Abstract

By means of spin-polarized density functional theory (DFT) computations, we unravel the reaction mechanisms of catalytic CO oxidation on B-doped fullerene. It is shown that O2 species favors to be chemically adsorbed via side-on configuration at the hex-C–B site with an adsorption energy of −1.07 eV. Two traditional pathways, Eley–Rideal (ER) and Langmuir–Hinshelwood (LH) mechanisms, are considered for the CO oxidation starting from O2 adsorption. CO species is able to bind at the B-top site of the B-doped fullerene with an adsorption energy of −0.78 eV. Therefore, CO oxidation that occurs starting from CO adsorption is also taken into account. Second reaction of CO oxidation occurs by the reaction of CO + O → CO2 with a very high energy barrier of 1.56 eV. A trimolecular Eley–Rideal (TER) pathway is proposed to avoid leaving the O atom on the B-doped fullerene after the first CO oxidation. These predictions manifest that boron-doped fullerene is a potential metal-free catalyst for CO oxidation.
Introduction
Catalytic CO oxidation reaction has attracted considerable attention due to its significance in purifying environmental pollution and eliminating electrode poisoning in fuel cells.1−5 CO oxidation is a simple and prototypical reaction and is suitable for unraveling the intrinsic reaction mechanism of heterogeneous catalysis.6,7 How to effectively detect and remove CO has been one of the most challenging subjects nowadays. Conversion of CO into nontoxic substances, CO2, by catalytic oxidation process has been extensively studied using noble metals (Au,8−12 Ag,13 Pd,14−16 Ru, Rh, Os, Ir,14 and Au–Pd17,18) and metal oxides (TiO2-,19 CeO2-,20−23 and ZrO2-based materials24) as excellent catalysts toward CO oxidation. However, their high cost and scarcity have limited their applications. As a result, development of nonmetal or metal-free catalysts to reduce or replace the use of those metals in a large-scale process has become urgent.
Carbon nanomaterials, such as fullerenes, carbon nanotubes, graphenes, etc., have been used widely as either catalyst supports or catalysts for heterogeneous catalytic reaction. There have been several studies on single-atom catalysts (SACs) supported on carbon nanomaterials, for example, Pt-,25−27 Cu-,28 Fe-,29 and Au-graphene,30 and Fe-BN-fullerene,31 which efficiently catalyze oxidation of CO. The activation energy for the rate-determining step (RDS) of CO oxidation on the above catalysts is predicted to be within 0.51–0.58 eV, which is comparable to those of metal catalysts. Doping heteroatoms (N, B, P, S, etc.) into carbon nanomaterials could result in electron modulation to modify the electronic properties and chemical activities that make heteroatom-doped carbon materials potential catalysts for CO oxidation.32−38 Previous investigations have shown that N-doped fullerene,39 N-doped carbon nanotube,40,41 N-doped penta-graphene,42−44 graphdiyne,45 B-doped and fullerene-like BN cage,46 B-doped carbon nanotube,47 and so on can serve as metal-free catalysts and exhibit superior catalytic activity toward CO oxidation. Unraveling the detailed reaction mechanisms of CO oxidation is essential to design novel metal-free catalysts. Boron-doped carbon nanotubes have been shown to exhibit high catalytic activity both experimentally and theoretically.47,48 Previous studies demonstrated that the energy barrier of RDS for CO oxidation on B-doped fullerene using the Eley–Rideal (ER) mechanism is predicted to be 0.48 eV.49,50 They also showed that second CO oxidation occurs by the reaction of CO with the remaining O atoms, which have an energy barrier of 1.47 eV, which limits its occurrence. To solve this issue, a trimolecular Eley–Rideal (TER) pathway is proposed to avoid leaving the O atom on the B-doped fullerene after the first CO oxidation, and it will be discussed below. Understanding the intrinsic mechanisms of CO oxidation is important to guide researchers to design the most effective catalyst.
Herein, we carry out spin-polarized first-principles calculations to unravel the catalytic mechanisms of CO oxidation on B-doped fullerene. In this study, diverse mechanisms of CO oxidation reaction on B-doped fullerene are revealed. The best reaction pathway of CO oxidation is verified by discussing the energy barrier of the RDS.
Results and Discussion
Interaction and Adsorption Behavior of O2 and CO on B-doped Fullerene
To unravel the reaction mechanisms of CO oxidation on B-doped fullerene, we first investigate the adsorption behavior of O2 and CO on B-doped fullerene. A polarized C––B+ bond, which becomes an active site for molecular adsorption, is created by doping the boron atom into fullerene (see Figure 1). Gao and Chen have demonstrated that B-doped fullerene can resist temperatures up to 1000 K, suggesting its structural stability.50 Bader charge calculation shows that the carbons bound to boron gain negative charges of 0.88–1.16 |e| from the substitutive boron. The O2 and CO species are placed above those C–B sites containing B-top, hex-C–B, and pent-C–B sites of the B-doped fullerene at a proper distance during the geometric optimizations.
Figure 1.

Optimized model structures, possible active sites, and Bader charge of B-doped fullerene. The gray and pink spheres represent C and B atoms, respectively. The charges are given in |e|.
The optimized configurations of O2 and CO adsorption on the B-doped fullerene as well as the related adsorption energies are displayed in Figure 2 and Table 1. It is found that O2 species chemisorbs on the C–B site of the B-doped fullerene via two types of configurations: side-on and end-on configurations. Calculations demonstrate that the O2 species prefers to strongly chemisorb at the hex-C–B site via the side-on configuration. The adsorption energy is computed to be −1.07 eV, while those are −0.61 and −0.63 eV for the O2 side-on configuration at the pent-C–B site and the O2 end-on configuration at the B-top site, respectively. The related distance of O–O bond is calculated to be elongated from 1.222 Å of free O2 molecule to 1.462, 1.461, and 1.318 Å, respectively, as depicted in Figure 2. Bader charge analysis (see Table 1) shows that the adsorbed O2 species gain −1.42, −1.33, and −1.28 |e| from the B-doped fullerene, indicating that the O2 species is chemisorbed on the B-doped fullerene. It is comparable to those on the catalysts of metals and SACs, Pt surfaces (−0.58 to −0.72 eV for (111),51 −1.48 eV for (110),52 and −1.30 eV for (001)53), Pt clusters (−0.72 eV for Pt2 and −1.08 eV for Pt3),54,55 Au clusters (−0.24 eV for Au4,55 ca. −0.50 eV for Au7 and Au9,30 −0.60 eV for Au29,10,56,57 −0.99 eV17,58 for Au38), Pt-graphene (−1.34 eV),25 Cu-graphene (−2.67 eV),28 Fe-graphene (−2.09 eV),29 Au-graphene (−1.34 eV).30
Figure 2.

Optimized structures of O2 and CO adsorption on B-doped fullerene. The gray, pink, and red spheres represent C, B, and O, atoms, respectively. The distances are given in angstroms.
Table 1. Computed Adsorption Energies (eV) and Bader Charges (|e|; in Parentheses) of O2 and CO Species on the B-doped Fullerene.
| adsorbate/site | hex-C–B | penta-C–B | B-top |
|---|---|---|---|
| O2 | –1.07 (−1.42) | –0.61 (−1.33) | –0.63 (−1.28) |
| CO | –0.78 (−0.49) |
For the CO adsorption, CO is placed at the mentioned active sites with a 1.5–2.0 Å distance away from the B-doped fullerene during structural optimization. One should note that CO species is able to adsorb at the B-top site of the B-doped fullerene with an adsorption energy of −0.78 eV. Compared to the cases of B-doped carbon nanotube47 and N-doped fullerene59 (−0.05 and −0.20 eV for B-doped carbon nanotube and N-doped fullerene, respectively), CO is chemisorbed on the B-doped fullerene. More charge transfer (−0.49 |e|) from the B-doped fullerene occurs, resulting in stronger interaction between the O2 species and the B-doped fullerene. This result indicates that CO is likely to bind to B-doped fullerene. The O adsorption, which is depicted in Figure S1, takes place by the first oxidation reaction of CO + O2 → CO2 + O. It is found that the atomic O prefers to bind at the C–B site with the computed adsorption energies of −5.10 and −5.24 eV at the penta-C–B and hex-C–B sites, respectively.
The minimum-energy paths of O2 and CO adsorption at B-doped fullerene have been illustrated by CINEB calculation, as depicted in Figure S2. The calculation shows that no energy barrier is required to surmount for the CO adsorption process, but it needs a barrier of 0.2 eV to surmount for the O2 adsorption process, indicating that the CO adsorption process is kinetically more favored than the O2 adsorption process, which is thermodynamically favored. Therefore, CO oxidation process occurs starting from both O2 and/or CO adsorption, as discussed the following section.
CO Oxidation Reaction Starting from O2 Adsorption and/or CO Adsorption on B-doped Fullerene
Here, we illustrate the mechanisms of CO oxidation beginning from O2 adsorption and/or CO adsorption on B-doped fullerene. Two traditional pathways, Eley–Rideal (ER) and Langmuir–Hinshelwood (LH) mechanisms, are considered for the CO oxidation.
CO Oxidation Starting from the O2 Side-On Adsorption
The potential energy surface (PES) of the CO oxidation reaction via the LH mechanism at the hex-C–B site of B-doped fullerene is mapped by the CINEB method, as described in Figure 3. The first step occurs by attachment of the gas-phase CO molecule to fullerene to achieve O2 and CO co-adsorption (intermediate C) on the fullerene. The step is found to be exothermic by 4.24 eV with an energy barrier of 0.51 eV. The O–O bond distance in the O2 and CO co-adsorption (intermediate C) is found to increase to 2.154 from 1.461 Å. Then, CO2 desorption takes place from the intermediate C with a small energy barrier of 0.08 eV and an exothermicity of 0.58 eV. The overall reaction of O2 + CO → CO2 + O on the B-doped fullerene is exothermic by ∼5.00 eV and requires to surmount a 0.51 eV barrier of the rate-determining step. The second CO oxidation takes place by the remaining O atom on fullerene, which will be studied below.
Figure 3.

Energy diagrams for CO oxidation starting from the O2 side-on adsorption via the LH mechanism at the hex-C–B site of B-doped fullerene.
Furthermore, the ER mechanism is also discussed beginning from the O2 side-on adsorption based on the PES depicted in Figure 4. The process begins by the attachment of a gas-phase CO molecule to insert into the O–O bond of the O2 end-on adsorption to form a carbonate-like CO3 (intermediate C) with a large exothermicity of 4.57 eV. In TS B, the O–O bond is slightly elongated to 1.472 from 1.461 Å with 0.34 eV higher in energy than the reactant A. Subsequently, CO2 desorption can occur by cleavage of Bfullerene–O and Cfullerene–O bonds with very high barriers of 2.36 and 1.95 eV, respectively. The high barrier of this process suggests that it is kinetically unfavored. To solve this issue, a trimolecular Eley–Rideal (TER) pathway, reaction of the second CO species with the CO3 intermediate, is proposed and will be discussed below.
Figure 4.

Energy diagrams for CO oxidation starting from the O2 side-on adsorption via the ER mechanism at the hex-C–B site of B-doped fullerene.
CO Oxidation Starting from the O2 End-On Adsorption
First, the LH mechanism is discussed beginning from the O2 end-on adsorption, as depicted in Figure 5. The CO species initially approaches the B-doped fullerene and interacts with the O2 end-on adsorption to form a peroxo-type (OOCO) intermediate with an exothermicity of 2.03 eV, as seen in Figure 5. This step only requires 0.24 eV to surmount. The O–O bond length is predicted to be elongated to 1.475 from 1.314 Å in the OOCO intermediate. Then, the cleavage of the O–O bond in the OOCO intermediate takes place to desorb CO2 via a transition structure (TS D), with a barrier of 0.89 eV and endothermicity of 1.64 eV. In TS D, the cleavage of the O–O and C–C distances is predicted to be 1.831 and 1.673 Å, respectively. The overall reaction beginning from the O2 end-on adsorption is exothermic by 3.67 eV and needs to overcome a 0.89 eV barrier of the rate-determining step.
Figure 5.

Energy diagrams for CO oxidation starting from the O2 end-on adsorption via the LH mechanism at the B-top site of B-doped fullerene.
The ER mechanism is further described beginning from the O2 end-on adsorption, as depicted in the PES (see Figure 6). CO species in reactant A achieves end-on O2 adsorption and extracts one O atom directly by surmounting a TS B with an energy barrier of 0.64 eV to produce the final state C (product). In TS B, the O–O bond increases from 1.318 to 1.330 Å, and the distance between the C atom of CO and the O atom of O2 decreases from 2.705 to 2.333 Å. The process is predicted to be highly exothermic by −5.01 eV. The energy barrier (0.64 eV) of this rate-determining step is slightly higher than that of the LH mechanism (0.51 eV).
Figure 6.

Energy diagrams for CO oxidation starting from the O2 end-on adsorption via the ER mechanism at the B-top site of B-doped fullerene.
CO Oxidation Starting from the CO Adsorption
As mentioned above, there is no energy barrier in the minimum-energy path for CO adsorption at the B-top site of the B-doped fullerene with an adsorption energy of −0.78 eV. Therefore, CO oxidation on the B-doped fullerene starting from the CO adsorption can be achieved more easily because this process is both kinetically and thermodynamically favored. As seen in Figure 7, the gas-phase O2 species approaches the B-doped fullerene and directly undergoes CO adsorption to form a peroxo-type (OOCO) intermediate with an exothermicity of 0.99 eV. In addition, only a low energy barrier of 0.05 eV is required for this step. The O–O bond is found to increase to 1.490 from 1.238 Å in the OOCO intermediate. Next, the cleavage of the O–O bond in the OOCO intermediate occurs to desorb CO2 via a transition structure (TS D), with a barrier of 0.39 eV and endothermicity of 3.40 eV. In TS D, the cleavage O–O and C–B distances are computed to be 1.707 and 1.608 Å, respectively. The overall reaction beginning from the CO adsorption is exothermic by 4.39 eV and needs to overcome a 0.39 eV barrier of the rate-determining step. Compared to the proposed reaction mechanisms beginning from the O2 adsorption as alluded above, the energy barrier of the rate-determining step is the lowest one (0.39 eV), indicating that CO oxidation prefers to occur via the LH mechanism starting from the preadsorption of CO on the B-doped fullerene.
Figure 7.

Energy diagrams for CO oxidation starting from the CO adsorption via the LH mechanism at the B-top site of B-doped fullerene.
Second CO Oxidation Occurring with the Remaining O Atom
As depicted in Figures S3 and S4, the gas-phase CO molecule approaches the adsorbed O species and undergoes O adsorption to produce a second CO2 species. The process via the ER mechanism (see Figure S3) is exothermic by 1.03 eV but needs to pass a high energy barrier of 1.56 eV, while the process via the LH mechanism (see Figure S4) needs to overcome a high energy barrier of 1.82 eV. The desorption energy of CO2 is predicted to be 0.03 eV, suggesting that CO2 species is easily released from the B-doped fullerene. However, the B-doped fullerene is readily poisoned by the remaining O species due to the very high barrier for removing the O atom. To solve this issue, a trimolecular Eley–Rideal (TER) pathway is proposed to avoid leaving the O atom on the B-doped fullerene after the first CO oxidation.
Trimolecular Eley–Rideal (TER) Mechanism
In this section, we consider the new TER mechanism of CO oxidation, which is called the self-promotion mechanism proposed by Lyalin et al.,60 to avoid leaving the O atom on the B-doped fullerene. The TER mechanism can occur starting from the carbonate-like (CO3) and the peroxo-type (OOCO) intermediates. As shown in Figure 8, the second CO molecule can approach the CO3 intermediate and extract one oxygen in the CO3 to directly form two CO2 species by passing barriers 1.37 eV (extracting the O atom bound to the B atom) and 1.79 eV (extracting the O atom bound to C atom), which are 0.99 and 0.18 eV lower than 2.36 and 1.95 eV, respectively. The process is more exothermic by 1.01 eV. The other TER mechanism can take place through the OOCO intermediate depicted in Figure 9; the second CO molecule can attach to the OOCO intermediate and extract the O bound to B and/or C to produce two CO2 species. The predicted energy barriers are 0.68 eV (extracting the O atom bound to the B atom) and/or 0.33 eV (extracting the O atom bound to the C atom), which are 0.21 and 0.06 eV lower than 0.89 and 0.39 eV for the traditional LH mechanism, respectively. Besides, the process is more exothermic by 2.33 and 1.37 eV, respectively. The calculation results demonstrate that the TER mechanism is found to be more kinetical and thermodynamical compared with the traditional LH and ER mechanisms.
Figure 8.

Energy diagrams for CO oxidation starting from the carbonate-like (CO3) intermediate via the TER mechanism.
Figure 9.
Energy diagrams for CO oxidation starting from the peroxo-type (OOCO) intermediate via the TER mechanism (a) starting from O2 end-on adsorption and (b) starting from CO adsorption.
The catalytic activity (energy barriers of the rate-determining step: 0.33–0.68 eV) of the B-doped fullerene via the TER mechanism is comparable to those on the most studied catalysts including metal surface system: Au(111) (1.97 eV)13 and Au(211) (0.59 and 0.65 eV);9 metal nanoparticle system: Au29 (0.69 eV),10 Au38 (0.37 eV),17 Pd38 (0.49 eV),17 and core–shell Au–Pd (0.10 and 0.59 eV for Au32/Pd6 and Pd32/Au6 clusters, respectively);17 SAC system: Pt- (0.33 eV),26,27 Cu- (0.54 eV),28 Fe- (0.58 eV),29 and Au-embedded graphene (0.31 eV),30 Fe-graphene oxide (0.61 eV);61 and metal-free catalysts: N-doped fullerene (0.20 eV),59 N-doped (0.45–0.58 eV),62 and B-doped CNTs (0.34–0.42 eV).47 Consequently, the calculations suggest that B-doped fullerene could be an effective metal-free catalyst for CO oxidation.
Conclusions
We have investigated the detailed mechanisms for catalytic CO oxidation catalyzed by the boron-doped fullerene by performing spin-polarized first-principles calculations. The calculations demonstrate that O2 is chemisorbed via side-on configuration with adsorption energies of −1.07 eV at the hex-C–B site of the boron-doped fullerene. It is found that CO species can also bind at the B-top site with an adsorption energy of −0.78 eV. There is no energy barrier required to surmount for the CO adsorption process, but it needs a barrier of 0.2 eV to surmount for the O2 adsorption process, indicating that the CO adsorption process is kinetically more favored than the O2 adsorption process. Therefore, two traditional pathways, Eley–Rideal (ER) and Langmuir–Hinshelwood (LH) mechanisms, are considered for the CO oxidation starting from both O2 and/or CO adsorption. The results show that CO oxidation takes place by the LH mechanism starting from the preadsorption CO with the lowest energy barrier of 0.39 eV for the rate-determining step. The remaining atomic O species continues to proceed second reaction of CO oxidation with a very high energy barrier of 1.56 eV, which hinders the catalytic activity of the B-doped fullerene. A TER pathway is proposed to avoid leaving the O atom on the B-doped fullerene after the first CO oxidation. The computations demonstrate that the TER mechanism is found to be more kinetical and thermodynamical compared with the traditional LH and ER mechanisms.
Computational Methods
All geometries are optimized using the periodic first-principles computations based on spin-polarized density functional theory (DFT) implemented in the Vienna Ab initio Simulation Package (VASP) code.63,64 The exchange–correlation functional of Perdew–Wang correlation (PW91) with the generalized gradient approximation (GGA) is described.65 The plane-wave basis set is set to be 400 eV of cutoff energy. The Brillouin zones are selected as Γ k-point mesh and 5 × 5 × 5 k-point mesh for geometric structure optimizations and electronic property calculations, respectively.66 A 25 × 25 × 25 Å3 periodic supercell is selected for the calculations, which is large enough to avoid the interaction of periodic images.59 The adsorption energy of adsorbates (O2 and CO) on B-doped fullerene is predicted according to the following equation
where Etot, EB-fullerene, and Eadsorbate represent the computed energies of adsorbed species on the B-doped fullerene, the isolated B-doped fullerene, and the gas-phase molecule, respectively. It is noted that the predicted O–O bond lengths of gas-phase O2, C–O bond length of gas-phase CO, and C–O bond length of gas-phase CO2 are 1.222, 1.137, and 1.169 Å, respectively, which agree with the experimental data of 1.207,67 1.128,68 and 1.193 Å.69 The minimum-energy paths (MEP) including transition states are constructed by the climbing image nudged elastic band (CINEB) method.70,71 The electron distribution and transfer are described by the Bader charge calculations.72−74
Acknowledgments
The authors thank the Chung Yuan Christian University (CYCU) and Ministry of Science and Technology (MOST), Taiwan, for supporting this work, under grant numbers MOST 109-2113-M-033-001, 108-2113-M-033-001, and 107-2113-M-033-004 and the use of facilities at the National Center for High-Performance Computing, Taiwan.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c04532.
Optimized structures of O adsorption on B-doped fullerene (Figure S1); calculated minimum-energy paths of O2 and CO adsorption processes at B-doped fullerene (Figure S2); and calculated minimum-energy paths for the reaction of CO + O at the C–B site on B-doped fullerene (Figures S3 and S4) (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Suh D. J.; Kwak C.; Kim J. H.; Kwon S. M.; Park T. J. Removal of Carbon Monoxide from Hydrogen-Rich Fuels by Selective Low-Temperature Oxidation over Base Metal Added Platinum Catalysts. J. Power Sources 2005, 142, 70–74. 10.1016/j.jpowsour.2004.09.012. [DOI] [Google Scholar]
- Chen L.; Chen B.; Zhou C.; Wu J.; Forrey R. C.; Cheng H. Influence of CO Poisoning on Hydrogen Chemisorption onto a Pt6 Cluster. J. Phys. Chem. C 2008, 112, 13937–13942. 10.1021/jp803504k. [DOI] [Google Scholar]
- Herzing A. A.; Kiely C. J.; Carley A. F.; Landon P.; Hutchings G. J. REPORT Identification of Active Gold Nanoclusters on Iron Oxide Supports for CO Oxidation. Science 2008, 321, 1331–1335. 10.1126/science.1159639. [DOI] [PubMed] [Google Scholar]
- Hornes A.; et al. Inverse CeO2/CuO Catalyst as an Alternative to Classical Direct Configurations for Preferential Oxidation of CO in Hydrogen-Rich Stream. J. Am. Chem. Soc. 2010, 132, 34–35. 10.1021/ja9089846. [DOI] [PubMed] [Google Scholar]
- Xie X. W.; Li Y.; Liu Z. Q.; Haruta M.; Shen W. J. Low-Temperature Oxidation of CO Catalysed by Co3O4 Nanorods. Nature 2009, 458, 746–749. 10.1038/nature07877. [DOI] [PubMed] [Google Scholar]
- Alavi A.; Hu P.; Deutsch T.; Silvestrelli P. L.; Hutter J. CO Oxidation on Pt(111): An Ab Initio Density Functional Theory Study. Phys. Rev. Lett. 1998, 80, 3650–3653. 10.1103/PhysRevLett.80.3650. [DOI] [Google Scholar]
- Freund H. J.; Meijer G.; Scheffler M.; Schlçgl R.; Wolf M. CO Oxidation as a Prototypical Reaction for Heterogeneous Processes. Angew. Chem., Int. Ed. 2011, 50, 10064–10094. 10.1002/anie.201101378. [DOI] [PubMed] [Google Scholar]
- Falsig H.; Hvolbak B.; Kristensen I. S.; Jiang T.; Bligaard T.; Christensen C.; Norskov J. K. Trends in the Catalytic CO Oxidation Activity of Nanoparticles. Angew. Chem., Int. Ed. 2008, 47, 4835–4839. 10.1002/anie.200801479. [DOI] [PubMed] [Google Scholar]
- Liu Z. P.; Hu P.; Alavi A. Catalytic Role of Gold in Gold-Based Catalysts: A Density Functional Theory Study on the CO Oxidation on Gold. J. Am. Chem. Soc. 2002, 124, 14770–14779. 10.1021/ja0205885. [DOI] [PubMed] [Google Scholar]
- Chen H.-T.; Chang J.-G.; Ju S.-P.; Chen H.-L. First-Principles Calculations on CO Oxidation Catalyzed by a Gold Nanoparticle. J. Comput. Chem. 2010, 31, 258–265. 10.1002/jcc.21314. [DOI] [PubMed] [Google Scholar]
- Fajín J. L. C.; Cordeiro N. D. S.; Gomes J. R. B. DFT Study of the CO Oxidation on the Au(321) Surface. J. Phys. Chem. C 2008, 112, 17291–17302. 10.1021/jp8031435. [DOI] [Google Scholar]
- Xu J.; White T.; Li P.; He C.; Yu J. G.; Yuan W. K.; Han Y. F. Biphasic Pd–Au Alloy Catalyst for Low-Temperature CO Oxidation. J. Am. Chem. Soc. 2010, 132, 10398–10406. 10.1021/ja102617r. [DOI] [PubMed] [Google Scholar]
- Su H. Y.; Yang M. M.; Bao X. H.; Li W. X. The Effect of Water on the CO Oxidation on Ag(111) and Au(111) Surfaces: A First-Principle Study. J. Phys. Chem. C 2008, 112, 17303–17310. 10.1021/jp803400p. [DOI] [Google Scholar]
- Gong X. Q.; Liu Z. P.; Raval R.; Hu P. A Systematic Study of CO Oxidation on Metals and Metal Oxides: Density Functional Theory Calculations. J. Am. Chem. Soc. 2004, 126, 8–9. 10.1021/ja030392k. [DOI] [PubMed] [Google Scholar]
- Hirvi J. T.; Kinnunen T. J.; Suvanto M.; Pakkanen T. A.; Norskov J. K. CO Oxidation on PdO Surfaces. J. Chem. Phys. 2010, 133, 084704 10.1063/1.3464481. [DOI] [PubMed] [Google Scholar]
- Oh S. H.; Hoflund G. G. Low-Temperature Catalytic Carbon Monoxide Oxidation over Hydrous and Anhydrous Palladium Oxide Powders. J. Catal. 2007, 245, 35–44. 10.1016/j.jcat.2006.09.016. [DOI] [Google Scholar]
- Chen H.-L.; Su C.-H.; Chen H.-T. Catalytic CO Oxidation by Au–Pd Core–Shell Nanoparticles: A First-Principles Study. Chem. Phys. Lett. 2012, 536, 100–103. 10.1016/j.cplett.2012.03.093. [DOI] [Google Scholar]
- Zhang J.; Jin H.; Sullivan M. B.; Lim F. C. H.; Wu P. Study of Pd–Au Bimetallic Catalysts for CO Oxidation Reaction by DFT Calculations. Phys. Chem. Chem. Phys. 2009, 11, 1441–1446. 10.1039/b814647k. [DOI] [PubMed] [Google Scholar]
- Liu Z. P.; Gong X. Q.; Kohanoff J.; Sanchez C.; Hu P. Catalytic Role of Metal Oxides in Gold-Based Catalysts: A First Principles Study of CO Oxidation on T iO2 Supported Au. Phys. Rev. Lett. 2003, 91, 266102 10.1103/PhysRevLett.91.266102. [DOI] [PubMed] [Google Scholar]
- Chen H.-T. First-Principles Study of CO Adsorption and Oxidation on Ru-Doped CeO2(111) Surface. J. Phys. Chem. C 2012, 116, 6239–6246. 10.1021/jp210864m. [DOI] [Google Scholar]
- Chen H.-T.; Chang J.-G. Computational Investigation of CO Adsorption and Oxidation on Iron-Modified Cerium Oxide. J. Phys. Chem. C 2011, 115, 14745–14753. 10.1021/jp201231d. [DOI] [Google Scholar]
- Hsu L.-C.; Tsai M.-K.; Lu Y.-H.; Chen H.-T. Computational Investigation of CO Adsorption and Oxidation on Mn/CeO2 (111) Surface. J. Phys. Chem. C 2013, 117, 433–441. 10.1021/jp310457g. [DOI] [Google Scholar]
- Huang M.; Fabris S. CO Adsorption and Oxidation on Ceria Surfaces from DFT+U Calculations. J. Phys. Chem. C 2008, 112, 8643–8648. 10.1021/jp709898r. [DOI] [Google Scholar]
- Wang C. M.; Fan K. N.; Liu Z. P. Origin of Oxide Sensitivity in Gold-Based Catalysts: A First Principle Study of CO Oxidation over Au Supported on Monoclinic and Tetragonal ZrO2. J. Am. Chem. Soc. 2007, 129, 2642–2647. 10.1021/ja067510z. [DOI] [PubMed] [Google Scholar]
- Liu X.; Sui Y. H.; Duan T.; Meng C. G.; Han Y. CO Oxidation on Pt-Graphene. Phys. Chem. Chem. Phys. 2014, 16, 23584–23593. 10.1039/C4CP02106A. [DOI] [PubMed] [Google Scholar]
- Tanaka K.-I.; Shou M.; Yuan Y. Low Temperature Prox Reaction of CO Catalyzed by Dual Functional Catalysis of the Pt Supported on Cnt, Cnf, Graphite, and Amorphous-C with Ni–Mgo, Fe, and Fe-Al2O3: Oxidation of Co Via Hcoo Intermediate. J. Phys. Chem. C 2010, 114, 16917–16923. 10.1021/jp103155g. [DOI] [Google Scholar]
- Yoo E. J.; Okata T.; Akita T.; Kohyama M.; Nakamura J.; Honma I. Enhanced Electrocatalytic Activity of Pt Subnanoclusters on Graphene Nanosheet Surface. Nano Lett. 2009, 9, 2255–2259. 10.1021/nl900397t. [DOI] [PubMed] [Google Scholar]
- Song E. H.; Wen Z.; Jiang Q. CO Catalytic Oxidation on Copper-Embedded Graphene. J. Phys. Chem. C 2011, 115, 3678–3683. 10.1021/jp108978c. [DOI] [Google Scholar]
- Li Y.; Zhou Z.; Yu G.; Chen W.; Chen Z. CO Catalytic Oxidation on Iron-Embedded Graphene: Computational Quest for Low-Cost Nanocatalysts. J. Phys. Chem. C 2010, 114, 6250–6254. 10.1021/jp911535v. [DOI] [Google Scholar]
- Lu Y. H.; Zhou M.; Zhang C.; Feng Y. P. Metal-Embedded Graphene: A Possible Catalyst with High Activity. J. Phys. Chem. C 2009, 113, 20156–20160. 10.1021/jp908829m. [DOI] [Google Scholar]
- Nigam S.; Majumder C. CO Oxidation by BN–Fullerene Cage: Effect of Impurity on the Chemical Reactivity. ACS Nano 2008, 2, 1422–1428. 10.1021/nn8001455. [DOI] [PubMed] [Google Scholar]
- Zheng Y.; Jiao Y.; Ge Y.; Jaroniec M.; Qiao S. Z. Two-Step Boron and Nitrogen Doping in Graphene for Enhanced Synergistic Catalysis. Angew. Chem. 2013, 125, 3192–3198. 10.1002/ange.201209548. [DOI] [PubMed] [Google Scholar]
- Wang S.; Iyyamperumal E.; Roy A.; Xue Y.; Yu D.; Dai L. Vertically Aligned Bcn Nanotubes as Efficient Metal-Free Electrocatalysts for the Oxygen Reduction Reaction: A Synergetic Effect by CO-Doping with Boron and Nitrogen. Angew. Chem., Int. Ed. 2011, 50, 11756–11760. 10.1002/anie.201105204. [DOI] [PubMed] [Google Scholar]
- Zhao Y.; Yang L.; Chen S.; Wang X.; Ma Y.; Wu Q.; Jiang Y.; Qian W.; Hu Z. Can Boron and Nitrogen CO-Doping Improve Oxygen Reduction Reaction Activity of Carbon Nanotubes. J. Am. Chem. Soc. 2013, 135, 1201–1204. 10.1021/ja310566z. [DOI] [PubMed] [Google Scholar]
- Mazzoni M. S. C.; Nunes R. W.; Azevedo S.; Chacham H. Electronic Structure and Energetics of Bxcynz Layered Structures. Phys. Rev. B 2006, 73, 073108 10.1103/PhysRevB.73.073108. [DOI] [Google Scholar]
- da Rocha Martins J.; Chacham H. Disorder and Segregation in B–C–Ngraphene-Type Layers and Nanotubes: Tuning the Band Gap. ACS Nano 2011, 5, 385–393. 10.1021/nn101809j. [DOI] [PubMed] [Google Scholar]
- da Rocha Martins J.; Chacham H. N-Rich Gr:Bn Layers: From Segregated Alloy to Solid Solution. Phys. Rev. B 2012, 86, 075421 10.1103/PhysRevB.86.075421. [DOI] [Google Scholar]
- Zabet-Khosousi A.; Zhao L.; Pálová L.; Hybertsen M. S.; Reichman D. R.; Pasupathy A. N.; Flynn G. W. Segregation of Sublattice Domains in Nitrogen-Doped Graphene. J. Am. Chem. Soc. 2014, 136, 1391–1397. 10.1021/ja408463g. [DOI] [PubMed] [Google Scholar]
- Lin I. H.; Lu Y. H.; Chen H. T. Nitrogen-Doped C60 as a Robust Catalyst for Co Oxidation. J. Comput. Chem. 2017, 38, 2041–2046. 10.1002/jcc.24851. [DOI] [PubMed] [Google Scholar]
- Lin I. H.; Hu Y. H.; Chen H. T. Nitrogen-Doped Carbon Nanotube as a Potential Metal-Free Catalyst for CO Oxidation. Phys. Chem. Chem. Phys. 2016, 18, 12093–12100. 10.1039/C6CP00162A. [DOI] [PubMed] [Google Scholar]
- Hu X. B.; Wu Y. T.; Zhang Z. B. CO Oxidation on Metal-Free Nitrogen-Doped Carbon Nanotubes and the Related Structure–Reactivity Relationships. J. Mater. Chem. 2012, 22, 15198–15205. 10.1039/c2jm33490a. [DOI] [Google Scholar]
- Krishnan R.; Su W.-S.; Chen H.-T. A. New Carbon Allotrope: Penta-Graphene as a Metal-Free Catalyst for CO Oxidation. Carbon 2017, 114, 465–472. 10.1016/j.carbon.2016.12.054. [DOI] [Google Scholar]
- Krishnan R.; Wu S.-Y.; Chen H.-T. Nitrogen-Doped Penta-Graphene as a Superior Catalytic Activity for CO Oxidation. Carbon 2018, 132, 257–262. 10.1016/j.carbon.2018.02.064. [DOI] [Google Scholar]
- Krishnan R.; Wu S.-Y.; Chen H.-T. Catalytic CO Oxidation on B-Doped and Bn CO-Doped Penta-Graphene: A Computational Study. Phys. Chem. Chem. Phys. 2018, 20, 26414–26421. 10.1039/C8CP04745F. [DOI] [PubMed] [Google Scholar]
- Wu P.; Du P.; Zhang H.; Cai C. Graphdiyne as a Metal-Free Catalyst for Low-Temperature CO Oxidation. Phys. Chem. Chem. Phys. 2014, 16, 5640–5648. 10.1039/C3CP55121K. [DOI] [PubMed] [Google Scholar]
- Sinthika S.; Kumar E. M.; Surya V. J.; Kawazoe Y.; Park N.; Iyakutti K.; Thapa R. Activation of CO and CO2 on Homonuclear Boron Bonds of Fullerene-Like Bn Cages: First Principles Study. Sci. Rep. 2015, 5, 17460 10.1038/srep17460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kao W.-H.; Wu S.-Y.; Chen H.-T. Exploring Reaction Mechanisms for CO Oxidation on Boron-Doped Carbon Nanotubes: A Computational Approach. Appl. Surf. Sci. 2019, 480, 63–69. 10.1016/j.apsusc.2019.02.223. [DOI] [Google Scholar]
- Yang L.; Jiang S.; Zhao Y.; Zhu L.; Chen S.; Wang X.; Wu Q.; Ma J.; Ma Y.; Hu Z. Boron-Doped Carbon Nanotubes as Metal-Free Electrocatalysts for the Oxygen Reduction Reaction. Angew. Chem. 2011, 123, 7270–7273. 10.1002/ange.201101287. [DOI] [PubMed] [Google Scholar]
- Esrafili M. D.; Heidari S. CO Catalytic Oxidation over C59X Heterofullerenes (X = B, Si, P, S): A Dft Study. Comput. Theor. Chem. 2019, 1151, 50–57. 10.1016/j.comptc.2019.02.007. [DOI] [Google Scholar]
- Gao B.; Chen G. CO Oxidization Catalyzed by B, N, and Their Codoped Fullerenes: A First-Principles Investigation. RSC Adv. 2019, 9, 21626–21636. 10.1039/C9RA02172H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nørskov J. K.; Rossmeisl J.; Logadottir A.; Lindqvist L.; Kitchin J. R.; Bligaard T.; Jónsson H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. 10.1021/jp047349j. [DOI] [Google Scholar]
- Petersen M. A.; Jenkins S. J.; King D. A. Ridge-Bridge Adsorption of Molecular Oxygen on Pt{110}(1 × 2) from First Principles. J. Phys. Chem. B 2006, 110, 11962–11970. 10.1021/jp057447k. [DOI] [PubMed] [Google Scholar]
- Escaño M. C. S.; Nakanishi H.; Kasai H. The Role of Ferromagnetic Substrate in the Reactivity of Pt/Fe Overlayer: A Density Functional Theory Study. J. Phys. Soc. Jpn. 2009, 78, 064603 10.1143/jpsj.78.064603. [DOI] [Google Scholar]
- Balbuena P. B.; Altomare D.; Agapito L.; Seminario J. M. Theoretical Analysis of Oxygen Adsorption on Pt-Based Clusters Alloyed with Co, Ni, or Cr Embedded in a Pt Matrix. J. Phys. Chem. B 2003, 107, 13671–13680. 10.1021/jp035729j. [DOI] [Google Scholar]
- Liu C.; Tan Y.; Lin S.; Li H.; Wu X.; Li L.; Pei Y.; Zeng X. C. CO Self-Promoting Oxidation on Nanosized Gold Clusters: Triangular Au3 Active Site and CO Induced O–O Scission. J. Am. Chem. Soc. 2013, 135, 2583–2595. 10.1021/ja309460v. [DOI] [PubMed] [Google Scholar]
- Chen H.-T.; Chang J.-G.; Ju S.-P.; Chen H.-L. Ethylene Epoxidation on a Au Nanoparticle Versus a Au(111) Surface: A Dft Study. J. Phys. Chem. Lett. 2010, 1, 739–742. 10.1021/jz900469f. [DOI] [Google Scholar]
- Chen H.-T.; Chan C.-W. Promoting Ethylene Epoxidation on Gold Nanoclusters: Self and CO Induced O2 Activation. Phys. Chem. Chem. Phys. 2015, 17, 22336–22341. 10.1039/C5CP02809D. [DOI] [PubMed] [Google Scholar]
- Lee C.-C.; Chen H.-T. Adsorption and Reaction of C2H4 and O2 on a Nanosized Gold Cluster: A Computational Study. J. Phys. Chem. A 2015, 119, 8547–8555. 10.1021/acs.jpca.5b04737. [DOI] [PubMed] [Google Scholar]
- Lin I.-H.; Lu Y.-H.; Chen H. -T. Nitrogen-Doped C60 as a Robust Catalyst for CO Oxidation. J. Comput. Chem. 2017, 38, 2041–2046. 10.1002/jcc.24851. [DOI] [PubMed] [Google Scholar]
- Gao M.; Lyalin A.; Taketsugu T. CO Oxidation on h-BN Supported Au Atom. J. Chem. Phys. 2013, 138, 034701 10.1063/1.4774216. [DOI] [PubMed] [Google Scholar]
- Li F.; Zhao J.; Chen Z. Fe-Anchored Graphene Oxide: A Low-Cost and Easily Accessible Catalyst for Low-Temperature CO Oxidation. J. Phys. Chem. C 2012, 116, 2507–2514. 10.1021/jp209572d. [DOI] [Google Scholar]
- Lin I.-H.; Lu Y.-H.; Chen H.-T. Nitrogen-Doped Carbon Nanotube as a Potential Metal-Free Catalyst for CO Oxidation. Phys. Chem. Chem. Phys. 2016, 18, 12093–12100. 10.1039/C6CP00162A. [DOI] [PubMed] [Google Scholar]
- Kresse G.; Furthmuller J.. Vienna Ab-initio Simulation Package, 2006, VASP 4.6.
- Kresse G.; Furthmu̅ller J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Wang Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B 1992, 45, 13244–13249. 10.1103/PhysRevB.45.13244. [DOI] [PubMed] [Google Scholar]
- Monkhorst H. J.; Pack J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. 10.1103/PhysRevB.13.5188. [DOI] [Google Scholar]
- Bielanski A.; Haber J., Oxygen in Catalysis; Marcel Dekker Inc.: New York, 1991. [Google Scholar]
- Huber K. P.; Herzberg G., Molecular Spectra and Molecular Structure. Iv. Constants of Diatomic Molecules; Van Nostrand Reinhold CO.: New York, 1979. [Google Scholar]
- Herzberg G., Electronic Spectra and Electronic Structure of Polyatomic Molecules; Van Nostrand: New York, 1966. [Google Scholar]
- Henkelman G. U.; 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. 10.1063/1.1329672. [DOI] [Google Scholar]
- 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. 10.1063/1.1323224. [DOI] [Google Scholar]
- Bader R. F. W.Atoms in Molecules - a Quantum Theory; Clarendon Press: Oxford, 1994. [Google Scholar]
- Bader R. F. W.; Beddall P. M. Virial Field Relationship for Molecular Charge Distributions and the Spatial Partitioning of Molecular Properties. J. Chem. Phys. 1972, 56, 3320–3328. 10.1063/1.1677699. [DOI] [Google Scholar]
- Henkelman G.; Arnaldsson A.; Jónsson H. A fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354–360. 10.1016/j.commatsci.2005.04.010. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.

