Oxygen coordination engineering in Fe–N–O–graphene single-atom catalysts for enhanced bifunctional oxygen electrocatalysis

Linlin Zhang *ac, Yanning Wang a and Kai Xiong b
aDepartment of Physics, Mathematics and Computer Science, Kunming Medical University, Kunming, China. E-mail: ynuzll@163.com
bMaterials Genome Institute, School of Materials and Energy, Yunnan University, Kunming 650091, China
cCenter for Regenerative Medicine Research, School of Basic Medical Sciences, Kunming Medical University, Kunming, China

Received 8th July 2025 , Accepted 1st October 2025

First published on 1st October 2025


Abstract

Developing efficient, durable, and low-cost bifunctional electrocatalysts is essential for renewable energy conversion. In this study, density functional theory (DFT) calculations are employed to systematically investigate Fe–N–O coordination structures anchored on graphene (Fe–N–O–gra), aiming to elucidate the structure–electronic–activity relationship in Fe-based single-atom catalysts. Among seven representative configurations, FeN3O exhibits the highest structural stability, optimal orbital hybridization, and the largest charge transfer, thereby effectively facilitating O2 activation. FeN3O exhibits a longer Fe–O bond and weaker electronic coupling during *OH adsorption, which moderates intermediate binding in a step-specific way—facilitating *OH desorption in the ORR and stabilizing *OOH formation in the OER. This balance leads to exceptionally low overpotentials of 0.41 V for the ORR and 0.55 V for the OER, outperforming other configurations. Furthermore, a volcano plot constructed based on adsorption energy scaling relations identifies the *OH adsorption free energy as an effective descriptor of bifunctional catalytic activity, with FeN3O located near the apex of the plot. These findings highlight the critical role of oxygen coordination in tuning the electronic structure and catalytic performance of Fe single-atom catalysts and provide theoretical guidance for the rational design of next-generation non-precious metal electrocatalysts.


1. Introduction

With the growing demand for sustainable energy systems, the development of efficient, stable, and cost-effective electrocatalysts has emerged as a central scientific challenge in energy conversion technologies.1 This is particularly critical for metal–air batteries (MABs) and proton exchange membrane fuel cells (PEMFCs), where the sluggish kinetics of the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) significantly hinder overall energy conversion efficiency.2–6 Although noble metal catalysts such as Pt and RuO2 exhibit excellent catalytic activity, their high cost, limited availability, and inability to simultaneously deliver high ORR and OER performance severely constrain their large-scale deployment in clean energy technologies.7–9 Consequently, the exploration of non-precious metal alternatives with high catalytic activity and long-term stability has become a focal point of current research.10–15

Among various non-precious metal catalysts, single-atom catalysts (SACs) have garnered widespread attention due to their maximum atom utilization, well-defined active sites, and tunable electronic structures.16–18 In particular, transition metal–nitrogen–carbon (M–N–C) moieties have demonstrated remarkable performance in diverse electrocatalytic reactions.12,17,19–21 Iron (Fe), owing to its earth abundance, low cost, and environmental benignity, is widely regarded as an ideal candidate for constructing non-precious metal electrocatalysts.22 Moreover, Fe-based M–N–C structures, especially Fe–N4 moieties, have shown ORR activity comparable to or even exceeding that of commercial Pt catalysts in alkaline and neutral media.23–26 The 3d electronic configuration of Fe enables effective orbital hybridization with oxygenated intermediates (e.g., O2, *OOH, *O, *OH), facilitating electron transfer during catalysis.27,28 Its electronic states are also highly susceptible to modulation by coordination environments, rendering Fe an ideal platform for exploring the structure–electronic–activity relationship in electrocatalysis.

Recent studies have revealed that tailoring the coordination environment of Fe atoms—especially by incorporating heteroatoms such as O, B, P, and S—can effectively modulate the electronic structure of the Fe center without compromising thermal stability.29–34 Among these, oxygen atoms, with their strong electronegativity and electron-withdrawing capability, offer significant advantages in inducing charge redistribution and enhancing the binding affinity of oxygenated intermediates. Notably, Fe–N–O SACs with N/O co-coordination have been experimentally synthesized and have exhibited promising bifunctional catalytic activity in both the ORR and OER, thus providing a solid foundation for theoretical investigations.35 However, despite preliminary progress, the fundamental understanding of the structure–activity relationship in Fe–N–O SACs remains limited. In particular, key aspects such as electronic structure modulation, intermediate adsorption behavior, and reaction pathways under different Fe–N–O coordination geometries lack systematic theoretical elucidation. The thermodynamic and electrochemical stability, mechanistic differences in reaction pathways, and variations in rate-determining steps (RDSs) across different coordination configurations remain largely unexplored.

In this context, we perform a comprehensive density functional theory (DFT) study to evaluate the bifunctional ORR/OER catalytic behavior of various Fe–N–O–gra configurations. By constructing representative Fe–N–O–gra coordination models, we first evaluate structural stability using ab initio molecular dynamics and then analyze electronic properties via PDOS, Bader charge, and charge-density difference to elucidate how coordination modulates the Fe center. Furthermore, adsorption energies of key intermediates and the corresponding reaction free energy profiles are analyzed to identify the reaction pathways, rate-determining steps, and theoretical overpotentials of each configuration, thereby allowing for a systematic evaluation of their bifunctional electrocatalytic performance. A volcano plot based on scaling relations is subsequently constructed to uncover the descriptor–activity correlation, revealing the coupling mechanism between coordination geometry and catalytic activity. This study not only deepens our understanding of the catalytic mechanisms in N/O co-coordinated Fe–N–O SACs but also provides valuable theoretical guidance for the rational design and structural optimization of next-generation non-precious metal bifunctional electrocatalysts.

2. Computational details

2.1 Computational methods

Density functional theory (DFT) calculations were performed using the Vienna Ab initio Simulation Package (VASP 6.1.0),36 with the projector augmented wave (PAW) method employed to describe the electron–ion interactions.37 The exchange–correlation potential was treated using the Perdew–Burke–Ernzerhof (PBE) functional within the generalized gradient approximation (GGA).38 A plane-wave energy cutoff of 500 eV was applied throughout all calculations. To more accurately describe the localized behavior of the 3d electrons in the Fe–N–O single-atom catalyst, a Hubbard U correction with Ueff = 3.29 eV was applied.39 Structural optimizations were carried out until the energy and force convergence thresholds were less than 10−5 eV and 0.02 eV Å−1, respectively.40 A Monkhorst–Pack 3 × 3 × 1 k-point mesh was used for geometry optimization, while a denser 11 × 11 × 1 k-point grid was applied for electronic structure calculations. van der Waals interactions were accounted for using the Grimme DFT-D3 correction scheme.41 Solvation effects were considered implicitly by applying the VASPsol continuum solvation model during all calculations, including geometry optimization, with a dielectric constant of 78.4 to simulate aqueous conditions.42,43

2.2 Structural models

A periodic 4 × 4 × 1 monolayer graphene supercell was constructed, with a vacuum space of 15 Å introduced along the z-direction to eliminate interactions between adjacent periodic images. Defective sites were created by removing two adjacent carbon atoms to form double vacancies, followed by substitution of neighboring carbon atoms with nitrogen or oxygen atoms. An Fe atom was subsequently embedded at the center of the vacancy, yielding seven distinct Fe–N–O coordination structures as illustrated in Fig. 1(a)–(g). Among them, the FeN2O2 configuration includes three geometric variants: (1) FeN2O2-p, in which the Fe atom and surrounding N/O atoms form a five-membered ring; (2) FeN2O2-h, a six-membered ring arrangement involving Fe, N, and O atoms; (3) FeN2O2-o, where the two nitrogen atoms occupy opposite positions relative to the Fe atom.
image file: d5re00292c-f1.tif
Fig. 1 (a) Optimized geometries of seven Fe–N–O–gra configurations, including FeN4, FeN3O, FeN2O2-pen (five-membered ring), FeN2O2-hex (six-membered ring), FeN2O2-oppo (opposite N atoms), FeNO3, and FeO4. Brown, red, light blue, and gold spheres represent C, O, N, and Fe atoms, respectively. (b–g) Time-dependent total energy profiles from ab initio molecular dynamics (AIMD) simulations at 500 K for FeN4, FeN3O, FeN2O2-pen, FeN2O2-hex, FeN2O2-oppo, and FeNO3 configurations. Insets show the top and side views of the structures at the 10 ps snapshot for each configuration.

2.3 Computational contents

(1) Adsorption energy (ΔEads). The adsorption energy of a reaction intermediate, denoted as ΔEads, was determined using the following expression:44
 
ΔEads = E*mEFe–N–O–GraEm(1)
where E*m is the total energy of the Fe–N–O–graphene system with the intermediate m adsorbed, Em is the energy of the isolated intermediate, and EFe–N–O–Gra is the energy of the pristine catalyst model. A negative ΔEads implies that adsorption is energetically favorable, indicating the catalyst's potential to effectively bind and activate the intermediate species.
(2) Gibbs free energy change(ΔG). The change in Gibbs free energy for each elementary reaction step was evaluated as:45,46
 
ΔG = ΔE + ΔZPE − TΔS + ΔGU + ΔGPH(2)
Here, ΔE represents the reaction energy obtained from DFT calculations, ΔZPE is the correction from zero-point vibrational energy, and ΔS is the entropy contribution at 298.15 K. The electrochemical term ΔGU is given by −neU, where n is the number of electrons transferred and U is the applied potential. The pH-dependent correction, ΔGpH, is calculated as κBT[thin space (1/6-em)]ln[thin space (1/6-em)]10 × pH,47 where κB is the Boltzmann constant; in this work, the pH value was defined to be 0 for acidic medium.

Under acidic media, the ORR proceeds via a four-electron mechanism as follows:47,48

 
O2 + * + H+ + e → *OOH(3)
 
*OOH + H+ + e → *O + H2O(4)
 
*O + H+ + e → *OH(5)
 
*OH + H+ + e → * + H2O(6)
Here, * denotes an active site on the catalyst surface. The overpotential η for both the ORR and OER is derived from the step with the highest Gibbs free energy change, according to:49
 
ηORR(+)/OER(−) = max(ΔG1, ΔG2, ΔG3, ΔG4)/e ± 1.23 V(7)

(3) Adsorption free energy (ΔGads). The Gibbs free energies of adsorption for the key oxygenated intermediates (*OOH, *O, and *OH) were computed using the following relationships:50
 
ΔG*OOH = G*OOHGFe–N–O–gra − (2GH2O − 3/2GH2)(8)
 
ΔG*O = G*OGFe–N–O–gra − (GH2OGH2)(9)
 
ΔG*OH = G*OHGFe–N–O–gra − (GH2O − 1/2GH2)(10)
In these expressions, G*OOH, G*O, and G*OH refer to the Gibbs free energies of the systems with the respective intermediates adsorbed. GH2 and GH2O correspond to the free energies of isolated hydrogen and water molecules, respectively. GFe–N–O–gra is the free energy of the Fe–N–O–graphene substrate with varying coordination environments.

3. Results and discussion

To systematically assess the thermal stability of Fe–N–O–gra coordination structures under practical electrocatalytic conditions, ab initio molecular dynamics (AIMD) simulations were performed on seven representative models: FeN4, FeN3O, FeN2O2-pen, FeN2O2-hex, FeN2O2-oppo, FeNO3, and FeO4. The simulations were conducted at 500 K using the Nosé–Hoover thermostat within the canonical (NVT) ensemble for a total duration of 10 ps, with a time step of 2 fs, ensuring reliable tracking of structural evolution under thermal perturbation. As shown in Fig. 1(b–g), all configurations except FeO4 maintained structural integrity throughout the simulation period. No signs of bond breakage, metal atom migration, or framework collapse were observed, and the total energy fluctuations remained within a reasonable range, indicating their favorable thermodynamic and geometric stability under ambient conditions. In contrast, the FeO4 model exhibited rapid structural degradation characterized by early-stage bond dissociation and collapse of the coordination framework, suggesting insufficient thermal robustness under catalytic conditions. Accordingly, FeO4 was excluded from subsequent electronic structure and catalytic performance analyses. Detailed AIMD results are provided in Fig. S1 of the SI. Compared with other representative configurations, the FeN3O structure exhibits slight bending deformation and asymmetric energy fluctuations during molecular dynamics simulations. These fluctuations remain within a reasonable range and do not lead to structural collapse, indicating overall dynamic stability. The observed deformation originates from the asymmetric Fe–N–O coordination, where the introduction of O ligands induces local stress release and structural adjustment, without compromising the catalytic performance.

Having established the thermal stability of the Fe–N–O–gra models, we next investigated their electronic structure and conductivity to evaluate their feasibility as electrocatalyst substrates. Projected density of states (PDOS) for the six thermally stable configurations are displayed in Fig. 2(a)–(f), clearly revealing strong hybridization between Fe 3d orbitals and the 2p orbitals of neighboring N/O atoms near the Fermi level. This orbital overlap not only stabilizes the metal–ligand bonding network but also facilitates efficient electron transfer and activation of reactive intermediates (e.g., O2, *OOH, *O, *OH) during catalytic processes. Additionally, the PDOS of all six structures exhibit continuous states across the Fermi level, indicative of good electronic conductivity, which is crucial for enhancing charge transport in the ORR and OER. These comprehensive thermodynamic and electronic analyses provide a solid theoretical foundation for subsequent investigations into adsorption behavior, charge redistribution, and catalytic mechanisms.


image file: d5re00292c-f2.tif
Fig. 2 Projected density of states (PDOS) of Fe–N–O–gra catalysts with different coordination environments: (a) FeN4, (b) FeN3O, (c) FeN2O2-p, (d) FeN2O2-h, (e) FeN2O2-o, and (f) FeNO3. Contributions from Fe 3d (green), N 2p (red), and O 2p (blue) orbitals are shown. The Fermi level is set to 0 eV (indicated by the dashed line).

In both the oxygen reduction reaction (ORR) and the oxygen evolution reaction (OER), the adsorption strength of key intermediates such as O2, *OOH, *O, and *OH critically governs the reaction kinetics and overall catalytic performance. To elucidate how the Fe–N–O–gra coordination environment influences the interaction with these intermediates, we systematically calculated the adsorption energies (ΔE) for each species across different configurations, as summarized in Table S1 of the SI. For molecular oxygen, two adsorption modes were considered: the “side-on” configuration, in which both O atoms interact with the metal center, and the “end-on” configuration, where only one O atom coordinates with the Fe site, as summarized in Table S1 of the SI. Upon geometry optimization, FeN4 and FeN3O show a slight thermodynamic preference for the end-on mode, whereas FeN2O2-o/h/p and FeNO3 favor the side-on configuration. This suggests that the end-on mode is more stable and likely more conducive to initial O2 capture and activation. Among all structures, the FeN4 configuration exhibits the weakest O2 affinity with an adsorption energy of −0.89 eV. In contrast, configurations with higher O coordination, such as FeN2O2 and FeNO3, show significantly stronger O2 binding, with adsorption energies of approximately −1.10 eV and −1.26 eV, respectively. A similar trend is observed for other intermediates. For example, FeNO3 binds *OOH with an energy of −1.75 eV, compared to −1.41 eV for FeN4. The difference becomes even more pronounced for *O adsorption, where FeNO3 achieves a ΔE as low as −6.85 eV, surpassing FeN4 (−5.98 eV), indicating the formation of a stronger Fe–O bond. Likewise, the *OH adsorption energy increases from −2.83 eV in FeN4 to −4.19 eV in FeNO3, indicating that oxygen coordination modulates the binding strength of intermediates depending on the coordination environment. These results underscore that increasing the degree of oxygen coordination in Fe–N–O–gra structures enhances the electronic interaction between the active Fe site and reaction intermediates, particularly evident in the strong *O and *OH binding observed in FeNO3. The variation in adsorption strength across different configurations reveals the profound influence of coordination geometry on the reaction pathway and provides valuable insights into the structure–activity relationship of Fe-based single-atom catalysts (SACs) in bifunctional oxygen electrocatalysis.

O2 adsorption is the first step of the ORR, and O2 desorption is the last step of the OER. Therefore, a detailed understanding of the electronic coupling between Fe 3d orbitals and O2 2p orbitals is critical for elucidating the underlying catalytic mechanism. Fig. 3 presents the projected density of states (PDOS) and charge density difference plots for six representative Fe–N–O–gra models (FeN4, FeN3O, FeN2O2-o, FeN2O2-h, FeN2O2-p, and FeNO3) in their O2-adsorbed states, highlighting how coordination structure affects electronic interactions and O2 activation capability. Across all structures, varying degrees of Fe 3d and O2 2p orbital overlap are observed near the Fermi level. Charge density difference plots further confirm significant charge redistribution within the Fe–O bonding region upon O2 adsorption, indicating electron donation from the Fe center. Bader charge analysis reveals that the amount of charge transferred from Fe to O2 ranges from 0.60 to 0.76 |e|. Notably, the FeN3O configuration exhibits the highest degree of orbital hybridization and maximum charge transfer (0.76 |e|), suggesting superior O2 adsorption and activation potential. The FeN2O2 series (Fig. 3c–e) show moderate Fe–O2 coupling and charge transfer values around 0.67–0.68 |e|, with charge accumulation localized beneath the O2 molecule, reflecting stable yet tunable adsorption strength. In contrast, the FeNO3 model (Fig. 3f) displays the weakest orbital overlap and minimal charge transfer (0.60 |e|), indicating limited electronic donation capability. This constraint may hinder O–O bond cleavage during the ORR, while simultaneously weakening the electronic coupling required for *OOH formation from *O and *OH in the OER, thereby rendering the O–O coupling step kinetically unfavorable. From an electronic structure perspective, the coordination environment around the Fe center governs its d-orbital distribution and hybridization capacity, ultimately affecting the adsorption strength and activation efficiency of O2. The FeN3O configuration, in particular, demonstrates optimal orbital interaction and charge redistribution, implying lower reaction barriers, faster kinetics, and enhanced catalytic performance in practical ORR/OER applications. These findings provide theoretical support for understanding its outstanding bifunctional electrocatalytic behavior.


image file: d5re00292c-f3.tif
Fig. 3 (a–f) Electronic structure analysis of O2-adsorbed Fe–N–O–gra configurations: (a) FeN4, (b) FeN3O, (c) FeN2O2-o, (d) FeN2O2-h, (e) FeN2O2-p, and (f) FeNO3. For each configuration, the left panel shows the projected density of states (PDOS), highlighting the Fe 3d (purple) and O2 2p (red) orbital contributions near the Fermi level (set to 0 eV, dashed line); the right panel displays the corresponding charge density difference map, where yellow and cyan regions represent electron accumulation and depletion, respectively. The numerical values indicate the amount of charge transferred from Fe to O2, as determined by Bader charge analysis (unit: |e|), reflecting the electron-donating ability of the Fe center.

To comprehensively evaluate the bifunctional catalytic activity of Fe–N–O–gra single-atom catalysts in the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER), we calculated and plotted the full reaction free energy profiles for six representative configurations: FeN4, FeN3O, FeN2O2-o, FeN2O2-h, FeN2O2-p, and FeNO3, as shown in Fig. 4. The black dashed lines represent the free energy changes at U = 0 V, while the black solid lines correspond to the standard electrode potential of U = 1.23 V. Key intermediates (*OOH, *O, and *OH) and their respective free energies are labeled at each reaction step, and the overpotentials (ηORR and ηOER) are marked with red and blue circles to facilitate comparison among the different configurations. Notably, all six structures exhibit a common rate-determining step (RDS) for the ORR: the reduction of *O to *OH, indicating that the desorption of *O plays a dominant role in controlling the overall ORR kinetics. For the OER, the shared RDS is the oxidation of *O to *OOH, emphasizing that further oxidation of the *O intermediate is the critical step governing OER activity. This trend highlights the central importance of *O adsorption strength in both reactions—excessive binding inhibits its reduction to *OH, while insufficient binding impairs its oxidation to *OOH. Quantitatively, the FeN3O configuration exhibits the most favorable bifunctional catalytic performance, with the lowest overpotentials of 0.41 V for the ORR and 0.55 V for the OER. FeN4 shows moderate activity, with ηORR and ηOER values of 0.63 V and 0.96 V, respectively. In contrast, FeNO3 suffers from overly strong intermediate adsorption, resulting in significantly higher overpotentials of 1.99 V (ORR) and 1.50 V (OER), which severely limits its catalytic kinetics. In summary, Fig. 4 clearly illustrates the distinct free energy landscapes and reaction barriers associated with each Fe–N–O–gra configuration. Among them, FeN3O demonstrates the lowest energy barriers and the best-matched ηORR/ηOER values, attributed to its optimal intermediate adsorption strength. These findings corroborate the earlier electronic structure analyses, reaffirming that rational engineering of the local coordination environment around the Fe center is a powerful strategy to achieve balanced energy profiles and enhanced bifunctional electrocatalytic performance.


image file: d5re00292c-f4.tif
Fig. 4 Free energy diagrams for the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) on (a) FeN4, (b) FeN3O, (c) FeN2O2-o, (d) FeN2O2-h, (e) FeN2O2-p, and (f) FeNO3 configurations. The black dashed lines represent the reaction pathways at U = 0 V, while the black solid lines correspond to the free energy changes at the standard electrode potential U = 1.23 V. Each diagram includes structural insets of key reaction intermediates (*OOH, *O, and *OH).

To further elucidate the adsorption characteristics between metal active sites and reaction intermediates in Fe–N–O–gra catalysts, we systematically analyzed the correlations among the adsorption free energies of three key intermediates (*OH, *O, and *OOH) and constructed volcano plots for both the ORR and OER, as shown in Fig. 5. A strong linear correlation was observed between ΔG*OH and ΔG*O (Fig. 5a), with a fitting equation of ΔG*O = 0.99 + 0.44ΔG*OH (R2 = 0.95). Similarly, ΔG*OOH exhibits a positive correlation with both ΔG*O(ΔG*OOH = 2.99 + 0.46ΔG*O, R2 = 0.91; Fig. 5b) and ΔG*OHG*OOH = 3.45 + 0.20ΔG*OH, R2 = 0.87; Fig. 5c). These results suggest intrinsic scaling relationships among intermediate adsorption energies, reflecting the coupled nature of active site interactions with different intermediates. Fig. 5d presents the volcano-type dependence of ηORR and ηOER on ΔG*OH. Both overpotentials exhibit a minimum near moderate *OH adsorption strengths, indicating that catalytic activity is highly sensitive to intermediate binding energies. Notably, the FeN3O configuration lies near the apex of both volcano plots, with low overpotentials of ηORR = 0.41 V and ηOER = 0.55 V—comparable to those of benchmark noble-metal catalysts such as Pt (ηORR ∼0.45 V) and RuO2 (ηOER ∼0.42 V). Overall, the adsorption energy correlations and volcano trends in Fig. 5 reveal a clear structure–performance relationship, highlighting that moderate *OH binding is a key design principle for achieving low-overpotential, high-efficiency bifunctional electrocatalysis.


image file: d5re00292c-f5.tif
Fig. 5 Scaling relationships between adsorption free energies of key intermediates and volcano-type trends of ORR/OER overpotentials for Fe–N–O–gra catalysts. (a–c) Linear correlations between ΔG*OH, ΔG*O, and ΔG*OOH. (d) Volcano plot of ηORR and ηOERversus ΔG*OH, highlighting its role as a bifunctional activity descriptor.

Building on the preceding free energy analysis, scaling relationships, and volcano plot trends, the FeN3O configuration is confirmed to exhibit significantly reduced overpotentials for both the ORR and OER, demonstrating outstanding bifunctional electrocatalytic performance. To further elucidate the underlying mechanism of its enhanced activity, we conducted a detailed comparison of the *OH adsorption states and electronic structures between FeN4 and FeN3O. As shown in the charge density difference maps (Fig. 6a and b), both configurations exhibit pronounced charge redistribution upon *OH adsorption, with electron accumulation primarily localized along the Fe–O adsorption bond, indicating stable bonding interactions. Notably, the Fe–O bond length in FeN3O (1.85 Å) is longer than that in FeN4 (1.81 Å), suggesting a weaker binding strength. Bader charge analysis further shows that FeN3O transfers slightly more charge to the *OH intermediate (0.72 |e| vs. 0.68 |e| for FeN4), despite the longer bond length, indicating that the incorporation of O coordination helps moderate the overly strong interaction, thus facilitating intermediate desorption. This observation is corroborated by the adsorption energy comparison in Fig. 6c, where *OH adsorption on FeN3O yields an energy of −2.48 eV, significantly weaker than the −2.83 eV observed for FeN4. The reduced binding strength places FeN3O closer to the optimal region predicted by the volcano plot, where catalytic activity is maximized. Such moderate adsorption ensures sufficient binding of intermediates without passivating the active site, thereby promoting faster reaction kinetics. The projected density of states (PDOS) analyses in Fig. 6d and e provide further insight into the electronic origins of the observed trends. In FeN4, strong overlap between Fe 3d and O 2p orbitals near the Fermi level indicates strong orbital hybridization, which can hinder intermediate desorption. In contrast, FeN3O exhibits weaker orbital overlap and more localized Fe 3d states near the Fermi level, reflecting a more moderate binding affinity. Overall, excessively strong Fe–O bonding leads to over-adsorption of *OH, hindering intermediate desorption and slowing reaction kinetics, whereas too weak bonding fails to stabilize intermediates. The FeN3O configuration features a moderately elongated Fe–O bond, with the *OH adsorption energy reduced from −2.83 eV in FeN4 to −2.48 eV, accompanied by enhanced charge transfer (0.72 |e| vs. 0.68 |e|). This coupled effect of the synergistic interaction between moderate adsorption and enhanced charge transfer ensures sufficient stabilization of intermediates while avoiding site passivation from overly strong binding, thereby accelerating *OH desorption in the ORR and *OOH formation in the OER, effectively lowering the reaction barriers. To provide quantitative insight into electronic regulation, we systematically analyzed the correlation between the Fe d-band center (εd) after O2 adsorption and the ORR/OER overpotentials of different Fe–N–O coordination structures. As shown in Fig. S2 (SI), an evident inverse correlation was observed: the closer the εd lies to the Fermi level, the lower the overpotential. This result is consistent with the PDOS and charge density analyses, further confirming that N3O coordination optimizes the electronic structure of the Fe active site, thereby facilitating a balanced adsorption–desorption process and enabling superior bifunctional electrocatalytic activity. In summary, the FeN3O structure achieves a well-coordinated modulation of adsorption strength, charge transfer, and band alignment through its optimized coordination environment and electronic structure. These features collectively result in the lowest overpotentials and enhanced reaction kinetics among the studied configurations. The results further validate that tailoring coordination heterogeneity is an effective strategy for boosting the activity of single-atom catalysts and provide both theoretical guidance and practical insights for the design of high-performance Fe–N–O SACs.


image file: d5re00292c-f6.tif
Fig. 6 Electronic structure comparison of FeN4 and FeN3O configurations after *OH adsorption. (a and b) Charge density difference maps (top) and top-view geometries (bottom) of FeN4 and FeN3O, showing Fe–O bond lengths and Bader charge transfer values. (c) Comparison of *OH adsorption energies on the two configurations.(d and e) Projected density of states (PDOS) of FeN4 and FeN3O with adsorbed *OH, illustrating orbital hybridization between Fe 3d and OH 2p states.

4. Conclusion

In this study, we employed density functional theory (DFT) to systematically investigate a series of Fe–N–O–gra single-atom catalyst configurations for their bifunctional electrocatalytic performance toward the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER). The aim was to clarify how the introduction of heteroatom O modulates the electronic structure and catalytic behavior of the Fe center. Seven representative coordination models, including FeN4, FeN3O, the FeN2O2 variants (pen, hex, and oppo), FeNO3, and FeO4, were constructed and evaluated. Ab initio molecular dynamics (AIMD) and projected density of states (PDOS) analyses confirmed that all structures except FeO4 exhibit good thermodynamic and dynamic stability. Charge density difference plots and Bader charge analysis revealed that FeN3O undergoes the most significant electron transfer during O2 adsorption, accompanied by strong d–p orbital hybridization at the Fe–O interface, indicating excellent O2 activation capability. Adsorption energy and reaction free energy pathway calculations further demonstrated that different coordination environments significantly impact the binding strengths of *OOH, *O, and *OH, thereby altering the rate-determining steps. Among all models, FeN3O exhibited the lowest overpotentials, with ηORR = 0.41 V and ηOER = 0.55 V, outperforming the conventional FeN4 configuration (ηORR = 0.63 V, ηOER = 0.96 V). This high activity is attributed to the FeN3O structure's moderate *OH adsorption free energy, optimized Fe–O bond length distribution, and effective electronic coupling. Moreover, scaling relations and volcano plots of adsorption energies further confirmed that FeN3O lies near the activity apex in the bifunctional catalytic landscape, showing simultaneously excellent ORR and OER performance. Collectively, this work not only establishes a clear structure–activity relationship for Fe–N–O coordination environments but also provides theoretical guidance and design strategies for developing efficient, stable, and cost-effective bifunctional non-precious metal single-atom catalysts.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

Supplementary information is available. See DOI: https://doi.org/10.1039/D5RE00292C.

All the data relevant to the study have been included in the manuscript and as part of the supplementary information (SI). The raw data will be provided by the corresponding authors upon request.

Acknowledgements

This work was supported by the Scientific Research Fund project of Education Department of Yunnan Province (grant no. 2024J0333), the Yunnan Fundamental Research Projects (grant no. 202401AY070001-059), and the High-level Talent Research Start-up Project of Kunming Medical University (Grant No. K132310559). Thanks to Weifang YuanSuan Technology Co., Ltd. for providing the computational platform (yuansuan.top).

References

  1. S. Chu and A. Majumdar, Nature, 2012, 488, 294–303 CrossRef PubMed.
  2. S. Liu, X. Wan, Y. Sun, S. Li, X. Guo, M. Li, R. Yin, Q. Kong, J. Kong and J. Zhang, Int. J. Miner., Metall. Mater., 2022, 29, 2212–2220 CrossRef.
  3. Q. Wang, Q. Feng, Y. Lei, S. Tang, L. Xu, Y. Xiong, G. Fang, Y. Wang, P. Yang, J. Liu, W. Liu and X. Xiong, Nat. Commun., 2022, 13, 3689 CrossRef PubMed.
  4. R. Wang, H. Yang, N. Lu, S. Lei, D. Jia, Z. Wang, Z. Liu, X. Wu, H. Zheng, S. Ali, F. Ma and S. Peng, Chem. Eng. J., 2022, 433, 134500 CrossRef.
  5. D. Qiu, H. Wang, T. Ma, J. Huang, Z. Meng, D. Fan, C. R. Bowen, H. Lu, Y. Liu and S. Chandrasekaran, ACS Nano, 2024, 18, 21651–21684 CrossRef PubMed.
  6. R. Song, L. Guan, L. Fan, X. Miao, H. Zhang, J. Cheng, T. Zhou, C. Ni and J. Fan, J. Energy Storage, 2024, 80, 110345 CrossRef.
  7. Y. Nie, L. Li and Z. Wei, Chem. Soc. Rev., 2015, 44, 2168–2201 RSC.
  8. X. Ren, Q. Lv, L. Liu, B. Liu, Y. Wang, A. Liu and G. Wu, Sustainable Energy Fuels, 2020, 4, 15–30 RSC.
  9. J. A. Zamora Zeledón, M. B. Stevens, G. T. K. K. Gunasooriya, A. Gallo, A. T. Landers, M. E. Kreider, C. Hahn, J. K. Nørskov and T. F. Jaramillo, Nat. Commun., 2021, 12, 620 CrossRef.
  10. J. Yang, Y. Wang, X. Zhao, J. Kang, X. Zhou, L. Ma, Y. Yang and R. Jiang, Adv. Funct. Mater., 2025, 35, 2416215 CrossRef CAS.
  11. Y. Wu, J. Chen, J. Liu, L. Zhang, R. Abazari, T.-T. Li and J. Qian, Chem. Eng. J., 2025, 503, 158343 CrossRef CAS.
  12. S. Wang, B. Zhu and L. Yan, Inorg. Chem. Front., 2025, 12, 2833–2843 RSC.
  13. J. Zheng, C. Lai, W. Chen, C. Liu, T. Yuan, J. Lv, B. Zhao, D. Dang, G. Long, T. Wang and X. Han, Small, 2025, 21, 2411894 CrossRef CAS.
  14. J. Jangra, S. Singh, J. Shah and R. K. Kotnala, Appl. Mater. Today, 2025, 43, 102652 CrossRef.
  15. P. Shao, Z. Ren, B. Zhao, X. Wang, J. Li, J. Xie, B. Wang and X. Feng, J. Am. Chem. Soc., 2025, 147, 8769–8777 CrossRef CAS.
  16. C. Tang, L. Chen, H. Li, L. Li, Y. Jiao, Y. Zheng, H. Xu, K. Davey and S.-Z. Qiao, J. Am. Chem. Soc., 2021, 143, 7819–7827 CrossRef CAS.
  17. H. Xu, D. Cheng, D. Cao and X. C. Zeng, Nat. Catal., 2024, 7, 207–218 CrossRef CAS.
  18. G. Wenxin, L. Yinghuan, Z. Huang and W. Yuen, Microstructures, 2025, 5, 2025031 Search PubMed.
  19. M. Zhang, H. Li, J. Chen, F.-X. Ma, L. Zhen, Z. Wen and C.-Y. Xu, Small, 2022, 18, 2202476 CrossRef CAS.
  20. H. Li, D. Wu, J. Wu, W. Lv, Z. Duan and D. Ma, Nanoscale, 2024, 16, 7058–7067 RSC.
  21. J. Pei, H. Shang, J. Mao, Z. Chen, R. Sui, X. Zhang, D. Zhou, Y. Wang, F. Zhang, W. Zhu, T. Wang, W. Chen and Z. Zhuang, Nat. Commun., 2024, 15, 416 CrossRef CAS.
  22. S. Cao, S. Zhou, H. Chen, S. Wei, S. Liu, X. Lin, X. Chen, Z. Wang, W. Guo and X. Lu, Energy Environ. Mater., 2023, 6, e12287 CrossRef CAS.
  23. Z. Lu, Z. Wang, Z. Yang, X. Jin, L. Tong, R. Xu, K. Kong, Y. Zhang, Y. Wang, Y. Liu, L. Meng, Z. Pan, S.-J. Hwang and L. Li, Adv. Funct. Mater., 2025, 35, 2418489 CrossRef CAS.
  24. Z. Wang, Z. Lu, Q. Ye, Z. Yang, R. Xu, K. Kong, Y. Zhang, T. Yan, Y. Liu, Z. Pan, Y. Huang and X. Lu, Adv. Funct. Mater., 2024, 34, 2315150 CrossRef CAS.
  25. X. Qu, Y. Yan, Z. Zhang, B. Tian, S. Yin, X. Cheng, R. Huang, Y. Jiang and S. Sun, Chem. – Eur. J., 2024, 30, e202304003 CrossRef CAS PubMed.
  26. A. Mehmood, M. Gong, F. Jaouen, A. Roy, A. Zitolo, A. Khan, M.-T. Sougrati, M. Primbs, A. M. Bonastre, D. Fongalland, G. Drazic, P. Strasser and A. Kucernak, Nat. Catal., 2022, 5, 311–323 CrossRef CAS.
  27. Z. Yang, Y. Wang, M. Zhu, Z. Li, W. Chen, W. Wei, T. Yuan, Y. Qu, Q. Xu, C. Zhao, X. Wang, P. Li, Y. Li, Y. Wu and Y. Li, ACS Catal., 2019, 9, 2158–2163 CrossRef CAS.
  28. T. Sun, W. Zang, H. Yan, J. Li, Z. Zhang, Y. Bu, W. Chen, J. Wang, J. Lu and C. Su, ACS Catal., 2021, 11, 4498–4509 CrossRef CAS.
  29. X. Lu, J. Li, S. Cao, Y. Hu, C. Yang, Z. Chen, S. Wei, S. Liu and Z. Wang, ChemSusChem, 2023, 16, e202300637 CrossRef CAS.
  30. Q. Lai, L. Zheng, Y. Liang, J. He, J. Zhao and J. Chen, ACS Catal., 2017, 7, 1655–1663 CrossRef CAS.
  31. K. Yuan, D. Lützenkirchen-Hecht, L. Li, L. Shuai, Y. Li, R. Cao, M. Qiu, X. Zhuang, M. K. H. Leung, Y. Chen and U. Scherf, J. Am. Chem. Soc., 2020, 142, 2404–2412 CrossRef CAS.
  32. N. Zhang, T. Zhou, M. Chen, H. Feng, R. Yuan, C. a. Zhong, W. Yan, Y. Tian, X. Wu, W. Chu, C. Wu and Y. Xie, Energy Environ. Sci., 2020, 13, 111–118 RSC.
  33. J. Zhang, J. Li, T. Liang, K. Gan and G. Nong, ACS Appl. Energy Mater., 2024, 7, 9930–9938 CrossRef CAS.
  34. J. Bai, J. Xu, Z. Sun, C. Zhang, H. Qin, Y. Lian and Y. Su, J. Phys. Chem. Lett., 2025, 6833–6840,  DOI:10.1021/acs.jpclett.5c01240.
  35. Y. Li, H. Sun, L. Ren, K. Sun, L. Gao, X. Jin, Q. Xu, W. Liu and X. Sun, Angew. Chem., Int. Ed., 2024, 63, e202405334 CrossRef CAS.
  36. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
  37. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef.
  38. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS.
  39. Z. Qiao, R. Jiang, H. Xu, D. Cao and X. C. Zeng, Angew. Chem., Int. Ed., 2024, 63, e202407812 CrossRef CAS PubMed.
  40. L. Zhang, N. Chen, Z. Gao, R. He, X. Zhang, Y. Wang and K. Xiong, Appl. Surf. Sci., 2025, 702, 163298 CrossRef CAS.
  41. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef.
  42. K. Mathew, R. Sundararaman, K. Letchworth-Weaver, T. A. Arias and R. G. Hennig, J. Chem. Phys., 2014, 140, 084106 CrossRef.
  43. X. Liu, Y. Zhang, W. Wang, Y. Chen, W. Xiao, T. Liu, Z. Zhong, Z. Luo, Z. Ding and Z. Zhang, ACS Appl. Mater. Interfaces, 2022, 14, 1249–1259 CrossRef CAS PubMed.
  44. J. Xu, A. Elangovan, J. Li and B. Liu, J. Phys. Chem. C, 2021, 125, 2334–2344 CrossRef CAS.
  45. X. Li, Z. Su, Z. Zhao, Q. Cai, Y. Li and J. Zhao, J. Colloid Interface Sci., 2022, 607, 1005–1013 CrossRef CAS.
  46. T. Hu, M. Wang, C. Guo and C. M. Li, J. Mater. Chem. A, 2022, 10, 8923–8931 RSC.
  47. J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard and H. Jónsson, J. Phys. Chem. B, 2004, 108, 17886–17892 CrossRef.
  48. H. Sun, X. Xu, Y. Song, W. Zhou and Z. Shao, Adv. Funct. Mater., 2021, 31, 2009779 CrossRef CAS.
  49. X. Zhao, X. Liu, B. Huang, P. Wang and Y. Pei, J. Mater. Chem. A, 2019, 7, 24583–24593 RSC.
  50. G. Xiao, R. Lu, J. Liu, X. Liao, Z. Wang and Y. Zhao, Nano Res., 2022, 15, 3073–3081 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.