DOI:
10.1039/C5RA11695C
(Paper)
RSC Adv., 2015,
5, 76651-76659
Catalytic oxidation of CO by N2O on neutral Y2MO5 (M = Y, Al) clusters: a density functional theory study†
Received
18th June 2015
, Accepted 4th September 2015
First published on 4th September 2015
Abstract
Density functional theory (DFT) calculations are employed to investigate the full catalytic cycle of CO oxidation by N2O on yttrium oxide clusters Y2MO5 (M = Y, Al) in the gas-phase. Extensive structural searches show that both the ground-state structures of Y3O5 and Y2AlO5 contain an oxygen radical (Ot˙) which plays an important role in CO oxidation. Energy profiles are calculated to determine the reaction mechanisms. Molecular electrostatic potential maps (MEPs) and natural bond orbital (NBO) analyses are employed to rationalize the reaction mechanisms. The results indicate that the whole catalytic cycle for the reaction CO + N2O → CO2 + N2, conducted by yttrium oxide clusters Y2MO5 (M = Y, Al), is favored both thermodynamically and kinetically. Moreover, compared with the previous report on di-nuclear YAlO3+˙ and Y2O3+˙, it's obvious we can conclude that tri-nuclear Y3O5 and Y2AlO5 exhibit greatly enhanced catalytic activity toward CO/N2O couples.
1. Introduction
Catalytic oxidation of harmful gases, such as carbon monoxide (CO) into carbon dioxide (CO2) and nitrous oxide (N2O) into nitrogen (N2), is of great importance both environmentally and economically.1–10 Even though oxidation of CO by N2O is exothermic, it does not occur directly at any measurable temperatures because of high energy barrier for the CO/N2O couple.11 It has been verified that some transition metal oxides are effective in catalytic oxidation of CO by N2O.12–18 Mechanistic investigation on the oxygen atom transfer process mediated by transition metal oxides which are involved in CO oxidation by N2O is meaningful to design better catalysts with high activity and selectivity. Gas-phase cluster study has been emerged as an alternative approach to provide molecular-level insights into the complex chemical reaction processes occurring on the bulk transition metal oxide.19–28
The catalytic oxidation of CO by N2O at room temperature (RT) in the gas phase was demonstrated for the first time by Kappes and Staley with FeO+ as a catalyst.12 In the ensuing decades, numerous relevant investigations were developed subsequently.13–16 For example, Castleman and co-workers have reported that a series of zirconium oxide (ZrO2)x+ (x = 2–5) cations as well as (ZrxO2x+1)− (x = 1–4) anions exhibit high activity in CO oxidation and the oxygen radical center with elongated zirconium–oxygen bond is the active site.13–15 They also have shown that the initial zirconium oxide ions can be regenerated by treating oxygen-deficient zirconium-oxide clusters with N2O. Besides, binary neutral metal oxide clusters ZrScO4 and ZrNbO5 were demonstrated to have similar reactivity with their isoelectronic ions Zr2O4+ and Zr2O5− for CO oxidation and N2O reduction.16 Some other metal-oxide clusters, such as CemO2m+ (m = 2–6), CenO2n+1− (n = 4–21), (TiO2)nO− (n = 3–25) and VO3, were also exploited to bring out their oxidizing capacity towards CO.29–32 These investigations have indicated the oxygen radical center plays a critical role in CO oxidation process.
Yttrium oxides are also promising for the catalytic application, yet the related studies on CO oxidation by yttrium oxides are relatively scarce.11,33 Many early studies have focused their attentions on the electronic structure and bonding of yttrium oxide clusters for their unique properties.34–38 The previous study has pointed out that the YAlO3+˙ cluster contains oxygen radical center and shows better reactivity than homonuclear yttrium oxide cluster Y2O3+˙ in CO oxidation.11 Recently, the terminal oxygen radical was also found in the neutral Y3O5 cluster.38
In this work, we present our theoretical study on CO oxidations by neutral Y2MO5 (M = Y, Al) clusters employing density functional calculations. Extensive structural searches show that both Y3O5 and Y2AlO5 clusters are doublet state with a single unpaired electron located on the terminal oxygen atom. The energy profiles are calculated to get the detail mechanisms underlying the oxygen atom transfer processes involved in CO oxidation reaction. Qualitative explanation for the reactions of Y2MO5 (M = Y, Al) clusters with CO can be understood from the molecular electrostatic potential maps and natural bond orbital analyses. The regenerations of Y2MO5 (M = Y, Al) clusters via the reactions of oxygen-deficient clusters Y2MO4 (M = Y, Al) with N2O are also calculated. A full catalytic cycle for the reaction of CO with N2O on Y2MO5 (M = Y, Al) clusters is thus complete.
2. Computational methods
The theoretical calculations were performed at the DFT level using the BP86 functional.39,40 The calculations were performed using analytical gradients with the Stuttgart relativistic small core basis set and efficient core potential41,42 augmented with two f-type and one g-type polarization functions for yttrium [ζ(f) = 0.144, 0.546; ζ(g) = 0.249] as recommended by Martin and Sundermann43 and the aug-cc-pvTZ basis set for oxygen, alumina, carbon and nitrogen.44,45 Scalar relativistic effects, that is, the mass velocity and Darwin effects, were taken into account via the quasi-relativistic pseudopotentials. All the structures presented in the calculated energy profiles were fully optimized. The initial structures of the transition state were obtained by potential energy surface scans with appropriate coordinates. Vibrational frequency calculations were performed at the same level of theory to verify the nature of the stationary points as minima (zero imaginary frequency) or transition states (one imaginary frequency). The pathways of the reaction mechanisms were further confirmed by intrinsic reaction coordinate (IRC) calculations.46,47
The relative energies of the low-lying Y2MOx (M = Y, Al; x = 4–5) structures were further evaluated via single-point calculations at the coupled cluster [CCSD(T)]48–52 level with the Y/Stuttgart+2f1g/O/Al/aug-cc-pVTZ basis sets at the BP86 geometries. The results are summarized and available in ESI (Table S1†). In our current studies, the BP86 gave superior results in terms of energies when directly compared to the results of CCSD(T) calculations. The previous studies on yttrium oxide clusters38,53 also have indicated that the BP86 calculations can give good consistency with the experimental results. As discussed below, we used the results with the BP86 functional for further discussion. All DFT calculations were performed with the Gaussian 09 software package54 and the CCSD(T) calculations were done using MOLPRO 2010.1 package.55 Natural bond orbital (NBO) analysis was performed by using the NBO 3.1 program.56 Three-dimensional molecular structures were visualized using the GaussView 4.1.57 Three-dimensional contours of the molecular orbitals were visualized using the VMD software.58
3. Results and discussions
3.1 The structures and spin density analyses of Y2MOx (M = Y, Al; x = 4–5)
The optimized ground state structures and low-lying isomers within 0.40 eV for the neutral Y3Ox and Y2AlOx (x = 4–5) clusters at the BP86 level are presented in Fig. 1 and 2. Alternative optimized geometries are shown in the ESI (Fig. S1 and 2†). In the following discussions, Ot, Ob and Oc stand for the terminal, bridging and capped oxygen atoms, respectively. The spin density analyses are performed to shed light on the nature of the radical character of these neutral ground states, as shown in Fig. 3.
 |
| Fig. 1 Optimized global minima and selected low-lying structures (within ∼0.40 eV) for Y3Ox (x = 4–5) clusters at the BP86 level. The bond lengths are in angstroms. | |
 |
| Fig. 2 Optimized global minima and selected low-lying structures (within ∼0.40 eV) for Y2AlOx (x = 4–5) clusters at the BP86 level. The bond lengths are in angstroms. | |
 |
| Fig. 3 Numerical electron spin density (in |e|) for the ground states of Y2MOx (M = Y, Al; x = 4–5) clusters. | |
3.1.1 Y3O4 and Y3O5. In the previous work, we have reported the structures of Y3O4 and Y3O5 clusters.38 The low-lying isomers within 0.40 eV for Y3O4 and Y3O5 are given in Fig. 1. The ground state of Y3O4 is a doublet state C3v (2A1) structure (Fig. 1a), which can be regarded as an umbellate structure with a capped oxygen atom and three bridging oxygen atoms. In our calculation, the bond lengths of Y–Oc and Y–Ob are 2.174 Å and 2.048 Å, respectively. Alternative optimized structures are located remarkably higher in energy (Fig. S1a–c†).The ground state of Y3O5 has a propensity to adopt a doublet (2A′′) state with Cs symmetry (Fig. 1b) at the BP86 level. The Y–Ot bond length is 2.155 Å and can be considered as Y–O single bond. Two kinds of connection between caped oxygen and yttrium atoms are found in this structure, where the Y–Oc distances are presented to be 2.134 Å and 2.309 Å, respectively. Two (2A′) state structures showing in Fig. 1c and d are calculated to be 0.11 eV and 0.21 eV above the ground state, respectively. We also calculated the higher symmetry structure (D3h) with two triply-bridging O atoms and three doubly-bridging O atoms for Y3O5, which converged to the Cs structure, as shown in Fig. 1d. The relative energies for these low-lying structures (within ∼0.20 eV) were further evaluated using single point CCSD(T) calculations. The results of CCSD(T) calculations are in line with those of DFT calculations.
3.1.2 Y2AlO4 and Y2AlO5. We carried out extensive structural searches for the ground-state structures on the potential energy surfaces of Y2AlO4 and Y2AlO5 clusters considering a variety of structural candidates including different spin states and geometries. The global minimum of Y2AlO4 is predicted to be a doublet state (2A′) with Cs symmetry (Fig. 2a), which shows resemblance with Y3O4 (C3v, 2A1) in geometry (Fig. 1a). The low-lying isomer C2v (2A1) (Fig. 2b) with four bridging O atoms is only 0.02 eV above the global minimum at BP86 level. Single-point CCSD(T) calculations were performed for these two isomers, revealing that the (Cs, 2A′) structure (Fig. 2a) is 0.12 eV more stable than the (C2v, 2A1) structure (Fig. 2b).As for Y2AlO5, selected optimized structures with their relative energies are given in Fig. 2c–f. The ground state of Y2AlO5 is a doublet state Cs (2A′′) as shown in Fig. 2c. It can be viewed that the Al atom replaces the yttrium atom which is link with the terminal oxygen atom in Y3O5. The Al–Ot distance is 1.767 Å and can be labeled as Al–O single bond.59 The (2A′) state with similar geometry is 0.12 eV above the ground state at the BP86 level. The C2v (2B2) (Fig. 2d) and C2v (2B1) (Fig. 2f), which own a terminal and four bridging oxygen atoms, are 0.08 eV and 0.31 eV higher in energy, respectively. These low-lying structures within 0.20 eV were further evaluated by CCSD(T) calculations and the isomer (Cs, 2A′′) (Fig. 2c) is still the lowest-energy structure.
3.1.3 Spin density analysis. All the ground states of Y2MOx (M = Y, Al; x = 4–5) clusters are doublet state. As mentioned above, the global minimum of Y3O4 is a C3v structure with a capped and three bridging oxygen atoms. The spin is distributed equally on three yttrium atoms, as shown in Fig. 3a. After doping an aluminum atom, the unpaired electron is mainly located on aluminum atom in Y2AlO4 (Fig. 3b). Both Y3O5 and Y2AlO5 clusters contain an oxygen radical (O˙) as depicted in Fig. 3c and d, respectively. The singly occupied molecular orbitals (SOMOs) of Y3O5 and Y2AlO5 (Fig. 4) indicate that the unpaired electron is located on the 2p orbitals of terminal oxygen atom.
 |
| Fig. 4 The singly occupied molecular orbitals for the ground states of Y2MO5 (M = Y, Al) clusters. | |
To our knowledge, the oxygen-centre radical is highly reactive. It's able to oxidize a variety of stable molecules including CO and many small organic species at low temperature, such as CH4, C2H4, C2H2 and so forth.14,60–63 He and co-workers have studied the reactions of CO with Y2O3+˙ and YAlO3+˙ by mass spectrometry and density functional theoretical.11 According to their results, oxidation of the CO by Y2O3+˙ is prevented by an energy barrier, while oxidation of the CO with YAlO3+˙ is favored. The differences are mainly produced by the spin distributions. In Y2O3+˙, the unpaired electron is delocalized over three bridging oxygen atoms. As for YAlO3+˙, the spin is mainly located on the terminal oxygen atom of Al–Ot˙ unit. However, the single unpaired electrons of Y3O5 and Y2AlO5 clusters both located on the terminal oxygen atom of M–Ot˙ (M = Y, Al) subunit. In the next section, the mechanisms for CO oxidation by Y3O5 and Y2AlO5 are presented.
3.2 The reactions of Y2MO5 (M = Y, Al) with CO
The detail mechanisms for the reactions of Y2MO5 (M = Y, Al) with CO are investigated at the BP86 level. In our calculation, all the intermediates and transition structures in doublet state are much more stable than those in quartet state. So DFT calculations are performed for the molecule reactions of Y2MO5 (M = Y, Al) towards CO on the doublet ground state potential energy surfaces. The products are proposed according to the following equations:
Y2MO5 (M = Y, Al) + CO → Y2MO4 (M = Y, Al) + CO2 |
We have considered the possible paths for the reactions and found the way in which CO reacts with the M–Ot˙ site is more preferred. Then we will focus on the mechanism for CO oxidation involving the oxygen radical center. Fig. 5 and 6 present the favored energy profiles calculated at the BP86 level in which IMn represent the intermediate structures and TSn correspond to the transition states, respectively. The detail of electron transfer procedures can be obtained by analyzing the spin density of the intermediates and transition states in the reactions as shown in Fig. S8 and 9.† The Cartesian coordinates for the optimized structures on the potential energy surfaces are given in Table S3.† Other possible reaction paths were also calculated and given in Fig. S3–S7.†
 |
| Fig. 5 DFT calculated energy profile for the reaction of Y3O5 cluster with CO2. The reactants, intermediates, transition states and produces of the reaction are denoted as R, IMn, TSn and P, respectively. The energies are given in eV and bond lengths are in angstroms. | |
 |
| Fig. 6 DFT calculated energy profile for the reaction of Y2AlO5 cluster with CO. The reactants, intermediates, transition states and produces of the reaction are denoted as R, IMn, TSn and P, respectively. The energies are given in eV and bond lengths are in angstroms. | |
3.2.1 The reaction profiles of Y2MO5 (M = Y, Al) with CO.
3.2.1.1 Y3O5 with CO. The calculated energy profile for the reaction between Y3O5 and CO is presented in Fig. 5. The oxidation reaction occurs through the exothermic (0.29 eV) binding of the carbon atom of CO to the radical oxygen and the adjacent yttrium atom of Y3O5 forming an intermediate (IM1). In IM1, the distances of carbon atom to yttrium atom and the radical oxygen are 2.699 Å and 2.749 Å, respectively. Formation of the IM2 requires a transfer of spin from the terminal oxygen to the nonlinear CO2 unit. This involves an easily surmountable barrier of 0.02 eV. In the process, the C–Ot distance decreases from 2.749 Å to 1.275 Å and Y–Ot length increases by 0.138 Å. From IM2, the nonlinear CO2 subunit rotates around the Y–Ot bond involving a transition state (TS2) which is 0.10 eV higher in energy. The resulting complex IM3 is lower in energy by 2.64 eV than the reactants. There is another transition state (TS2′) from IM2 to IM3, which needs to overcome 0.54 eV of barrier, as shown in Fig. S3.† The unpaired electron transfers from the curving bent CO2 motif back to the yttrium oxide cluster needs to overcome a barrier of 0.61 eV through the transition state (TS3). In this process, the distance of the terminal oxygen atom to the adjacent yttrium atom is elongated, that is, 2.325 Å for IM3, 2.510 Å for TS3 and 2.589 Å for IM4, respectively. The IM4 is 2.05 eV lower in energy than the reactants with a linear CO2. Finally, dissociation of CO2 from IM4 to form the products Y3O4 and CO2 requires 0.25 eV of energy. The overall oxidation reaction is barrierless and calculated to be exothermic by 1.80 eV, indicating the oxidation reaction is favorable both kinetically and thermodynamically.
3.2.1.2 Y2AlO5 with CO. The calculated energy profile of Fig. 6 shows that the reaction of Y2AlO5 with CO proceeds according to a general mechanism involving the initial binding of the carbon atom of CO to the radical oxygen of Y2AlO5. The initial encounter complex IM1 is 2.11 eV more stable than the reactants and contains a bent CO2 subunit that results from the transfer of single unpaired electron to CO. The bond lengths of C–Ot and Al–Ot are 1.314 Å and 1.800 Å, respectively. In the following steps, the CO2 moiety is transferred from the Al site to the Y site of Y2AlO4 subunit, involving the transitions which are 0.54 eV and 0.53 eV higher in energy for TS1 and TS2, respectively. IM2 and IM3 are 2.34 eV and 2.20 eV lower in energy than the separated reactants, respectively. Spin density is appeared to be mainly located on the Y2AlO4 subunit for the IM2 and Al atom for IM3, respectively (Fig. S9†). With the increase of the distance from the carbon atom to the bridge oxygen atom and the decrease of the distance from the carbon atom to the capped oxygen atom, the CO2 subunit is passed to the Y–Oc site forming IM4, requiring an energy barrier of 0.53 eV through the TS3. The CO2-absorbing (CO2 ready-desorbing) IM4 is 1.70 eV lower in energy than the reactants. Besides, the electron located on CO2 unit of IM1 can directly transfer to Al atom forming the IM4 with a barrier of 1.28 eV, as depicted in Fig. S5.† Finally, loss of the CO2 from IM4 requires 0.31 eV of energy, and the overall process is exothermic by 1.39 eV, resulting in Y2AlO4 and CO2 products.
3.2.2 The molecular electrostatic potential and natural bond orbital analyses. An additional supporting finding for the reactions is acquired by considering the molecular electrostatic potential maps (MEPs) and natural bond orbital (NBO) analyses. As can be seen in Fig. S10† for Y3O5 cluster, the MEP exhibits positive values (the blue areas) which are preferred for CO-adsorption. In our calculations, the way that CO adsorbs on the yttrium atom featured by terminal oxygen atom is preferred both kinetically and thermodynamically. In contrast, CO binding to the yttrium atoms which connect with the capped oxygen and bridging oxygen atoms is impeded by an energy barrier, as shown in Fig. S4.† Just as many researchers have reported, metal atom adjacent to the oxygen-centered radical is more reactive toward CO.16,29 The MEP of Y2AlO5 is positive over two yttrium atoms surface. However, the way that CO adsorbs on the yttrium atoms of the Y2AlO5 cluster is not favored as depicted in Fig. S7.† In Y2AlO5, the oxygen radical center area is relatively negative. Attacking at the oxygen radical center leads to the oxidation reaction, which is a preferable path without any energy barrier as shown previously.14,16,17The MEP of CO2 molecular, shown in Fig. S10,† is positive over carbon atom and negative over oxygen atoms, respectively. In Y3O4, the yttrium atoms are the most positive (1.74|e|) sites and the capped oxygen is more negative (−1.34|e|) than bridging oxygen sites. Therefore, the linear CO2 has a tendency to split from the Y–Oc bond. Analogously, the yttrium site (1.91|e|) is more positive than aluminum site (1.47|e|) in Y2AlO4, which may explain why the CO2 molecular dissociates from the Y–Oc bond rather than Al–Oc bond.
3.3 The reaction of Y2MO4 (M = Y, Al) with N2O
The Y2MO5 (M = Y, Al) clusters can be retrieved by treating Y2MO4 (M = Y, Al) clusters with N2O. Then a full cycle is attainable in the reactions of CO with N2O by neutral Y2MO5 (M = Y, Al) clusters (Fig. 7). DFT calculations at the BP86 level are performed for the reactions of Y2MO4 (M = Y, Al) with N2O. The favorite paths for Y3O4 and Y2AlO4 towards N2O are presented in Fig. 8 and 9, respectively. The others listed in the ESI (Fig. S11 and 12†).
 |
| Fig. 7 Proposed full catalytic cycle reaction involving CO and N2O mediated by Y2MO5 (M = Y, Al) clusters. | |
 |
| Fig. 8 DFT calculated for the reaction of N2O with Y3O4 cluster. The reactants, intermediates, transition states and produces of the reaction are denoted as R, IMn, TSn and P, respectively. The energies are given in eV and bond lengths are in angstroms. | |
 |
| Fig. 9 DFT calculated for the reaction of N2O with Y2AlO4 cluster. The reactants, intermediates and produces of the reaction are denoted as R, IM1 and P, respectively. The energies are given in eV and bond lengths are in angstroms. | |
3.3.1 Y3O4 with N2O. The calculated energy profile for the reaction between Y3O4 and N2O is presented in Fig. 8. For Y3O4, the spin is delocalized equally over the three yttrium atoms. The adsorption of N2O on the equivalent yttrium atom leads to the intermediate (IM1), which is 1.35 eV more stable than the reactants. With the increase of the distance between O atom and N atom (1.399 to 2.789 Å) and the decrease of the distance from O atom to Y atom (2.275 to 2.172 Å), the O atom is passed on smoothly to bind with Y atom forming IM2. The activation barrier of the oxygen transfer is calculated to be 0.05 eV, which corresponds primarily to the breaking of the bond of N–O. The IM2 is 1.68 eV lower in energy than the reactants. Dissociation of the N2 unit generates the products Y3O5 and N2, which requires 0.01 eV of energy. The total oxidation reaction is exothermic by 1.67 eV.
3.3.2 Y2AlO4 with N2O. The possible paths for the Y2AlO4 cluster reducing N2O are displayed in Fig. 9, S11 and 12.† The current study shows that the unpaired electron in Y2AlO4 cluster mainly located on the main-group metal aluminum atom. As the N2O molecule attacking the aluminum atom, the unpaired electron will occupy the LUMO anti-bonding orbital of N2O and the N–O distance will be lengthened to 2.824 Å. The interaction of the Al atom and N2 unit is weak as their distance is 2.643 Å. The complex IM1 is 1.99 eV more stable than the separated reactants. Releasing the N2 from IM1 is rather easy without any barrier. The overall process is barrier-less and calculated to be exothermic by 2.09 eV.The different reactivity of Y3O4 versus Y2AlO4 in their reactions with N2O is just another example of the role spin distributions often played in chemical reactions. In the doped Y2AlO4, the unpaired electron is mainly localized on the Al atom, while in the homonuclear Y3O4 cluster, the spin delocalized equally over the three bridging Y atoms. Consequently, the reaction patterns of Y3O4 and Y2AlO4 towards N2O are different. Furthermore, the different spin distributions also induce a more negative change of Gibbs free energy for Y2AlO4 (−2.14 eV) in reacting with N2O than that of Y3O4 (−1.79 eV). In other words, while doping a main group Al atom the reaction is favorable both thermodynamically and kinetically.
3.4 Comparison with di-nuclear yttrium oxide clusters
As mentioned above, the reactions of CO with Y2MO5 (M = Y, Al) are barrierless and exothermic, and the other half of the catalytic cycles for regenerations of Y2MO5 (M = Y, Al) by N2O are favored both thermodynamically and kinetically. Thus the full catalytic cycle is attainable in this work. The catalytic ability of di-nuclear yttrium oxides clusters Y2O3+˙ and YAlO3+˙ for CO/N2O couple also have been studied by He and co-workers.11 The oxidation of CO with Y2O3+˙ is impeded by an energy barrier. As one yttrium atom of Y2O3+˙ was replaced by one aluminum atom, the oxygen radical center was generated leading to the reaction barrier-less. Regeneration of YAlO3+˙ from YAlO2+˙ with N2O dose not take place and the catalytic cycle cannot be closed. Combined our calculations with He et al. report, it's obvious to conclude that the tri-nuclear yttrium oxides clusters exhibit greatly enhanced catalytic activity in the reaction of CO with N2O.
4. Conclusion
Extensive DFT calculations are employed to investigate the reaction mechanism of CO oxidation by Y2MO5 (M = Y, Al) clusters and the regeneration of Y2MO5 (M = Y, Al) clusters using N2O. Theoretical investigations reveal that both the Y3O5 and Y2AlO5 clusters are doublet state with a single unpaired electron located on the terminal oxygen atom. Calculated energy profiles for the reactions of Y3O5 and Y2AlO5 with CO are energetically favorable and barrierless. The high reactivity for Y2MO5 clusters is ascribed to the oxygen radical center. The reaction mechanisms for Y2MO5 (M = Y, Al) clusters with CO may be understood qualitatively from the molecular electrostatic potential maps and natural bond orbital analyses. Regenerations of the initial species Y2MO5 (M = Y, Al) by N2O constitute the other half of the catalytic cycle. Based on the previous study of di-nuclear YMO2+˙ (M = Y, Al), we propose that by controlling the cluster size (or doping ratio), it's possible to generate neutral oxide cluster with localized radical oxygen center, which is highly active in redox reactions.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (21371034, 21373048 and 21301030), the Natural Science Foundation of Fujian Province for Distinguished Young Investigator Grant (2013J06004), and the Independent Research Project of State Key Laboratory of Photocatalysis on Energy and Environment (2014A02).
References
- A. R. Ravishankara, J. S. Daniel and R. W. Portmann, Science, 2009, 326, 123 CrossRef CAS PubMed.
- B. K. Min and C. M. Friend, Chem. Rev., 2007, 107, 2709 CrossRef CAS PubMed.
- S. Royer and D. Duprez, ChemCatChem, 2011, 3, 24 CrossRef CAS PubMed.
- S. J. Lin, J. Cheng, C. F. Zhang, B. Wang, Y. F. Zhang and X. Huang, Phys. Chem. Chem. Phys., 2015, 17, 11499 RSC.
- I. Lopes, A. Davidson and C. Thomas, Catal. Commun., 2007, 8, 2105 CrossRef CAS PubMed.
- G. W. Peng, L. R. Merte, J. Knudsen, R. T. Vang, E. Lagsgaard, F. Besenbacher and M. Marvrikakis, J. Phys. Chem. C, 2010, 114, 21579 CAS.
- X. W. Xie, Y. Li, Z. Q. Liu, M. Haruta and W. J. Shen, Nature, 2009, 458, 746 CrossRef CAS PubMed.
- H. J. Freund, G. Meijer, M. Scheffler, R. Schlogl and M. Qolf, Angew. Chem., Int. Ed., 2011, 50, 10064 CrossRef CAS.
- W. E. Kaden, W. A. Kunkel, F. S. Roberts, M. Kane and S. L. Andersona, J. Chem. Phys., 2012, 136, 204705 CrossRef.
- O. P. Balaj, I. Balteanu, T. T. J. Roβteuscher, M. K. Beyer and V. E. Bondybey, Angew. Chem., Int. Ed., 2004, 43, 6519 CrossRef CAS PubMed.
- J. B. Ma, Z. C. Wang, M. Schlangen, S. G. He and H. Schwarz, Angew. Chem., Int. Ed., 2013, 52, 1226 CrossRef CAS PubMed.
- M. M. Kappes and R. H. Staley, J. Am. Chem. Soc., 1981, 103, 1286 CrossRef CAS.
- G. E. Johnson, R. Mitrić, E. C. Tyo, V. Bonačić-Koutecký and A. W. Castleman Jr, J. Am. Chem. Soc., 2008, 130, 13912 CrossRef CAS PubMed.
- G. E. Johnson, R. Mitrić, M. Nössler, E. C. Tyo, V. Bonačić-Koutecký and A. W. Castleman Jr, J. Am. Chem. Soc., 2009, 131, 5460 CrossRef CAS PubMed.
- G. E. Johnson, R. Mitrić, V. Bonačić-Koutecký and A. W. Castleman Jr, Chem. Phys. Lett., 2009, 475, 1 CrossRef CAS PubMed.
- M. Nößler, R. Mitrić, V. Bonačić-Koutecký, G. E. Johnson, E. C. Tyo and A. W. Castleman Jr, Angew. Chem., Int. Ed., 2010, 49, 407 CrossRef PubMed.
- Z. C. Wang, N. Dietl, R. Kretschmer, T. Weiske, M. Schlangen and H. Schwarz, Angew. Chem., Int. Ed., 2011, 50, 12351 CrossRef CAS PubMed.
- E. C. Tyo, M. Nößler, R. Mitrić, V. Bonačić-Koutecký and A. W. Castleman Jr, Phys. Chem. Chem. Phys., 2011, 13, 4243 RSC.
- A. W. Castleman Jr, Catal. Lett., 2011, 141, 1243 CrossRef.
- S. Yin and E. R. Bernstein, Int. J. Mass Spectrom., 2012, 321–322, 49 CrossRef CAS PubMed.
- S. M. Lang and T. M. Bernhardt, Phys. Chem. Chem. Phys., 2012, 14, 9255 RSC.
- K. R. Asmis, Phys. Chem. Chem. Phys., 2012, 14, 9270 RSC.
- H. Schwarz, Angew. Chem., Int. Ed., 2011, 50, 10096 CrossRef CAS PubMed.
- M. Schlangen and H. Schwarz, Catal. Lett., 2012, 142, 1265 CrossRef CAS.
- Z. C. Wang, N. Dietl, R. Kretschmer, J. B. Ma, T. Weiske, M. Schlangen and H. Schwarz, Angew. Chem., Int. Ed., 2012, 51, 3703 CrossRef CAS PubMed.
- Z. C. Wang, X. N. Wu, Y. X. Zhao, J. B. Ma, X. L. Ding and S. G. He, Chem.–Eur. J., 2011, 17, 3449 CrossRef CAS.
- Z. C. Wang, T. Weiske, R. Kretschmer, M. Schlangen, M. Kaupp and H. Schwarz, J. Am. Chem. Soc., 2011, 133, 16930 CrossRef CAS PubMed.
- Z. C. Wang, X. N. Wu, Y. X. Zhao, J. B. Ma, X. L. Ding and S. G. He, Chem. Phys. Lett., 2010, 489, 25 CrossRef CAS PubMed.
- X. N. Wu, Y. X. Zhao, W. Xue, Z. C. Wang, S. G. He and X. L. Ding, Phys. Chem. Chem. Phys., 2010, 12, 3984 RSC.
- X. N. Wu, X. L. Ding, S. M. Bai, B. Xu, S. G. He and Q. Shi, J. Phys. Chem. C, 2011, 115, 13329 CAS.
- J. B. Ma, B. Xu, J. H. Meng, X. N. Wu, X. L. Ding, X. N. Li and S. G. He, J. Am. Chem. Soc., 2013, 135, 2991 CrossRef CAS PubMed.
- Z. C. Wang, W. Xue, Y. P. Ma, X. L. Ding, S. G. He, F. Dong, S. Heinbuch, J. J. Rocca and E. R. Bernstein, J. Phys. Chem. A, 2008, 112, 5984 CrossRef CAS PubMed.
- M. R. Sievers and P. B. Armentrout, Inorg. Chem., 1999, 38, 397 CrossRef CAS.
- M. Knickelbein, Chem. Phys., 1995, 102, 1 CAS.
- A. Pramann, Y. Nakamura and A. Nakajima, J. Phys. Chem. A, 2001, 105, 7534 CrossRef CAS.
- G. Gu, B. Dai, X. Ding and J. Yang, Eur. Phys. J. D, 2004, 29, 27 CrossRef CAS.
- B. Dai, K. Deng and J. Yang, Chem. Phys. Lett., 2002, 364, 188 CrossRef CAS.
- L. Xu, C. J. Xia, L. F. Wang, L. Xie, B. Wang, Y. F. Zhang and X. Huang, RSC Adv., 2014, 4, 60270 RSC.
- A. D. Becke, Phys. Rev. A: At., Mol., Opt. Phys., 1988, 38, 3098 CrossRef CAS.
- J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8822 CrossRef.
- W. Küchle, M. Dolg, H. Stoll and H. Preuss, Pseudopotentials of the Stuttgart/Dresden group 1998, revision August 11, 1998, http://www.theochem.uni-stuttgart.de/pseudopotentiale.
- D. Andrae, U. Häeußermann, M. Dolg, H. Stoll and H. Preuss, Theor. Chim. Acta, 1990, 77, 123 CrossRef CAS.
- J. M. L. Martin and A. Sundermann, J. Chem. Phys., 2001, 114, 3408 CrossRef CAS PubMed.
- R. A. Kendall, T. H. Dunning Jr and R. J. Harrison, J. Chem. Phys., 1992, 96, 6796 CrossRef CAS PubMed.
- T. H. Dunning Jr, J. Chem. Phys., 1989, 90, 1007 CrossRef PubMed.
- C. Gonzalez and H. B. Schlegel, J. Chem. Phys., 1989, 90, 2154 CrossRef CAS PubMed.
- C. Gonzalez and H. B. Schlegel, J. Phys. Chem., 1990, 94, 5523 CrossRef CAS.
- G. D. Purvis III and R. J. Bartlett, J. Chem. Phys., 1982, 76, 1910 CrossRef PubMed.
- G. E. Scuseria, C. L. Janssen and H. F. Schaefer III, J. Chem. Phys., 1988, 89, 7382 CrossRef CAS PubMed.
- K. Raghavachari, G. W. Trucks, J. A. Pople and M. Head-Gordon, Chem. Phys. Lett., 1989, 157, 479 CrossRef CAS.
- J. D. Watts, J. Gauss and R. J. Bartlett, J. Chem. Phys., 1993, 98, 8718 CrossRef CAS PubMed.
- R. J. Bartlett and M. Musial, Rev. Mod. Phys., 2007, 79, 291 CrossRef CAS.
- Y. Gong, C. F. Ding and M. F. Zhou, J. Phys. Chem. A, 2009, 113, 8569 CrossRef CAS.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylow, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian 09, Revision A.1, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
- H. J. Werner, P. J. Knowles, F. R. Manby, M. Schütz, P. Celani, G. Knizia, T. Korona, R. Lindh, A. Mitrushenkov, G. Rauhut, T. B. Adler, R. D. Amos, A. Bernhardsson, A. Berning, D. L. Cooper, M. J. O. Deegan, A. J. Dobbyn, F. Eckert, E. Goll, C. HamPel, A. Hesselmann, G. Hetzer, T. Hrenar, G. Jansen, C. Köppl, Y. Liu, A. W. Lloyd, R. A. Mata, A. J. May, S. J. McNicholas, W. Meyer, M. E. Mura, A. Nicklass, D. P. O'Neill, P. Palmieri, K. Pflüger, R. Pitzer, M. Reiher, T. Shiozaki, H. Stoll, A. J. Stone, R. Tarroni, T. Thorsteinsson, M. Wang and A. Wolf, MOLPRO, version 2010.1, a package of ab initio programs, http://www.molpro.net Search PubMed.
- E. D. Glendening, A. E. Reed, J. E. Carpenter and F. Weinhold, NBO Version 3.1 Search PubMed.
- R. Dennington II, T. Keith, J. Millam, K. Eppinnett, W. L. Hovell and R. Gilliland, GaussView: Version 4.1, Semichem Inc., Shawnee Mission, KS, 2007 Search PubMed.
- W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics, 1996, 14, 33 CrossRef CAS.
- A. Martínez, F. J. TenorioJ and V. Ortiz, J. Phys. Chem. A, 2001, 105, 11291 CrossRef.
- Y. X. Zhao, X. N. Wu, J. B. Ma, S. G. He and X. L. Ding, Phys. Chem. Chem. Phys., 2011, 13, 1925 RSC.
- J. B. Ma, Z. C. Wang, M. Schlangen, S. G. He and H. Schwarz, Angew. Chem., Int. Ed., 2012, 51, 5991 CrossRef CAS PubMed.
- N. Dietl, M. Schlangen and H. Schwarz, Angew. Chem., Int. Ed., 2012, 51, 5544 CrossRef CAS PubMed.
- X. L. Ding, X. N. Ding, Y. X. Zhao and S. G. He, Acc. Chem. Res., 2012, 45, 382 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: The others possible pathways for the redox reaction are presented in Fig. S3–S7, S11 and 12. Numerical electron spin density (in |e|) for the structures presented in the energy profiles of Fig. 5 and 6 are shown in Fig. S8 and 9. The Cartesian coordinates for the reactants, intermediates, transition states and the products shown in the energy profiles of Fig. 5, 6, 8 and 9 are listed in Table S3. See DOI: 10.1039/c5ra11695c |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.