Next Article in Journal
Acid Mine Drainage Precipitates from Mining Effluents as Adsorbents of Organic Pollutants for Water Treatment
Previous Article in Journal
Advanced Application of Polymer Nanocarriers in Delivery of Active Ingredients from Traditional Chinese Medicines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting the Electrocatalytic Oxygen Reduction Activity of MnN4-Doped Graphene by Axial Halogen Ligand Modification

1
College of Physics and Electronic Information, Inner Mongolia Normal University, Hohhot 010022, China
2
Inner Mongolia Key Laboratory for Physics and Chemistry of Functional Materials, Hohhot 010022, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(15), 3517; https://doi.org/10.3390/molecules29153517
Submission received: 4 July 2024 / Revised: 23 July 2024 / Accepted: 25 July 2024 / Published: 26 July 2024

Abstract

:
Exploring highly active electrocatalysts as platinum (Pt) substitutes for the oxygen reduction reaction (ORR) remains a significant challenge. In this work, single Mn embedded nitrogen-doped graphene (MnN4) with and without halogen ligands (F, Cl, Br, and I) modifying were systematically investigated by density functional theory (DFT) calculations. The calculated results indicated that these ligands can transform the dyz and dxz orbitals of Mn atom in MnN4 near the Fermi-level into d z 2 orbital, and shift the d-band center away from the Fermi-level to reduce the adsorption capacity for reaction intermediates, thus enhancing the ORR catalytic activity of MnN4. Notably, Br and I modified MnN4 respectively with the lowest overpotentials of 0.41 and 0.39 V, possess superior ORR catalytic activity. This work is helpful for comprehensively understanding the ligand modification mechanism of single-atom catalysts and develops highly active ORR electrocatalysts.

Graphical Abstract

1. Introduction

With the ever-increasing energy crisis and environmental issues, a major challenge facing mankind is to develop eco-friendly and renewable energy conversion and storage devices, such as fuel cells and rechargeable batteries [1,2,3]. However, due to the sluggish kinetics of the cathodic oxygen reduction reaction (ORR), these devices are heavily dependent on Pt catalysts to accelerate the reaction process [4,5]. Unfortunately, the natural scarcity and prohibitive cost of Pt greatly increased the large-scale commercialization cost of fuel cells and rechargeable batteries [6,7]. Therefore, it is imperative to explore inexpensive and high-performance materials as alternatives to Pt catalysts.
In this case, graphene-supported nitrogen-doped transition metal (M−N−G) single atom catalysts (SACs) have attracted increasing attention because of their efficient use of atoms and their outstanding mechanical flexibility, and which are considered as promising alternative material to Pt catalysts [8,9,10]. However, their catalytic activities are greatly impeded by the unsatisfactory adsorption strength for ORR intermediates [11,12]. Optimizing the electronic structure of metal active center can further improve their catalytic activities, including axial ligand coordination engineering, heteroatom doping, and the construction of diatomic or polyatomic active sites [13,14,15]. Among which, axial ligand coordination engineering was proved to be a novel and feasible strategy to boost the catalytic activity of M−N−Gs, especially for halogen ligands [13,16,17,18,19,20,21]. For example, Han et.al. successfully prepared an FeN4Cl catalyst with a half-wave potential of 0.921 V, which showed remarkably enhanced ORR catalytic activity compared to FeN4. Furthermore, density functional theory (DFT) calculations indicated that the axial Cl ligand can effectively modulate the adsorption capacity of FeN4 for reaction intermediates and transform the rate-determination step (RDS) of *OH reduction into *OOH formation, resulting in the superior ORR activity of FeN4Cl [22]. Both experiments and DFT calculations have confirmed that FeN4I possesses superior ORR catalytic activity, and the corresponding half-wave potential can reach to 0.948 V, which should be attributed to the appearance of Fe d z 2  orbital with a lower energy level caused by I modification [23]. Chen et al. studied a series of MN4X (M = Fe, Co, and Ni; X = F, Cl, Br, and I) catalysts by performing DFT calculations, and found that axial coordination of X can modulate the electronic structure of M active center to alter the binding of MN4 with reaction intermediates. In particular, CoN4Cl and CoN4Br, respectively, with ultralow overpotentials of 0.25 and 0.26 V possess excellent ORR catalytic activity [24]. In addition, the axial halogen ligand can regulate the selectivity of the ORR pathway. For example, different from the four-electron (4e) ORR pathway on CuN4 and ZnN4, CuN4Cl and ZnN4Br are more favorable with a two-electron (2e) ORR pathway to form H2O2, and the overpotentials are only 0.07 and 0.05 V, respectively [25]. Although certain M−N−Gs modified by halogen ligands have shown excellent catalytic activity, a comprehensive examination of single Mn−embedded and N−doped graphene (MnN4) modified by various halogen ligands (F, Cl, Br, and I) is still lacking. According to available studies, the overpotential of ORR on MnN4 is about 0.90 V [26], which is significantly higher than those for Pt (100) (0.47 V) [27] and Pt (111) (0.44 V) [28]. Therefore, it is expected to reduce the ORR overpotential and enhance the catalytic activity of MnN4 by halogen ligand modification. Moreover, how the halogen ligands regulate the electronic structures, reaction intermediates adsorption characteristics, ORR catalytic activity, and reaction mechanism of MnN4 remain to be further explored.
Herein, halogen ligands X (X = F, Cl, Br, and I) were introduced to axially modify the active center of MnN4, named MnN4−X. The structural stabilities, electronic properties, and catalytic characters of MnN4 and MnN4−X were investigated by DFT calculations. Our calculated results showed that axial halogen ligands can significantly improve the ORR activity of MnN4. In particular, the overpotentials of MnN4−Cl, MnN4−Br, and MnN4−I are 0.44, 0.41, and 0.39 V, respectively, which are comparable to or even better than those of Pt catalysts. This work provides great significance for the design and development of low cost and high efficiency graphene-based electrocatalysts.

2. Results and Discussion

2.1. Structural Model

A 5 × 5 graphene supercell containing 50 carbon atoms was used to construct a MnN4 catalyst, as shown in Figure 1. It can be seen that all of the atoms are kept in the same plane, and the average bond length of Mn−N bonds, the formation energy (Ef1), and the binding energy between Mn and N4−G (Eb1) are 1.916 Å, −3.16 eV, and 6.90 eV, respectively. These values are consistent with those (1.920 Å, −3.07 eV, and 6.84 eV) in available studies [26,29,30], implying the reliability of our calculations. Moreover, the negative Ef1 and the Eb1 larger than the cohesive energy (3.87 eV) of bulk Mn indicate the good structural stability of MnN4. Hence, both sides of MnN4 can adsorb ORR intermediates, potentially realizing an axial ligand coordination with the Mn atom [8,31]. Thus, axial halogen ligand X was introduced into MnN4 to form MnN4−X catalysts (see Figure 1). Due to the effect of X, the average bond lengths of Mn−N bonds in MnN4−F, MnN4−Cl, MnN4−Br, and MnN4−I are, respectively, stretched into 1.975 Å, 1.975 Å, 1.973 Å, and 1.971 Å. Meanwhile, the Mn atom is deviated from the plane of graphene. To evaluate the structural stabilities MnN4−X, the binding energies (Eb2) between X and MnN4 and the formation energy (Ef2) for MnN4−X were calculated by Equations (10) and (12) and are presented in Figure 2. One can find that all MnN4−X structures have smaller formation energy than that of MnN4, indicating the better structural stability of MnN4−X than MnN4. Moreover, the positive values of Eb2 suggest the good combinations between X and MnN4. Furthermore, the AIMD simulations show that the energy of each MnN4 and MnN4−X structure smoothly oscillates around a certain value, and there is no bond breakage, although their structures were wrinkled to some extent, as indicated by Figure S1. These results indicate that all structures mentioned above are thermodynamically stable. In addition, according to the calculated positive values of Udiss in Table S1, all structures are verified to be electrochemically stable. Therefore, all MnN4 and MnN4−X structures are thermodynamically and electrochemically stable and are considered in the following discussions. From the Bader charge analyses in Table S1, the Mn atom loses 1.28 e to N4−G in MnN4. After introducing X into MnN4, the ligands of F, Cl, Br, and I, respectively, obtained 0.65, 0.59, 0.53, and 0.47 e from the Mn atom, in agreement with the electronegative order of F (3.98) > Cl (3.16) > Br (2.96) > I (2.66). The Mn atom loses more charges than that in MnN4, presenting the order of MnN4−F (1.50 e) > MnN4−Cl (1.40 e) > MnN4−Br (1.36 e) > MnN4−I (1.30 e). Therefore, the X ligand modification would lead to different charge densities of Mn atoms and affect the adsorption activities for ORR intermediates.

2.2. Adsorption of Intermediates

According to the Sabatier’s principle [32], the catalytic activity of catalysts is strongly dependent on the adsorption strength for reaction intermediates. Thus, the adsorption characteristics of ORR intermediates on MnN4 and MnN4−X were explored. After optimizing all horizontal and vertical configurations of *O2, *OOH, *O, and *OH with different orientations on all catalysts, the most stable adsorption structures were obtained and are presented in Figure 3. It can be seen from this figure that all intermediates are preferably adsorbed on the Mn site through an O atom, indicating that the Mn atom is the active center in MnN4 and MnN4−X. This is understandable since the net spin charges mainly distribute on the Mn atom in MnN4 and MnN4−X (see Figure 4). In theory, the adsorption of the O2 molecule is regarded as a prerequisite for triggering the ORR, so it is firstly considered. As shown in Figure 3, the O2 molecule tends to adsorb on the Mn site through side-on configuration in MnN4 but through end-on configuration in MnN4−X. The adsorption energies of O2 on MnN4, MnN4−F MnN4−Cl, MnN4−Br, and MnN4−I, respectively, are −1.34, −0.51, −0.25, −0.26, and −0.44 eV (see Table S2), and MnN4 exhibits the strongest adsorption capability to O2. Among this, the adsorption energy of O2 on MnN4 is in agreement with the values of −1.33 and −1.43 eV in previous studies [9,26,29]. Meanwhile, the O−O bonds are, respectively, elongated from 1.232 Å in gas-phase O2 to 1.392, 1.302, 1.305, 1.304, and 1.295 Å. Since the spin charges of the active center and the Bader charges analysis can be used to describe the binding strength between the metal center and adsorbate in the catalysts [33,34,35], the spin charges of the Mn atom and the charges transfer between O2 and the catalysts were calculated to explore the difference of O2 adsorption strength on MnN4 and MnN4−X. Figure 4 and Table S3 show that the spin charge of the Mn atom in MnN4 is significantly higher than that in MnN4−X, and the Mn atom in MnN4 can transfer more charges (0.70 e) to O2 than that in MnN4−X (0.50, 0.53, 0.51, and 0.44 e, respectively, for MnN4−F, MnN4−Cl, MnN4−Br, and MnN4−I), indicating the strongest interactions between O2 and MnN4. However, the adsorption energy of MnN4 for O2 is much stronger than that on Pt (−1.10 eV) [27], FeN4 (−1.13 eV), and CoN4 (−0.90 eV) [36], which may hinder the subsequent reaction processes. In contrast, MnN4−X is likely to possess better ORR activity than Pt, FeN4, CoN4, and MnN4. Similar to the adsorption of the O2 molecule, the X ligand can weaken the interactions between *OOH, *O, and *OH and MnN4 to different extents, as shown in Table S2. Therefore, the X ligand modification can effectively alter the adsorption capacity of MnN4 towards reaction intermediates, and then optimize the ORR catalytic activity.

2.3. Catalytic Performance

According to the 4e reaction pathway (Equations (1)–(4)) of ORR, the free energy change for all catalysts was calculated by Equation (7) and presented in Table S2 and Figure 5 to evaluate the catalytic performance. At zero electrode potential, the free energy in the whole reaction progress is gradually decreased for all catalysts, which means that each step of the reaction is exothermic and can spontaneously proceed. When increasing electrode potential, MnN4, MnN4−F, MnN4−Cl, MnN4−Br, and MnN4−I, respectively, present working potentials of 0.34, 0.59, 0.79, 0.82, and 0.84 V for maintaining the exothermic reaction process (see Figure 5). Among this, MnN4−Br and MnN4−I with larger working potentials are more prone to promote ORR compared with other catalysts. Under the standard ORR potential of 1.23 V, some reaction steps become uphill and are thermodynamically unfavorable. In particular, the hydrogenation of *OH to H2O with the maximum free energy barrier is the RDS for the ORR on all catalysts. In order to intuitively compare the catalytic activity, the theoretical overpotential is calculated through Equation (8) and illustrated in Figure 5. Note that the lower overpotential represents the higher ORR activity. The overpotentials of MnN4, MnN4−F, MnN4−Cl, MnN4−Br, and MnN4−I are 0.89, 0.64, 0.44, 0.41, and 0.39 V, respectively. Evidently, MnN4−I with the lowest overpotential of 0.39 V exhibits the best catalytic activity. Compared with the overpotentials of Pt (100) (0.47 V) [27] and Pt (111) (0.43 V) [28], MnN4−Cl and MnN4−Br/I, respectively, show comparable and higher catalytic activity, indicating that MnN4−Cl, MnN4−Br, and MnN4−I should be promising alternatives to Pt catalysts. Thus, X ligand modification is an effective strategy to improve the ORR catalytic activity of MnN4. Besides the 4e reaction pathway, the 2e reaction pathway (Equations (5) and (6)) [37,38] is also investigated to identify the ORR selectivity of all catalysts. The optimized adsorption structures of H2O2 on MnN4 and MnN4−X in Figure S2 indicate that the H2O2 molecules cannot be stably adsorbed on MnN4, MnN4−Cl, and MnN4−Br and decompose into *O and H2O, which is consistent with the hydrogenation products of *OOH in the 4e reaction pathway, implying the better 4e reaction pathway selectivity. Although MnN4−F and MnN4−I can physically adsorb H2O2 with the adsorption energies of −0.08 and −0.06 eV, the free energy changes of *OOH to *O+H2O at zero electrode potential for MnN4−F (−1.70 eV) and MnN4−I (−1.86 eV) are significantly smaller than that (−0.20 and −0.39 eV) for *OOH to H2O2, indicating the 4e reaction pathway is easier to occur. Therefore, the ORR tends to follow the 4e reaction pathway on all catalysts.

2.4. Origin of the Catalytic Activity

To specifically analyze the origin of excellent catalytic activity of MnN4−X, the projected density of states (PDOS), d-band center, and the local density of states (LDOS) of Mn in MnN4 and MnN4−X were presented in Figure 6. It can be seen from this figure that that the dyz and dxz orbitals of Mn atoms in MnN4 are mainly located near the Fermi-level. When incorporating X ligand into MnN4, the d-orbital of the Mn atom moves towards the Fermi-level, and the main contribution near the Fermi-level is transformed into the d z 2 orbital, resulting in the enhanced reaction activity. Therefore, the O2 adsorption configuration is transformed from side-on for MnN4 into end-on for MnN4−X. Moreover, the d z 2 orbital occupations near the Fermi-level are significantly weaker than dyz and dxz orbitals in MnN4, and the d-band center of the Mn atom in MnN4 downshift from −0.81 eV to −1.10, −1.13, −1.11, and −1.07 eV in MnN4−F, MnN4−Cl, MnN4−Br, and MnN4−I, respectively. These changes result in the weakened adsorption capacity of MnN4−X for reaction intermediates as compared with MnN4. In particular, MnN4−Cl, MnN4−Br, and MnN4−I present appropriate adsorption strength for reaction intermediates and excellent ORR catalytic activity. In addition, the electronic band structures play an important role in determining the electronic transmission properties of catalysts [39]. Hence, we further examine the total density of states (TDOS) of MnN4 and MnN4−X. As shown in Figure S3, after modifying with X, both spin-up and spin-down orbitals of MnN4 cross through the Fermi-level, achieving the transition from semiconductor to metallic behaviors. This is conducive to the electron transfer during the reaction process, thereby enhancing the ORR catalytic activity. Therefore, halogen ligand modification can be an effective strategy to regulate the electronic structure of metal active center and improve the ORR catalytic activity of graphene-based SACs.

3. Computational Method

All spin-polarized DFT calculations were performed using the Vienna ab initio Simulation Package (VASP) [40]. The Perdew–Burke–Ernzerhof (PBE) within the generalized gradient approximate functional (GGA) was used to deal with exchange–correlation interactions [40,41]. For all calculations, the kinetic energy cutoff was chosen to be 500 eV, thus achieving good energy convergence. The thickness of the vacuum layer was set to be 15 Å to avoid interactions between mirror images. All structures were allowed to converge until Feynman Hallman forces were smaller than 0.02 eV/Å, and the total energy fluctuation was set to be 10−5 eV. The Brillouin zone was sampled with 5 × 5 × 1 k-points for structural optimization, and 11 × 11 × 1 k-points were sampled for the electronic structure calculations. The DFT-D3 method with the Becke–Johnson damping function was used to correct the van der Waals force dispersion [42]. Ab initio molecular dynamics (AIMD) simulations were performed using the Nosé–Hoover thermostat at 300 K in the canonical ensemble (NVT) with a time step of 1 fs and a total time scale of 10 ps.
ORR includes two different competing reaction pathways, namely, the 4e reaction pathway for the reduction of O2 to H2O and the 2e reaction pathway for the reduction of O2 to H2O2.The 4e reaction process under an acidic medium is listed as follows
* + O 2 + H + + e = * OOH
* OOH + H + + e = * O + H 2 O
* O + H + + e = * OH
* OH + H + + e = H 2 O
The 2e reaction process under an acidic medium is presented as follows
* + O 2 + H + + e = * OOH
* OOH + H + + e = H 2 O 2
The change in Gibbs free energy for each step of the reaction is defined as
Δ G = Δ E DFT + Δ E ZPE T Δ S + Δ G U + G pH
where Δ E DFT represents the total energy calculated by DFT. Δ E ZPE and Δ S , respectively, are the changes in zero-point energy and entropy. T stands for the room temperature of 298.15 K.   Δ G U = neU , where n is the number of electrons required to complete the reaction, and U is the electrode potential.   Δ G PH = k B T × ln 10 × pH , in which k B stands for the Boltzmann’s constant, and pH is set to 0 in an acidic medium.
For the 4e reaction pathway, the step with the maximum free energy barrier is considered to be the RDS. The overpotential of the whole reaction progress is calculated by the following equation
η = 1.23 + max ( Δ G 1 ,   Δ G 2 ,   Δ G 3 ,   Δ G 4 ) / e
The binding energy (Eb1) between Mn and N−coordinated graphene (N4−G) and the binding energy (Eb2) between MnN4 and X are, respectively, given by the following equations
E b 1 = E Mn + E N 4 G E Mn N 4
E b 2 =   E Mn N 4 + 1 / 2 E X 2   E Mn N 4 X
where E Mn ,   E N 4 G , E Mn N 4 , and E Mn N 4 X represent the total energies of a single Mn atom, N4−G, MnN4, and X modified MnN4, respectively. E X 2 represents the energy of a single halogen molecule.
The formation energies (Ef1 and Ef2) of MnN4 and MnN4−X are, respectively, defined as
E f 1 =   E Mn N 4     E Mn   4 μ N   44 μ C
E f 2 = E Mn N 4 X E Mn   4 μ N 44 μ C μ X
where μ C , μ N , and μ X denote the chemical potentials of C, N, and halogen atoms, which can be obtained from perfect graphene, nitrogen gas, and halogen molecules, respectively.
The adsorption energy (Eads) of catalysts for reaction intermediates is defined as the following equation
Eads = Esup+interEsupEinter
where Esup+inter, Esup, and Einter represent the energies of the adsorption system, the individual support, and the reaction intermediates, respectively.
In addition, the electrochemical stability of catalysts can be evaluated by the solution potential (Udiss) [43], which is defined as
U diss = U diss ° ( metal ,   bulk ) E f ne
where U diss ° ( metal ,   bulk ) and n represent the standard dissolution potential of bulk metal and the number of electrons involved in the dissolution, respectively. Udiss > 0 vs. the standard hydrogen electrode means that the system is electrochemically stable.

4. Conclusions

In summary, the structural stabilities, electronic structures, and ORR catalytic performances of MnN4 with and without halogen ligand X (X = F, Cl, Br, and I) modification were systemically investigated by DFT calculations. The calculated results show that the MnN4 and MnN4−X catalysts are thermodynamically and electrochemically stable, and follow the four-electron ORR pathway. After introducing axial X into MnN4, the X can obtain some electrons from the Mn atom in MnN4 and regulate the electronic structures of the Mn atom. The dyz and dxz orbitals of the Mn atom near the Fermi-level are transformed into the weaker d z 2 orbital, and the d-band center of the Mn atom is shifted away from the Fermi-level, resulting in the reduced adsorption strength of MnN4 for *OH. Thus, all MnN4−X catalysts present enhanced ORR catalytic activity compared with MnN4. In particular, MnN4−Br and MnN4−I exhibit excellent ORR catalytic activity, respectively, with the overpotentials of 0.41 and 0.39 V, which are superior to MnN4 and Pt catalysts. Halogen ligand modification was identified to be an effective strategy for improving the ORR catalytic activity. This study is expected to provide a new perspective for the design and development of highly active graphene-based single-atom ORR catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29153517/s1. Figure S1. Evolution of energy with the elapsed time for AIMD simulations of MnN4 and MnN4−X at 300 K. Inserts are top and side views of the snapshot of atomic configuration at 10 ps; Figure S2. Optimized structures of *HOOH adsorbed on MnN4 and MnN4−X; Figure S3. The total density of states (TDOS) of MnN4 and MnN4−X.; Table S1. The average Mn−N bond length (dMn−N), binding energy (Eb), formation energy (Ef), dissolution potential (Udiss), and Bader charge transfer for MnN4 and MnN4−X.; Table S2. Adsorption energy (Eads) and Gibbs free energy (ΔG) for reaction intermediates; Table S3. Bader charge transfer for intermediates adsorbed MnN4 and MnN4−X systems.

Author Contributions

Calculation and writing—original draft, S.W.; investigation and formal analysis, R.Z. and W.Y.; writing—review and editing, M.Z. and L.L.; visualization, L.L.; funding acquisition, M.Z. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Inner Mongolia Autonomous Region (Grant Nos. 2023QN01010 and 2024MS01004), the Research Program of Science and Technology at Universities of Inner Mongolia Autonomous Region (Grant No. NJZZ23019), the Fundamental Research Funds for the Inner Mongolia Normal University (Grant Nos. 2023JBYJ016 and 2023JBZD006), and the Innovation and Entrepreneurship Training Program for College Students at Inner Mongolia Normal University (Grant No. 202310135010).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, H.F.; Xu, Q. Materials design for rechargeable metal-air batteries. Matter 2019, 1, 565–595. [Google Scholar] [CrossRef]
  2. Zhang, Q.N.; Dong, S.D.; Shao, P.P.; Zhu, Y.H.; Mu, Z.J.; Sheng, D.F.; Zhang, T.; Jiang, X.; Shao, R.W.; Ren, Z.; et al. Covalent organic framework-based porous ionomers for high-performance fuel cells. Science 2022, 378, 181–186. [Google Scholar] [CrossRef]
  3. Brett, D.J.L.; Aguiar, P.; Brandon, N.P.; Bull, R.N.; Galloway, R.C.; Hayes, G.W.; Lillie, K.; Mellors, C.; Millward, M.; Smith, C.; et al. Project absolute: A zebra battery/Intermediate temperature solid oxide fuel cell hybrid for automotive applications. J. Fuel Cell Sci. Technol. 2006, 3, 254–262. [Google Scholar] [CrossRef]
  4. Mukerjee, S.; Srinivasan, S.; Soriaga, M.P.; Mcbreen, J. Role of structural and electronic properties of Pt and Pt alloys on electrocatalysis of oxygen reduction: An in situ XANES and EXAFS investigation. J. Electrochem. Soc. 1995, 142, 1409–1422. [Google Scholar] [CrossRef]
  5. Ammal, S.C.; Heyden, A. Origin of the unique activity of Pt/TiO2 catalysts for the water–gas shift reaction. J. Catal. 2013, 306, 78–90. [Google Scholar] [CrossRef]
  6. Debe, M.K. Electrocatalyst approaches and challenges for automotive fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  7. Huang, L.; Zaman, S.; Tian, X.L.; Wang, Z.T.; Fang, W.S.; Xia, B.Y. Advanced platinum-based oxygen reduction electrocatalysts for fuel cells. Acc. Chem. Res. 2021, 54, 311–322. [Google Scholar] [CrossRef]
  8. Niu, W.J.; He, J.Z.; Gu, B.N.; Liu, M.C.; Chueh, Y.L. Opportunities and challenges in precise synthesis of transition metal single-atom supported by 2D materials as catalysts toward oxygen reduction reaction. Adv. Funct. Mater. 2021, 31, 2103558. [Google Scholar] [CrossRef]
  9. Li, L.; Huang, R.; Cao, X.R.; Wen, Y.H. Computational screening of efficient graphene-supported transition metal single atom catalysts toward the oxygen reduction reaction. J. Mater. Chem. A 2020, 8, 19319–19327. [Google Scholar] [CrossRef]
  10. Zhuo, H.Y.; Zhang, X.; Liang, J.X.; Yu, Q.; Xiao, H.; Li, J. Theoretical understandings of graphene-based metal single-atom catalysts: Stability and catalytic performance. Chem. Rev. 2020, 120, 12315–12341. [Google Scholar] [CrossRef]
  11. Wang, J.; Liu, W.; Luo, G.; Li, Z.; Zhao, C.; Zhang, H.R.; Zhu, M.; Xu, Q.; Wang, X.Q.; Zhao, C.M.; et al. Synergistic effect of well-defined dual sites boosting the oxygen reduction reaction. Energy Environ. Sci. 2018, 11, 3375–3379. [Google Scholar] [CrossRef]
  12. Luo, E.G.; Chu, Y.Y.; Liu, J.; Shi, Z.P.; Zhu, S.Y.; Gong, L.Y.; Ge, J.J.; Choi, C.H.; Liu, C.P.; Xing, W. Pyrolyzed M–Nx catalysts for oxygen reduction reaction: Progress and prospects. Energy Environ. Sci. 2021, 14, 2158–2185. [Google Scholar] [CrossRef]
  13. Zhou, Y.; Xing, Y.F.; Wen, J.; Ma, H.B.; Wang, F.B.; Xia, X.H. Axial ligands tailoring the ORR activity of cobalt porphyrin. Sci. Bull. 2019, 64, 1158–1166. [Google Scholar] [CrossRef]
  14. Sun, J.K.; Pan, Y.W.; Xu, M.Q.; Sun, L.; Zhang, S.L.; Deng, W.Q.; Zhai, D. Heteroatom doping regulates the catalytic performance of single-atom catalyst supported on graphene for ORR. Nano Res. 2024, 17, 1086–1093. [Google Scholar] [CrossRef]
  15. Li, R.Z.; Wang, D.S. Superiority of dual-atom catalysts in electrocatalysis: One step further than single-stom catalysts. Adv. Energy Mater. 2022, 12, 2103564. [Google Scholar] [CrossRef]
  16. Wang, G.L.; Yang, X.L.; Wang, R.Y.; Jia, J.F. A theoretical study on the ORR electrocatalytic activity of axial ligand modified cobalt polyphthalocyanine. Phys. Chem. Chem. Phys. 2023, 25, 27342–27351. [Google Scholar] [CrossRef]
  17. She, X.L.; Gao, J.H.; Gao, Y.; Tang, H.; Li, K.; Wang, Y.; Wu, Z.J. Axial ligand engineering for highly efficient oxygen reduction catalysts in transition metal–N4 doped graphene. New J. Chem. 2022, 46, 16138–16150. [Google Scholar] [CrossRef]
  18. Shakir, R.; Komsa, H.P.; Sinha, A.S.K.; Karthikeyan, J. Electronic origin of enhanced selectivity through the halogenation of a single Mn atom on graphitic C3N4 for electrocatalytic reduction of CO2 from first-principles calculations. J. Phys. Chem. C 2023, 127, 14694–14703. [Google Scholar] [CrossRef]
  19. Liu, X.J.; Liu, Y.R.; Yang, W.X.; Feng, X.; Wang, B. Controlled modification of axial coordination for transition-metal single-atom electrocatalyst. Chem.-A Eur. J. 2022, 28, e202201471. [Google Scholar] [CrossRef]
  20. Liu, Y.N.; Wang, D.S.; Yang, B.; Li, Z.J.; Peng, X.Y.; Liu, Z.B.; Zeng, L.B.; Zhang, T.; Rodriguez, R.; Lei, L.; et al. Efficiently electrochemical CO2 reduction on molybdenum-nitrogen-carbon catalysts with optimized p-block axial ligands. Chem. Eng. Sci. 2023, 273, 118638. [Google Scholar] [CrossRef]
  21. Nagaprasad Reddy, S.; Krishnamurthy, C.B.; Grinberg, I. First-principles study of the ligand substituent effect on ORR catalysis by metallocorroles. J. Phys. Chem. C 2020, 124, 11275–11283. [Google Scholar] [CrossRef]
  22. Han, Y.H.; Wang, Y.G.; Xu, R.R.; Chen, W.X.; Zheng, L.R.; Han, A.J.; Zhu, Y.Q.; Zhang, J.; Zhang, H.B.; Luo, J.; et al. Electronic structure engineering to boost oxygen reduction activity by controlling the coordination of the central metal. Energy Environ. Sci. 2018, 11, 2348–2352. [Google Scholar] [CrossRef]
  23. Zhao, K.M.; Liu, S.Q.; Li, Y.Y.; Wei, X.L.; Ye, G.Y.; Zhu, W.W.; Su, Y.K.; Wang, J.; Liu, H.T.; He, Z.; et al. Insight into the mechanism of axial ligands regulating the catalytic activity of Fe–N4 sites for oxygen reduction reaction. Adv. Energy Mater. 2022, 12, 2103588. [Google Scholar] [CrossRef]
  24. Chen, X.; Zhang, Y.Z.; Zhao, X.; Yu, H.; Zhang, H. Effect of the axial halogen ligand on the oxygen reduction reaction performance of transition metal–nitrogen–carbon catalysts. J. Phys. Chem. C 2023, 127, 14107–14116. [Google Scholar] [CrossRef]
  25. Liu, Q.Q.; Bian, X.Q.; Xie, S.Y.; Ruan, W.Q.; Chen, W.K.; Guo, X.Y.; Ding, K.N. Axial halogen coordinated metal-nitrogen-carbon moiety enables efficient electrochemical oxygen reduction to hydrogen peroxide. Int. J. Hydrog. Energy 2024, 51, 1413–1420. [Google Scholar] [CrossRef]
  26. Li, L.; Li, Y.M.; Huang, R.; Cao, X.R.; Wen, Y.H. Single Mn atom anchored on nitrogen-doped graphene as a highly efficient electrocatalyst for oxygen reduction reaction. Chem. Eur. J. 2021, 27, 9686–9693. [Google Scholar] [CrossRef]
  27. Duan, Z.Y.; Wang, G.F. Comparison of reaction energetics for oxygen reduction reactions on Pt(100), Pt(111), Pt/Ni(100), and Pt/Ni(111) surfaces: A first-principles study. J. Phys. Chem. C 2013, 117, 6284–6292. [Google Scholar] [CrossRef]
  28. Tripković, V.; Skúlason, E.; Siahrostami, S.; Nørskov, J.K.; Rossmeisl, J. The oxygen reduction reaction mechanism on Pt(111) from density functional theory calculations. Electrochim. Acta 2010, 55, 7975–7981. [Google Scholar] [CrossRef]
  29. Orellana, W. Catalytic properties of transition metal–N4 moieties in graphene for the oxygen reduction reaction: Evidence of spin-dependent mechanisms. J. Phys. Chem. C 2013, 117, 9812–9818. [Google Scholar] [CrossRef]
  30. Lu, Z.S.; Xu, G.L.; He, C.Z.; Wang, T.X.; Yang, L.; Yang, Z.X.; Ma, D.W. Novel catalytic activity for oxygen reduction reaction on MnN4 embedded graphene: A dispersion-corrected density functional theory study. Carbon 2015, 84, 500–508. [Google Scholar] [CrossRef]
  31. Xu, H.X.; Cheng, D.J.; Cao, D.P.; Zeng, X.C. A universal principle for a rational design of single-atom electrocatalysts. Nat. Catal. 2018, 1, 339–348. [Google Scholar] [CrossRef]
  32. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier principle to a predictive theory of transition-metal heterogeneous catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  33. Gerber, I.C.; Serp, P. A theory/experience description of support effects in carbon-supported catalysts. Chem. Rev. 2020, 120, 1250–1349. [Google Scholar] [CrossRef] [PubMed]
  34. Guan, D.Q.; Zhong, J.; Xu, H.Y.; Huang, Y.C.; Hu, Z.W.; Chen, B.; Zhang, Y.; Ni, M.; Xu, X.M.; Zhou, W.; et al. A universal chemical-induced tensile strain tuning strategy to boost oxygen-evolving electrocatalysis on perovskite oxides. Appl. Phys. Rev. 2022, 9, 011422. [Google Scholar] [CrossRef]
  35. Li, L.; Wu, X.X.; Du, Q.Y.; Bai, N.S.; Wen, Y.H. Boosting the oxygen reduction reaction activity of dual-atom catalysts on N-doped graphene by regulating the N coordination environment. Phys. Chem. Chem. Phys. 2024, 26, 628–634. [Google Scholar] [CrossRef] [PubMed]
  36. Li, L.; Li, Y.; Huang, R.; Cao, X.R.; Wen, Y.H. Boosting the electrocatalytic activity of Fe−Co dual-atom catalysts for oxygen reduction reaction by ligand-modification engineering. ChemCatChem 2021, 13, 4645–4651. [Google Scholar] [CrossRef]
  37. Guo, X.Y.; Lin, S.R.; Gu, J.X.; Zhang, S.L.; Chen, Z.F.; Huang, S.P. Simultaneously achieving high activity and selectivity toward two-electron O2 electroreduction: The power of single-atom catalysts. ACS Catal. 2019, 9, 11042–11054. [Google Scholar] [CrossRef]
  38. Zhao, X.H.; Liu, Y.Y. Origin of selective production of hydrogen peroxide by electrochemical oxygen reduction. J. Am. Chem. Soc. 2021, 143, 9423–9428. [Google Scholar] [CrossRef] [PubMed]
  39. Guan, D.Q.; Zhou, J.; Huang, Y.C.; Dong, C.L.; Wang, J.Q.; Zhou, W.; Shao, Z.P. Screening highly active perovskites for hydrogen-evolving reaction via unifying ionic electronegativity descriptor. Nat. Commun. 2019, 10, 3755. [Google Scholar] [CrossRef]
  40. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  41. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B-Condens. Matter 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  42. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurateab initioparametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  43. Greeley, J.; Nørskov, J.K. Electrochemical dissolution of surface alloys in acids: Thermodynamic trends from first-principles calculations. Electrochim. Acta 2007, 52, 5829–5836. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of MnN4 and MnN4−X structures. Grey, blue, green, and purple balls represent C, N, Mn, and X (X = F, Cl, Br, and I) atoms, respectively.
Figure 1. Schematic diagram of MnN4 and MnN4−X structures. Grey, blue, green, and purple balls represent C, N, Mn, and X (X = F, Cl, Br, and I) atoms, respectively.
Molecules 29 03517 g001
Figure 2. The binding and formation energies for MnN4 and MnN4−X.
Figure 2. The binding and formation energies for MnN4 and MnN4−X.
Molecules 29 03517 g002
Figure 3. Schematic diagram of the most stable adsorption structures of intermediates on MnN4 and MnN4−X. Grey, blue, green, purple, red, and yellow balls represent C, N, Mn, X, O, and H atoms, respectively. “*” indicates the clean surface.
Figure 3. Schematic diagram of the most stable adsorption structures of intermediates on MnN4 and MnN4−X. Grey, blue, green, purple, red, and yellow balls represent C, N, Mn, X, O, and H atoms, respectively. “*” indicates the clean surface.
Molecules 29 03517 g003
Figure 4. Spin-resolved charge densities of Mn atom in MnN4 and MnN4−X with an isosurface value of 0.06 e/Å3. Green and blue balls represent Mn and N atom, respectively.
Figure 4. Spin-resolved charge densities of Mn atom in MnN4 and MnN4−X with an isosurface value of 0.06 e/Å3. Green and blue balls represent Mn and N atom, respectively.
Molecules 29 03517 g004
Figure 5. Free-energy diagrams and overpotentials of ORR on all catalysts. “*” indicates the clean surface.
Figure 5. Free-energy diagrams and overpotentials of ORR on all catalysts. “*” indicates the clean surface.
Molecules 29 03517 g005
Figure 6. The PDOS and the LDOS within the range of −0.6 to 0 eV for Mn in MnN4 and MnN4−X.
Figure 6. The PDOS and the LDOS within the range of −0.6 to 0 eV for Mn in MnN4 and MnN4−X.
Molecules 29 03517 g006
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

Wei, S.; Zhao, R.; Yu, W.; Li, L.; Zhang, M. Boosting the Electrocatalytic Oxygen Reduction Activity of MnN4-Doped Graphene by Axial Halogen Ligand Modification. Molecules 2024, 29, 3517. https://doi.org/10.3390/molecules29153517

AMA Style

Wei S, Zhao R, Yu W, Li L, Zhang M. Boosting the Electrocatalytic Oxygen Reduction Activity of MnN4-Doped Graphene by Axial Halogen Ligand Modification. Molecules. 2024; 29(15):3517. https://doi.org/10.3390/molecules29153517

Chicago/Turabian Style

Wei, Shaoqiang, Ran Zhao, Wenbo Yu, Lei Li, and Min Zhang. 2024. "Boosting the Electrocatalytic Oxygen Reduction Activity of MnN4-Doped Graphene by Axial Halogen Ligand Modification" Molecules 29, no. 15: 3517. https://doi.org/10.3390/molecules29153517

Article Metrics

Back to TopTop