Next Article in Journal
Density Functional Theory Simulations of Skaergaardite (CuPd) with a Self-Consistent Hubbard U-Correction
Previous Article in Journal
Study on the Effect of Different Cathodic Protection Potentials on the Growth of Mixed Bacteria and Cathodic Protection Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Carbon-Rich Selenide Monolayers as Metal-Free Catalysts for Oxygen Reduction Reactions: A First-Principles Investigation

Key Laboratory of Semiconductor Photovoltaic Technology and Energy Materials at Universities of Inner Mongolia Autonomous Region, School of Physical Science and Technology, Inner Mongolia University, Hohhot 010021, China
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(2), 55; https://doi.org/10.3390/chemistry7020055
Submission received: 24 February 2025 / Revised: 15 March 2025 / Accepted: 26 March 2025 / Published: 1 April 2025
(This article belongs to the Section Theoretical and Computational Chemistry)

Abstract

:
Carbon-based materials have garnered significant attention for electrocatalysis applications in fuel cells due to their unique structural and electronic properties, but rapid oxygen reduction reaction (ORR) at the cathode of fuel cells is challenging. Dopants are typically used as active sites for ORR, and increasing the number of active sites for carbon-based catalysts remains a challenge. Here, we carried out first-principles calculations for the electrocatalytic ORR performance of the recently reported monolayer superconductors of carbon-rich selenides. Remarkably, the abundant C atoms serve as the active centers instead of the foreign atoms (Se). All the free energy changes during the ORR process are downhill, suggesting that these carbon-rich selenides hold promise as metal-free electrocatalysts for ORR. Note that the promising electrocatalytic performance of carbon-rich selenides is theoretically predicted; validation is encouraged for experimental efforts.

1. Introduction

Energy plays the most important role in the development of society and industry. As the storage of fossil fuels is limited, the search for renewable and sustainable clean energy is increasingly urgent. Proton exchange membrane fuel cells (PEMFCs) converting the chemical energy from hydrogen into electrical energy through an electrochemical reaction provide a very attractive alternative [1]. However, the reaction in which oxygen reacts with protons and electrons to produce water (oxygen reduction reaction, ORR) at the cathode is kinetically sluggish and has a significant energy barrier, requiring efficient catalysts to facilitate the reaction [2]. Pt-based catalysts have long been the leading ORR catalysts, but the high cost and scarcity of Pt have severely hindered their large-scale commercial application [3]. Therefore, it is of paramount importance to develop low-cost, highly durable, and highly active ORR electrocatalysts for the large-scale application of PEMFCs, and seeking efficient and inexpensive metal-free ORR catalysts has become a hot topic in recent years [4].
Carbon-based materials are a new class of metal-free ORR catalysts that are expected to replace platinum for efficiently catalyzing the ORR in fuel cells due to their large surface area, good electrical conductivity, tunable morphology, simple and economically feasible preparation, etc. [5]. Among them, N-doped graphene and graphitic flakes have been extensively synthesized, and most reported samples can efficiently electrocatalyze the ORR via a four-electron pathway in fuel cells, showing high catalytic activity, stability, and tolerance to methanol crossover effects [6,7,8,9,10,11]. Moreover, various carbon materials doped with heteroatoms, such as B-doped carbon nanotubes (CNTs) or grapheme [12,13], sulfur-doped graphene [14], phosphorus-doped graphite layers [15], iodine-doped graphene [16], and graphene nanosheets with edge halogenation (Cl, Br, or I), have also been investigated [17]. In addition, the binary or ternary doping of carbon materials with different heteroatoms was found to be more electrocatalytically active toward ORR than the single-doped counterparts [5]. Typically, dopants serve as the primary active sites but are constrained by low doping concentrations in the carbon matrix. This limitation hinders further improvements in catalytic efficiency. Therefore, developing strategies to fully utilize the abundant intrinsic carbon atoms as active sites is crucial for enhancing catalytic performance. Given that the synthesized selenium-doped carbon materials are utilized in a range of sustainable technologies [18], we would like to know whether the large number of intrinsic C atoms in carbon-rich selenide monolayers can serve as the active sites for ORR. This consideration is based on the large surface-to-volume ratio in two-dimensional (2D) materials facilitating tremendous contact with reactants and intermediates.
In this work, we examined the ORR performance of three 2D carbon-rich selenide monolayers, namely C4Se, C5Se, and C6Se, which were recently reported to be the global minima and to exhibit superconductivity [19], by means of first-principles calculations. Our calculations revealed that carbon atoms (C) are the active centers that boost ORR, largely because of the electron transfer from Se to C. Moreover, both C4Se and C5Se superconductors are promising candidates for metal-free electrocatalysts for ORR, as all elementary reaction steps are downhill in free energy. Though the carbon-rich selenide monolayers are not experimentally available yet, these lowest-energy and highly stable structures suggest they are highly promising for experimental synthesis. There are many predictions realized by experimentalists. For example, our predicted Cu2N global minimum in 2D space [20] was successfully synthesized by Wu’s group [21], and the promising ORR performance of the iron phthalocyanine moiety [22] was experimentally validated [23]. Our theoretical work provides guidelines for developing 2D metal-free electrocatalysts with carbon atoms serving as active sites.

2. Computational Methodology

All the spin-polarized density functional theory (DFT) calculations were carried out in the Vienna Ab initio Simulation Package (VASP) [24]. Electron exchange and correlation interactions were treated by the generalized gradient approximation (GGA) method with the Perdew–Burke–Ernzerhof (PBE) functional, whereas the interactions between ion nuclei and valence electrons were described using the projector-augmented wave (PAW) method [25]. D2 dispersion corrections were employed for calculations [26]. The wavefunction for valence electrons was expanded by the Kohn–Sham plane-wave basis set with an energy cutoff of 500 eV. All the structures were fully relaxed until the convergence tolerances for the energy and force on each atom were less than 1.0 × 10−4 eV and −0.01 eV/Å, respectively. The three monolayers, C4Se, C5Se, and C6Se, were sampled in the Brillouin zone at k points using a scheme of a 3 × 3 × 1 k-point Monkhorst-Pack grid [27]. To avoid image interaction, 2 × 5 × 1, 3 × 3 × 1, and 2 × 4 × 1 supercells were used in the calculations for C4Se, C5Se, and C6Se monolayers, respectively, and a vacuum layer of ~20 Å was applied in the z-direction. The free energy was calculated according to the computational hydrogen electrode (CHE) proposed by Nørskov et al. [28]. The zero-point energy (ZPE) and entropic corrections were obtained from DFT calculations. Bader charge analysis was used to assess charge transfer [29]. The crystal orbit Hamilton population (COHP) analysis was performed by using LOBSTER-5.0.0 software [30,31].
The overall reaction of four-electron ORR is
O 2 g + 4 H + + 4 e 2 H 2 O ( l )
The four elementary reaction steps during ORR in acidic media are as follows:
O 2 g + H + + e + *   OOH *
OOH * + H + + e O * + H 2 O ( l )
O   * + H + + e O H *
OH * + H + + e H 2 O l + *
where “ * ” denotes the adsorption site. The adsorbed energy is calculated by the following equations [32]:
Δ E ( * O ) = E ( * O ) E ( * ) ( E H 2 O   E H 2 )
Δ E ( * OH ) = E ( * OH ) E ( * ) ( E H 2 O 1 2 E H 2 )
Δ E ( * OOH ) = E ( * OOH ) E ( * ) ( 2 E H 2 O 3 2 E H 2 )
where E( * ) is the ground state energy of the catalyst. Δ E ( * O   ) , Δ E ( * OH   )   , and Δ E ( * OOH   ) represent the energies of the intermediates O, OH, and OOH adsorbed onto the catalysts, respectively. E H 2 O and E H 2 are the calculated energies of the gas-phase water and hydrogen molecules, respectively, in a 15 Å × 15 Å × 15 Å cube.
The free energy of each reaction step is determined by the following equation:
Δ G = Δ E DFT + Δ ZPE T Δ S + Δ G U 0 Δ G pH
where Δ E D F T is the reaction energy directly from the DFT calculation. Δ Z P E and T Δ S are the zero-point energy and entropy contributions, respectively [33]. The term related to pH is Δ G p H , determined by Δ G p H = k B   T × ln10 × pH, where “kB” is Boltzmann’s constant. In this paper, pH is set to 0; therefore, Δ G p H = 0 eV. U0 is the potential relative to the standard hydrogen electrode (SHE) under standard conditions (U0 = 0 V, pH = 0, P = 1 bar, T = 298.15 K). In this paper, the free energy of a free O2 is determined by the following equation [34]:
G O 2 = 2 G H 2 O 2 G H 2 + 4.92   e V
where G H 2 O and G H 2 are the free energies of H2 and H2O, respectively. The Δ G 1 ~ Δ G 4 of each ORR step can be obtained by the following expressions.
Δ G 1 = Δ G ( * OOH ) 4.92   eV
Δ G 2 = Δ G ( * O )     Δ G ( * OOH )
Δ G 3 = Δ G ( * OH )     Δ G ( * O )
Δ G 4 = Δ G ( * OH )
where 4.92 eV is the total reaction energy. The limiting potential ( U L ) is used to evaluate the catalytic performance [35], representing the maximum free energy change in the four steps of (11)–(14) and having the expression:
U L = max   ( Δ G 1 ,   Δ G 2 ,   Δ G 3 ,   Δ G 4 ) e
The overpotential ( η ORR ) for the whole process of the oxygen reduction reaction can be obtained via the following formula:
η ORR = 1.23 U L

3. Results and Discussion

The optimized structures of the three carbon-rich selenides are shown in Figure 1. The lattice parameters are a = 6.16 Å and b = 2.55 Å in C4Se, a = b = 4.55 Å in C5Se, and a = 9.69 Å and b = 4.54 Å in C6Se. The corresponding space group symmetries are Pmm2, P31m, and P21212, respectively, where the lattice constant c was fixed at 20 Å. The C−Se bond lengths of the three monolayers were measured to be 1.94, 1.91, and 1.98~2.00 Å for C4Se, C5Se, and C6Se, respectively. These structural parameters are in good agreement with the results reported in the literature [19].

3.1. Stability

The three C-Se monolayers (C4Se, C5Se, and C6Se) represent the global minimum of the same stoichiometry in the two-dimensional space. Their stabilities were confirmed through multiple criteria, including formation energy, ab initio molecular dynamics simulations, phonon dispersion curves, and elastic constants [19], indicating a high likelihood of experimental realization. In addition, we calculated the integrated crystal orbital Hamilton population (ICOHP) of Se−C bonding (Se−CI, Se−CII, and Se−CIII in C4Se, C5Se, and C6Se, respectively, as exemplified and illustrated in Figure 1). The ICOHP values for the three monolayers are −6.13, −6.35, and −5.64 eV, respectively (Figure 2). The negative ICOHP values further indicate the excellent structural stability of these C-Se monolayers.

3.2. O2 Adsorption

To investigate the ORR performance, we first examined the adsorption behavior of O2. Both Se and C sites were considered, and O2 prefers to adsorb at C sites on all three carbon-rich selenides. The O−O bond length of the adsorbed O2 is elongated to 1.38 Å on C4Se, compared to the calculated value of 1.25 Å for free O2. Interestingly, O2 dissociates spontaneously on C5Se and C6Se, and with O−O bond distances of 2.57 and 2.20 Å, respectively. Among them, the adsorption energies of oxygen on the three systems are −0.80, −3.08, and −3.60 eV, respectively. We found that oxygen molecules and subsequent intermediates are adsorbed on the C atoms that surround the Se atoms, with the C atoms serving as active sites, which significantly increases the number of available active sites.

3.3. Gibbs Free Energy

We calculated the Gibbs free energy change (ΔG) of each elemental step during the four-electron ORR process. The free energy profile and the side view of each intermediate are shown in Figure 3. The rate-determining step (RDS) is the formation of the second water molecule (*OH → *H2O) on C4Se, the formation of the intermediate (*O → *OH) on C5Se, and the formation of the first water molecule (*O + *OH → *O + H2O) on C6Se, with values for the vales of limiting potential U L of 0.09, 0.08, and −0.34 V and values for the overpotential η ORR of 1.14, 1.15, and 1.57 V for C4Se, C5Se, and C6Se, respectively. As a metal-free catalyst, the η ORR values of C4Se and C5Se are both lower than that of the 2D carbon materials of the Haeckelite and N-doped Haeckelite structures (1.26 and 1.20 eV) [36]. Notably, the free energies are downhilled during the four-electron process on both C4Se and C5Se, suggesting their high ORR activity.

3.4. Origin of ORR Performance

Furthermore, we compared the charge transfer between C and Se atoms to gain insight into the origin of the ORR activity of abundant C sites. Bader charge analysis revealed that Se atoms transfer electrons to C (Figure 1), in line with the higher electronegativity of C compared to Se, and the extra electrons on C activate O2, facilitating ORR. Note that C atoms have more extra electrons on average in both C4Se and C5Se than those of C6Se, since each Se has fewer coordinates with C (two and three) in the former two monolayers than that (four) in the latter one. In addition, the projected density of states (PDOS) was analyzed to further investigate the electronic structure. As shown in Figure 4, the electronic density of states diagrams reveal that C4Se, C5Se, and C6Se exhibit metallic properties, indicating that all of these materials are suitable for use as electrocatalysts to promote the oxygen reduction reaction. From the analysis of PDOS, the electronic states are primarily contributed by the p orbitals of carbon and selenium, and the p-band center of the C atoms in the monolayers of C4Se, C5Se, and C6Se is closer to the Fermi energy level compared to that of the Se atoms (0.57 vs. 0.62, 0.11 vs. 1.34, and −0.20 vs. −0.65 eV, respectively). It is further confirmed that the adsorption capacity of carbon atoms toward O2 and intermediates is stronger than that of Se atoms, leading to C atoms being the active sites and to the increase in the number of active sites.

4. Conclusions

In summary, we performed DFT calculations to investigate the electrocatalytic performance of superconducting C4Se, C5Se, and C6Se monolayers for an oxygen reduction reaction (ORR). O2 was effectively activated on all three monolayers, as evidenced by the elongated or dissociated O−O bonds. It is noteworthy that the free energy changes for the four-electron process on both C4Se and C5Se are downhill, indicating excellent catalytic activity. Specifically, carbon atoms serve as active sites, benefiting from the electron transfer from Se to the adjacent C atoms, which effectively addresses the issue of insufficient active sites and enhances the overall reaction efficiency. Though our theoretical study provides valuable insights for the development of metal-free electrocatalysts with intrinsic carbon atoms as active centers for oxygen reduction reactions, our predictions require experimental validation. Given the superconductivity and global stability of these carbon-rich selenide monolayers, our findings are of broad interest to both theoretical and experimental research communities.

Author Contributions

Y.X.: methodology, data and formal analysis, investigation, and writing—original draft; F.L.: conceptualization, project administration, supervision, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (12364038), the Industrial Technology Innovation Projects of the Inner Mongolia Academy of Science and Technology of China (2023JSYD01002), and the Science and Technology Plan Projects of the Inner Mongolia Autonomous Region of China (2023KYPT0012).

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hu, C.; Paul, R.; Dai, Q.; Dai, L. Carbon-based Metal-free Electrocatalysts: From Oxygen Reduction to Multifunctional Electrocatalysis. Chem. Soc. Rev. 2021, 50, 11785–11843. [Google Scholar] [CrossRef] [PubMed]
  2. Erable, B.; Féron, D.; Bergel, A. Microbial Catalysis of the Oxygen Reduction Reaction for Microbial Fuel Cells: A Review. ChemSusChem 2012, 5, 975–987. [Google Scholar] [CrossRef] [PubMed]
  3. Santos, E.; Lundin, A.; Pötting, K.; Quaino, P.; Schmickler, W. Hydrogen Evolution and Oxidation—A Prototype for an ElectroCatalytic Reaction. J. Solid State Electrochem. 2009, 13, 1101–1109. [Google Scholar] [CrossRef]
  4. Yuan, Y.; Wu, S.; Ai, H.; Lee, Y.J.; Kang, B. γ-Graphyne Nanotubes as Defect-free Catalysts of the Oxygen Reduction Reaction: A DFT Investigation. Phys. Chem. Chem. Phys. 2020, 28, 633–8638. [Google Scholar] [CrossRef]
  5. Dai, L.; Xue, Y.; Qu, L.; Choi, H.J.; Baek, J.B. Metal-free Catalysts for Oxygen Reduction Reaction. Chem. Rev. 2015, 115, 4823–4892. [Google Scholar] [CrossRef]
  6. Zhang, J.; Xia, Z.; Dai, L. Carbon-based Electrocatalysts for Advanced Energy Conversion and Storage. Sci. Adv. 2015, 1, e1500564. [Google Scholar] [CrossRef]
  7. Singh, R.K.; Douglin, J.C.; Kumar, V.; Tereshchuk, P.; Santori, P.G.; Ferreira, E.B.; Jerkiewicz, G.; Ferreira, P.J.; Natan, A.; Jaouen, F.; et al. Metal-free advanced energy materials for the oxygen reduction reaction in anion-exchange membrane fuel cells. Appl. Catal. B Environ. 2024, 357, 124319. [Google Scholar] [CrossRef]
  8. Yu, D.; Zhang, Q.; Dai, L. Highly Efficient Metal-free Growth of Nitrogen-doped Single-walled Carbon Nanotubes on Plasma-etched Substrates for Oxygen Reduction. J. Am. Chem. Soc. 2010, 132, 15127–15129. [Google Scholar] [CrossRef]
  9. Ma, R.; Ren, X.; Bao, Y.X.; Zhou, Y.; Sun, C.; Liu, Q.; Liu, J.; Wang, J. Novel Synthesis of N-doped Graphene as an Efficient Electrocatalyst Towards Oxygen Reduction. Nano Res. 2016, 9, 808–819. [Google Scholar] [CrossRef]
  10. Gu, D.; Zhou, Y.; Ma, R.; Wang, F.; Liu, Q.; Wang, J. Facile Synthesis of N-doped Graphene-like Carbon Nanoflakes as Efficient and Stable Electrocatalysts for the Oxygen Reduction Reaction. Nano-Micro Lett. 2017, 10, 29. [Google Scholar] [CrossRef]
  11. Ly, A.; Murphy, E.; Wang, H.; Huang, Y.; Ferro, G.; Guo, S.; Asset, T.; Liu, Y.; Zenyuk, I.V.; Atanassov, P. Electrochemical Trends of a Hybrid Platinum and Metal-nitrogen-carbon Catalyst Library for the Oxygen Reduction Reaction. EES. Catal. 2024, 2, 624–637. [Google Scholar] [CrossRef]
  12. 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. Int. Ed. 2011, 123, 7270–7273. [Google Scholar] [CrossRef]
  13. Sheng, Z.H.; Gao, H.L.; Bao, W.J.; Wang, F.B.; Xia, X.H. Synthesis of Boron Doped Graphene for Oxygen Reduction Reaction in Fuel Cells. J. Mater. Chem. 2012, 22, 390–395. [Google Scholar] [CrossRef]
  14. Yang, Z.; Yao, Z.; Li, G.; Fang, G.; Nie, H.; Liu, Z.; Zhou, X.; Chen, X.; Huang, S. Sulfur-doped Graphene as an Efficient Metal-free Cathode catalyst for oxygen reduction. ACS Nano 2011, 6, 205–211. [Google Scholar] [CrossRef]
  15. Liu, Z.W.; Peng, F.; Wang, H.J.; Yu, H.; Zheng, W.X.; Yang, J. Phosphorus-doped Graphite Layers with High Electrocatalytic Activity for the O2 Reduction in an Alkaline Medium. Angew. Chem. Int. Ed. 2011, 50, 3257–3261. [Google Scholar] [CrossRef]
  16. Yao, Z.; Nie, H.; Yang, Z.; Zhou, X.; Liu, Z.; Huang, S. Catalyst-free Synthesis of Iodine-doped Graphene via a Facile Thermal Annealing Process and its use for Electrocatalytic Oxygen Reduction in an Alkaline Medium. Chem. Commun. 2012, 48, 1027–1029. [Google Scholar] [CrossRef]
  17. Jeon, I.Y.; Choi, H.J.; Choi, M.; Seo, J.M.; Jung, S.M.; Kim, M.J.; Zhang, S.; Zhang, L.; Xia, Z.; Dai, L.; et al. Facile, Scalable Synthesis of Edge-halogenated Graphene Nanoplatelets as Efficient Metal-free Eletrocatalysts for Oxygen Reduction Reaction. Sci. Rep. 2013, 3, 1810. [Google Scholar] [CrossRef]
  18. Dyjak, S.; Jankiewicz, J.B.; Kaniecki, S.; Kiciński, W. Selenium-doped Carbon Materials: Synthesis and Applications for Sustainable Technologies. Green Chem. 2024, 26, 2985–3020. [Google Scholar] [CrossRef]
  19. Kong, P.; Zhang, X.; Wang, J.; Tian, W.; Ni, Y.; Sun, B.; Wang, H.; Wang, H.; Feng, Y.P.; Chen, Y. Electron-phonon Coupling Superconductivity and Tunable Topological State in Carbon-rich Selenide Monolayers. Phys. Rev. B 2023, 107, 184115. [Google Scholar] [CrossRef]
  20. Jia, J.; Wang, Z.; Liu, Y.; Li, F.; Shang, Y.; Liu, Y.; Cai, Q.; Zhao, J. A Metallic Cu2N Monolayer with Planar Tetracoordinated Nitrogen as a Promising Catalyst for CO2 Electroreduction. J. Mater. Chem. A 2022, 10, 1560–1568. [Google Scholar] [CrossRef]
  21. Hu, X.; Zhang, R.W.; Ma, D.S.; Cai, Z.; Geng, D.; Sun, Z.; Zhao, Q.; Gao, J.; Cheng, P.; Chen, L.; et al. Realization of a Two-dimensional Checkerboard Lattice in Monolayer Cu2N. Nano Lett. 2023, 23, 5610–5616. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Yuan, H.; Li, Y.; Chen, Z. Two-dimensional Iron-phthalocyanine (Fe-Pc) Monolayer as a Promising Single-atom-catalyst for Oxygen Reduction Reaction: A Computational Study. Nanoscale 2015, 7, 11633–11641. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, X.; Wang, B.; Zhong, J.; Zhao, F.; Han, N.; Huang, W.; Zeng, M.; Fan, J.; Li, Y. Iron Polyphthalocyanine Sheathed Multiwalled Carbon Nanotubes: A High-performance Electrocatalyst for Oxygen Reduction Reaction. Nano Res. 2016, 9, 1497–1506. [Google Scholar] [CrossRef]
  24. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Iinitio Total-energy Calculations Using a Plane-wave Basis Set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  25. Blöchl, P.E. Projector Augmented-wave Method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef]
  26. Chowdhury, C.; Datta, A. Silicon-doped Nitrogen-coordinated Graphene as Electrocatalyst for Oxygen Reduction Reaction. J. Phys. Chem. C 2018, 122, 27233–27240. [Google Scholar] [CrossRef]
  27. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  28. 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. [Google Scholar] [CrossRef]
  29. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  30. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal Orbital Hamilton Population (COHP) Analysis as Projected from Plane-wave Basis Sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef]
  31. Maintz, S.; Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. LOBSTER: A Tool to Extract Chemical Bonding from Plane-wave Based DFT. J. Comput. Phys. 2016, 37, 1030–1035. [Google Scholar] [CrossRef]
  32. Man, I.C.; Su, H.Y.; Calle-Vallejo, F.; Hansen, H.A.; Martínez, J.I.; Inoglu, N.G.; Kitchin, J.; Jaramillo, T.F.; Nørskov, J.K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces. ChemCatChem 2011, 3, 1159–1165. [Google Scholar] [CrossRef]
  33. Gao, M.R.; Cao, X.; Gao, Q.; Xu, Y.F.; Zheng, Y.R.; Jiang, J.; Yu, S.H. Nitrogen-doped Graphene Supported CoSe2 Nanobelt Composite Catalyst for Efficient Water Oxidation. ACS Nano 2014, 8, 3970–3978. [Google Scholar] [CrossRef] [PubMed]
  34. Sun, J.; Guo, N.; Shao, Z.; Huang, K.; Li, Y.; He, F.; Wang, Q. A Facile Strategy to Construct Amorphous Spinel-based Electrocatalysts with Massive Oxygen Vacancies Using Ionic Liquid Dopant. Adv. Energy Mater. 2018, 8, 1800980. [Google Scholar] [CrossRef]
  35. Meng, Y.; Yin, C.; Li, K.; Tang, H.; Wang, Y.; Wu, Z. Improved Oxygen Reduction Activity in Heteronuclear FeCo-codoped Graphene: A Theoretical Study. ACS Sustain. Chem. Eng. 2019, 7, 17273–17281. [Google Scholar] [CrossRef]
  36. Wang, Y.; Sun, X.; He, F.; Li, K.; Wu, Z. Haeckelite and N-doped Haeckelite as Catalysts for Oxygen Reduction Reaction: Theoretical Studies. J. Phys. Chem. C 2017, 121, 28339–28347. [Google Scholar] [CrossRef]
Figure 1. The top and side views of (a) C4Se, (b) C5Se, and (c) C6Se monolayers. The unit cell is given by a red tetragonum with two lattice vectors a and b. The orange and gray balls represent Se and C atoms, respectively. The data represent the number of valence electrons on the atom.
Figure 1. The top and side views of (a) C4Se, (b) C5Se, and (c) C6Se monolayers. The unit cell is given by a red tetragonum with two lattice vectors a and b. The orange and gray balls represent Se and C atoms, respectively. The data represent the number of valence electrons on the atom.
Chemistry 07 00055 g001
Figure 2. The projected COHP for (a) C4Se, (b) C5Se, and (c) C6Se. The bonding and anti-bonding states are shown on the right and the left of the vertical zero line, respectively. The Fermi level is set at 0 eV and indicated by the horizontal line.
Figure 2. The projected COHP for (a) C4Se, (b) C5Se, and (c) C6Se. The bonding and anti-bonding states are shown on the right and the left of the vertical zero line, respectively. The Fermi level is set at 0 eV and indicated by the horizontal line.
Chemistry 07 00055 g002
Figure 3. A free energy diagram of ORR on models of (a) C4Se, (b) C5Se, and (c) C6Se. The data denote the free energy of each state (in eV). The black and red lines represent the potentials of 0 and 1.23 V, respectively. “*” denotes the adsorption site.
Figure 3. A free energy diagram of ORR on models of (a) C4Se, (b) C5Se, and (c) C6Se. The data denote the free energy of each state (in eV). The black and red lines represent the potentials of 0 and 1.23 V, respectively. “*” denotes the adsorption site.
Chemistry 07 00055 g003
Figure 4. The PDOS of monolayer (a) C4Se, (b) C5Se, and (c) C6Se. The black, red, blue, green, and purple curves denote the total, Se-s, Se-p, C-s, and C-p orbitals, respectively. The p-band centers of C and Se and the Fermi level (set at 0 eV) are marked by pink, brown, and light-blue dashed lines, respectively.
Figure 4. The PDOS of monolayer (a) C4Se, (b) C5Se, and (c) C6Se. The black, red, blue, green, and purple curves denote the total, Se-s, Se-p, C-s, and C-p orbitals, respectively. The p-band centers of C and Se and the Fermi level (set at 0 eV) are marked by pink, brown, and light-blue dashed lines, respectively.
Chemistry 07 00055 g004
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

Xu, Y.; Li, F. Carbon-Rich Selenide Monolayers as Metal-Free Catalysts for Oxygen Reduction Reactions: A First-Principles Investigation. Chemistry 2025, 7, 55. https://doi.org/10.3390/chemistry7020055

AMA Style

Xu Y, Li F. Carbon-Rich Selenide Monolayers as Metal-Free Catalysts for Oxygen Reduction Reactions: A First-Principles Investigation. Chemistry. 2025; 7(2):55. https://doi.org/10.3390/chemistry7020055

Chicago/Turabian Style

Xu, Yao, and Fengyu Li. 2025. "Carbon-Rich Selenide Monolayers as Metal-Free Catalysts for Oxygen Reduction Reactions: A First-Principles Investigation" Chemistry 7, no. 2: 55. https://doi.org/10.3390/chemistry7020055

APA Style

Xu, Y., & Li, F. (2025). Carbon-Rich Selenide Monolayers as Metal-Free Catalysts for Oxygen Reduction Reactions: A First-Principles Investigation. Chemistry, 7(2), 55. https://doi.org/10.3390/chemistry7020055

Article Metrics

Back to TopTop