Strain-controlled spin regulation in Fe–N–C catalysts for enhanced oxygen reduction reaction activity

Mingyuan Yuabc, Jiaxiang Wuabc, Yashi Chend, Yongping Duabc, Ang Liabc, Erjun Kan*abc and Cheng Zhan*abc
aSchool of Physics, Nanjing University of Science and Technology, Nanjing 210094, China
bMIIT Key Laboratory of Semiconductor Microstructure and Quantum Sensing, Nanjing University of Science and Technology, Nanjing 210094, China
cEngineering Research Center of Semiconductor Device Optoelectronic Hybrid Integration in Jiangsu Province, Nanjing 210094, China
dSchool of Materials Science and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China

Received 27th June 2024 , Accepted 6th August 2024

First published on 7th August 2024


Abstract

Fe-embedded N-doped graphene (Fe–N–C) stands out as a promising electrocatalyst for the oxygen reduction reaction (ORR), a critical process in fuel cell technologies. The role of spin multiplicity at the Fe center and its influence on ORR intermediates, particularly the spin crossover effect, is a pivotal yet underexplored aspect of spin-related electrocatalysis. Unraveling the nuances of spin multiplicity and its manipulation holds the key to the next generation of Fe-based spin electrocatalysts for renewable energy. Our study employs a first-principles density functional theory (DFT) approach to dissect the electronic configurations across various spin states of ORR intermediates. Our wavefunction and bond analysis unveil a critical mechanism: the spin electron transfer from the nonbonding dx2–y2 orbital to the antibonding image file: d4ta04458d-t1.tif orbital, which dominantly dictates the spin state transition in ORR intermediates on the Fe–N–C catalyst. This insight is further complemented by the discovery that the electronic occupation of d–p π bonds, formed by X–Fe–Y (X/Y: *OH, *O, and *OOH), is insensitive to the IS–HS transition of adsorbates. Another novel finding of this research is the strain-induced modulation of spin states within Fe–N–C. We demonstrate that increasing strain drives a transition from an intermediate spin (IS) to a high spin (HS) state in FeN4 and its ORR intermediates, effectively suppressing ligand-induced spin crossover. Notably, a unique behavior is observed in five-coordinate FeN4OH and FeN4OOH, where strain prompts a return to IS due to electron injection into the dxy orbital. Our simulation of polarization current density reveals a direct correlation between applied strain and ORR onset potential. Specifically, at +8% strain, the onset potential is remarkably enhanced to 0.98 V vs. SHE due to the dominant active moiety of FeNHS4*OH, underscoring the role of spin state modulation in weakening the Fe–O bond and reducing the desorption energy of *OH. This work not only elucidates the origin of spin multiplicity in ORR intermediates on Fe–N–C but also establishes strain-controlled spin regulation as a viable strategy for tuning ORR activity. Our findings offer a significant insight into the design of spin-related electrocatalysts, paving the way for more efficient and sustainable energy technologies.


Introduction

Efficient energy conversion is the key principle in the design of electrocatalysts for renewable energy applications.1–4 Single-atom-catalysts (SACs) with isolated single metal sites dispersed on substrates have become a new frontier in catalysis science, owing to their high efficiency of atomic utilization and activity.5–10 Among different SACs, Fe-embedded N-doped graphene (Fe–N–C) is the most representative case that has shown excellent ORR activity in fuel cell catalysis.11–15 However, partial occupation of d orbitals in Fe could bring diverse spin states in electrocatalysis, which make mechanistic understanding more difficult and challenging.16–18

In recent years, the spin effects in SACs have shown significant impact on the catalytic mechanism and activity, which has been both experimentally and theoretically studied.19–21 For instance, Yang et al. introduced Mn single atoms around Fe sites in the Fe–N–C system by pre-polymerization and pyrolysis processes, which caused the Fe3+ spin state transition from the low spin (LS) to intermediate spin (IS) state, leading to the half-wave potential of 0.93 V at 0.1 M KOH.22 Zhang et al. also observed the spin state transition of Fe from LS to IS in Fe–N–C by introducing Fe atomic clusters, leading to the spin electron filling in the Fe–O σ* orbital and favors OH desorption.23 However, Cheng et al. and Hou et al. found that Fe could change from HS to IS by introducing Fe and Ru clusters and they proposed that spin transition enhances the interaction with the π* antibonding orbitals of oxygen.24,25 Zhong et al. reported a nearly linear relationship between magnetism and the oxygen reduction reaction (ORR) activity of an Fe single atom supported on C2N (C2N–Fe), suggesting that the electronic spin is a promising descriptor for Fe single-atom-based electrocatalysts.26 Hu et al. found that the ORR catalytic performance of Fe–N–C is related to the spin state, which can be controlled by external fields.27 Our previous work revealed the impact of spin multiplicity on ORR/OER catalysis and proposed the spin-switching pathway with the consideration of potential-dependent spin crossover.28 This work also emphasized that the spin state preference is jointly determined by the valence (electron number) and ligand (coordination number), but the latter one is more dominative in spin manipulation.

According to the crystal field theory, the most preferable spin state of a complex is determined by its coordination environment and metal–ligand (M–L) interaction.18,29 Thus, to manipulate the spin state, extra ligand introduction and strain engineering are the commonly used strategies. Xin et al. investigated the electronic and magnetic properties of 3d transition metals in TMN4 SACs using density functional theory calculations and found two different spin states that can be switched over via external tensile strain.30 Wang et al. demonstrated that the introduction of XO2 groups with different polarity (SeO2 > TeO2 > SO2) into the FeN4 site leads to the spin electron transfer and spin state change due to the symmetry breaking.31 However, early research mainly focuses on the spin state transition and the effect of magnetic moment on activity, while the origin of diverse spin states was rarely discussed in early studies. In particular, the correlation between spin multiplicity, orbital hybridization, coordination geometry and frontier orbital information in SACs and reaction intermediates is not well understood at the current stage.

To reveal the origin of spin multiplicity in single atom electrocatalysis, in this work we take Fe–N–C as an example and employ density functional theory (DFT) to calculate the high-spin (HS) and intermediate-spin (IS) structures of all ORR intermediates. We analyze the bonding modes of adsorbates, the orbital occupation of spin states through the projected density of state (PDOS), and the bonding information by Crystal Orbital Hamilton Population (COHP) to explain the observed magnetic moment of each ORR intermediate and its corresponding spin state. In addition, we also investigate how the spin state is modulated by mechanical strain and its impact on the ORR activity, which brings a feasible way to improve the electrocatalytic activity via spin manipulation.

Methods

Spin-polarized DFT calculations were performed by using the Vienna ab initio simulation package (VASP).32–34 In our DFT calculation, the exchange-correlation effect was treated by the general gradient approximation with the Perdew–Burke–Ernzerhof formalism (GGA-PBE).35 The van der Waals interaction of adsorbates was described by the empirical correction in Grimme's scheme (DFT-D3).36 The DFT + U method was employed to deal the strongly correlated 3d orbital of Fe and the value of Ueff = UJ was set to 4.03 eV according to our previous research.28 The electron–nuclei interaction was described using the Projector-Augmented Wave (PAW) pseudopotential.32 The energy cutoff for the plane-wave basis was set to 550 eV. The k-point sampling in the Brillouin zone was implemented with a gamma-centered 2 × 2 × 1 mesh for geometry relaxation and 10 × 6 × 1 mesh for electronic structure calculation, which has been fully tested for energy convergence in our previous study.28 The Fe–N–C SAC was modeled by a FeN4 site embedded into a monolayer graphene in an orthorhombic cell with the lattice parameters of a = 9.94 Å and b = 17.04 Å, and the surface was modeled in contact with a 30 Å vacuum layer. The convergence criteria in our DFT calculation were 1 × 10−6 eV and 0.02 eV Å−1 for electronic and ionic optimization. VASPKIT was used for electronic structure and density of states (DOS) analysis.37 The LOBSTER package was used to analyze the bonding orbital and binding strength between Fe and adsorption via COHP analysis.38,39 The visualized orbitals were calculated using the wannier90 package.40

The adsorption free energy and geometry relaxation at constant electrode potential were computed by VASPsol integrated with the optimization algorithm developed by Xiao et al.41–44 The implicit solvation based on the linearized Poisson–Boltzmann model (PBM) was adopted to describe the electrochemical response behavior of the aqueous electrolyte with a dielectric constant of 80 and Debye length of 3 Å in our simulation.45 The potential-dependent free energy can be expressed as:

 
E = EvaspEF × Δq (1)
where Evasp is DFT energy determined by VASPsol, EF is the Fermi energy and Δq is the extra number of electrons. The electrode potential referenced to the standard hydrogen electrode U is:
 
U = −4.6V − (EFEsol) (2)
where Esol is electrostatic potential in the bulk of solution. Then, the calculated potential-dependent free energies can be fitted using a quadratic function:46
 
E(U) = −1/2 × C × (UU0)2 + E0 (3)
where C is the interfacial capacitance of the electrode, U0 is the potential of zero charge (PZC), and E0 is the energy at the PZC. For each structure, we calculated at charges ranging from 0e to +3e with the increment of +0.5e; corresponding fit parameters are shown in Tables S1–S3. The energy of the reacting molecules is given in Table S4. The potential-dependent free energy was used to construct the free energy profile of the ORR and set up the micro-kinetics simulation; all the detailed information is provided in the ESI.

The ab initio molecular dynamics (AIMD) simulation was performed to validate the stability of Fe spin states in contact with explicit water solvation. Specifically, six H2O molecules were displaced on each side of FeNIS4, and AIMD simulations were carried out using the Nosé-thermostat at 300 K for 1 ps with a time step of 0.5 fs. The out-of-plane displacement of the Fe and spin states were analyzed and compared with the implicit solvent model.

Results and discussion

Spin rearrangement of image file: d4ta04458d-t2.tif under external strain

In the FeN4 model, Fe is coordinated with four N atoms to form a square-planar complex. Four N atoms in sp2 hybridization donate six electrons in total and form four σ bonds with the dsp2 hybridized Fe center,47 as shown in Fig. 1a. Except two electrons from the 4s orbital of Fe that have bonded to N, the five 3d orbitals could be selectively filled with the remaining six 3d electrons, generating three possible spin states: S = 0, 1, or 2. Here we note that, in the case of HS (S = 2), the dxy orbital filled with spin-up electrons is actually the antibonding orbital σ* of Fe–N bonds, as shown in Fig. 1a. This attribution has been clearly validated by the COHP analysis (Fig. S2) of the Fe–N bond in FeN4. We obtained four structures of three spin states, among which S = 2 has two configurations with out-of-plane Fe and in-plane Fe, as marked by FeNHS4-1 and FeNHS4-2 in Fig. S1 and S3. Their stability follows the order: FeNIS4 (S = 1) > FeNHS4-1 (S = 2) > FeNHS4-2 (S = 2) > FeNLS4 (S = 0). Then, we applied a monoaxial strain from −6% to +8% along the y direction to these four structures as shown in Fig. 1b. The relationship of energy vs. strain in Fig. 1c shows that spin state transition from S = 1 to S = 2 occurs when the applied strain is beyond +6% or below −4%. To understand this strain-induced spin state transition, we compared the structural parameters under different strains (Fig. S4). The Fe–N bond length increases with strain in FeNIS4, FeNHS4-1 and FeNLS4 due to the expanding N4 vacancy site (Fig. 1d), which weakens the σFe–N bond and reduces the energy of the antibonding image file: d4ta04458d-t3.tif orbital. By comparing PDOS of FeNIS4 under different strains, we found that the energy of the dxy orbital decreases when the strain increases but other d orbitals are less affected by strain (Fig. S5). Partial spin density was found on N atoms in FeNHS4-1 and FeNHS4-2 due to the electron filling in the dxy orbital, which refers to the antibonding orbital image file: d4ta04458d-t4.tif, as shown in Fig. S6. This anomalous electron filing in the antibonding orbital also explains why FeN4 energetically prefers IS at equilibrium in Fig. 1c. To lower the energy of the dxy image file: d4ta04458d-t5.tif orbital and stabilize the HS state, application of tensile strain is a feasible way. As shown in Fig. 1c and e, under 6% tensile strain, FeN4 prefers the HS electronic configuration rather than IS. When the strain is between −4% and +6%, the four nonbonding d orbitals (the dxy orbital forms four σFe–N bonds) are filled with six d electrons. According to Hund's rule, the electrons should occupy as many different orbitals as possible, which lead to the most preferable IS electronic configuration, as shown in Fig. 1e. For FeNHS4-1, tensile or compressive strain increases the Fe–N bond length, but large tensile strain could drag the out-of-plane Fe back to the graphene plane, resulting in the transition to FeNHS4-2 (Fig. 1e and S4). When the applied strain is below −2%, due to the bending of the graphene substrate under compressive strain (Fig. 1d), the Fe–N bond in the out-of-plane configuration continues to increase and weakens the σFe–N bond, making the HS state more favorable. We also calculated the strain applied on the y-axis in bare FeN4. As shown in Fig. S7, when tensile strain is applied to the y-axis, the same spin state transition characteristics are shown as when strain is applied to the x-axis. Specifically, a spin state transition from S = 1 to S = 2 occurs when the applied strain is beyond +6%. However, under compressive strain, the IS state remains the most stable spin state instead of turning to out-of-plane configuration of the HS state. This can be attributed to the different compressibility of graphene in different directions. As shown in Fig. S8, under −6% x-axis strain, the compression ratio of the hexagonal ring is only 4.2%, while it increases to 5.5% with the strain along the y axis. This means that the compression ratio of the FeN4 center under x-axis strain is larger than that of the y-axis strain, resulting in the insensitive response of the spin transition under y-axis strain.
image file: d4ta04458d-f1.tif
Fig. 1 (a) Schematic diagram of the Fe–N bond and energy levels in FeN4. (b) Fe–N–C structure and the strain diagram. (c) Free energy of different spin states of Fe–N–C under strain from −6% to 8%. (d) In-plane (down) and out-plane (up) of Fe–N–C structure evolution under different strains. (e) Electron occupation diagram of high spin (red) and intermediate spin (blue) with the d orbital diagram of Fe.

To demonstrate the feasibility of applying strain, we calculated the formation energy of FeN4 with different spin states under different strains along the x-axis. As shown in Fig. S9, the formation energy of IS and HS states of FeN4 with no strain is negative, while the formation energy of LS is slightly positive, indicating good stability of IS and HS FeN4 under zero strain. Under the tensile strain, the formation energies of all spin states decrease, suggesting that the formation of FeN4 becomes more favorable in stretched graphene. In contrast to the tensile strain, the formation energy of all spin states increases under compressive strain, except for the HS state with an out-of-plane configuration (FeNHS4-1), which can be attributed to the out-of-plane displacement of Fe that weakens the strong interatomic repulsion. These results indicate that it is feasible to tune the spin state of FeN4 by tensile strain in the experiment from a thermodynamic perspective. Since the tensile strain is experimentally more feasible for graphene-based SACs,48–50 we will mainly focus on the x-axis tensile strain of 4% and 8% in ORR intermediates.

Strain-induced spin regulation in five-coordinate FeN4–X (X = *OH/*O/*OOH)

Considering the strain-induced spin state transition in bare FeN4, the possible spin state transition in ORR intermediates should be taken into consideration. According to early studies, within the operating potential of the ORR, the reaction intermediates can be adsorbed on one side of the Fe–N–C monolayer and form a new active site with five-fold coordination on Fe.28,51–53 We first constructed three possible active sites in HS/IS states without lattice strain: FeN4*OH (S = 5/2, 3/2), FeN4*O (S = 2, 1), and FeN4*OOH (S = 5/2, 3/2). Their corresponding electronic configuration is determined by PDOS analysis, as shown in Fig. S10. By comparing the spin density and COHP analysis results of Fe–O/Fe–N bonds, we interestingly found that the Fe center is still in dsp2 hybridization, the same as that in the bare FeN4. This finding is different from the prediction by conventional Valence Shell Electron Pair Repulsion (VSEPR) theory which should predict d2sp3 hybridization in the Fe center based on the number of valence electrons. The electronic configuration of IS/HS FeN4–X (X: *OH/*O/*OOH) is similar to that of bare FeN4. Two electrons are either paired in the nonbonding dx2–y2 orbital for the IS configuration or individually filled in dxy image file: d4ta04458d-t6.tif and dx2–y2 orbitals for the HS configuration because the adsorbate does not change the bonding type of Fe–N, as shown in Fig. S11. However, besides the formation of σFe–N, the Fe center also bonds to the absorbed OH and O via dxz, dyz, and dz2 orbitals, leading to the presence of spin density in σFe–O and πFe–O, as shown in Fig. 2. The sp2 of O and dz2 of Fe form a σ bond with three electrons, leading to the bond order of 0.5 due to one electron in the image file: d4ta04458d-t7.tif orbital.23 This has been validated by COHP analysis in Fig. 2c, which shows that the –ICOHP value of Fe(dz2)–O(pz) for spin-up electrons is close to 0 (Fig. 2c and S12). Besides, the px orbital of O forms a π23 bond with the dxz orbital of Fe with the bond order of 0.5, while the bonding interaction of py–dyz is relatively much weaker, as shown in Fig. 2c and S12. By scanning the rotational potential energy surface of O–H orientation in Fig. 2d, we found that the OH bond tends to align with the y axis (pointing to a six-membered ring) rather than the x axis (pointing to a five-membered ring) with the energy difference of 0.03 eV. This observation indicated that the coordination of Fe–N4 is not a rigid D4h symmetry. An identical bonding type was also found in FeN4*OOH which includes the σFe–O and πFe–O both with the bond order of 0.5 (Fig. S13) and the O–O bond also tends to align with the y axis (Fig. S14). Due to the partial filling of electrons in the image file: d4ta04458d-t8.tif and image file: d4ta04458d-t9.tif orbitals, we found obvious spin density on O atoms (Fig. S15a–d). In FeN4*O, the empty O(pz) orbital bonds to the Fe(dz2) orbital, generating a σFe–O bond with the bond order of 1. The dxz and dyz orbitals in Fe form two π32 bonds with the px and py orbitals of O in side-by-side mode (Fig. 2b and S16) with obvious spin density observed on O atoms in Fig. S15e–f. In the HS state of FeN4*O, partial electron transfer to the antibonding image file: d4ta04458d-t10.tif orbital (Fig. S16b) was observed. The COHP analysis shows that the spin-up –ICHOP for the σ bond in Fe–O decreases from 0.79 (IS) to 0.24 (HS), suggesting that the valence of O is between −2 and −1. This result also agrees with the experimental finding by Frevel et al., who reported the coexistence of μ2-O and μ1-O oxyl species in OER intermediates through O K-edge X-ray absorption spectroscopy.54 In summary, through the electronic structure and wavefunction analysis, we find that FeN4 and FeN4–X possess the same dsp2 hybridization mode on Fe, generating four stable σFe–N bonds. The IS–HS transition is dominated by the spin electron transfer from the nonbonding dx2–y2 orbital to the antibonding image file: d4ta04458d-t11.tif (dxy) orbital. With the adsorption of *OH and *OOH, Fe–O forms one σFe–O bond (dz2–sp2) and one π32 bond (dxz–px) both with the bond order of 0.5. With the *O adsorption, Fe–O forms one σFe–O bond (dz2–pz) and two π32 bonds (dxz–pz and dyz–py) with the bond order of 1 and 0.5 (for each π bond) respectively. Different from FeN4*OH and FeN4*OOH, the antibonding image file: d4ta04458d-t12.tif (dz2–pz) orbital also partially contributes to the spin electron transfer in the IS–HS transition.
image file: d4ta04458d-f2.tif
Fig. 2 Schematic diagram of the Fe–O bond and electron occupation in (a) FeN4–X(*OH/*OOH) and (b) FeN4*O (red/blue arrows on the right part represent electron occupation under HS/IS) (c) Orbital resolved COHP analysis for the Fe–O bond in FeNHS4*OH. (d) Energy variation as a function of the rotation of the H–O–Fe–N dihedral angle; the labeled θH–O,y denotes the angle between the H–O bond and the y-axis within the xy plane.

As we mentioned before, since the IS–HS transition is mainly dominated by the spin electron transfer from the nonbonding dx2–y2 orbital to the antibonding image file: d4ta04458d-t13.tif (dxy) orbital, tuning the Fe–N bond length is a feasible way to manipulate the spin state of Fe. The energy and structural parameters of FeN4–X under different strains are listed in Table 1. Comparing the structural parameters in Table 1, *OH/*O/*OOH adsorption brings an out-of-plane displacement (0.46–0.8 Å) on Fe and increases the Fe–N bond length from 1.90–1.97 Å (FeN4) to 1.95–2.06 Å (FeN4–X). This observation explains why the HS of five-coordinate FeN4–X is energetically more favorable than IS.

Table 1 The structural parameter and energy of HS and IS FeN4–X (X: *OOH, *O, *OH) in the neutral state under 0%, 4%, and 8% strain. The most stable spin state is marked by a bold font
  Spin E (eV) Fe–N (Å) dFe (Å) Spin E (eV) Fe–N (Å) dFe (Å)
0% strain
FeN4*OOH 5/2 −590.27 2.05 0.78 3/2 −590.09 1.97 0.55
FeN4*O 2 −581.10 2.06 0.80 1 −580.81 1.95 0.46
FeN4*OH 5/2 −586.03 2.05 0.78 3/2 −585.80 1.96 0.50
[thin space (1/6-em)]
4% strain
FeN4*OOH 5/2 −587.55 2.08 0.60 2 −587.64 2.12 0.69
FeN4*O 2 −578.40 2.10 0.65 1 −577.97 2.00 0.36
FeN4*OH 5/2 −583.31 2.08 0.63 3/2 −583.04 2.01 0.38
[thin space (1/6-em)]
8% strain
FeN4*OOH 5/2 −580.79 2.13 0.41 2 −581.03 2.17 0.49
FeN4*O 2 −571.71 2.16 0.52 1 −570.99 2.07 0.30
FeN4*OH 5/2 −576.54 2.14 0.45 2 −576.77 2.17 0.49


With the introduction of tensile strain of 4% and 8%, an unexpected spin state transition was observed in FeN4–X. Under 4% and 8% tensile strain, the spin state of FeN4*OOH and FeN4*OH varies from S = 5/2 to S = 2, while the FeN4*O still prefers HS (S = 2) under strain, as shown in Table 1. Unlike the regular spin state transition between dxy and dx2–y2 orbitals, our PDOS analysis in Fig. 3 indicates that the transition from S = 5/2 to S = 2 originates from external spin electron injection in the dxy orbital, while electronic occupation of other d orbitals is not affected. Bader charge analysis shows that the strain-induced electron filling in the dxy orbital comes from the neighboring N and C atoms, as shown in Fig. S17 and S18. In Fig. 3, the peak position of the dxy (spin-up component) in FeNIS4*OOH/*OH is slightly beyond the Fermi level at 0% strain. With the increase of strain, the energy of dxy gradually drops down below the Fermi level, eventually leading to the electron filling and spin state transition. Besides the charge density analysis, the wavefunction analysis could provide a deeper understanding of this spin state change. We performed the COHP analysis of Fe–N bonds in FeNIS4*OH and FeNHS4*OOH at different strains, as shown in Fig. S19. As tensile strain increases, the –ICOHP of the Fe–N bond gradually decreases, indicating that the tensile strain weakens the Fe–N bond. Additionally, the –ICOHP value of the spin-down in the IS state is consistently larger than that of HS until 8% strain. This indicates that the overlapping part of the wavefunction in N and Fe is significantly weakened, leading to the partial bond breaking and backward electron transfer. Different from FeN4*OH and FeN4*OOH, the spin state transition was not observed in FeN4*O due to the relatively shorter Fe–N bond, as shown in Table 1. The presence of two π32 bonds (dxz–pz and dyz–py) in FeN4*O could inhibit the out-of-plane shift of Fe, thus shortening the Fe–N bond and preventing the spin state transition.


image file: d4ta04458d-f3.tif
Fig. 3 The PDOS of (a) FeNIS4*OH and (b) FeNIS4*OOH under 0%, 4% and 8% strain.

Since the four- and five-coordinate Fe are both ORR-active, it's important to track how the strain-induced spin state transition affects the partition of different active moieties. To evaluate the potential-dependent evolution of active sites with HS/IS states, we calculated constant-potential free energy of FeN4–X at different strains, as shown in Fig. S20–S22. At 0% strain, potential-dependent spin state crossover was observed in the ORR reaction potential window (0–1.23 V vs. SHE) for both bare FeN4 and FeN4*O, and the corresponding potential was approximately 1.1 V vs. SHE. This is attributed to the effect of electric double layer capacitance of solution in different spin states. With the increase of strain, the crossover potential increases owing to the larger energy difference at the potential of zero charge. At the same time, FeN4*OOH also shows spin crossover at 4% and 8% strains. The lower energy of the IS state below the crossover potential can be attributed to electron injection in the dxy orbital, but increasing potential could lead to electron loss in FeN4. When the electrode potential reaches the crossover potential, an electron loss in the dxy orbital is observed, resulting in the spin state turning back to S = 3/2, as illustrated in Fig. S23. However, the FeN4*OH under 8% strain exhibits the same energy in HS and IS when U > 0.50 V vs. SHE due to the electron loss in the dx2–y2 orbital instead of the dxy orbital (Fig. S24).

Then, we calculated the coverage of each configuration and spin states at different potentials and different strains based on formation energy, as illustrated in Fig. S25. Two dominant active sites (FeNIS4 and FeNHS4*OH) at 0% strain are found in our calculation. As shown in Fig. 4a, the active site is dominated by bare FeNIS4 when U < 0.45 V vs. SHE and by FeNHS4*OH when U > 0.47 V vs. SHE. As strain increases to 4%, FeNHS4-2 becomes the dominant active site with the coverage potential range increasing to U < 0.62 V vs. SHE, while the coverage of FeNIS4 at U = 0 V vs. SHE is only 30% and gradually decreases as the potential increases. When U > 0.62 V vs. SHE, the active site is dominated by FeNHS4*OH. With further increasing the strain to 8%, the coverage of FeNIS4 is very low and can be ignored. The active site is dominated by FeNHS4, and the coverage potential range further increases to U < 0.86 V vs. SHE. According to our discussion on spin stability of bare FeN4 under different strains, FeNHS4-2 becomes the main active site in the low potential range at 4% and 8% strain because tensile strain enhanced the stability of the HS state. Furthermore, the coverage range of FeNHS4*OH active sites shifts toward a higher potential region with strain, which indicates that tensile strain reduces the adsorption strength between *OH and Fe. We used LOBSTER to analyze the strength of the Fe–O bond in FeNHS4*OH under different strains and found that the –ICOHP value decreases as strain increases, suggesting the weakened Fe–O bond as the strain increases (Fig. S26).


image file: d4ta04458d-f4.tif
Fig. 4 Coverage vs. potential of dominant active sites under (a) 0%, (b) 4% and (c) 8% strain.

Unusual spin regulation in the six-coordinate X–FeN4–Y (X/Y = *OH/*O/*OOH)

Subsequently, we calculated all the possible ORR intermediates with a six-coordinate Fe center with the HS/IS state at 0% strain, and the corresponding results are listed in Table 2. Interestingly, the spin state of six-coordinate Fe is either S = 3 or 1 in X–FeN4–Y (X/Y: *OOH/*OH) and S = 2 or 1 in O*FeN4–X (X: *OOH/*O/*OH). We calculated the PDOS to identify the orbital occupation of HS/IS intermediates in Fig. S27. Interestingly, it shows that no matter what ligands are bonded to Fe, the d orbital of Fe always retains at least 4 electrons, which means the highest valence of Fe is +4. In order to reveal the bonding information of six-coordinate ORR intermediates, according to the PDOS and COHP analysis, we provided a schematic of electronic occupation and bonding types in X–FeN4–Y, as shown in Fig. 5. In the six-coordinate model, the Fe center is bonded with four N atoms and two axial ligands to form an octahedral complex with d2sp3 hybridization.47 Four hybridized d2sp3 orbitals of Fe are bonded to the sp2 orbital of N in the xy plane and the other two d2sp3 orbitals are bonded to O atoms in the z direction, as shown in Fig. 5a. In X–FeN4–Y (X/Y: *OH/*OOH), besides the σFe–N and σFe–O formed by d2sp3 hybridized orbitals, the px orbital of adsorbates (OH/OOH) and the dxz orbital form a delocalized π53 bond, as described in Fig. 5a. Different from the five-coordinate cases, we obtained the S = 1 and S = 3 spin states in the HO*FeN4*OH complex. In the S = 1 state, the nonbonding dyz orbital of Fe and the π53 bond (dxz orbital) are filled with two unpaired electrons. To convert the S = 1 to S = 3 state, the spin electron transfer from nonbonding dx2–y2 to image file: d4ta04458d-t14.tif and from σFe–O to image file: d4ta04458d-t15.tif is observed in Fig. 5a. According to the schematic description of orbital occupations, the bond order of Fe–O is 1 in the S = 1 state, indicating a stronger OH binding in HO*FeN4*OH than that in FeN4*OH. This also can be confirmed by the fact that the Fe–O bond length in HO*FeN4*OH is shorter than that in FeN4*OH/*OOH (Table S5). However, in the HS (S = 3) state, PDOS shows a spin electron filling in the dz2 orbital due to the electron transfer from σFe–O to image file: d4ta04458d-t16.tif. According to the COHP analysis of the Fe–O bond in Fig. S28, the –ICOHP value of the spin up state of Fe(dz2)–O(pz) bond shows a significant difference in the HS state (0.44) and IS state (0.72), indicating the typical spin electron transfer in the HS–IS transition. In addition, the spin charge density on O also can be clearly observed in the HS state, as shown in Fig. S29a. Dumbbell-shaped spin charge density on O in parallel to the y-axis in the IS state is observed in Fig. S29b, indicating the spin up electron filling in the π53(O–Fe–O) bond. Bader charge analysis of Fe shows that the HS and IS states hold 6.17|e| and 6.24|e| electronic charges, suggesting that no external charge transfer exists in the HS–IS transition. This conclusion was also supported by the observed increasing Fe–O bond length from 1.85 Å (IS) to 2.00 Å (HS), as shown in Fig. S29.
Table 2 The structural parameter and energy of HS and IS X–FeN4–Y (X/Y: *OOH, *O, and *OH) in the neutral state under 0%, 4%, and 8% strain. The most stable spin state is marked by a bold font
  Spin E (eV) Fe–N (Å) dFe (Å) Spin E (eV) Fe–N (Å) dFe (Å)
0% strain
HOO*FeN4*OOH 3 −604.07 1.98 0.03 1 −604.34 1.93 0.01
HOO*FeN4*O 2 −594.79 2.00 0.33 1 −595.27 1.94 0.15
HOO*FeN4*OH 3 −599.76 2.01 0.01 1 −599.92 1.93 0.02
O*FeN4*O 2 −584.57 1.99 0.17 1 −585.05 1.94 0.04
O*FeN4*OH 2 −590.31 1.99 −0.15 1 −590.83 1.94 −0.13
HO*FeN4*OH 3 −595.36 2.00 0.01 1 −595.58 1.94 0.01
[thin space (1/6-em)]
4% strain
HOO*FeN4*OOH 3 −601.78 2.05 0.00 3/2 −602.01 2.05 0.01
HOO*FeN4*O 2 −592.45 2.06 0.16 1 −592.60 2.01 0.10
HOO*FeN4*OH 3 −597.49 2.06 0.00 1 −597.27 2.00 0.00
O*FeN4*O 2 −582.31 2.05 0.01 1 −582.44 2.00 0.01
O*FeN4*OH 2 −588.02 2.05 −0.08 1 −588.21 2.00 −0.08
HO*FeN4*OH 3 −593.19 2.06 0.00 1 −592.89 2.01 0.00
[thin space (1/6-em)]
8% strain
HOO*FeN4*OOH 3 −595.40 2.13 0.00 3/2 −595.56 2.12 0.00
HOO*FeN4*O 2 −585.97 2.13 0.12 1 −585.76 2.09 0.07
HOO*FeN4*OH 3 −591.12 2.15 0.09 3/2 −591.04 2.13 0.04
O*FeN4*O 2 −575.95 2.13 0.00 1 −575.63 2.13 0.00
O*FeN4*OH 2 −581.58 2.12 −0.07 1 −581.39 2.08 0.00
HO*FeN4*OH 3 −586.83 2.15 0.00 3/2 −586.73 2.13 0.00



image file: d4ta04458d-f5.tif
Fig. 5 (a) Schematic diagram of the sp3d2 hybrid orbital of Fe. Schematic diagram of the Fe–O bond and electron occupation in (a) X–FeN4–Y (X/Y: *OH/*OOH), (b) O*FeN4–X (X: *OH/*OOH), and (c) O*FeN4*O (red/blue arrows on the right part represent electron occupation under HS/IS).

With the replacement of *OH/*OOH by *O, the σFe–O is formed by the empty O(pz) orbital and the Fe(d2sp3) orbital. The px and py orbitals of O could form the π bonds with the dxz and dyz orbitals of Fe, as illustrated in Fig. 5b and c. The presence of πFe–O bonds can be clearly confirmed by COHP analysis in Fig. S30. Taking the O*FeN4*OH as an example, the –ICOHP value of the dxz–pz orbital is 0.00 and 0.14 for spin up and spin down states, indicating that the πFe–O in Fe–OH is very weak. Compared to the Fe–OH bond, the –ICOHP value of the dxz–px orbital (Fe–O bond) is 0.00 and 0.66 with a relatively higher peak in COHP analysis, indicating that there is a typical πFe–O bond formed by the Fe(dxz) and O(px) in Fe–O. Thus, the *OH ligand only provides the py orbital to form the π bond with Fe, while the *O ligand could provide both px and py orbitals to form the π bond. Therefore, there is a π32 and a π53 in the HO*FeN4*O, but two π53 in the O*FeN4*O. This bonding scenario of three-centered delocalized π bonds is similar to that of the HN3 and N3 anion. In addition, the COHP analysis in Fig. S28 and Fig. S31 shows that the HS–IS transition doesn't change the electronic occupation of Fe–O bonds in HO*FeN4*O and O*FeN4*O. In summary, the six-coordinate X–FeN4–Y adopts the d2sp3 hybridization in Fe favoring the IS in all intermediates (Table 2) due to the unshifted Fe atom (along the z axis) in the graphene plane. The *OH and *O ligands could provide py and py(px) orbitals to form three-centered π bonds, leading to highly stable O–Fe–O bonding that is not affected by the HS–IS transition in HO*FeN4*O and O*FeN4*O.

With the introduction of +4% tensile strain on X–FeN4–X (X/Y: *OOH/*O/*OH), a strain-induced spin state transition from IS (S = 1) to HS (S = 3) was found in HO*FeN4*OOH and HO*FeN4*OH due to the expanded N4 vacancy as we discussed before, but the IS–HS transition was not observed in HOO*FeN4*OOH and O*FeN4–X (X: *OOH/*O/*OH). However, we interestingly found the IS(S = 1)–IS(S = 3/2) transition in HOO*FeNIS4*OOH at +4% strain due to an extra electron injection from N atoms, as shown in Fig. S32. This electron injection-induced S = 1 to S = 3/2 spin state transition was also observed in HO*FeN4*OOH and HO*FeN4*OH (Fig. S33) when the strain increases to +8%. This phenomenon can be explained by the effect of *O/*OH ligands on the energy of the dxy orbital. Owing to the presence of π, the Fe–O bond in *O-containing adsorbates is stronger than that in *OH-containing adsorbates, leading to the larger orbital splitting energy in the octahedral crystal field and higher energy of the dxy orbital. Thus, it is energetically more difficult to induce the electron filling of the dxy orbital in *O-containing intermediates. This analysis also explains why the HOO*FeN4*O, O*FeN4*O and HO*FeN4*O still retain the IS state under +4% strain while HOO*FeN4*OH and HO*FeN4*OH converted to HS, as shown in Table 2.

Enhanced ORR kinetics via strain-induced spin regulation

After a detailed discussion on the spin states and electronic configuration, we calculated the energy vs. potential relation of all ORR intermediates under 0%, +4%, and +8% strain, as shown in Fig. S34–S36. Based on potential-dependent partition of different active sites in Fig. 4, the ORR pathways and corresponding reaction free energy under different strains were obtained as shown in Fig. 6 and S37. According to the reaction free energy results, the ORR only occurs on the FeN4 and FeN4OH sites, and we defined it as path A and path B respectively. At 0% strain, the bare FeN4 and HO*FeN4–X (X: *OOH/*O/*OH) prefer IS, while FeN4–X prefer HS, leading to the ligand-dependent IS/HS switching ORR mechanism (Fig. 6a). Under +4% strain, the bare FeN4 and HO*FeN4–Y turn to the HS state, while the HO*FeN4–Y (Y: *OOH/*OH) goes back to the IS at low overpotential. FeN4*OOH exhibits the IS state at high overpotential owing to the electron injection in the dxy orbital and its induced spin state transition, but it turns to HS at low overpotential (Fig. 6c). With the increase of strain to +8%, all the reaction intermediates prefer the HS state at low overpotential (Fig. 6e), leading to the steady HS Fe center along the ORR path. According to potential-dependent reaction free energy in Fig. S37, we confirmed that whether there is applied strain or not, *OH desorption and *OOH adsorption are the potential-limiting step in path A and path B respectively. In addition, we also found that OH desorption is easier in HO*FeN4*OH than in FeN4OH despite the orbital occupation suggesting a higher Fe–O bond order in HO*FeN4*OH because the image file: d4ta04458d-t17.tif orbital is unoccupied. The limiting potential of path A is 0.4 V at 0% strain, 0.7 V at +4% strain and 0.75 V at +8% strain since the strain-weakened Fe–O bond promoted *OH desorption (Fig. S26), which significantly enhances the ORR activity.48 For path B, the limiting potential is 0.8 V at 0% strain, 0.7 V at 4% strain and 1.05 V at 8% strain due to the spin state transition of active sites. According to the polarization current simulation, the overall onset potential is reported as 0.8 V at 0% strain, 0.7 V at 4% strain, and 1.05 V at 8% strain as shown in Fig. 6. These results suggest that the strain-induced spin manipulation is a feasible strategy to improve the ORR performance of Fe–N–C in fuel cell applications.
image file: d4ta04458d-f6.tif
Fig. 6 ORR reaction path with the spin state for intermediate and corresponding polarization curves at (a and b) 0% strain, (c and d) 4% strain and (e and f) 8% strain.

Finally, to validate the impact of solvation on the Fe spin state in our simulation, we also employed the AIMD simulation with explicit water solvation in contact with FeN4. Our simulation shows that, within 1 ps, the Fe atoms shift slightly in the vertical direction (less than 0.05 Å), and the spin state of FeN4 does not change, which indicates that explicit solvation does not significantly affect the position of Fe in the z direction and does not affect the preferable spin state, as shown in Fig. S38. In addition, we also calculated the free energy of IS and out-of-plane HS of FeN4 with a H2O adsorption in the implicit solvent model, as shown in Fig. S39. After adsorption of H2O, the IS state is still the most preferable state in FeN4 and the out-of-plane displacement of Fe is 0.05 Å, which is much smaller than that in FeNHS4 (0.53 Å). These results suggest that the spin state of Fe is only dominated by chemical factors, such as valence or ligand change, while the electrostatic interaction from solvation has very little influence.

Summary and conclusion

In our quest to unravel the intricacies of spin multiplicity in ORR intermediates on Fe–N–C, we conducted a systematic exploration of the electronic configuration and bonding paradigms of both high-spin (HS) and intermediate-spin (IS) states across the ORR process. Our comprehensive analysis, based on Projected Density of States (PDOS), Crystal Orbital Hamilton Population (COHP), and spin density visualization, has unveiled the nuanced electronic configurations of HS/IS states in these ORR intermediates. We discovered that within the four-coordinate FeN4, the nonbonding orbitals dyz, dz2, dxz, and dx2–y2 are filled with six d electrons, culminating in the most stable IS state according to Hund's rule. A pivotal finding of our work is the dominance of spin electron transfer from the nonbonding dx2–y2 orbital to the antibonding image file: d4ta04458d-t18.tif orbital in governing the IS–HS transition, a process that can be modulated by adjusting the Fe–N bond length. The application of tensile strain has been shown to drive an IS–HS transition, attributed to the decrease of the image file: d4ta04458d-t19.tif orbital energy with concurrent elongation of the Fe–N bond. In five-coordinate FeN4–X species (X = *OH/*OOH), our findings reveal the complex electronic configuration of σFe–O and πFe–O bonds, with implications for understanding the IS–HS transition under strain. Notably, the *O adsorption introduces two π bonds between Fe–O formed by Fe(dxz)–O(px) and Fe(dyz)–O(py), leading to the steady bonding of Fe and O with the bond order of 2. In six-coordinate X–FeN4–Y complexes, we observed a d2sp3 hybridization pattern in Fe, with the highest oxidation state of +4, and the formation of both σ and π bonds in *OH/*OOH and *O adsorbates. Intriguingly, the bonding mode in HO–Fe–O and O–Fe–O complexes parallels that of hydrazoic acid and the azide anion, featuring three-centered π bonds. The tensile strain effect extends to these six-coordinate X–FeN4–Y complexes, prompting a spin state transition from S = 1 to S = 3/2 due to electron transfer from N/C to Fe. At +8% tensile strain, the spin states of most ORR intermediates turn to S = 3 or S = 2, underscoring the critical role of spin manipulation in tuning catalytic activity. Our potential-dependent formation energy calculations reveal a dynamic conversion between FeNIS/HS4 and FeNHS4*OH as active sites across different potentials, with FeNHS4 gaining dominance under increased strain. By pinpointing the most stable spin states for each intermediate, we demonstrate that tensile strain effectively suppresses ligand-related spin transitions, stabilizing the HS state on the Fe center across the whole ORR process. Finally, the simulated current density indicates that +8% strain significantly improves the ORR activity, pushing the onset potential to 1.05 V vs. SHE. Our study not only elucidates the origins of spin multiplicity in Fe–N–C but also proposes a viable strategy for enhancing ORR activity through spin manipulation. Our work could significantly facilitate the development of spin-related electrocatalysis from both experimental and theoretical perspectives.

Data availability

The authors confirm that the data supporting the findings of this study are available within the article and its ESI. The code and input parameters of this simulation work are available on request from the authors.

Author contributions

C. Z. designed the project. M. Y. performed the simulation, data analysis, and manuscript writing. M. Y., Y. D., J. W., Y. C., A. L., E. K. and C. Z. participated in the technical discussion and manuscript revision.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was funded by the Funding of NJUST (No. TSXK2022D002), Natural Science Foundation of Jiangsu Province (No. BK20220929), Natural Science Foundation of China (No. 22303039), and Startup Grant of NJUST.

References

  1. B. C. Steele and A. Heinzel, Nature, 2001, 414, 345–352 CrossRef CAS PubMed .
  2. M. Shao, Q. Chang, J.-P. Dodelet and R. Chenitz, Chem. Rev., 2016, 116, 3594–3657 CrossRef CAS PubMed .
  3. M. K. Debe, Nature, 2012, 486, 43–51 CrossRef CAS PubMed .
  4. J. Shui, M. Wang, F. Du and L. Dai, Sci. Adv., 2015, 1, e1400129 CrossRef PubMed .
  5. Y. Wang, H. Su, Y. He, L. Li, S. Zhu, H. Shen, P. Xie, X. Fu, G. Zhou, C. Feng, D. Zhao, F. Xiao, X. Zhu, Y. Zeng, M. Shao, S. Chen, G. Wu, J. Zeng and C. Wang, Chem. Rev., 2020, 120, 12217–12314 CrossRef CAS PubMed .
  6. X. Cui, W. Li, P. Ryabchuk, K. Junge and M. Beller, Nat. Catal., 2018, 1, 385–397 CrossRef CAS .
  7. W. H. Lee, Y.-J. Ko, J.-Y. Kim, B. K. Min, Y. J. Hwang and H.-S. Oh, Chem. Commun., 2020, 56, 12687–12697 RSC .
  8. H. Fei, J. Dong, D. Chen, T. Hu, X. Duan, I. Shakir, Y. Huang and X. Duan, Chem. Soc. Rev., 2019, 48, 5207–5241 RSC .
  9. Y. Wang, J. Mao, X. Meng, L. Yu, D. Deng and X. Bao, Chem. Rev., 2018, 119, 1806–1854 CrossRef PubMed .
  10. M. B. Gawande, P. Fornasiero and R. Zbořil, ACS Catal., 2020, 10, 2231–2259 CrossRef CAS .
  11. Y. Dai, F. Kong, X. Tai, Y. Zhang, B. Liu, J. Cai, X. Gong, Y. Xia, P. Guo and B. Liu, Electrochem. Energy Rev., 2022, 5, 22 CrossRef CAS .
  12. W. Song, C. Xiao, J. Ding, Z. Huang, X. Yang, T. Zhang, D. Mitlin and W. Hu, Adv. Mater., 2024, 36, 2301477 CrossRef CAS PubMed .
  13. S. Liu, C. Li, M. J. Zachman, Y. Zeng, H. Yu, B. Li, M. Wang, J. Braaten, J. Liu, H. M. Meyer III, M. Lucero, A. J. Kropf, E. E. Alp, Q. Gong, Q. Shi, Z. Feng, H. Xu, G. Wang, D. J. Myers, J. Xie, D. A. Cullen, S. Litser and G. Wu, Nat. Energy, 2022, 7, 652–663 CrossRef CAS .
  14. X. Zhang, L. Truong-Phuoc, T. Asset, S. Pronkin and C. Pham-Huu, ACS Catal., 2022, 12, 13853–13875 CrossRef CAS .
  15. H. Shen, T. Thomas, S. A. Rasaki, A. Saad, C. Hu, J. Wang and M. Yang, Electrochem. Energy Rev., 2019, 2, 252–276 CrossRef CAS .
  16. K. Song, B. Yang, X. Zou, W. Zhang and W. T. Zheng, Energy Environ. Sci., 2023, 17, 27–48 RSC .
  17. J. Li, J. Ma, Z. Ma, E. Zhao, K. Du, J. Guo and T. Ling, Adv. Energy Sustainability Res., 2021, 2, 2100034 CrossRef CAS .
  18. Z. Zhang, P. Ma, L. Luo, X. Ding, S. Zhou and J. Zeng, Angew. Chem., Int. Ed., 2023, 62, e202216837 CrossRef CAS PubMed .
  19. L. Scarpetta-Pizo, R. Venegas, P. Barrías, K. Muñoz-Becerra, N. Vilches-Labbé, F. Mura, A. M. Méndez-Torres, R. Ramírez-Tagle, A. Toro-Labbé, S. Hevia, J. H. Zagal, R. Oñate, A. Aspée and I. Ponce, Angew. Chem., Int. Ed., 2024, 136, e202315146 CrossRef .
  20. Y. Sun, S. Sun, H. Yang, S. Xi, J. Gracia and Z. J. Xu, Adv. Mater., 2020, 32, 2003297 CrossRef CAS PubMed .
  21. D. Xia, X. Yang, L. Xie, Y. Wei, W. Jiang, M. Dou, X. Li, J. Li, L. Gan and F. Kang, Adv. Funct. Mater., 2019, 29, 1906174 CrossRef CAS .
  22. G. Yang, J. Zhu, P. Yuan, Y. Hu, G. Qu, B.-A. Lu, X. Xue, H. Yin, W. Cheng, J. Cheng, W. Xu, J. Li, J. Hu, S. Wu and J.-N. Zhang, Nat. Commun., 2021, 12, 1734 CrossRef CAS PubMed .
  23. H. Zhang, H.-C. Chen, S. Feizpoor, L. Li, X. Zhang, X. Xu, Z. Zhuang, Z. Li, W. Hu, R. Snyders, D. Wang and C. Wang, Adv. Mater., 2024, 36, 2400523 CrossRef CAS PubMed .
  24. C. Chen, Y. Wu, X. Li, Y. Ye, Z. Li, Y. Zhou, J. Chen, M. Yang, F. Xie, Y. Jin, C. Jones, N. Wang, H. Meng and S. Chen, Appl. Catal., B, 2024, 342, 123407 CrossRef CAS .
  25. J. Hou, Y. Jian, C. Chen, D. Zhang, F. Xie, J. Chen, Y. Jin, N. Wang, X. Yu and H. Meng, Mater. Chem. Front., 2024, 8, 2592–2598 RSC .
  26. W. Zhong, Y. Qiu, H. Shen, X. Wang, J. Yuan, C. Jia, S. Bi and J. Jiang, J. Am. Chem. Soc., 2021, 143, 4405–4413 CrossRef CAS PubMed .
  27. X. Hu and N. Q. Su, J. Phys. Chem. Lett., 2023, 14, 9872–9882 CrossRef CAS PubMed .
  28. M. Yu, A. Li, E. Kan and C. Zhan, ACS Catal., 2024, 14, 6816–6826 CrossRef CAS .
  29. D. Benito-Garagorri, I. Lagoja, L. F. Veiros and K. A. Kirchner, Dalton Trans., 2011, 40, 4778–4792 RSC .
  30. X. Xin, C. Guo, R. Pang, X. Shi and Y. Zhao, Appl. Surf. Sci., 2021, 570, 151126 CrossRef CAS .
  31. R. Wang, L. Zhang, J. Shan, Y. Yang, J. F. Lee, T. Y. Chen, J. Mao, Y. Zhao, L. Yang, Z. Hu and L. Tao, Adv. Sci., 2022, 9, 2203917 CrossRef CAS PubMed .
  32. P. E. Blöchl, Phys. Rev. B, 1994, 50, 17953 CrossRef PubMed .
  33. G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS .
  34. G. Kresse and D. Joubert, Phys. Rev. B, 1999, 59, 1758 CrossRef CAS .
  35. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed .
  36. S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed .
  37. V. Wang, N. Xu, J.-C. Liu, G. Tang and W.-T. Geng, Comput. Phys. Commun., 2021, 267, 108033 CrossRef CAS .
  38. V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, J. Phys. Chem. A, 2011, 115, 5461–5466 CrossRef CAS PubMed .
  39. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, J. Comput. Chem., 2016, 37, 1030–1035 CrossRef CAS PubMed .
  40. G. Pizzi, V. Vitale, R. Arita, S. Blügel, F. Freimuth, G. Géranton, M. Gibertini, D. Gresch, C. Johnson, T. Koretsune, J. Ibañez-Azpiroz, H. Lee, J.-M. Lihm, D. Marchand, A. Marrazzo, Y. Mokrousov, J. I. Mustafa, Y. Nohara, L. Paulatto, S. Poncé, T. Ponweiser, J. Qiao, F. Thöle, S. S. Tsirkin, M. Wierzbowska, N. Marzari, D. Vanderbilt, I. Souza, A. A. Mostofi and J. R. Yates, J. Phys.: Condens. Matter, 2020, 32, 165902 Search PubMed .
  41. K. Mathew, V. Kolluru, S. Mula, S. N. Steinmann and R. G. Hennig, J. Chem. Phys., 2019, 151, 234101 CrossRef PubMed .
  42. K. Mathew, R. Sundararaman, K. Letchworth-Weaver, T. Arias and R. G. Hennig, J. Chem. Phys., 2014, 140, 084106 CrossRef PubMed .
  43. J. A. Gauthier, S. Ringe, C. F. Dickens, A. J. Garza, A. T. Bell, M. Head-Gordon, J. K. Nørskov and K. Chan, ACS Catal., 2018, 9, 920–931 CrossRef .
  44. Z. Duan and P. Xiao, J. Phys. Chem. C, 2021, 125, 15243–15250 CrossRef CAS .
  45. A. Hagopian, M.-L. Doublet, J.-S. Filhol and T. Binninger, J. Chem. Theory Comput., 2022, 18, 1883–1893 CrossRef CAS PubMed .
  46. Y. Huang, R. J. Nielsen and W. A. Goddard III, J. Am. Chem. Soc., 2018, 140, 16773–16782 CrossRef CAS PubMed .
  47. Y. Cui, C. Ren, Q. Li, C. Ling and J. Wang, J. Am. Chem. Soc., 2024, 15640–15647 CrossRef CAS PubMed .
  48. S. Zhu, L. Ding, X. Zhang, K. Wang, X. Wang, F. Yang and G. Han, Angew. Chem., Int. Ed., 2023, 62, e202309545 CrossRef CAS PubMed .
  49. L. Li, J. Zhu, F. Kong, Y. Wang, C. Kang, M. Xu, C. Du and G. Yin, Matter, 2024, 7, 1517–1532 CrossRef CAS .
  50. Z. Miao, X. Wang, Z. Zhao, W. Zuo, S. Chen, Z. Li, Y. He, J. Liang, F. Ma and H. L. Wang, Adv. Mater., 2021, 33, 2006613 CrossRef CAS PubMed .
  51. Z. Huang and Q. Tang, J. Phys. Chem. C, 2022, 126, 21606–21615 CrossRef CAS .
  52. F. Sun, F. Li and Q. Tang, J. Phys. Chem. C, 2022, 126, 13168–13181 CrossRef CAS .
  53. Y. Wang, Y.-J. Tang and K. Zhou, J. Am. Chem. Soc., 2019, 141, 14115–14119 CrossRef CAS PubMed .
  54. L. J. Frevel, R. Mom, J.-J. s. Velasco-Vélez, M. Plodinec, A. Knop-Gericke, R. Schlögl and T. E. Jones, J. Phys. Chem. C, 2019, 123, 9146–9152 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: Structural parameters; PDOS and electron occupation; COHP analysis and spin density; chemical reactions and free energy calculation; constant-potential free energy diagrams; occupation calculation. See DOI: https://doi.org/10.1039/d4ta04458d

This journal is © The Royal Society of Chemistry 2024