Next Article in Journal
Multifunctional NADES-Based Extracts from Paeonia lactiflora Pall. Flowers for Potential Cosmetic and Pharmaceutical Applications
Previous Article in Journal
Cork By-Products as Bioactive Ingredients: From Waste Valorization to Pharmaceutical Prototypes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxygen Bridge Governs OER via Deep Self-Reconstruction in Fe–Co Oxyhydroxides

1
College of Chemistry and Chemical Engineering, China University of Petroleum (East China), Qingdao 266580, China
2
School of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252059, China
*
Authors to whom correspondence should be addressed.
Molecules 2026, 31(1), 96; https://doi.org/10.3390/molecules31010096 (registering DOI)
Submission received: 29 October 2025 / Revised: 3 December 2025 / Accepted: 19 December 2025 / Published: 25 December 2025
(This article belongs to the Special Issue Advanced Nanomaterials for Electrocatalysis)

Abstract

The oxygen evolution reaction (OER) in water splitting involves complex multi-electron–proton transfer processes and represents the rate-determining step limiting overall electrolysis efficiency. Developing non-noble-metal catalysts with high activity and stability is therefore essential. Herein, a heterogeneous synthesis strategy was employed to in situ construct an iron-rich layered sulfate precursor (Fe0.42Co0.58-SO4/NF) on nickel foam, which underwent deep self-reconstruction in alkaline electrolyte to form nanoflower-like Fe0.42Co0.58OOH/NF. The optimized catalyst maintained its iron-rich composition and hierarchical structure, delivering outstanding OER performance with an overpotential of 220 mV at 10 mA·cm−2, a Tafel slope of 31.9 mV·dec−1, and stability exceeding 12 h at 600 mA·cm−2. Synchrotron analyses revealed dynamic transitions between mono-μ-O and di-μ-O Fe–M (M = Fe, Co) oxygen bridges during reconstruction, which enhanced both structural robustness and active-site density. The Fe-rich environment promoted the formation of Fe3+–O–Fe3+ units that synergized with Co4+ species to activate the lattice oxygen mechanism (LOM), thereby accelerating OER kinetics. This work elucidates the key role of oxygen-bridge geometry in optimizing catalytic activity and durability, providing valuable insights into the rational design of Fe–Co-based non-noble-metal catalysts with high iron content for efficient water oxidation.

Graphical Abstract

1. Introduction

Hydrogen production via water splitting is one of the key pathways toward establishing a sustainable hydrogen energy system [1]. Among the half-reactions involved, the oxygen evolution reaction (OER) occurs at the anode and often serves as the bottleneck limiting the overall energy efficiency of water splitting, owing to its complex multi-step electron–proton transfer process and the high energy barrier for O–O bond formation [2]. Developing efficient, stable, and cost-effective electrocatalysts therefore remains a central challenge for improving OER performance. At present, although noble metal oxides such as IrO2 and RuO2 exhibit excellent catalytic activity, their high cost and limited availability severely restrict large-scale commercial applications [3].
Iron-group (Fe, Co, Ni) based oxides and layered double hydroxides (LDHs) have attracted extensive attention due to their outstanding electrocatalytic performance toward the OER in alkaline media, among which FeCo LDH materials are particularly representative [4]. Studies have shown that the appropriate incorporation of Fe not only induces the reconstruction of the crystal structure in Co-based hydroxides [5], but also generates Fe–M (M = Co, Fe) active centers with μ-O (oxo-bridge) structures, thereby providing more highly active sites and significantly enhancing the intrinsic catalytic activity. However, excessive Fe content tends to reduce the electrical conductivity of the material and aggravate its dissolution under alkaline conditions, thus compromising stability. Conversely, higher Co content helps maintain structural integrity but often at the expense of catalytic activity [6,7]. Therefore, achieving a reproducible and scalable synergistic balance between activity and stability—while overcoming the Fe/Co molar ratio limitation (≤0.33) caused by the intrinsic electrostatic repulsion between Fe3+–O–Fe3+ pairs—remains one of the key challenges in the rational design of FeCo LDH catalysts [8].
Beyond compositional optimization, understanding structural self-reconstruction under operating potentials and identifying the true active phase are crucial for elucidating reaction mechanisms and guiding rational catalyst design [7]. In situ Raman and synchrotron X-ray absorption spectroscopy (XAS) studies have revealed that, under anodic polarization, LDH precursors undergo progressive deintercalation or substitution of interlayer anions (e.g., SO42−, NO3, Cl), transforming into oxyhydroxide phases [8,9,10,11]. This phase evolution is often accompanied by lattice-oxygen activation, oxygen-vacancy formation, and dynamic reorganization of μ-oxo bridges, generating additional active sites. Notably, in situ XAS and electrochemical analyses have confirmed μ-oxo-bridged Fe–M (M = Co, Fe) units as key OER centers [12,13]. Ou et al. further demonstrated that adjacent Fe sites on transition metal (oxy)hydroxides act cooperatively during OER, where Fe-Fe (or Fe-M) coupling via bridging oxygen species facilitates access to high-valence Fe states and stabilizes critical reaction inter-mediates, leading to enhanced OER kinetics compared to isolated Fe sites [14]. Simultaneously, potential-driven Co3+ → Co4+ oxidation is accompanied by μ-oxo rearrangements, while Fe–Co μ-oxo motifs—through the strong Lewis acidity of Fe3+—modulate the Co3+/Co4+ redox potential, facilitate lattice-oxygen participation, and accelerate OER dynamics [15]. Therefore, precisely identifying the reconstruction-derived active phase and μ-oxo configurations is fundamental to advancing mechanistic understanding and enabling the rational design of efficient OER catalysts.
Nevertheless, previous studies have shown that electrochemical reconstruction typically occurs only within a few nanometers of the surface, forming an ultrathin active layer that is challenging to unambiguously resolve using conventional characterization techniques [16,17]. This shallow reconstruction limits the identification of the true active phase and hinders the establishment of reliable structure–performance correlations. Recently, however, several studies have revealed that, under specific precursor compositions and operating conditions, deep self-reconstruction can be induced, leading to a complete transformation into a uniform new phase. Such a process not only overcomes the limitations of surface-confined reconstruction but also provides a unique opportunity to directly probe the intrinsic active phase and elucidate its dynamic evolution [18].
Against this background, we construct an Fe-rich layered sulfate precursor in situ (Fe0.42Co0.58-SO4) on nickel foam and subsequently induce deep self-reconstruction under alkaline galvanostatic/potentiostatic conditions to yield Fe0.42Co0.58OOH/NF while preserving the nanoflower morphology. In situ spectroscopic and electrochemical analyses reveal that, during reconstruction, Fe–M (M = Fe, Co) oxygen-bridge motifs dynamically interconvert between mono-μ-O and di-μ-O configurations, which stabilize the framework and substantially increase the number of accessible active sites. Remarkably, Raman spectra confirm the abundant formation of Fe3+–O–Fe3+ linkages under Fe-rich conditions, exceeding the commonly cited Fe/Co compositional threshold of ≤0.33. Further investigation demonstrates that these Fe-rich bridging motifs cooperate with potential-induced Co4+ species to activate the lattice oxygen mechanism (LOM), thereby accelerating OER kinetics. Overall, this study underscores the pivotal role of oxygen-bridge geometry in simultaneously enhancing catalytic activity and stability, offering a new strategy for the rational design of Fe-rich, FeCo-based non-noble OER catalysts.

2. Results and Discussion

2.1. Structural Characterization of the Precursor Fe0.42Co0.58-SO4/NF

The precursor of Fe0.42Co0.58-SO4/NF was prepared by dissolving Co(NO3)2·6H2O and FeSO4·7H2O at a molar ratio of 10:1 in different solvents. Owing to the solubility difference between the two salts, Fe–Co nanoparticles were in situ-deposited onto the nickel foam (NF) surface, forming a nanoflower-like morphology. As observed in the SEM images (Figure 1a,b and Figure S1), the FeCo–SO4 nanosheets self-assembled into well-defined nanoflowers on NF. TEM analysis (Figure 1c) revealed that each nanoflower consisted of ultrathin nanosheet subunits, while the HRTEM image (Figure 1d and Figure S2) displayed lattice fringes with a spacing of 0.259 nm, corresponding to the (104) plane of layered double hydroxide (LDH) [19]. Furthermore, EDS mapping (Figure 1e–i, Table S1) confirmed the homogeneous distribution of Fe, Co, O, and S elements throughout the structure, with sulfur accounting for approximately 4 at%.
X-ray diffraction (XRD) analysis (Figure S3) of the precursor exhibited characteristic diffraction peaks in the 10–70° range. The broad peaks indicated relatively low crystallinity, and several reflections matched well with the standard card of poorly crystalline FeCo-LDH (PDF#00-050-0235). Notably, the FeCo-LDH pattern exhibited a low-angle (001) reflection at approximately 10°, while in Fe0.42Co0.58-SO4/NF, the presence of large interlayer SO42− ions resulted in a weaker diffraction peak at 8.1° [20]. Inductively coupled plasma atomic emission spectroscopy (ICP-AES, Figure 1j) further confirmed that the Co:Fe molar ratio in Fe0.42Co0.58-SO4/NF was approximately 3:2.
To further verify the chemical structure of the precursor, X-ray absorption fine structure spectroscopy (XAFS) was employed. The Fe K-edge X-ray absorption near-edge structure spectrum (XANES) of Fe0.42Co0.58-SO4/NF showed an absorption edge position close to that of Fe2O3, indicating that Fe predominantly existed in the +3 oxidation state (Figure 1k). The chemical state of cobalt was examined by X-ray photoelectron spectroscopy (XPS, Figure S4). The Co 2p spectrum exhibited clear spin–orbit splitting into Co 2p3/2 and Co 2p1/2 components at binding energies of 780.5 eV and 796.6 eV, respectively. Despite some peak overlap, qualitative analysis of the oxidation states was still possible. The Co 2p spectrum of Fe0.42Co0.58-SO4/NF indicated that cobalt predominantly existed as Co2+, as evidenced by the main peaks and the presence of shake-up satellite (“Sat”) peaks. Additional weak satellite features at 787.0 eV and 803.2 eV further confirmed the typical spectral characteristics of Co2+, consistent with previous reports [21,22,23].

2.2. Composition and Local Structure of Self-Reconstructed Fe0.42Co0.58OOH/NF

The precursor Fe0.42Co0.58-SO4/NF was electro-oxidized in 1 M KOH within a potential range of 0.87–1.27 V (vs. RHE). As shown in the cyclic voltammetry (CV) curves (Figure 2a), the electrochemically induced surface reconstruction gradually stabilized after ten activation cycles. SEM images confirmed that the nanoflower-like morphology was well preserved throughout the process (Figure 2b,c and Figure S5). Meanwhile, SEM-EDS analysis revealed a drastic decrease in sulfur content to only 0.06%, suggesting substantial structural reconstruction accompanied by electron redistribution during the OER process (Figure 2d–h and Table S2). To further probe the local environment and oxidation state of Fe, Fe K-edge XAFS analysis was performed. The XANES of Fe0.42Co0.58OOH/NF exhibited an absorption edge position comparable to that of FeOOH (Figure 1k), indicating that Fe mainly existed in the +3 oxidation state. Consistently, the EXAFS spectrum also closely resembled that of FeOOH (Figure 3a), corroborating the formation of a Fe3+-dominated oxyhydroxide phase. Additionally, ICP–AES showed negligible leaching of Fe and Co, and the final Co/Fe molar ratio remained nearly identical to that of the precursor (Figure 3a), confirming the successful formation of Fe0.42Co0.58OOH/NF.
The local structure of Fe was further probed by Fe K-edge XAFS. Previous studies have shown that Fe–O–M (M = Fe, Co) linkages can involve not only stable di-μ-O (edge-sharing) bridges but also a fraction of mono-μ-O (corner-sharing) bridges formed during electrochemical self-reconstruction [12,24,25,26]. The di-μ-O bridges, originating from two shared corner oxygen atoms between MO6 octahedra, help maintain the lattice stability, whereas mono-μ-O bridges, connected by a single oxygen atom, more readily induce local coordination unsaturation and oxygen vacancy formation, thus facilitating lattice oxygen participation in OER. Their cooperative effect enables FeMOOH (M = Ni, Co) to achieve both structural robustness and high catalytic activity. The EXAFS spectrum exhibited a weak peak at 3.2 Å for FeOOH, Fe2O3, and Fe0.42Co0.58OOH/NF (Figure 3a), which was absent in FeO and the precursor Fe0.42Co0.58-SO4/NF, and was attributed to Fe–M linkages coordinated by mono-μ-O(H) [24].
Further structural refinement using Fourier-transformed EXAFS and fitting analysis (Figure 3d,e, Table S3) revealed good consistency between experimental and modeled spectra in both K-space and R-space (Figure S6), confirming the reliability of the fitted local structural model. Compared with the precursor, the Fe–O bond length in Fe0.42Co0.58OOH/NF shortened from 2.00 Å to 1.96 Å, reflecting enhanced M–O covalency under operating potentials. The coordination number increased from 5 to nearly 6, suggesting a more complete FeO6 octahedron. Meanwhile, the Fe–O–M distance increased from 3.06 Å to 3.25 Å with a decrease in coordination number from 6 to 5, implying local structural defects. This was further corroborated by electron paramagnetic resonance (EPR, Figure 3c), which showed an increased intensity of oxygen vacancy (Vo) signal at g = 2.003 [27]. Another notable observation was that the Fe–M (2.8–3.0 Å) di-μ-O contribution weakened and split in the EXAFS fitting, while new signals appeared at 3.3–3.5 Å, corresponding to mono-μ-O(H) FeO6 units [24]. Wavelet transform EXAFS (WT-EXAFS, Figure 3f,g) showed that the precursor Fe0.42Co0.58-SO4/NF featured pronounced Fe–O first-shell coordination and concentrated metal second-shell signals, along with weak long-range features [28]. After reconstruction, Fe0.42Co0.58OOH/NF retained the Fe–O nearest-neighbor shell, while the second-shell metal-related peaks broadened and attenuated at higher R, consistent with anion deintercalation and deep self-reconstruction from LDH to OOH [29].
Previous studies have shown that strong electrostatic repulsion inhibits the formation of Fe3+–O–Fe3+ bridge units, typically limiting the Fe/Co ratio in FexCo1−xOOH to no more than 1/2 (0 ≤ x ≤ 0.33) [30]. Interestingly, the Fe0.42Co0.58OOH/NF synthesized in this work exceeded this conventional Fe content limit. Raman spectroscopy of the precursor revealed characteristic Co–O vibrational modes (A1g and Eg) (Figure 3b) [31,32]. Upon reconstruction, Fe0.42Co0.58OOH/NF displayed a new peak at ~533 cm−1, corresponding to Fe3+–O–Fe3+ bridging vibrations [24]. This observation confirms that the elevated Fe content facilitated the formation of abundant Fe3+–O–Fe3+ structural units in the reconstructed catalyst.

2.3. Oxygen Evolution Reaction Performance

To directly verify the occurrence of the oxygen evolution reaction, we employed in situ differential electrochemical mass spectrometry (DEMS) to monitor gas products in real time during the catalytic process (Figure S7). The experiment was conducted under an argon atmosphere, beginning with a 10-min baseline acquisition to eliminate interference from dissolved oxygen and residual air. Upon initiating linear sweep voltammetry at 600 s, a pronounced increase in signal intensity at m/z = 32 was detected by DEMS, corresponding to the applied potential. This directly confirms the generation of oxygen on the working electrode surface, unequivocally demonstrating the occurrence of the oxygen evolution reaction and thereby providing direct evidence for the catalytic process. To further evaluate the electrocatalytic performance, Fe0.26Co0.74OOH/NF with a lower Fe content and FeOOH/NF without Co were synthesized on nickel foam (NF) as control samples using the same procedure (Figures S8 and S9). The OER activity of these catalysts was investigated in 1.0 M KOH electrolyte. The small redox peaks observed between 1.25 and 1.40 V in the inset of Figure 4a could be ascribed to the Co3+/Co4+ transition [12]. Therefore, linear sweep voltammetry (LSV) was recorded in the reverse scan mode to avoid interference.
As shown in the iR-corrected polarization curves (Figure 4a), Fe0.42Co0.58OOH/NF exhibited the lowest overpotential of only 220 mV at 10 mA·cm−2, which was significantly lower than Fe0.26Co0.74OOH/NF (280 mV), FeOOH/NF (279 mV), bare NF (370 mV), and even IrO2/NF (291 mV) [24], demonstrating its outstanding OER activity. It is worth noting that FeOOH/NF performed reasonably well at low current densities but showed a rapid decline in catalytic activity under high currents, which is consistent with the poor conductivity of FeOOH at elevated currents [6].
The Tafel slope, derived from LSV curves (Figure 4b), was used to further investigate the reaction kinetics. Fe0.42Co0.58OOH/NF exhibited the lowest Tafel slope (31.9 mV·dec−1), indicating more favorable kinetics compared to other catalysts. The electrochemical double-layer capacitance (Cdl) was extracted from CVs recorded at different scan rates (Figure S10), and the electrochemical surface area (ECSA) was calculated accordingly (Figure 4e). The ECSA of Fe0.42Co0.58OOH/NF reached as high as 89.5 cm2, which was significantly larger than Fe0.26Co0.74OOH/NF (47.75 cm2) and bare NF (27.75 cm2), suggesting the provision of more accessible active sites. Electrochemical impedance spectroscopy (EIS) was employed to evaluate the interfacial charge-transfer ability. The Nyquist plots revealed that Fe0.42Co0.58OOH/NF exhibited a much smaller charge-transfer resistance (Rct) compared to the control samples (Figure 4d), demonstrating its superior interfacial charge transport capability. Based on quantitative gas measurements, we calculated the turnover number (TON) of the as-prepared catalysts. As shown in Table S4, the Fe0.42Co0.58OOH/NF exhibited the highest TON value of 1.31, while the reference samples Fe0.26Co0.74OOH/NF and FeOOH/NF showed TON values of 1.14 and 1.05, respectively. All catalysts demonstrated TON values exceeding unity, confirming their intrinsic catalytic nature, with the TON trend aligning consistently with their electrochemical activity. Moreover, Table S4 compares the over-potentials and Tafel slopes of high-performance Co-based catalysts reported in previous studies, highlighting that Fe0.42Co0.58OOH/NF remains at the forefront.
The durability of the catalysts was further evaluated through long-term electrochemical testing. As shown in Figure 4c, Fe0.42Co0.58OOH/NF maintained stable activity after 1000 continuous CV cycles. Furthermore, chronopotentiometric testing at a high current density of 600 mA·cm−2 confirmed that Fe0.42Co0.58OOH/NF retained its excellent catalytic activity for over 12.6 h of continuous operation (Figure 4f). Taken together, these results demonstrate that Fe0.42Co0.58OOH/NF combines outstanding catalytic activity with robust structural stability.

2.4. Mechanistic Investigation of the OER

The intrinsic relationship among the structural, chemical, and kinetic properties of Fe–M (M = Fe/Co) oxygen-bridged motifs was further examined to clarify the origin of the OER activity and catalytic mechanism of Fe0.42Co0.58OOH/NF (Figure 5a). In general, the OER can proceed through two possible pathways: the conventional adsorbate evolution mechanism (AEM) or the lattice oxygen oxidation mechanism (LOM). As shown in Figure S11, the LOM pathway involves O–O bond formation via the coupling of lattice oxygen with adsorbed oxygen species, with peroxide intermediates (O22−) serving as characteristic markers. Tetramethylammonium cations (TMA+) are known to interact with O22− species and thereby suppress the LOM process [33]. Accordingly, tetramethylammonium hydroxide (TMAOH) was employed in this study as a chemical probe to discern the dominant reaction pathway. When OER was carried out in 1 M TMAOH electrolyte (Figure 5b), the catalytic activity of Fe0.42Co0.58OOH/NF decreased markedly compared with that in 1 M KOH, with the overpotential at 10 mA cm−210) increasing from 270 to 300 mV. This inhibition effect confirms that the OER on Fe0.42Co0.58OOH/NF proceeds primarily via the LOM pathway.
Further structural and defect-related evidence revealed the strong correlation between LOM promotion and Fe–M oxygen-bridged structures. Electron paramagnetic resonance (EPR) showed that the oxygen vacancy (Vo) concentration in Fe0.42Co0.58OOH/NF was significantly higher than that of its precursor Fe0.42Co0.58-SO4/NF (Figure 3c). EXAFS fitting showed an increased Fe–O coordination number coupled with a decreased Fe–O–M (M = Fe/Co) coordination number, indicating partial deconstruction of the octahedral framework and the formation of additional boundary and unsaturated sites. This observation aligns with the EXAFS evidence that some di-μ-O bridges dissociate and convert into mono-μ-O bridges. As discussed in Section 3.2, di-μ-O motifs help preserve lattice stability, whereas mono-μ-O motifs are more prone to inducing oxygen vacancy formation and lattice oxygen activation, thereby triggering the LOM pathway. Thus, their dynamic inter-conversion is considered a critical factor that enables the catalyst to simultaneously maintain high stability and high activity under large current densities.
The structural evolution governs the resulting chemical properties, while changes in valence states and surface adsorption behaviors further support this correlation. Previous studies have shown that Fe-participated di-μ-O Fe–Co bridges play a crucial role in promoting the formation of high-valence Co species. Meanwhile, the in situ generated Co4+ species during the electrochemical process are widely recognized as the primary active centers [34]. The linear sweep voltammetry (LSV) curve exhibited a pronounced reduction peak in the range of 1.2–1.35 V, indicating the participation of Co4+-based reactive oxygen species (e.g., Co(VI), CoO2) (Figure 4a). Moreover, CV comparisons revealed that the Fe3+/Fe4+ transition potential of FeOOH/NF was around 1.45 V (Figure S12), consistent with previous literature [35], whereas in Fe0.26Co0.74OOH/NF and Fe0.42Co0.58OOH/NF, the Co3+/Co4+ transition potentials decreased to 1.05 V and 0.9 V, respectively, much lower than the reported value for pure CoOOH (≈1.25 V) [35]. In addition, the peak area of Fe0.42Co0.58OOH/NF increased markedly, indicating that Fe–Co interactions and the presence of (di-μ-O) Fe–Co bridges not only lower the Co3+/Co4+ transition potential but also enhance the generation of active Co4+ species. Nocera et al. proposed that the strong Lewis acidity of Fe3+ can reduce the energy barrier for the Ni3+/Ni4+ transition [36]; a similar effect has been observed in Co-based oxyhydroxides [12]. Therefore, Fe and Co act cooperatively through (di-μ-O) bridging to reduce the activation barrier for Co to reach higher oxidation states, thereby facilitating the generation of more Co4+ species at lower potentials. This accelerates the LOM pathway and enhances the overall OER kinetics.
At the adsorption level, density functional theory (DFT) calculations (Figure 5e, Figures S14 and S15) indicate that, under conditions where the O2 desorption energy is similar, OH preferentially adsorbs at the Fe3+–O–Fe3+ sites rather than the Fe–O–Co sites. This suggests that the Fe3+–O–Fe3+ oxygen bridge effectively adsorbs and stabilizes OH, and its higher electron density and strong electron-accepting ability contribute to accelerating the oxygen evolution reaction (OER). This phenomenon further demonstrates that the Fe3+–O–Fe3+ oxygen bridge enhances the catalyst’s reaction activity by promoting the adsorption and activation of intermediates during the catalytic process.
From a kinetic standpoint, the pH-dependent activities of the three samples revealed that the overall catalytic performance increased with rising OH concentration within the pH range of 13.45–14.00 (Figure 5c and Figure S13). Combined with the current–pH response and Tafel slope analysis, the apparent reaction orders of FeOOH/NF and Fe0.26Co0.74OOH/NF were determined to be 0.89 and 0.86, respectively (Figure 5d). These values indicate that their catalytic activities are relatively insensitive to OH concentration and tend toward lattice oxygen depletion during the reaction. In contrast, Fe0.42Co0.58OOH/NF exhibited an apparent reaction order close to 1, demonstrating an approximately first-order dependence of current on OH concentration. This behavior suggests that lattice oxygen consumption was efficiently compensated by OH replenishment from the electrolyte, thereby enhancing its kinetic stability. Overall, these findings highlight that an optimized Fe/Co ratio can effectively regulate reaction kinetics, achieving a balance between catalytic activity and stability through pronounced synergistic effects.
In summary, Fe–M oxygen-bridged motifs play a pivotal role in bond formation and the rate-determining step by modulating defect states and the local electronic/valence environment, thereby synergistically optimizing the adsorption of *OH intermediates. This structural tuning lowers the energy barrier of the rate-limiting step and enhances apparent kinetics, all while preserving lattice stability. Consequently, Fe–M oxygen bridges serve as the core contributors to the exceptional OER performance of this system.

3. Experimental Section

3.1. Materials and Chemicals

The reagents and materials used in this study include: potassium hydroxide (KOH, analytical grade, purity 95%, Macklin Reagent Co., Ltd., Shanghai, China); cobalt nitrate hexahydrate (Co(NO3)2·6H2O, reagent grade, purity 99%, Macklin Reagent Co., Ltd. Shanghai, China); ferrous sulfate heptahydrate (FeSO4·7H2O, analytical grade, Shanghai Hushi Reagent Co., Ltd., Shanghai, China); tetramethylammonium hydroxide pentahydrate (TMAOH·5H2O, purity 97%, Aladdin Reagent Co., Ltd., Shanghai, China); anhydrous sodium sulfate (Na2SO4, analytical grade, Shanghai Hushi Reagent Co., Ltd., Shanghai, China); nickel foam (pore density 95 ppi, Kunshan Xingzhenghong Electronic Materials Co., Ltd., Kunshan, China). All chemicals were used as received without further purification. Ultrapure water (18.2 MΩ·cm) was used in all experiments.

3.2. Catalyst Preparation

3.2.1. Synthesis of Fe0.42Co0.58OOH/NF

As shown in Scheme 1, first, 5.093 g of Co(NO3)2·6H2O was dissolved in 60 mL of isopropanol to form a homogeneous solution; then 0.4875 g of FeSO4·7H2O was dissolved in 20 mL of deionized water. Under vigorous stirring, the iron salt solution was slowly dropped into the cobalt salt solution, generating a stable nanoparticle dispersion. Pre-cleaned nickel foam substrates (3 × 3 cm2, treated by ultrasonic cleaning in HCl–ethanol solution and dried) were immersed completely into the mixed solution and left at room temperature for 24 h. The obtained product was thoroughly rinsed with deionized water, followed by drying in a vacuum oven at 30 °C for 12 h, affording the precursor Fe0.42Co0.58-SO4/NF. Electrochemical activation was performed in 1 M KOH electrolyte by subjecting the precursor to 10 cyclic voltammetry (CV) scans within 0.87–1.27 V (vs. RHE), using a graphite rod as the counter electrode, ultimately yielding the target catalyst Fe0.42Co0.58OOH/NF.

3.2.2. Synthesis of Fe0.26Co0.74OOH/NF

The synthesis procedure was essentially identical to that of Fe0.26Co0.74OOH/NF, except for the precursor composition: 10.053 g of Co(NO3)2·6H2O was dissolved in 60 mL of isopropanol, while the Fe precursor remained unchanged at 0.4875 g of FeSO4·7H2O in 20 mL of water. The subsequent mixing, substrate immersion, washing, drying, and electrochemical activation steps were performed in the same way, yielding Fe0.26Co0.74OOH/NF.

3.2.3. Synthesis of FeOOH/NF

In this case, the synthesis followed the same procedure as Fe0.42Co0.58OOH/NF, except that 5.093 g of Co(NO3)2·6H2O was replaced with 1.487 g of NaNO3 (molar equivalent to Co(NO3)2·6H2O), dissolved in 60 mL of isopropanol. Meanwhile, 0.4875 g of FeSO4·7H2O was dissolved in 20 mL of water and slowly added dropwise under stirring to form a uniform mixture. The pre-treated nickel foam substrate was immersed, allowed to react at room temperature, rinsed, dried, and then subjected to identical CV activation in 1 M KOH to yield FeOOH/NF.

3.3. Catalyst Characterization

The morphology of the catalysts was examined using field-emission scanning electron microscopy (FE-SEM, JSM-6330F, ZESSI sigma300, Jena, Germany) and transmission electron microscopy (TEM, JEM-2010HR, JEOL, Tokyo, Japan). The crystallographic structure of the materials was analyzed by X-ray diffraction (XRD, Rigaku D/max 2500/PC, smartlabSE, Tokyo, Japan), while local structural information was obtained from Raman spectroscopy (Renishaw, inVia, Thermo Scientific, Waltham, MA, USA). The chemical states of elements were determined by X-ray photoelectron spectroscopy (XPS, ESCA KAB 250, Thermo Fisher Scientific, K-Alpha, Waltham, MA, USA), with the C 1s peak calibrated to 284.8 eV. The metal loading and elemental ratios were quantified by inductively coupled plasma atomic emission spectroscopy (ICP-AES, TJA IRIS (HR), Agilent 720ES, Santa Clara, CA, USA). X-ray absorption fine structure (XAFS) measurements were performed at the 1W1B beamline of the Singapore Synchrotron Light Source (SSLS).

3.4. Electrochemical Measurements

Electrochemical measurements were carried out on a DH7003 electrochemical workstation (Donghua Instruments, Jingjiang, China). A graphite rod and a Hg/HgO electrode were used as the counter and reference electrodes, respectively. All potentials were converted to the reversible hydrogen electrode (RHE) scale according to the following equation:
ERHE = EHg/HgO + 0.098 + 0.059 × pH
Prior to measurements, the electrolyte was saturated with O2, and all electrochemical data were corrected for 90%iR compensation. The ECSA could be estimated by measuring the non-Faradaic capacitive current associated with double-layer charging from the scan-rate dependence of cyclic voltammetry (CV), according to below Equation.
E C S A = C d l C s
where Cdl is the double-layer capacitance, and Cs is the specific capacitance of any investigated electrode material. Here we used general specific capacities of Cs = 0.040 mF·cm−2 in 1 M KOH for estimating our surface area. Cdl could be extracted from recording CVs at various scan rates. In specific, the potential range of the CV measurements was selected between 0.60 and 0.70 V (versus RHE), in order to avoid the Faradaic and redox processes taking place. Also, the scan rates varied from 0.02, 0.04, 0.06, 0.08, 0.1 mV·s−1. These tests were all conducted in a static electrolyte. Here, the slope of the resulting charging current vs. scan rate plot was approximately regarded as Cdl.

3.5. pH-Dependent Electrochemical Measurements

To investigate the effect of alkalinity on catalytic behavior, linear sweep voltammetry (LSV) was performed in KOH electrolytes with concentrations ranging from 0.1 to 5.0 M. The polarization curves obtained at different pH values were analyzed in the kinetic region, where the E–log j plots were linearly fitted to extract the corresponding Tafel slopes. The current densities at fixed potentials were further plotted as log j versus pH, and the apparent reaction order with respect to OH concentration was calculated using the Tafel slope, thereby quantitatively describing the dependence of current on OH activity.

3.6. TurnOver Number (TON)

The intrinsic activity of the catalyst was evaluated through TON calculation by performing chronopotentiometry at a constant current density of 600 mA·cm−2 for 60 s in a sealed two-compartment electrochemical cell, where the evolved gases were collected via a water-filled syringe connected to the anode outlet. The oxygen volume was determined from the total collected gas volume based on the water electrolysis stoichiometric ratio (H2:O2 = 2:1), and the moles of oxygen produced were calculated using the ideal gas law.
TON was then obtained using the equation:
TON = (nO2 × NA)/nmetal
where nO2 is the moles of oxygen produced, NA is Avogadro’s constant (6.022 × 1023 mol−1), and nmetal represents the total moles of metal atoms (Co + Fe) on the working electrode, determined from the catalyst loading mass (2.83 mg) and its chemical composition.

3.7. Computational Methods

All quantum chemical calculations were performed using the CP2K v2024.3 package, with geometric optimization and energy calculations conducted via the xTB-GFN1 method under the Quickstep module. The Ewald summation was enabled to handle the electrostatic interactions of the periodic system; the net charge of the system was set to 0 and the spin multiplicity to 1. Meanwhile, geometric structure minimization optimization was carried out using the BFGS optimizer (with a trust radius of 0.2 Å and a maximum number of iterations of 500), and the convergence criteria were defined as follows: maximum atomic displacement ≤ 3.0 × 10−3 Å, root-mean-square (RMS) atomic displacement ≤ 1.5 × 10−3 Å, maximum force ≤ 4.5 × 10−4 a.u., and RMS force ≤ 3.0 × 10−4 a.u. Additionally, fixed constraints were applied to the XYZ Cartesian coordinates of some atoms. For the self-consistent field (SCF) calculations, the orbital transformation (OT) algorithm was adopted, with the maximum number of inner loop steps set to 25 and the convergence threshold to 1.0 × 10−6, as well as the maximum number of outer loop steps set to 20 and the convergence threshold to 1.0 × 10−6. Structural trajectories were output in XYZ format for subsequent analysis.

4. Conclusions

This study highlights the pivotal role of Fe–M oxygen-bridged motifs in the OER process. During the electrochemical self-reconstruction process, dynamic interconversion between mono-μ-O and di-μ-O bridging units occurs, which not only provides abundant active sites but also preserves the structural stability. In addition, Fe3+–O–Fe3+ motifs generated under the Fe-rich environment act synergistically with high-valence Co species to promote the LOM, thereby accelerating the overall reaction kinetics. Our findings emphasize the importance of oxygen bridges in the synergistic regulation of catalytic activity and stability, providing structural insights for the rational design of Fe-rich Fe–Co-based catalysts. Future studies combining multi-scale simulations with advanced in situ characterization are expected to further elucidate the evolution of oxygen-bridged structures and explore their stability mechanisms under long-term operation and large-scale applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules31010096/s1, Figure S1: Fe0.42Co0.58-SO4/NF SEM. Figure S2: High-Resolution Lattice Spacing Distribution of Fe0.42Co0.58-SO4/NF. Figure S3: XRD pattern of Fe0.42Co0.58-SO4/NF. Figure S4: The XPS spectra of Co 2p of Fe0.42Co0.58-SO4/NF. Figure S5: The SEM of Fe0.42Co0.58OOH/NF. Figure S6: K-space (photoelectron wave vector space). Figure S7. In situ DEMS monitoring of the oxygen evolution reaction over the Fe0.42Co0.58OOH/NF catalyst (m/z = 32). Figure S8: Fe0.26Co0.74OOH/NF SEM. Figure S9: FeOOH/NF SEM. Figure S10: CVs at various scan rates Figure S11: Schematic illustration of different reaction pathways for the oxygen evolution reaction (LOM and AEM). Figure S12: CV tests of different catalysts in the redox potential region. Figure S13: (a) Fe0.26Co0.74OOH (b)FeOOH pH-Dependent test. Figure S14: Adsorption energy of OH intermediate at different sites (a) Fe-O-Co(b) Fe-O-Fe. Figure S15: Adsorption energy of O2 at different sites (a) Fe-O-Co(b) Fe-O-Fe. Table S1: Elemental Composition of Sample Fe0.42Co0.58-SO4/NF (SEM Mapping). Table S2: Elemental Composition of Sample Fe0.42Co0.58OOH/NF (SEM Mapping) Table S3: EXAFS fitting parameters at the Fe K-edge various samples (S02 = 0.70). Table S4. Turnover number (TON) calculations of the catalysts based on quantitative gas measurements. Table S5: Comparison of OER activities of other catalysts in KOH solution. References [37,38,39,40,41,42,43,44,45,46] are cited in the Supplementary Materials.

Author Contributions

Writing—original draft preparation, M.L.; Software, formal analysis, B.P.; Validation and data curation, H.B. and W.N.; Formal analysis, H.Z.; Writing—review and editing, S.C., J.Z. (Jiamin Zhao) and J.Z. (Jinsheng Zhao). All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Natural Science Foundation of China (22402072, 22172069), Natural Science Foundation of Shandong Province (ZR2023QB025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors. This study does not involve humans, and thus no informed consent from subjects is required.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Wu, L.; Wang, Q.; Yuan, S.; Mei, X.; Wang, Q.; Zou, X.; Zhang, K.; Huo, X.; Shi, X.; Pan, Z.; et al. Unrevealing the interaction between electrode degradation and bubble behaviors in an anion exchange membrane water electrolyzer. Adv. Sci. 2025, 12, 2412962. [Google Scholar] [CrossRef]
  2. Kurtz, D.A.; Hunter, B.M. Forming O-O bonds. Joule 2022, 6, 2272–2292. [Google Scholar] [CrossRef]
  3. Zhu, J.Y.; Xue, Q.; Xue, Y.Y.; Ding, Y.; Li, F.M.; Jin, P.; Chen, P.; Chen, Y. Iridium nanotubes as bifunctional electrocatalysts for oxygen evolution and nitrate reduction reactions. ACS Appl. Mater. Interfaces 2020, 12, 14064–14070. [Google Scholar] [CrossRef]
  4. Zhang, K.; Zou, R. Advanced transition metal-based OER electrocatalysts: Current status, opportunities, and challenges. Small 2021, 17, 2100129. [Google Scholar] [CrossRef]
  5. Wang, Y.; Zhang, Y.; Wang, S.; Li, L.; Lv, T.; Liu, X.; Zhao, H.; Han, Z.; Tan, X.; Mu, Y.; et al. Fe coordination environment modulating oxygen evolution reaction properties of cobalt site. Small 2025, 21, 2500121. [Google Scholar] [CrossRef] [PubMed]
  6. Zhu, M.; Xu, H.; Dai, J.; Guan, D.; Hu, Z.; She, S.; Chen, C.T.; Ran, R.; Zhou, W.; Shao, Z. A dynamically stable self-assembled CoFe (oxy)hydroxide-based nanocatalyst with boosted electrocatalytic performance for the oxygen-evolution reaction. J. Mater. Chem. A 2024, 12, 24308–24317. [Google Scholar] [CrossRef]
  7. Badreldin, A.; Abusrafa, A.E.; Abdel-Wahab, A. Oxygen-deficient cobalt-based oxides for electrocatalytic water splitting. ChemSusChem 2021, 14, 10–32. [Google Scholar] [CrossRef] [PubMed]
  8. Deng, Y.; Wang, J.; Zhang, S.F.; Zhang, Z.J.; Sun, J.F.; Li, T.T.; Kang, J.L.; Liu, H.; Bai, S. In situ constructing lamella-heterostructured nanoporous CoFe/CoFe2O4 and CeO2−x as bifunctional electrocatalyst for high-current-density water splitting. Rare Met. 2025, 44, 1053–1066. [Google Scholar] [CrossRef]
  9. Kim, D.H.; Lee, Y.K. Understanding highly active and durable Fe-decorated Co(OH)2 catalysts in alkaline oxygen evolution reaction by in situ XANES studies. Chem. Eng. J. 2024, 490, 151701. [Google Scholar] [CrossRef]
  10. Bo, X.; Li, Y.; Chen, X.; Zhao, C. Operando Raman spectroscopy reveals Cr-induced-phase reconstruction of NiFe and CoFe oxyhydroxides for enhanced electrocatalytic water oxidation. Chem. Mater. 2020, 32, 4303–4311. [Google Scholar] [CrossRef]
  11. Dionigi, F.; Zeng, Z.; Sinev, I.; Merzdorf, T.; Deshpande, S.; Lopez, M.B.; Kunze, S.; Zegkinoglou, I.; Sarodnik, H.; Fan, D.; et al. In-situ structure and catalytic mechanism of NiFe and CoFe layered double hydroxides during oxygen evolution. Nat. Commun. 2020, 11, 2522. [Google Scholar] [CrossRef]
  12. Smith, R.D.; Pasquini, C.; Loos, S.; Chernev, P.; Klingan, K.; Kubella, P.; Mohammadi, M.R.; Gonzalez-Flores, D.; Dau, H. Spectroscopic identification of active sites for the oxygen evolution reaction on iron-cobalt oxides. Nat. Commun. 2017, 8, 2022. [Google Scholar] [CrossRef] [PubMed]
  13. Zhou, Q.; Xue, W.; Cui, X.; Wang, P.; Zuo, S.; Mo, F.; Li, C.; Liu, G.; Ouyang, S.; Zhan, S.; et al. Oxygen-bridging Fe, Co dual-metal dimers boost reversible oxygen electrocatalysis for rechargeable Zn-air batteries. Proc. Natl. Acad. Sci. USA 2024, 121, e2404013121. [Google Scholar] [CrossRef] [PubMed]
  14. Ou, Y.; Twight, L.P.; Samanta, B.; Liu, L.; Biswas, S.; Fehrs, J.L.; Sagui, N.A.; Villalobos, J.; Morales-Santelices, J.; Antipin, D.; et al. Cooperative Fe sites on transition metal (oxy) hydroxides drive high oxygen evolution activity in base. Nat. Commun. 2023, 14, 7688. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, Y.; Dongfang, N.; Huang, C.; Erni, R.; Li, J.; Zhao, H.; Pan, L.; Iannuzzi, M.; Patzke, G.R. Operando monitoring of the functional role of tetrahedral cobalt centers for the oxygen evolution reaction. Nat. Commun. 2025, 16, 580. [Google Scholar] [CrossRef]
  16. Zhan, Y.; Yang, T.; Liu, S.; Yang, L.; Wang, E.; Yu, X.; Wang, H.; Chou, K.C.; Hou, X. In situ characterization method to reveal the surface reconstruction process of an electrocatalyst. Nanomaterials 2025, 15, 917. [Google Scholar] [CrossRef]
  17. Grimaud, A.; Demortière, A.; Saubanère, M.; Dachraoui, W.; Duchamp, M.; Doublet, M.L.; Tarascon, J.M. Activation of surface oxygen sites on an iridium-based model catalyst for the oxygen evolution reaction. Nat. Energy 2016, 2, 16189. [Google Scholar] [CrossRef]
  18. Li, X.; Zhao, J.; Zhou, J.; Wang, Q.; Han, J. Deep sulfur doping induces the rapid electrochemical self-reconstruction of Ni–Fe hydroxide to drive water oxidation. Green Chem. 2023, 25, 10684–10692. [Google Scholar] [CrossRef]
  19. Sheng, J.; Kang, J.; Jiang, P.; Meinander, K.; Hong, X.; Jiang, H.; Nonappa; Ikkala, O.; Komsa, H.P.; Peng, B.; et al. Guided heterostructure growth of CoFe LDH on Ti3C2Tx MXene for durably high oxygen evolution activity. Small 2025, 21, 2404927. [Google Scholar] [CrossRef]
  20. Li, Z.; Lin, G.; Wang, L.; Lee, H.; Du, J.; Tang, T.; Ding, G.; Ren, R.; Li, W.; Cao, X.; et al. Seed-assisted formation of NiFe anode catalysts for anion exchange membrane water electrolysis at industrial-scale current density. Nat. Catal. 2024, 7, 944–952. [Google Scholar] [CrossRef]
  21. Yu, M.; Weidenthaler, C.; Wang, Y.; Budiyanto, E.; Sahin, E.O.; Chen, M.; DeBeer, S.; Rüdiger, O.; Tüysüz, H. Surface boron modulation on cobalt oxide nanocrystals for electrochemical oxygen evolution reaction. Angew. Chem. Int. Ed. 2022, 61, e202211543. [Google Scholar] [CrossRef]
  22. El-Seidy, A.M.; El-Okaily, M.S.; Nabil, I.M.; Mostafa, A.A. New binary and ternary SiO2 composites with Fe2O3 and Co2.74O4 and the evaluation of their γ-radiation shielding properties. Sci. Rep. 2025, 15, 13714. [Google Scholar] [CrossRef] [PubMed]
  23. Vinothkumar, V.; Koventhan, C.; Chen, S.M.; Abinaya, M.; Kesavan, G.; Sengottuvelan, N. Preparation of three dimensional flower-like cobalt phosphate as dual functional electrocatalyst for flavonoids sensing and supercapacitor applications. Ceram. Int. 2021, 47, 29688–29706. [Google Scholar] [CrossRef]
  24. Ye, S.; Lei, Y.; Xu, T.; Zheng, L.; Chen, Z.; Yang, X.; Ren, X.; Li, Y.; Zhang, Q.; Liu, J. Deeply self-reconstructing CoFe (H3O)(PO4)2 to low-crystalline Fe0.5Co0.5OOH with Fe3+–O–Fe3+ motifs for oxygen evolution reaction. Appl. Catal. B Environ. 2022, 304, 120986. [Google Scholar] [CrossRef]
  25. Zhang, T.; Jiang, J.; Sun, W.; Gong, S.; Liu, X.; Tian, Y.; Wang, D. Spatial configuration of Fe–Co dual-sites boosting catalytic intermediates coupling toward oxygen evolution reaction. Proc. Natl. Acad. Sci. USA 2024, 121, e2317247121. [Google Scholar] [CrossRef] [PubMed]
  26. Singh, B.; Sherman, D.M.; Gilkes, R.J.; Wells, M.A.; Mosselmans, J.F.W. Incorporation of Cr, Mn and Ni into goethite (α-FeOOH): Mechanism from extended X-ray absorption fine structure spectroscopy. Clay Miner. 2002, 37, 639–649. [Google Scholar] [CrossRef]
  27. Zhang, X.; Zhong, H.; Zhang, Q.; Zhang, Q.; Wu, C.; Yu, J.; Ma, Y.; An, H.; Wang, H.; Zou, Y.; et al. High-spin Co3+ in cobalt oxyhydroxide for efficient water oxidation. Nat. Commun. 2024, 15, 1383. [Google Scholar] [CrossRef]
  28. Zuo, S.; Wu, Z.; Zhang, G.; Chen, C.; Ren, Y.; Liu, Q.; Zheng, L.; Zhang, J.; Han, Y.; Zhang, H. Correlating structural disorder in metal (oxy)hydroxides and catalytic activity in electrocatalytic oxygen evolution. Angew. Chem. Int. Ed. 2024, 63, e202316762. [Google Scholar] [CrossRef]
  29. Chen, S.; Ma, L.; Huang, Z.; Liang, G.; Zhi, C. In situ/operando analysis of surface reconstruction of transition metal-based oxygen evolution electrocatalysts. Cell Rep. Phys. Sci. 2022, 3, 100729. [Google Scholar] [CrossRef]
  30. Cai, Z.; Zhou, D.; Wang, M.; Bak, S.-M.; Wu, Y.; Wu, Z.; Tian, Y.; Xiong, X.; Li, Y.; Liu, W.; et al. Introducing Fe2+ into nickel-iron layered double hydroxide: Local structure modulated water oxidation activity. Angew. Chem. Int. Ed. 2018, 130, 9536–9540. [Google Scholar] [CrossRef]
  31. Sui, M.; Wang, F.; Liu, C.; Hao, Q.; Li, Y.; Liu, H.; Liang, L. Interfacial electron coupling in Au nanoparticle/CoFe layered double hydroxide nanosheet heterostructures as water oxidation catalysts. ACS Appl. Nano Mater. 2024, 7, 22444–22452. [Google Scholar] [CrossRef]
  32. Zhao, N.; Ou, Z.; Chen, J.; Tang, C.; Yu, D. Facile synthesis of Co3O4 nanocrystals with superior lithium storage properties from bis (8-hydroxyquinoline) cobalt complex as anode materials for lithium-ion batteries. Scripta Mater. 2024, 238, 115766. [Google Scholar] [CrossRef]
  33. Grimaud, A.; Diaz-Morales, O.; Han, B.; Hong, W.T.; Lee, Y.-L.; Giordano, L.; Stoerzinger, K.A.; Koper, M.T.M.; Shao-Horn, Y. Activating lattice oxygen redox reactions in metal oxides to catalyse oxygen evolution. Nat. Chem. 2017, 9, 457–465. [Google Scholar] [CrossRef]
  34. Reith, L.; Hausmann, J.N.; Mebs, S.; Mondal, I.; Dau, H.; Driess, M.; Menezes, P.W. In situ detection of iron in oxidation states ≥ IV in cobalt-iron oxyhydroxide reconstructed during oxygen evolution reaction. Adv. Energy Mater. 2023, 13, 2203886. [Google Scholar] [CrossRef]
  35. Rao, R.R.; Corby, S.; Bucci, A.; García-Tecedor, M.; Mesa, C.A.; Rossmeisl, J.; Giménez, S.; Lloret-Fillol, J.; Stephens, I.E.L.; Durrant, J.R. Spectroelectrochemical analysis of the water oxidation mechanism on doped nickel oxides. J. Am. Chem. Soc. 2022, 144, 7622–7633. [Google Scholar] [CrossRef] [PubMed]
  36. Li, N.; Bediako, D.K.; Hadt, R.G.; Hayes, D.; Kempa, T.J.; von Cube, F.; Bell, D.C.; Chen, L.X.; Nocera, D.G. Influence of iron doping on tetravalent nickel content in catalytic oxygen-evolving films. Proc. Natl. Acad. Sci. USA 2017, 114, 1486–1491. [Google Scholar] [CrossRef] [PubMed]
  37. Diao, F.; Kraglund, M.R.; Cao, H.; Yan, X.; Liu, P.; Engelbrekt, C.; Xiao, X. Moderate heat treatment of CoFe Prussian blue analogues for enhanced oxygen evolution reaction performance. J. Energy Chem. 2023, 78, 476–486. [Google Scholar] [CrossRef]
  38. Li, M.; Gu, Y.; Chang, Y.; Gu, X.; Tian, J.; Wu, X.; Feng, L. Iron doped cobalt fluoride derived from CoFe layered double hydroxide for efficient oxygen evolution reaction. Chem. Eng. J. 2021, 425, 130686. [Google Scholar] [CrossRef]
  39. Yao, R.; Wu, J.; Kansara, S.; Sun, Z.; Kang, H.; Liu, F.; Wang, K.; Hu, J.; Li, X.; Wu, D.; et al. Recycling defunct lithium-ion battery cathodes to quaternary layered double hydroxides for efficient oxygen evolution reaction. Adv. Sci. 2025, 12, e2501957. [Google Scholar] [CrossRef]
  40. Li, X.; Xiao, L.; Zhou, L.; Xu, Q.; Weng, J.; Xu, J.; Liu, B. Adaptive bifunctional electrocatalyst of amorphous CoFe oxide@2D black phosphorus for overall water splitting. Angew. Chem. Int. Ed. 2020, 132, 21292–21299. [Google Scholar] [CrossRef]
  41. Li, J.; Liu, Z.; Li, W.; Ma, H.; Fang, P.; Xiong, R.; Pan, C.; Wei, J. In situ interfacial engineering of MnIn2S4@In2S3 hollow nanotubes for enhanced photocatalytic production of H2O2 and antibiotic degradation. J. Colloid Interface Sci. 2025, 682, 1–10. [Google Scholar] [CrossRef]
  42. Thangasamy, P.; Oh, S.; Nam, S.; Randriamahazaka, H.; Oh, I.K. Ferrocene-incorporated cobalt sulfide nanoarchitecture for superior oxygen evolution reaction. Small 2020, 16, 2001665. [Google Scholar] [CrossRef]
  43. Yang, S.; Lu, L.; Zhan, P.; Si, Z.; Chen, L.; Zhuang, Y.; Qin, P. Amorphous hetero-structure iron/cobalt oxyhydroxide with atomic dispersed palladium for oxygen evolution reaction. Appl. Catal. B Environ. Energy 2024, 355, 124213. [Google Scholar] [CrossRef]
  44. Wu, L.; Qin, H.; Ji, Z.; Zhou, H.; Shen, X.; Zhu, G.; Yuan, A. Nitrogen-Doped Carbon Dots Modified Fe–Co Sulfide Nanosheets as High-Efficiency Electrocatalysts toward Oxygen Evolution Reaction. Small 2024, 20, 2305965. [Google Scholar] [CrossRef]
  45. Huo, J.; Wang, Y.; Yan, L.; Xue, Y.; Li, S.; Hu, M.; Jiang, Y.; Zhai, Q.-G. In situ semi transformation from heterometallic MOFs to Fe-Ni LDH/MOF hierarchical architectures for boosted oxygen evolution reaction. Nanoscale 2020, 12, 14514–14523. [Google Scholar] [CrossRef]
  46. Li, S.; Liu, J.; Duan, S.; Wang, T.; Li, Q. Tuning the oxygen evolution electrocatalysis on NiFe-layered double hydroxides via sulfur doping. Chin. J. Catal. 2020, 41, 847–852. [Google Scholar] [CrossRef]
Figure 1. The precursor catalyst Fe0.42Co0.58-SO4/NF’s: (a,b) SEM image; (c) TEM image; (d) HRTEM image; (ei) SEM-EDS elemental maps of Co, Fe, O, and S; (j) Variation in the content of Co and Fe before and after the OER process; (k) XANES spectra for Fe K-edge.
Figure 1. The precursor catalyst Fe0.42Co0.58-SO4/NF’s: (a,b) SEM image; (c) TEM image; (d) HRTEM image; (ei) SEM-EDS elemental maps of Co, Fe, O, and S; (j) Variation in the content of Co and Fe before and after the OER process; (k) XANES spectra for Fe K-edge.
Molecules 31 00096 g001
Figure 2. The working catalyst Fe0.42Co0.58OOH/NF’s: (a) Electrochemical activation by CV (10 cycles); (b,c) SEM image; (dh) SEM-EDS elemental maps of Co, Fe, O, and S.
Figure 2. The working catalyst Fe0.42Co0.58OOH/NF’s: (a) Electrochemical activation by CV (10 cycles); (b,c) SEM image; (dh) SEM-EDS elemental maps of Co, Fe, O, and S.
Molecules 31 00096 g002
Figure 3. (a) EXAFS spectra for Fe K-edge; (b) Raman spectra of the Fe0.42Co0.58OOH/NF; (c) EPR spectra of Fe0.42Co0.58-SO4/NF and Fe0.42Co0.58OOH/NF recorded; (d) Fe0.42Co0.58-SO4/NF’s FT k2χ(R) EXAFS spectra for Fe K-edge; (e) Fe0.42Co0.58OOH/NF’s FT k2χ(R) EXAFS spectra for Fe K-edge; (f) Fe0.42Co0.58-SO4/NF’s WT transformed EXAFS spectra for Fe K-edge; (g) Fe0.42Co0.58OOH/NF’s WT transformed EXAFS spectra for Fe K-edge.
Figure 3. (a) EXAFS spectra for Fe K-edge; (b) Raman spectra of the Fe0.42Co0.58OOH/NF; (c) EPR spectra of Fe0.42Co0.58-SO4/NF and Fe0.42Co0.58OOH/NF recorded; (d) Fe0.42Co0.58-SO4/NF’s FT k2χ(R) EXAFS spectra for Fe K-edge; (e) Fe0.42Co0.58OOH/NF’s FT k2χ(R) EXAFS spectra for Fe K-edge; (f) Fe0.42Co0.58-SO4/NF’s WT transformed EXAFS spectra for Fe K-edge; (g) Fe0.42Co0.58OOH/NF’s WT transformed EXAFS spectra for Fe K-edge.
Molecules 31 00096 g003
Figure 4. (a) LSV curves; (b) Tafel plots; (c) LSV curves of Fe0.42Co0.58OOH/NF before and after 1000 cycles CV durability test; (d) EIS plots; (e) Cdl plots derived from CV curve; (f) E-t curves of Fe0.42Co0.58OOH/NF at a current density of 600 mA cm−2.
Figure 4. (a) LSV curves; (b) Tafel plots; (c) LSV curves of Fe0.42Co0.58OOH/NF before and after 1000 cycles CV durability test; (d) EIS plots; (e) Cdl plots derived from CV curve; (f) E-t curves of Fe0.42Co0.58OOH/NF at a current density of 600 mA cm−2.
Molecules 31 00096 g004
Figure 5. (a) Schematic diagram of Fe-M bonds promoting the OER mechanism; (b) LSV curves of Fe0.42Co0.58OOH/NF in 1M KOH and 1M TMAOH solution; (c) OER activities of Fe0.42Co0.58OOH/NF in KOH solutions with different pH values (14.00, 13.82, 13.66, and 13.45); (d) Fitting plot of the Overpotentials at 10 mA cm−2 vs. the pH value of the electrolytes; (e) Fe-O-Fe and Fe-O-Co sites’ adsorption energies and desorption energies of OH and O2.
Figure 5. (a) Schematic diagram of Fe-M bonds promoting the OER mechanism; (b) LSV curves of Fe0.42Co0.58OOH/NF in 1M KOH and 1M TMAOH solution; (c) OER activities of Fe0.42Co0.58OOH/NF in KOH solutions with different pH values (14.00, 13.82, 13.66, and 13.45); (d) Fitting plot of the Overpotentials at 10 mA cm−2 vs. the pH value of the electrolytes; (e) Fe-O-Fe and Fe-O-Co sites’ adsorption energies and desorption energies of OH and O2.
Molecules 31 00096 g005
Scheme 1. Formation process of Fe0.42Co0.58OOH/NF.
Scheme 1. Formation process of Fe0.42Co0.58OOH/NF.
Molecules 31 00096 sch001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, M.; Pei, B.; Ba, H.; Ni, W.; Zhao, H.; Chen, S.; Zhao, J.; Zhao, J. Oxygen Bridge Governs OER via Deep Self-Reconstruction in Fe–Co Oxyhydroxides. Molecules 2026, 31, 96. https://doi.org/10.3390/molecules31010096

AMA Style

Liu M, Pei B, Ba H, Ni W, Zhao H, Chen S, Zhao J, Zhao J. Oxygen Bridge Governs OER via Deep Self-Reconstruction in Fe–Co Oxyhydroxides. Molecules. 2026; 31(1):96. https://doi.org/10.3390/molecules31010096

Chicago/Turabian Style

Liu, Mingyu, Bowen Pei, Hongyu Ba, Wei Ni, Huaheng Zhao, Shuang Chen, Jiamin Zhao, and Jinsheng Zhao. 2026. "Oxygen Bridge Governs OER via Deep Self-Reconstruction in Fe–Co Oxyhydroxides" Molecules 31, no. 1: 96. https://doi.org/10.3390/molecules31010096

APA Style

Liu, M., Pei, B., Ba, H., Ni, W., Zhao, H., Chen, S., Zhao, J., & Zhao, J. (2026). Oxygen Bridge Governs OER via Deep Self-Reconstruction in Fe–Co Oxyhydroxides. Molecules, 31(1), 96. https://doi.org/10.3390/molecules31010096

Article Metrics

Back to TopTop