Next Article in Journal
Efficient Degradation of Tetracycline via Cobalt Phosphonate-Activated Peroxymonosulfate: Mechanistic Insights and Catalytic Optimization
Previous Article in Journal
Effects of Mg on CO2 Hydrogenation on PtCoAl Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

MOF-Derived Electrocatalysts for High-Efficiency Hydrogen Production via Water Electrolysis

School of Chemistry and Chemical Engineering, Xi’an University of Science and Technology, 58 Yanta Road, Yanta District, Xi’an 710054, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(6), 579; https://doi.org/10.3390/catal15060579
Submission received: 9 April 2025 / Revised: 31 May 2025 / Accepted: 6 June 2025 / Published: 10 June 2025
(This article belongs to the Section Catalytic Materials)

Abstract

Water electrolysis for hydrogen production has garnered significant attention in the context of increasing global energy demands and the “dual-carbon” strategy. However, practical implementation is hindered by challenges such as high overpotentials, high catalysts costs, and insufficient catalytic activity. In this study, three mono and bimetallic metal−organic framework (MOFs)-derived electrocatalysts, Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs, were synthesized via a one-step hydrothermal method, using nitro-terephthalic acid (NO2-BDC) as the ligand and N,N-dimethylacetamide (DMA) as the solvent. Electrochemical tests demonstrated that the Fe/Mn-MOFs catalyst exhibited superior performance, achieving an overpotential of 232.8 mV and a Tafel slope of 59.6 mV·dec−1, alongside the largest electrochemical active surface area (ECSA). In contrast, Fe/Co-MOFs displayed moderate catalytic activity, while Fe-MOFs exhibited the lowest efficiency. Stability tests revealed that Fe/Mn-MOFs retained 92.3% of its initial current density after 50 h of continuous operation, highlighting its excellent durability for the oxygen evolution reaction (OER). These findings emphasize the enhanced catalytic performance of bimetallic MOFs compared to monometallic counterparts and provide valuable insights for the development of high-efficiency MOF-based electrocatalysts for sustainable hydrogen production.

1. Introduction

The growing global demand for sustainable energy, coupled with the urgent need to achieve “carbon peaking and carbon neutrality” has intensified the focus on developing clean and efficient energy technology [1,2]. Hydrogen energy, a promising clean energy carrier, can be produced through various methods, including fossil fuel reforming and biological processes. However, these conventional techniques often face challenges such as high carbon emissions or complex operational processes. In contrast, water electrolysis, using renewable electricity to split water into hydrogen and oxygen, has emerged as a highly attractive and environmentally friendly hydrogen production method, offering a zero-carbon footprint and utilizing abundant and renewable feedstocks [3,4,5].
Water electrolysis holds significant promise for large-scale hydrogen production. Overall, water splitting consists of two key reactions: the oxygen evolution reaction (OER) at the anode (4OH → O2 + 2H2O + 4e in alkaline media) produces oxygen, and the hydrogen evolution reaction (HER) at the cathode (2H2O + 2e → H2 + 2OH in alkaline media) produces hydrogen. However, its practical implementation faces notable challenges, particularly the high overpotential during the oxygen evolution reaction (OER), which severely limits energy efficiency and hampers industrial scalability [6,7,8]. To overcome these barriers, the development of advanced electrocatalysts that enhance reaction kinetics is crucial for the advancement of this technology [9,10]. Electrocatalysts are typically classified into two categories: noble-metal-based materials (e.g., Pt, RuO2, and IrO2) and transition-metal-based compounds (e.g., oxides, hydroxides, and sulfides of Ni, Co, and Fe). While noble metal catalysts exhibit outstanding catalytic activity, their scarcity and high cost restrict their widespread use [10,11,12]. In contrast, transition-metal-based catalysts are more cost-effective, but often suffer from lower activity and poor durability [13,14]. Therefore, the design of novel electrocatalysts that combine high efficiency, low cost, and robust stability remains a critical area of focus in water electrolysis research [15,16,17,18].
Non-metallic catalysts have recently garnered significant attention due to their unique physicochemical properties. Among them, metal-organic frameworks (MOFs) have emerged as promising candidates for electrocatalysis, owing to their structural versatility, tunable porosity, high surface area, and customizable metalligand coordination [19,20,21]. MOF-derived electrocatalysts, through rational design and tailored synthesis strategies, offer precise control over structural and compositional parameters, thereby optimizing the catalytic performance. For example, S. Rajasekaran et al. synthesized a Ni-Cu MOFs film via a solvothermal method, achieving an OER overpotential of 340 mV at 10 mA cm−2 with remarkable stability [22]. Li et al. developed Ni-Co-Fe trimetallic MOF nanosheets through microwave-assisted synthesis, demonstrating an ultralow overpotential of 243 mV at 10 mA cm−2 and a Tafel slope of 48.1 mV·dec−1 [23]. Danning Xing et al. reported a π–d conjugated trimetallic FeCoNi-BHT MOF (BHT stood for benzenehexathiol), which exhibited an overpotential of 266 mV at 10 mA cm−2 for OER [24]. These studies underscore that MOF-derived electrocatalysts not only reduce overpotentials, but also enhance the current density and stability, providing a robust material foundation for efficient water splitting. However, the role of electron-withdrawing ligands in modulating the electronic structure of bimetallic MOFs remains underexplored for optimizing oxygen evolution reaction (OER) kinetics. Furthermore, while trimetallic systems (e.g., Ni-Co-Fe and FeCoNi-BHT) have demonstrated high activity, their complex synthesis limits the scalability. Addressing these gaps, our work pioneers a one-step hydrothermal synthesis of Fe/Mn-MOFs using nitro-terephthalic acid (NO2-BDC), an electron-deficient ligand, to create asymmetric metal sites. This design uniquely enhances in situ charge transfer between Fe3+ and Mn2+, achieving an ultralow overpotential of 232.8 mV at 10 mA cm−2 and 92.3% stability retention over 50 h. Crucially, we resolve two key questions: (1) How electron-withdrawing ligands optimize bimetallic synergy in MOFs beyond conventional trimetallic systems; (2) Why Fe/Mn pairs outperform Fe/Co in OER energetics, as validated by Tafel kinetics and active-site exposure.
In this study, three MOF-based catalysts, Fe-MOFs, Fe/Mn-MOFs, and Fe/Co- MOFs, were synthesized through a one-step hydrothermal method, using nitro-terephthalic acid (NO2-BDC) as the ligand; N, N-dimethylacetamide (DMA) as the solvent; and FeCl3·6H2O, Co(NO3)2·6H2O, and Mn(NO3)2·4H2O as metal precursors. As shown in Figure 1, the detailed synthetic route is depicted. The morphological characteristics, elemental composition, crystallinity, and chemical structures of the catalysts were thoroughly characterized using scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), X-ray diffraction (XRD), and Fourier-transform infrared spectroscopy (FT-IR). The electrocatalytic performance for OER was assessed by linear sweep voltammetry (LSV), and Tafel slope analysis was performed to investigate the reaction kinetics and rate-determining steps. The electrochemical active surface area (ECSA) was estimated from double-layer capacitance (Cdl) values obtained through cyclic voltammetry (CV) at various scan rates. Additionally, chronoamperometry was employed to evaluate the long-term stability under constant potential conditions.

2. Experiments and Methods

2.1. Experimental Materials

The following analytical grade chemicals were obtained from commercial suppliers and used without further purification: anhydrous ethanol (≥99.5%) and N, N-dimethylacetamide (>99.0%) (Tianjin Comio Chemical Reagent Co., Ltd., Tianjin, China); nitroterephthalic acid (≥98.0%), iron (III) chloride hexahydrate (>99.0%), cobalt (II) nitrate hexahydrate (≥99.0%), and manganese (II) nitrate tetrahydrate (≥98.0%) (Aladdin Reagent (Shanghai) Co., Ltd., Shanghai, China).

2.2. Preparation of Metal-Organic Frameworks Nanocatalysts

The Fe/Mn-MOFs catalyst was synthesized via the hydrothermal method [25]. In a typical procedure, 0.15 mmol (40.50 mg) of FeCl3·6H2O, 0.15 mmol (43.00 mg) of Mn(NO3)2·4H2O, and 0.30 mmol (63.30 mg) of NO2-BDC were accurately weighed and dissolved in 35.00 mL (376.00 mmol) of DMA. The mixture was sonicated to ensure uniformity, then transferred to a hydrothermal reactor and heated at 150 °C for 4 h. After the reaction, the product was washed with absolute ethanol, separated by centrifugation, and dried at 60 °C for 24 h to obtain the Fe/Mn-MOFs catalyst.
The preparation methods for Fe/Co-MOFs and Fe-MOFs catalysts are identical to that of Fe/Mn-MOFs, with the only difference being the use of different reactants. For the Fe/Co-MOFs catalyst, 0.15 mmol (40.50 mg) of FeCl3·6H2O and 0.15 mmol (43.40 mg) of Co(NO3)2·6H2O were used. For the Fe-MOFs catalyst, 0.30 mmol (81.00 mg) of FeCl3·6H2O was used. Given the use of nitro-functionalized terephthalic acid (NO2-BDC) as the organic linker and the adopted hydrothermal conditions, the synthesized catalysts are classified as MIL-type MOFs. The specific synthetic route is shown in Figure 2.

2.3. Electrode Preparation

First, we weighed 5 mg of the sample, added 950 µL of isopropanol and 50 μL of Nafion, and ultrasonicated for 30 min to achieve uniform dispersion. Then, we measured 10 μL of the uniformly dispersed sample and dropped it onto a clean glassy carbon electrode (5 mm). The sample was allowed to dry naturally for testing.

2.4. Characterization and Electrochemical Evaluation of Synthesized MOFs Electrocatalysts

The chemical structure of the synthesized MOFs electrocatalysts was characterized using Fourier transform infrared spectroscopy (FT-IR, iS20, Thermo Fisher Scientific, Waltham, MA, USA). The surface morphology and microstructural features were examined by field-emission scanning electron microscopy (FESEM, Verios G4, FEI Company, Hillsboro, OR, USA). The crystallinity and phase structure were analyzed using X-ray diffraction (XRD, D8 Advance, Bruker, Karlsruhe, BW, Germany). All electrochemical measurements were performed using an electrochemical workstation (Model CHI660D, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) with a standard three-electrode configuration. The working electrode was a glassy carbon electrode (GCE) loaded with the MOFs catalysts, while a platinum plate served as the counter electrode and an Ag/AgCl electrode was used as the reference electrode. The electrolyte consisted of 1 M KOH. All potentials were referenced to the reversible hydrogen electrode (RHE): EVS RHE = EVS Ag/AgCl + 0.059pH + 0.197. The oxygen evolution activity of the MOFs catalysts was evaluated by LSV at a scan rate of 5 mV/s (potential range: 0.1–0.7 V vs. Ag/AgCl). The Tafel slope was calculated to assess the reaction kinetics of the OER. To determine the Cdl, CV was performed at various scan rates (5, 10, 20, 30, 40, and 50 mV/s), which were then used to estimate the ECSA of the MOFs catalysts and to evaluate their active sites. Finally, the long-term stability of the MOFs catalysts was examined via chronopotentiometry. The decay of the catalytic performance over a 50 h electrolysis period was monitored to assess the durability and stability of the catalysts [26].

3. Results and Discussion

3.1. SYNTHESIS and Characterization of MOF-Derived Catalysts

The surface morphologies of the as-synthesized MOFs catalysts (Fe-MOFs, Fe/Co- MOFs, and Fe/Mn-MOFs) were systematically investigated using SEM, as shown in Figure 3a–c. All three catalysts exhibited densely packed granular morphologies with uniform particle sizes in the range of 50–80 nm and interparticle voids of 10–30 nm. Notably, minimal morphological variations were observed among the different metal-incorporated MOFs, indicating that the introduction of secondary metal ions (Co2+ or Mn2+) under identical synthetic conditions had little effect on the surface architecture. Infrared spectroscopy (Figure 3d) revealed significant spectral changes compared to the pristine NO2-BDC ligand. A characteristic splitting of the C=O stretching vibration band around 1600 cm−1, corresponding to carboxylate asymmetric vibrations, was consistently observed across all catalysts. This spectral modification provides direct evidence of successful metal-ligand coordination between the carboxylate groups and metal centers [27,28].
The phase characterization via XRD (Figure 3e) indicated that the three catalysts had low crystallinity, yet a distinct diffraction peak appeared at 8.81°. Quantitative analysis of peak sharpness revealed changes in crystallinity: Fe/Mn-MOFs exhibited the highest crystallinity (FWHM = 2.9°), followed by Fe/Co-MOFs (full width at half maximum (FWHM) of 3.2°), and Fe-MOFs showed the lowest crystallinity (FWHM = 3.7°). This hierarchical trend in crystallinity can be attributed to the differences in ionic radii and coordination kinetics among the combined transition metals. The small hump at 15.33° in Fe/Mn-MOFs might result from the altered crystal packing mode or unit cell parameters after introducing Mn into Fe-MOFs [29]. The small humps in Fe/Co-MOFs and Fe-MOFs are basically caused by some mechanical and instrumental noises. Taken together, these results demonstrate the successful synthesis of three distinct MOFs catalysts, as evidenced by comprehensive morphological, functional group, and crystalline structure characterization. The preservation of framework integrity despite metal substitution further underscores the robustness of the employed synthesis protocol.
XPS was performed to probe the electronic structure and bimetallic interactions (Figure 4). The survey spectra confirmed the presence of Fe, Mn/Co, O, C, and N, and EDS analysis further validated this finding (Tables S1–S3). In Fe/Mn-MOFs, the Fe 2p3/2 peak at 711.8 eV shifted ~0.5 eV lower than in Fe-MOFs, indicating electron delocalization from Mn2+ to Fe3+, which optimizes OH adsorption [26]. The Mn 2p3/2 peak at 641.5 eV suggested hybridized Mn2+ states, enhancing charge transfer kinetics. Conversely, Fe/Co-MOFs showed negligible shifts in Fe 2p and Co 2p peaks, implying weaker synergistic effects. The O 1s spectrum of Fe/Mn- MOFs featured a higher hydroxyl component (531.5 eV), aligning with its larger electrochemical active surface area (ECSA). These results validate that the Fe3+–Mn2+ interaction via electron-withdrawing NO2-BDC ligands boosts OER activity by modulating electronic structure and active site accessibility.

3.2. Electrochemical Performance and Characterization

LSV is widely employed to evaluate the catalytic activity in electrocatalytic reactions [4]. To assess the catalytic performance of the synthesized MOFs, LSV measurements were conducted for three types of MOFs catalysts in a 1 M KOH solution using a standard three-electrode system. The results are shown in Figure 5a. As shown in Figure 5a, the onset potentials of the three catalysts were as follows: Fe/Mn-MOFs (1439.8 mV), Fe/Co-MOFs (1449.8 mV), and Fe-MOFs (1458.8 mV). Among these, Fe/Mn-MOFs exhibited the lowest onset potential (1439.8 mV), indicating that this catalyst initiated the reaction more readily and exhibited superior catalytic activity [18]. At a current density of 10 mA·cm−2, the overpotentials for the three catalysts were Fe/Mn-MOFs (232.8 mV), Fe/Co-MOFs (248.8 mV), and Fe-MOFs (253.8 mV), respectively. These results suggest that Fe/Mn MOFs require a significantly lower voltage to achieve the same current density compared to Fe- MOFs and Fe/Co-MOFs catalysts. This enhanced performance can be attributed to the superior crystallinity of Fe/Mn-MOFs, which is higher than that of Fe-MOFs and Fe/Co -MOFs catalysts. The improved crystallinity facilitates more efficient electron conduction within the lattice, thereby enhancing the electrical conductivity of the material and, consequently, improving the catalytic activity of the MOFs catalysts.
The Tafel slope is a key parameter for evaluating the performance of electrocatalysts. It not only reflects the kinetic rate of the catalyst in OER, but also provides insight into the rate-determining step of the reaction process [10,26,30]. To further investigate the catalytic mechanism of the synthesized MOFs catalysts, we analyzed the data from the LSV curves and plotted the corresponding Tafel curves (Figure 5b). As shown in Figure 5b, the Tafel curves of the different MOFs catalysts were derived from the polarization curves after coordinate transformation. By fitting the Tafel curves, the Tafel slopes for the OER were determined. The Tafel slopes for Fe-MOFs, Fe/Mn-MOFs, and Fe/Co-MOFs were 72.2 mV·dec−1, 59.6 mV·dec−1, and 64.2 mV·dec−1, respectively. The Fe/Mn-MOFs catalyst exhibited the smallest Tafel slope, indicating that this catalyst has more excellent kinetics in catalyzing the oxygen evolution reaction. In contrast, the Fe-MOFs catalyst showed the largest Tafel slope, suggesting significant hindrance in the adsorption process of OER. Furthermore, by extrapolating the Tafel curves to the point where the current density is zero, it is evident that Fe/Mn-MOFs has the highest exchange current, demonstrating superior reaction kinetics compared to the other two catalysts. The Tafel slopes for all three catalysts are close to 60 mV·dec−1, which suggests that the OER for these catalysts was limited by the second stage of the reaction (M–OH + OH → M–O + H2O + e). The rate of the OER at this stage is governed by the equilibrium between the adsorption of hydroxyl groups and the formation of O–O bonds during the oxygen evolution process [24].
ECSA was employed to evaluate the catalytic activity of materials. As a crucial parameter reflecting the number of accessible active sites for electrochemical reactions, a higher ECSA value typically indicates larger reaction interfaces and more abundant catalytic sites, thereby significantly enhancing electrocatalytic efficiency [9,31,32]. The quantitative determination of ECSA was achieved through its proportional relationship with the Cdl [26,33,34]. To calculate ECSA for different MOFs catalysts, CV measurements were conducted within a potential window of 0–0.06 V vs. RHE at varying scan rates (5, 10, 20, 50, and 100 mV/s). The corresponding CV curves of Fe-MOFs, Fe/Mn-MOFs, and Fe/Co-MOFs catalysts are presented in Figure 6a–c. Analysis of current responses in the non-faradaic region enabled calculation of Cdl values (Figure 6d) [22]. The obtained Cdl values for Fe-MOFs, Fe/Co-MOFs, and Mn/Fe-MOFs catalysts were 0.381, 0.409, and 0.429 mF cm−2, respectively. Notably, Fe/Mn-MOFs catalyst demonstrated a significantly higher Cdl value compared to other catalysts, indicating its superior ECSA. This observation suggests that Fe/Mn-MOFs catalysts possess more abundant exposed active sites, which facilitates efficient participation in the OER [29].
A comprehensive analysis highlights three key advantages of Fe/Mn-MOFs catalysts: (1) The lowest onset potential and minimal overpotential (η) at 10 mA cm−2, which can be attributed to the enhanced electrical conductivity and catalytic activity from its high crystallinity; (2) the smallest Tafel slope (56.3 mV dec−1), indicating accelerated charge transfer kinetics and stronger interactions between active sites and reactants; (3) the highest Cdl value and ECSA, confirming the optimal exposure of catalytically active centers. These findings collectively demonstrate that bimetallic MOFs catalysts outperform their monometallic counterparts in OER performance, with Fe/Mn dual-metal systems exhibiting superior catalytic properties compared to Fe/Co systems.
The stability of electrocatalysts is a critical factor in evaluating their practicality for real-world applications [31]. Building on the impressive catalytic performance of Fe/Mn-MOFs observed in preliminary tests, we systematically assessed their long-term durability through a 50 h chronoamperometric test at a fixed potential of 1.482 V vs. RHE, with the results shown in Figure 7. As depicted in Figure 7, the Fe/Mn-MOFs catalyst exhibited excellent current density stability during the first 25 h of operation, maintaining a steady current density of approximately 25 mA·cm−2. After this period, a gradual decrease in current density was observed, although the decline was minimal throughout the remainder of the test. Importantly, the total current density loss after 50 h of continuous operation was less than 10%, highlighting the catalyst’s exceptional stability in the OER process. This result strongly supports the potential of Fe/Mn-MOFs as a highly stable and efficient electrocatalyst, making it a promising candidate for practical applications in sustainable energy conversion systems.
To contextualize the performance of the Fe/Mn-MOFs catalysts developed in this work, a comparative analysis with several recently reported state-of-the-art catalysts for the oxygen evolution reaction (OER) is presented in Table 1 (see below). Notably, the Fe/Mn-MOFs exhibit a remarkably low overpotential of 233 mV to achieve a current density of 10 mA cm−2, surpassing many prominent catalysts including Ir-MSC-Co3O4 (248 mV in acid) [35], FeCoNiCuCr HEA NPs (272 mV in alkaline) [36], FeCo/FePx/MNx-CNTs/CNF (~266 mV in alkaline) [37], and Co0.3Fe-ZP (284 mV in alkaline) [38]. While Ta-RuO2 demonstrates an exceptional overpotential of 178 mV, it operates under acidic conditions and utilizes precious Ru [39]. The Tafel slope of Fe/Mn-MOFs (59.6 mV dec−1) indicates favorable kinetics, comparable to FeCoNiCuCr HEA NPs (58.3 mV dec−1) and significantly lower than Co0.3Fe-ZP (71.2 mV dec−1) and FeCo/FePx/MNx-CNTs/CNF (67.5 mV dec−1), although slightly higher than the highly active Ir-MSC-Co3O4 (42.1 mV dec−1) and Ta-RuO2 (40.2 mV dec−1). Crucially, the Fe/Mn-MOFs achieve this high performance using the common glassy carbon (GC) substrate, demonstrating its effectiveness without requiring specialized conductive supports like carbon cloth (CC), carbon fiber (CF), or nickel foam (NF) used by some competitors. This comparison underscores the competitive edge of the Fe/Mn-MOFs catalyst, combining low overpotential, favorable kinetics, and substrate versatility, positioning it as a highly promising non-precious metal OER electrocatalyst.

4. Conclusions

In conclusion, three distinct MOF-derived electrocatalysts, Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs, were successfully synthesized using a one-step hydrothermal method with NO2-BDC as the ligand and DMAC as the solvent. A thorough investigation of their surface morphology, elemental composition, crystal structure, chemical properties, and electrocatalytic performance was conducted. The results showed that the MOFs catalysts exhibited a surface morphology consisting of uniformly packed and closely adhered particles, with the introduction of bimetals having minimal effect on the surface structure. Electrochemical testing in a 1 M KOH electrolyte revealed that Fe/Mn-MOFs catalyst exhibited the best OER performance, requiring an overpotential of only 232.8 mV to achieve a current density of 10 mA·cm−2 and demonstrating a notably low Tafel slope of 59.6 mV·dec−1. Mechanistic analysis suggested that the OER rates for all three MOFs catalysts were limited by the second step (M–OH + OH → M– O + H2O + e), where the rate-determining step was influenced by the equilibrium between hydroxyl group adsorption and O−O bond formation. Furthermore, Fe/Mn-MOFs catalysts exhibited the largest electrochemical active surface area and maintained a high current density even after 50 h of stability testing. The advantages of this MOFs catalyst include its simple synthesis method, readily available raw materials, the use of common metals, and the presence of an electron-withdrawing nitro group that enhances catalytic activity. However, its stability remains insufficient, and the material is challenging to produce on a large scale. In summary, compared to single-metal catalysts, the bimetallic MOFs catalysts demonstrated significantly improved catalytic activity and reduced overpotentials. This enhancement was attributed to the synergistic effects of bimetals and the excellent structural characteristics of the bimetallic catalysts. This study provides essential experimental data and theoretical insights for the development of efficient and stable MOF-based electrocatalysts, which is crucial for advancing water electrolysis technology for hydrogen production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15060579/s1. Table S1. Component content of Fe-MOFs catalyst. Table S2. Component content of Fe/Mn-MOFs catalyst. Table S3. Component content of Fe/Co-MOFs catalyst.

Author Contributions

N.Z.: (First Author & Corresponding Author) Conceptualization, Funding Acquisition, Resources, Formal Analysis, Writing—Original Draft; P.C.: Supervision Methodology, Investigation, Writing—Review & Editing; J.Z.: Data Curation, Validation; Y.Q.: Investigation; All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Yulin Science and Technology Project (No. CXY-2022-147), And the APC was funded by the National Natural Science Foundation of China (No. 22205180).

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 author(s).

Acknowledgments

The authors acknowledge financial support from the Yulin Science and Technology Project (No. CXY-2022-147), the National Natural Science Foundation of China (No. 22205180), the Scientific Research Start-up Fund of Xi’an University of Science and Technology (XUST), and the XUST Outstanding Youth Project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, J.; Long, Q.; Xiao, K.; Ouyang, T.; Li, N.; Ye, S.; Liu, Z.-Q. Vertically-interlaced NiFeP/MXene electrocatalyst with tunable electronic structure for high-efficiency oxygen evolution reaction. Sci. Bull. 2021, 66, 1063–1072. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, H.; Ci, S.; Ding, Y.; Wang, G.; Wen, Z. Recent advances in precious metal-free bifunctional catalysts for electrochemical conversion systems. J. Mater. Chem. A 2019, 7, 8006–8029. [Google Scholar] [CrossRef]
  3. Li, Y.; Lu, M.; Wu, Y.; Ji, Q.; Xu, H.; Gao, J.; Qian, G.; Zhang, Q. Morphology regulation of metal-organic framework-derived nanostructures for efficient oxygen evolution electrocatalysis. J. Mater. Chem. A 2020, 8, 18215–18219. [Google Scholar] [CrossRef]
  4. Shi, Y.; Wu, H.; Chang, J.; Tang, Z.; Lu, S. Progress on the mechanisms of Ru-based electrocatalysts for the oxygen evolution reaction in acidic media. J. Energy Chem. 2023, 85, 220–238. [Google Scholar] [CrossRef]
  5. Guo, B.; Ding, Y.; Huo, H.; Wen, X.; Ren, X.; Xu, P.; Li, S. Recent Advances of Transition Metal Basic Salts for Electrocatalytic Oxygen Evolution Reaction and Overall Water Electrolysis. Nano-Micro Lett. 2023, 15, 57. [Google Scholar] [CrossRef]
  6. Bhanja, P.; Kim, Y.; Paul, B.; Kaneti, Y.V.; Alothman, A.A.; Bhaumik, A.; Yamauchi, Y. Microporous nickel phosphonate derived heteroatom doped nickel oxide and nickel phosphide: Efficient electrocatalysts for oxygen evolution reaction. Chem. Eng. J. 2021, 405, 126803. [Google Scholar] [CrossRef]
  7. Septiani, N.L.W.; Kaneti, Y.V.; Fathoni, K.B.; Guo, Y.; Ide, Y.; Yuliarto, B.; Jiang, X.; Nugraha; Dipojono, H.K.; Golberg, D.; et al. Tailorable nanoarchitecturing of bimetallic nickel-cobalt hydrogen phosphate via the self-weaving of nanotubes for efficient oxygen evolution. J. Mater. Chem. A 2020, 8, 3035–3047. [Google Scholar] [CrossRef]
  8. Septiani, N.L.W.; Kaneti, Y.V.; Fathoni, K.B.; Kani, K.; Allah, A.E.; Yuliarto, B.; Nugraha; Dipojono, H.K.; Alothman, Z.A.; Golberg, D.; et al. Self-Assembly of Two-Dimensional Bimetallic Nickel-Cobalt Phosphate Nanoplates into One-Dimensional Porous Chainlike Architecture for Efficient Oxygen Evolution Reaction. Chem. Mater. 2020, 32, 7005–7018. [Google Scholar] [CrossRef]
  9. Chen, I.W.P.; Chen, W.-Y.; Liu, T.Y. Pioneering ultra-efficient oxygen evolution reaction: A breakthrough in tri-metallic organic frameworks synthesis. Mater. Today Chem. 2024, 35, 101873. [Google Scholar] [CrossRef]
  10. Hu, W.-C.; Shi, Y.; Zhou, Y.; Wang, C.; Younis, M.R.; Pang, J.; Wang, C.; Xia, X.-H. Plasmonic hot charge carriers activated Ni centres of metal- organic frameworks for the oxygen evolution reaction. J. Mater. Chem. A 2019, 7, 10601–10609. [Google Scholar] [CrossRef]
  11. Song, J.; Wei, C.; Huang, Z.-F.; Liu, C.; Zeng, L.; Wang, X.; Xu, Z.J. A review on fundamentals for designing oxygen evolution electrocatalysts. Chem. Soc. Rev. 2020, 49, 2196–2214. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, Z.-P.; Lu, X.F.; Zang, S.-Q.; Lou, X.W. Non-Noble-Metal-Based Electrocatalysts toward the Oxygen Evolution Reaction. Adv. Funct. Mater. 2020, 30, 1910274. [Google Scholar] [CrossRef]
  13. Tang, Y.-J.; Zhu, H.-J.; Dong, L.-Z.; Zhang, A.M.; Li, S.-L.; Liu, J.; Lan, Y.-Q. Solid-phase hot-pressing of POMs-ZIFs precursor and derived phosphide for overall water splitting. Appl. Catal. B-Environ. 2019, 245, 528–535. [Google Scholar] [CrossRef]
  14. Zhao, M.; Li, H.; Yuan, W.; Li, C.M. Tannic Acid-Mediated In Situ Controlled Assembly of NiFe Alloy Nanoparticles on Pristine Graphene as a Superior Oxygen Evolution Catalyst. Acs Appl. Energy Mater. 2020, 3, 3966–3977. [Google Scholar] [CrossRef]
  15. Ding, S.; Zhang, Y.; Lou, F.; Aslam, M.K.; Sun, Y.; Li, M.; Duan, J.; Li, Y.; Chen, S. “Uncapped” metal-organic framework (MOF) dispersions driven by O2 plasma towards superior oxygen evolution electrocatalysis. J. Mater. Chem. A 2022, 10, 20813–20818. [Google Scholar] [CrossRef]
  16. Zhao, C.-X.; Liu, J.-N.; Wang, J.; Ren, D.; Li, B.-Q.; Zhang, Q. Recent advances of noble-metal-free bifunctional oxygen reduction and evolution electrocatalysts. Chem. Soc. Rev. 2021, 50, 7745–7778. [Google Scholar] [CrossRef]
  17. Xia, B.Y.; Yan, Y.; Li, N.; Wu, H.B.; Lou, X.W.; Wang, X. A metal-organic framework-derived bifunctional oxygen electrocatalyst. Nat. Energy 2016, 1, 15006. [Google Scholar] [CrossRef]
  18. Ding, S.; Sun, Y.; Lou, F.; Yu, L.; Xia, B.; Duan, J.; Zhang, Y.; Chen, S. Plasma-regulated two-dimensional high entropy oxide arrays for synergistic hydrogen evolution: From theoretical prediction to electrocatalytic applications. J. Power Sources 2022, 520, 230873. [Google Scholar] [CrossRef]
  19. Cai, G.; Zhang, W.; Jiao, L.; Yu, S.-H.; Jiang, H.-L. Template-Directed Growth of Well-Aligned MOF Arrays and Derived Self-Supporting Electrodes for Water Splitting. Chem 2017, 2, 791–802. [Google Scholar] [CrossRef]
  20. Peng, X.; Ye, L.; Ding, Y.; Yi, L.; Zhang, C.; Wen, Z. Nanohybrid photocatalysts with ZnIn2S4 nanosheets encapsulated UiO-66 octahedral nanoparticles for visible-light-driven hydrogen generation. Appl. Catal. B-Environ. 2020, 260, 118152. [Google Scholar] [CrossRef]
  21. Yang, M.; Jiao, L.; Dong, H.; Zhou, L.; Teng, C.; Yan, D.; Ye, T.-N.; Chen, X.; Liu, Y.; Jiang, H.-L. Conversion of bimetallic MOF to Ru-doped Cu electrocatalysts for efficient hydrogen evolution in alkaline media. Sci. Bull. 2021, 66, 257–264. [Google Scholar] [CrossRef] [PubMed]
  22. Rajasekaran, S.; Reghunath, B.S.; Devi, K.R.S.; Pinheiro, D. Designing coordinatively unsaturated metal sites in bimetallic organic frameworks for oxygen evolution reaction. Mater. Today Chem. 2023, 31, 101616. [Google Scholar] [CrossRef]
  23. Li, Q.; Liu, Y.; Niu, S.; Li, C.; Chen, C.; Liu, Q.; Huo, J. Microwave-assisted rapid synthesis and activation of ultrathin trimetal-organic framework nanosheets for efficient electrocatalytic oxygen evolution. J. Colloid Interface Sci. 2021, 603, 148–156. [Google Scholar] [CrossRef]
  24. Xing, D.; Wang, H.; Cui, Z.; Lin, L.; Liu, Y.; Dai, Y.; Huang, B. A Conductive Two-dimensional Trimetallic FeCoNi-Benzenehexathiol π-d Conjugated Metal-organic Framework for Highly Efficient Oxygen Evolution Reaction. J. Colloid Interface Sci. 2024, 656, 309–319. [Google Scholar] [CrossRef] [PubMed]
  25. Zhai, X.; Yu, Q.; Liu, G.; Bi, J.; Zhang, Y.; Chi, J.; Lai, J.; Yang, B.; Wang, L. Hierarchical microsphere MOF arrays with ultralow Ir doping for efficient hydrogen evolution coupled with hydrazine oxidation in seawater. J. Mater. Chem. A 2021, 9, 27424–27433. [Google Scholar] [CrossRef]
  26. Cheng, Y.; Luo, Y.; Zheng, Y.; Pang, J.; Sun, K.; Hou, J.; Wang, G.; Guo, W.; Guo, X.; Chen, L. Self-supporting one-dimensional ZnFe-BDC for electrocatalysis oxygen evolution reaction in alkaline and natural seawater. Int. J. Hydrog. Energy 2022, 47, 35655–35665. [Google Scholar] [CrossRef]
  27. Dou, S.; Tao, L.; Wang, R.; El Hankari, S.; Chen, R.; Wang, S. Plasma-Assisted Synthesis and Surface Modification of Electrode Materials for Renewable Energy. Adv. Mater. 2018, 30, e1705850. [Google Scholar] [CrossRef]
  28. Xu, L.; Jiang, Q.; Xiao, Z.; Li, X.; Huo, J.; Wang, S.; Dai, L. Plasma-Engraved Co3O4 Nanosheets with Oxygen Vacancies and High Surface Area for the Oxygen Evolution Reaction. Angew. Chem. -Int. Ed. 2016, 55, 5277–5281. [Google Scholar] [CrossRef]
  29. Doughty, T.; Zingl, A.; Wunschek, M.; Pichler, C.M.; Watkins, M.B.; Roy, S. Structural Reconstruction of a Cobalt- and Ferrocene-Based Metal-Organic Framework during the Electrochemical Oxygen Evolution Reaction. Acs Appl. Mater. Interfaces 2024, 16, 40814–40824. [Google Scholar] [CrossRef]
  30. Li, J.; Liu, P.; Mao, J.; Yan, J.; Song, W. Structural and electronic modulation of conductive MOFs for efficient oxygen evolution reaction electrocatalysis. J. Mater. Chem. A 2021, 9, 11248–11254. [Google Scholar] [CrossRef]
  31. Hu, Q.; Huang, X.; Wang, Z.; Li, G.; Han, Z.; Yang, H.; Ren, X.; Zhang, Q.; Liu, J.; He, C. Unconventionally fabricating defect-rich NiO nanoparticles within ultrathin metal-organic framework nanosheets to enable high-output oxygen evolution. J. Mater. Chem. A 2020, 8, 2140–2146. [Google Scholar] [CrossRef]
  32. Kumar, S.; Weng, P.H.; Fu, Y.P. Core-shell-structured CuO@Ni-MOF: Bifunctional electrode toward battery-type supercapacitors and oxygen evolution reaction. Mater. Today Chem. 2022, 26, 101159. [Google Scholar] [CrossRef]
  33. Yu, R.; Wang, C.; Liu, D.; Wu, Z.; Li, J.; Du, Y. Bimetallic sulfide particles incorporated in Fe/Co-based metal-organic framework ultrathin nanosheets toward boosted electrocatalysis of the oxygen evolution reaction. Inorg. Chem. Front. 2022, 9, 3130–3137. [Google Scholar] [CrossRef]
  34. Ren, T.; Wang, J.; Yu, X.; Chen, Y.; Wu, Y.; Che, G.; Jiang, W.; Teng, H.; Liu, C. The electrocatalytic self-reconstruction of ultrathin 2D MOF nanoarrays supported on alloy foam improves the oxygen evolution reaction. Colloids Surf. A-Physicochem. Eng. Asp. 2024, 684, 133136. [Google Scholar] [CrossRef]
  35. Wang, Y.; Qin, Y.; Liu, S.; Zhao, Y.; Liu, L.; Zhang, D.; Zhao, S.; Liu, J.; Wang, J.; Liu, Y.; et al. Mesoporous Single-Crystalline Particles as Robust and Efficient Acidic Oxygen Evolution Catalysts. J. Am. Chem. Soc. 2025, 14, 13345–13355. [Google Scholar] [CrossRef] [PubMed]
  36. Zhu, X.; Huang, W.; Tan, L.; Yao, Z.; Yang, X.; Song, R.; Lan, S. Ultrafast Synthesis of Tetragonal-Distorted FeCoNiCuCr High-Entropy Alloy Nanoparticles for Enhanced OER Performance. Chin. Chem. Lett. 2025, 36, 110852. [Google Scholar] [CrossRef]
  37. Ding, W.; Saad, A.; Wu, Y.; Wang, Z.; Li, X. CNTs/CNF-Supported Multi-Active Components as Highly Efficient Bifunctional Oxygen Electrocatalysts and Their Applications in Zinc-Air Batteries. Nano Res. 2023, 16, 4793–4802. [Google Scholar] [CrossRef]
  38. Ni, H.; Xu, S.; Lin, R.; Ding, Y.; Qian, J. Ligand-Induced Hollow Binary Metal-Organic Framework Derived Fe-Doped Cobalt-Carbon Nanomaterials for Oxygen Evolution. J. Colloid Interface Sci. 2024, 671, 100–109. [Google Scholar] [CrossRef]
  39. Zhang, J.; Fu, X.; Kwon, S.; Chen, K.; Liu, X.; Yang, J.; Kang, Y. Tantalum-Stabilized Ruthenium Oxide Electrocatalysts for Industrial Water Electrolysis. Science 2025, 387, 48–55. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthesis of Fe-MOFs, Fe/Mn-MOFs, and Fe/Co-MOFs catalysts.
Figure 1. Schematic illustration of the synthesis of Fe-MOFs, Fe/Mn-MOFs, and Fe/Co-MOFs catalysts.
Catalysts 15 00579 g001
Figure 2. Schematic diagram of the preparation method for MOF-derived catalysts (using Fe/Mn- MOFs as an example).
Figure 2. Schematic diagram of the preparation method for MOF-derived catalysts (using Fe/Mn- MOFs as an example).
Catalysts 15 00579 g002
Figure 3. (ac) SEM images, (d) FT-IR spectra, and (e) XRD patterns of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts.
Figure 3. (ac) SEM images, (d) FT-IR spectra, and (e) XRD patterns of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts.
Catalysts 15 00579 g003
Figure 4. XPS spectra of Fe/Mn-MOFs and Fe/Co-MOFs (a) Survey. (b) C 1s XPS spectra. (c) N 1s XPS spectra. (d) Fe 2p XPS spectra. (e) Mn 2p XPS spectrum. (f) Co 2p3/2 XPS spectrum.
Figure 4. XPS spectra of Fe/Mn-MOFs and Fe/Co-MOFs (a) Survey. (b) C 1s XPS spectra. (c) N 1s XPS spectra. (d) Fe 2p XPS spectra. (e) Mn 2p XPS spectrum. (f) Co 2p3/2 XPS spectrum.
Catalysts 15 00579 g004
Figure 5. (a) OER polarization curves and (b) Tafel curves of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts.
Figure 5. (a) OER polarization curves and (b) Tafel curves of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts.
Catalysts 15 00579 g005
Figure 6. Cyclic voltammetry curves of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts at different sweep velocities (ac). (d) Electric double-layer capacitance fitting curve.
Figure 6. Cyclic voltammetry curves of Fe-MOFs, Fe/Co-MOFs, and Fe/Mn-MOFs catalysts at different sweep velocities (ac). (d) Electric double-layer capacitance fitting curve.
Catalysts 15 00579 g006
Figure 7. Time-current density curve of Mn/Fe-MOFs catalysts.
Figure 7. Time-current density curve of Mn/Fe-MOFs catalysts.
Catalysts 15 00579 g007
Table 1. Comparison of OER performance between Fe/Mn-MOFs and other electrocatalysts.
Table 1. Comparison of OER performance between Fe/Mn-MOFs and other electrocatalysts.
CatalystOverpotential
@10 mA cm−2
Tafel SlopeSubstrate
Fe/Mn-MOFs233 mV (alkaline)59.6 mV dec−1Glassy Carbon
Ir-MSC-Co3O4248 mV (acidic)42.1 mV dec−1Glassy Carbon
FeCoNiCuCr HEA NPs272 mV (alkaline)58.3 mV dec−1Carbon Cloth
FeCo/FePx/MNx-CNTs/CNF~266 mV (alkaline)67.5 mV dec−1Carbon Fiber
Co0.3Fe-ZP284 mV (alkaline)71.2 mV dec−1Nickel Foam
Ta-RuO2178 mV (acidic)40.2 mV dec−1Ti Mesh
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

Zhang, N.; Cui, P.; Zhang, J.; Qiao, Y. MOF-Derived Electrocatalysts for High-Efficiency Hydrogen Production via Water Electrolysis. Catalysts 2025, 15, 579. https://doi.org/10.3390/catal15060579

AMA Style

Zhang N, Cui P, Zhang J, Qiao Y. MOF-Derived Electrocatalysts for High-Efficiency Hydrogen Production via Water Electrolysis. Catalysts. 2025; 15(6):579. https://doi.org/10.3390/catal15060579

Chicago/Turabian Style

Zhang, Nan, Pengfei Cui, Jinrong Zhang, and Yang Qiao. 2025. "MOF-Derived Electrocatalysts for High-Efficiency Hydrogen Production via Water Electrolysis" Catalysts 15, no. 6: 579. https://doi.org/10.3390/catal15060579

APA Style

Zhang, N., Cui, P., Zhang, J., & Qiao, Y. (2025). MOF-Derived Electrocatalysts for High-Efficiency Hydrogen Production via Water Electrolysis. Catalysts, 15(6), 579. https://doi.org/10.3390/catal15060579

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop