Yixin
Bi‡
a,
Yuhao
Wang‡
a,
Yufei
Song
d,
Qing
Chen
*a and
Francesco
Ciucci
*abc
aDepartment of Mechanical and Aerospace Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong S.A.R., China. E-mail: chenqing@ust.hk
bUniversity of Bayreuth, Chair of Electrode Design for Electrochemical Storage, Weiherstraße 26, 95448 Bayreuth, Germany. E-mail: francesco.ciucci@uni-bayreuth.de
cUniversity of Bayreuth, Bavarian Center for Battery Technology (BayBatt), Universitätsstraße 30, 95447 Bayreuth, Germany
dState Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing, 210009, China
First published on 4th June 2025
Developing cost-effective, efficient oxygen evolution reaction (OER) catalysts is critical for sustainable hydrogen production through water electrolysis. While noble metal-based catalysts like RuO2 and IrO2 show high activity, their widespread adoption is limited by cost. Fe-based perovskite oxides present a more abundant alternative but typically exhibit inferior OER activities. In this study, we achieved systematic dual-site modulation by incorporating Ba at the A-site and Ni at the B-site of NdFeO3−δ, transforming it into a double perovskite structure. The resulting Nd0.8Ba1.2Fe1.6Ni0.4O6−δ catalyst achieved an overpotential of 320 mV at 10 mA cm−2 in 0.1 M KOH, significantly lower than typical Fe-based perovskites and noble metals. Ab initio simulations revealed that A-site modulation reduces the band gap, which enhances electronic conductivity. Meanwhile, B-site Ni incorporation strengthens metal–oxygen covalency and decreases charge-transfer energy. The synergistic effects between enhanced electronic conductivity and metal–oxygen covalency led to a significantly reduced Tafel slope of 63.23 mV dec−1, compared to 114.85 mV dec−1 for single-site modified Nd0.8Ba1.2Fe2O6−δ and 154.34 mV dec−1 for unmodified NdFeO3−δ. This work provides a framework for understanding and improving performance in Fe-based perovskite OER catalysts through dual-site modulation, paving the way for more cost-effective and sustainable water electrolysis technologies.
Transition-metal (TM)-containing perovskite oxides are an alternative to noble metal-based catalysts as they enable strong electrocatalytic activity through flexible compositional adjustments.4,5 Co-based perovskites, particularly Ba0.5Sr0.5Co0.8Fe0.2O3−δ, have demonstrated exceptional catalytic performance.4,6,7 However, growing concerns about Co's toxicity and increasing demand in battery applications have intensified the search for safer, more abundant alternatives. Fe-based perovskites present a promising direction, given Fe's natural abundance, cost-effectiveness, and reduced environmental impact. Despite these advantages, their development as primary catalysts has been hampered by relatively low activity compared to their Co and Ni-based counterparts.8,9 Besides, Fe-based oxides are thermodynamically unstable in alkaline solutions under anodic potentials, resulting in poor OER stability.8 Therefore, Fe has primarily been limited to dopant roles in Co or Ni-based perovskites.
Researchers have pursued various optimization strategies to enhance the OER performance of Fe-based perovskites. The base compound NdFeO3−δ (NF) exhibits relatively weak catalytic activity, requiring an overpotential of 420 mV to achieve a current density of 10 mA cm−2 in 1 M KOH.10 Omari et al.10 demonstrated that introducing A-site deficiency in NF creates oxygen vacancies, reducing the required overpotential to 370 mV for Nd0.9FeO3−δ. Their findings highlight NF's potential for further optimization as an electrochemically active catalyst.
LaFeO3−δ, a material widely investigated for applications in the fields of sensors, photocatalysis, and solid oxide fuel cells,11 showed limited OER activity with a 510 mV overpotential at 10 mA cm−2 in 0.1 M KOH. She et al.11 enhanced its catalytic performance through A-site Sr substitution to make La1−xSrxFeO3−δ (x = 0, 0.2, 0.5, 0.8 and 1). This series of materials underwent a structural transformation from orthorhombic to cubic at x = 0.5, and this transformation was accompanied by a change in surface oxygen content. La0.2Sr0.8FeO3−δ had an optimal performance at a maximum surface O22−/O− content of 31.3%, achieving an overpotential of 370 mV at a current density of 10 mA cm−2.
BaFeO3−δ (BF) is not stable at room temperature, leading to the collapse of the cubic structure.12 Zhu et al.13 demonstrated that B-site Zr doping stabilized a cubic phase that promoted OER activity, achieving a 412 mV overpotential at 10 mA cm−2 in 0.1 M KOH. Zhao et al.14 leveraged the synergy between the perovskite and graphene support to make graphene-supported Ba0.95La0.05FeO3−δ, which achieved 23 mA cm−2 at 0.6 V versus mercury/mercury oxide electrode.
Furthermore, extensive research focused on Fe-based perovskites. Shang et al.15 utilized La0.5Sr0.5FeO3−δ as a foundational material, applying multi-element doping at the A-site to produce high-entropy Ba0.2Sr0.2La0.2Pr0.2Sm0.2FeO3−δ and medium-entropy Ba0.35Sr0.35La0.1Pr0.1Sm0.1FeO3−δ for use as cathode materials in SOFCs. Chen et al.16 doped Pr into the A-site of La0.5Sr0.5FeO3−δ, aiming to improve activity and tolerance to CO2 in SOFC cathodes. Inoue et al.17 incorporated Co as a B-site dopant in both Ba0.5Sr0.5FeO3−δ and LaFeO3 to exploit the synergistic effects of Co and Fe in catalyzing the OER. Omari et al.18 introduced A-site defects into GdFeO3 to enhance its OER catalytic performance. Chu et al.19 doped Pr at the A-site of Bi0.8Ca0.2FeO3−δ to boost activity and CO2 tolerance in SOFC cathodes.
Previous studies have primarily focused on enhancing performance through modifications of either the A-site or the B-site individually; however, the improvements from these single-site adjustments have been limited, resulting in suboptimal performance. This situation highlights the need for a more comprehensive and effective approach to materials optimization. Zhao et al.20 addressed this by exploring the potential of simultaneously adjusting both the A-site and B-site PrCoO3, which appeared to be a more promising strategy than single-site modulation. They implemented dual-site substitution in PrCoO3 by substituting Sr and Ru at the A and B sites, respectively. Their experimental results showed that Pr0.7Sr0.3Co1−xRuxO3 exhibited superior catalytic activity compared to both Pr0.7Sr0.3CoO3 and PrCoO3. However, a critical aspect missing from their analysis is a direct comparison between Pr0.7Sr0.3Co1−xRuxO3 and PrCo1−xRuxO3. This lack of direct comparison means that while the effectiveness of Ru substitution at the B site is evident, the individual contributions of A-site and B-site modifications remain unclear. Therefore, a comprehensive investigation into the effects of dual-site substitution is essential to quantify the individual and combined impacts on performance enhancement and to elucidate the underlying mechanisms.
In this study, we selected NF as the base compound due to its capability for compositional tuning and the specific properties of Nd atoms at the A-site, which allows for the adjustment of the double perovskite structure. We aimed to investigate dual-site modulation on NF to overcome the limitations of single-site modulation for OER catalysis. A site substitution of large diameter Ba in NF allows structure evolution from single perovskite to double perovskite. Furthermore, substituting low-valence Ni at the B site is expected to optimize the surface oxygen content and the electronic structure, further benefiting OER activity.
First-principles calculations validated the effectiveness of dual-site modification through electronic structure and bonding characteristics analysis. A-site Ba substitution reduced the band gap, enhancing electronic conductivity, while B-site Ni incorporation strengthened metal–oxygen covalency and reduced charge-transfer energy, creating optimal pathways for oxygen evolution. The optimized Nd0.8Ba1.2Fe1.6Ni0.4O6−δ (NBFN) catalyst demonstrated superior performance with a Tafel slope of 63.23 mV dec−1, significantly outperforming both single-site modified Nd0.8Ba1.2Fe2O6−δ (NBF) (114.85 mV dec−1) and unmodified NF (154.34 mV dec−1).
These findings establish dual-site modulation as an effective method for enhancing double perovskite catalysts, positioning NBFN among the most efficient Fe-based perovskite OER catalysts. Our work provides a rational framework for designing high-performance perovskite catalysts through precise electronic structure engineering.
To evaluate catalytic activity, we performed linear sweep voltammetry (LSV) by scanning from 0.2 to 1 V versus Ag/AgCl at 5 mV s−1 while maintaining the RDE rotation at 1600 rpm. All LSV curves were IR-corrected to account for solution resistance. Double-layer capacitance (Cdl) measurements employed cyclic voltammetry within a non-faradaic window of 0.2–0.3 V versus Ag/AgCl, using scan rates from 40 to 100 mV s−1. We calculated Cdl from the linear relationship between current density and scan rate at 0.25 V. Catalyst stability was assessed through chronopotentiometry at a constant current density of 10 mA cm−2.
High-resolution imaging and elemental mapping were conducted using a JEOL JEM 2010F transmission electron microscope (TEM) equipped with energy-dispersive X-ray spectroscopy (EDS). Surface morphology was examined using a JSM-6390 scanning electron microscope (SEM). We probed the electronic structure using soft X-ray absorption spectroscopy (sXAS) at the MCD-A and MCD-B beamlines of the National Synchrotron Radiation Laboratory (NSRL).
For OER pathway analysis, we constructed symmetric slab models based on experimentally observed (100) surfaces on NBF and NBFN, incorporating a 15 Å vacuum layer along the c direction to prevent periodic image interactions. The energetics of OER intermediates were evaluated using the Rossmeisl–Nørskov free energy framework,25 with Gibbs free energy changes calculated as follows:
ΔG = ΔE + ΔZPE − TΔS | (1) |
The O 2p band center energy has been used as a descriptor for OER activity in perovskites.27 Prior research by Zhao et al.28 established a direct correlation between the position of the O 2p band center and catalytic performance; the closer this band center is to the Fermi level, the higher the OER activity. This occurs because the closer the O 2p band center is to the Fermi level, the easier it becomes for lattice oxygen to participate in the OER process by undergoing oxidation. Our calculations show that modifying both A-site and B-site can move the O 2p band center closer to the Fermi level, with simultaneous modifications at both sites resulting in a higher O 2p band center compared to modifications at individual sites. The original NF displayed an O 2p band center of −3.74 eV, while modifications at the A-site raised it to −1.84 eV, and B-site modifications raised it to −2.14 eV. In contrast, simultaneous modifications at both the A-site and B-site increased the O 2p band center to −1.6 eV, highlighting the synergistic benefits of dual-site modifications. This indicates that the O 2p band center in NBFN is closest to the Fermi level, suggesting that NBFN is likely to exhibit superior catalytic performance.
In addition to analyzing the O 2p band center, we examined charge-transfer energy (Δ) as a complementary descriptor of OER activity. This parameter, defined as the energy difference between the O 2p band center and unoccupied transition metal 3d orbital, has also been shown to correlate with catalytic performance.29 A lower Δ facilitates easier electron transfer from oxygen to the transition metal centers during the OER process, thus enhancing activity. Our analysis of the Fe 3d–O 2p orbital interactions showed that the original NF had the highest ΔFe–O at 4.83 eV. In contrast, A-site adjustment (NBF) reduced ΔFe–O to 4.48 eV, and B-site adjustment (NFN) decreased it to 4.76 eV. Notably, simultaneous A/B-site adjustments resulted in an even lower ΔFe–O of 4.28 eV, making NBFN the most active catalyst for OER, which underscores the synergistic effect of A/B-site modifications. Furthermore, the investigation of Ni 3d–O 2p orbital interactions in NBFN showed a significantly lower charge-transfer energy (ΔNi–O = 2.87 eV), indicating that the Ni sites possess a more favorable electronic structure for OER in comparison to the Fe sites. For NFN, the ΔNi–O (3.36 eV) is also less than ΔFe–O (4.76 eV), suggesting that the electronic structure of Ni is optimized compared to that of Fe.
To investigate metal–oxygen bonding in NBFN, we analyzed the integrated projected density of states (IPDOS) for Fe and Ni d-orbitals. In the spin-up channel (Fig. 1a), Fe d-orbitals exhibited pronounced PDOS peaks below the Fermi level (0 eV), indicating complete orbital filling. IPDOS analysis confirmed this observation quantitatively (Fig. 1b), with electronic state density exceeding 0.5 at the Fermi level.30 In contrast, the spin-down (spin-dw) channel exhibited Fe d-orbital PDOS peaks above the Fermi level, with IPDOS occupancy below 0.5, demonstrating the presence of unoccupied states. This electronic configuration, characterized by five d-electrons in the spin-up channel and none in the spin-down channel, suggested that the valence state in NBFN is 3+. Conversely, Ni displayed different electronic occupation in NBFN (Fig. 1b). The Ni d-orbitals showed full occupancy in the spin-up channel, while the spin-down channel exhibited partial filling, with dxy, dyz, and dxz orbitals showing occupancy rates near 1. This electronic distribution, with five d-electrons in the spin-up channel and three in the spin-down channel, indicated that the valence state of Ni is 2+.
Analysis of TM–O covalency follows the methodology established by Jiang et al.,31 which quantifies orbital interactions through the energy difference between TM 3d and oxygen 2p band centers. In NBFN, this analysis revealed distinct bonding characteristics: the Fe–O interaction showed a band center difference of 4.587 eV, while the Ni–O interaction exhibited a smaller difference of 2.901 eV. The reduced energy separation for Ni–O indicated a stronger covalent character, enabling more efficient electron transfer between Ni sites and oxygen adsorbates.32 This enhanced Ni–O covalency identified Ni sites as more active for OER, which is consistent with our charge transfer energy calculations. These simulations suggested that by combining double perovskite formation with Ni substitution, we can reduce the band gap, shift the O 2p band center towards the Fermi energy, enhance metal–oxygen covalency, and minimize charge transfer energy (Fig. 1c). These electronic structure modifications work synergistically to enhance catalytic performance.
To understand the mechanistic basis of catalyst performance, we investigated the oxygen evolution reaction (OER) pathway using the adsorbate evolution mechanism (AEM) framework for perovskite oxides in alkaline conditions (Fig. 1d and e). The reaction proceeded through four elementary steps, each involving the formation and transformation of key oxygen-containing intermediates (*O, OH, and OOH, where * represents the active site). An analysis of all possible reaction pathways across Fe and Ni sites (structures detailed in Fig. S1 and S2†) identified the deprotonation of OH to O as the rate-determining step (RDS) for both NBF and NBFN catalysts. Notably, the energy barrier for this critical step was lower at the Ni sites in NBFN (1.01 eV) compared to the Fe sites in NBF (1.10 eV), suggesting optimized reaction pathways on the Ni sites. This observation regarding optimized reaction pathways on the Ni sites is consistent with the study of charge transfer energy and TM–O covalency. The lower RDS energy barrier at Ni sites in NBFN, as predicted by DFT, directly accounts for the superior OER activity of NBFN compared to NBF.
Rietveld refinement analysis (Fig. 2b) confirmed that NBFN crystallized in the cubic space group Pmm with lattice parameters a = b = c = 3.92 Å. The XRD pattern is similar to double perovskite NdBaFe2−xMnxO5+δ reported by Mao et al.33 Scanning transmission electron microscopy (STEM) and energy-dispersive X-ray (EDX) mapping (Fig. 2c) revealed uniform element distribution within individual NBFN particles, suggesting the successful dual-site substitution and homogeneity of the NBFN catalyst. High-resolution transmission electron microscopy (HR-TEM, Fig. 2d) showed a crystalline cubic structure with clear lattice fringes, with a lattice spacing of 0.398 nm, corresponding to (100) lattice planes from Rietveld refinement results. The combination of SEM-EDS and ICP results (Fig. S4 and Table S1†) confirmed that the composition of NBFN elements aligned with the intended stoichiometry. SEM images (Fig. S5†) showed that NF had a smaller grain size compared to BF, with NBF and NBFN having smaller grain size than BF.
We measured surface oxygen vacancies and transition metal valence state using X-ray photoelectron spectroscopy (XPS) and soft X-ray absorption spectroscopy (sXAS). Analysis of the O 1s XPS spectra (Fig. 3a and Table S2†) reveals that pristine NF had the highest lattice oxygen content. In contrast, both A-site modified NBF and B-site modified NFN exhibited decreased lattice oxygen levels, suggesting that adjustments at both sites promote the creation of oxygen vacancies. Interestingly, the NBFN with simultaneous A/B-site adjustments displayed the lowest lattice oxygen content, indicating that this combined modification is more effective for forming oxygen vacancies than modifications at individual sites. Oxygen vacancies significantly change how electrocatalysts interact with reactants (both in absorption and desorption) and also alter the bulk properties like electronic structure and conductivity, benefiting OER.34
A detailed analysis of the Fe 2p and Ni 2p XPS spectra (Fig. 3b and S6,† respectively), along with the corresponding peak deconvolution results (Tables S3†), provided insights into the oxidation states of the transition metals. The Fe ions in NBFN exhibited a valence state of +3, while the Ni ions displayed a valence state of +2.36, which aligns broadly with the trends predicted by our DFT calculations concerning electron occupancy in the Fe and Ni d-orbitals. It is significant that Ni demonstrated a lower oxidation state compared to Fe, aligning with previous findings.2,8 The addition of low valence Ni to Fe-based perovskites encourages the development of oxygen vacancies, which subsequently reduces the oxidation state of Fe. This occurrence is illustrated by the shifts from NF to NFN and from NBF to NBFN (Table S3†). Moreover, the presence of Ba has minimal impact on the oxidation states of B-site transition metals, mainly affecting the control of the crystal structure.
We employed soft X-ray absorption spectroscopy (sXAS) at the Fe and Ni L-edges to measure the electronic structures of NF, NBF, and NBFN catalysts and to confirm the XPS findings (Fig. 3c and d). The Fe L3-edge sXAS spectra (Fig. 3c) indicate that NBFN exhibited a characteristic peak shift toward lower energies compared to NBF, suggesting a reduction in the valence state of Fe, consistent with the XPS results. The similar peak positions of NF and NBF suggest that A-site modification had minimal effect on the oxidation state of Fe, which aligns with the XPS results. Analysis of the Ni L3 edge sXAS spectra (Fig. 3d) revealed a peak shift toward higher energies for NBFN in comparison to the NiO reference, suggesting that the Ni in NBFN had a valence state higher than +2. Furthermore, the higher peak intensity of NBFN compared to NiO at around 856 eV indicates the presence of Ni3+ (Fig. 3d).35 This observation aligns with XPS characterization and provides additional evidence for the mixed-valence character of Ni in NBFN (Fig. S6†).
We also examined the effects of varying the substitution levels of Ba at the A site and Ni at the B site. As shown in Fig. S7 and S8,† changes in A-site Ba concentration did not affect the double perovskite structure, and all series of double perovskites exhibited higher performance compared to the single perovskite NF and BF (0.81 mA cm−2 at 1.7 V). Additionally, our analysis of Ni concentration indicated that when the Ni:
Fe ratio is greater than 2
:
8, it results in the degradation of both the perovskite structure and catalytic performance. We also investigated other potential dopant levels such as NdxBa2−xFe1.6Ni0.4O6−δ (Fig. S9†) or NdxBa2−xFe1.2Ni0.8O6−δ (Fig. S10†) to see if NBFN is the optimized composition. The results show that there are no double perovskites with higher performance than the NBFN.
The electrochemical active surface area (ECSA) was quantified through double-layer capacitance (Cdl) analysis (Fig. 4c and S11†).36 Compared to the pristine NF, the NBF and NFN showed modest ECSA enhancements of 1.03 and 1.16 times, respectively. In contrast, NBFN displayed a substantial 1.74-fold increase in ECSA. The larger ECSA for NBFN suggested a greater number of electrochemically active sites available for OER, contributing to its improved performance.37,38
We assessed the stability through chronopotentiometry at a constant current density of 10 mA cm−2 (Fig. 4e). NF exhibited rapid potential increase after only 500 s at a current density of 10 mA cm−2, which agrees with the argument in ref. 8 about the thermodynamically unstable characteristics of Fe based oxides in alkaline OER conditions. In contrast, NBF demonstrated improved stability, functioning effectively for approximately 5 hours before failing, while NFN operated steadily for about 7.8 hours before ceasing to work. Notably, following an initial activation period, NBFN maintained a stable operating voltage of 1.55 V for over 20 hours. To confirm the robustness of NBFN, XRD pattern (Fig. S12†) and HR-TEM image (Fig. S13†) were obtained after stability testing, which showed no evidence of phase transformation. The XRD pattern did not detect any secondary phases, and the HR-TEM image indicated that the lattice structure remained stable, with no collapse or distortion, thereby demonstrating the robustness of NBFN. The performance of NBFN was benchmarked against state-of-the-art catalysts, including both transition metal perovskites and precious metal-based materials2,36,39–46 (Fig. 4d and Table S4†). This comparison showed that the overpotential of NBFN is significantly lower than that of other perovskite catalysts and even surpasses the precious metal benchmarks IrO2 and RuO2.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta01754h |
‡ Co-first authors. |
This journal is © The Royal Society of Chemistry 2025 |