corresponding optimized intermediates and transition states involved in N2O decomposition and CO oxidation on the Ni13@Cu42 cluster (black numbers represent the bond length).
Figure 4. (a) N2O decomposition to N2 and O* and (b) CO oxidation by the remaining O* on the Ni13@Cu42 cluster. (c) Corresponding optimized intermediates and transition states involved in N2O decomposition and CO oxidation on the Ni13@Cu42 cluster (black numbers represent the bond length).
Molecules 28 04485 g004
Figure 5. Energy barriers (TS) for the CO oxidation via the L-H and E-R mechanisms on Cu13/Co13/Ni13/Ru13@Cu42 clusters (orange circle) and the barriers for the CO oxidation in some reported studies (green triangle).
Figure 5. Energy barriers (TS) for the CO oxidation via the L-H and E-R mechanisms on Cu13/Co13/Ni13/Ru13@Cu42 clusters (orange circle) and the barriers for the CO oxidation in some reported studies (green triangle).
Molecules 28 04485 g005
Figure 6. Gibbs free energies (at 1 atm pressure and a temperature of 298.15 K) of the N2O reduction by CO reactions via the L-H mechanism for the M13@Cu42 clusters.
Figure 6. Gibbs free energies (at 1 atm pressure and a temperature of 298.15 K) of the N2O reduction by CO reactions via the L-H mechanism for the M13@Cu42 clusters.
Molecules 28 04485 g006
Figure 7. (a) d orbital centers as a function of ΔEN2O for different M13@Cu42 clusters. (b) Local density of states (LDOSs) of Ni-doped Ni13@Cu42 and Cu55 clusters. The black dashed lines and numbers next to them indicate the d orbital centers of each system.
Figure 7. (a) d orbital centers as a function of ΔEN2O for different M13@Cu42 clusters. (b) Local density of states (LDOSs) of Ni-doped Ni13@Cu42 and Cu55 clusters. The black dashed lines and numbers next to them indicate the d orbital centers of each system.
Molecules 28 04485 g007
Table 1. Structural, energetic, and electronic properties of M13@Cu42 clusters, including the average M–M and M–Cu bond lengths (dCu–Cu, dM–Cu), binding energies (Eb), magnetic moment (Mag), average charges transferred from M to Cu atoms (CTM), average charges transferred from the shell-Cu atoms to the core-M metal atoms (CTCu), d orbital centers (εd), and HOMO–LUMO gaps (Eg).
Table 1. Structural, energetic, and electronic properties of M13@Cu42 clusters, including the average M–M and M–Cu bond lengths (dCu–Cu, dM–Cu), binding energies (Eb), magnetic moment (Mag), average charges transferred from M to Cu atoms (CTM), average charges transferred from the shell-Cu atoms to the core-M metal atoms (CTCu), d orbital centers (εd), and HOMO–LUMO gaps (Eg).
MdCu-M (Å)dM-M (Å)Eb (kcal/mol)Mag (μB)CTM (e)CTCu (e)εd (eV)Eg (eV)
Cu2.422.49−70.103+0.06−0.0172.420.1
Co2.422.45−77.4821−0.050.0172.210.01
Ni2.432.45−77.488−0.010.0032.070.01
Zn2.442.59−55.812−0.160.0523.510.59
Ru2.622.61−90.404+0.01−0.0022.450.02
Rh2.482.69−81.869+0.12−0.0352.220.02
Pd2.482.70−70.792+0.18−0.0542.110.17
Pt2.652.82−80.710+0.28−0.0862.260.16
Table 2. Adsorption energies of N2O and CO molecules on M13@Cu42 clusters (ΔEN2O, ΔECO), adsorption energies of co-adsorbed N2O and CO molecules on M13@Cu42 clusters (ΔEN2O-CO, ΔEO-CO), charge transfer between the molecule and cluster (CT), and energy barriers of the CO oxidation (TS).
Table 2. Adsorption energies of N2O and CO molecules on M13@Cu42 clusters (ΔEN2O, ΔECO), adsorption energies of co-adsorbed N2O and CO molecules on M13@Cu42 clusters (ΔEN2O-CO, ΔEO-CO), charge transfer between the molecule and cluster (CT), and energy barriers of the CO oxidation (TS).
E-RL-H
M13@Cu42ΔEN2O (kcal/mol)ΔECO (kcal/mol)CT (e)TS (kcal/mol)ΔEN2O-CO (kcal/mol)ΔEO-CO (kcal/mol)CT (e)TS (kcal/mol)
Cu−53.27−27.440.9711.64−72.87−7.611.0910.58
Co−50.04−28.591.0714.75−77.48−2.771.109.80
Ni−46.12−32.511.0810.97−74.02−5.531.089.68
Zn−50.27−28.360.9717.35−78.40−2.311.0919.58
Ru−50.04−28.361.1112.90−86.71−5.771.1211.99
Rh−50.96−27.671.1013.99−84.40−2.541.1312.53
Pd−70.33−20.061.0922.37−97.77−0.921.1818.22
Pt−57.88−20.751.1114.93−75.41−4.841.1212.47
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Z.; Wang, H.; Gao, Y.; Zhao, J. Mechanisms in the Catalytic Reduction of N2O by CO over the M13@Cu42 Clusters of Aromatic-like Inorganic and Metal Compounds. Molecules 2023, 28, 4485. https://doi.org/10.3390/molecules28114485

AMA Style

Liu Z, Wang H, Gao Y, Zhao J. Mechanisms in the Catalytic Reduction of N2O by CO over the M13@Cu42 Clusters of Aromatic-like Inorganic and Metal Compounds. Molecules. 2023; 28(11):4485. https://doi.org/10.3390/molecules28114485

Chicago/Turabian Style

Liu, Ziyang, Haifeng Wang, Yan Gao, and Jijun Zhao. 2023. "Mechanisms in the Catalytic Reduction of N2O by CO over the M13@Cu42 Clusters of Aromatic-like Inorganic and Metal Compounds" Molecules 28, no. 11: 4485. https://doi.org/10.3390/molecules28114485

Article Metrics

Back to TopTop