E. S. Sowbakkiyavathiad,
Preethi Dhandapanib,
Senthilkumar Ramasamyc,
Ju Hyun Ohd,
Insik Ind,
Seung Jun Lee*d and
A. Subramania
*bd
aDepartment of Physics, Sathyabama Institute of Science and Technology, Chennai 600 119, India
bElectro-Materials Research Laboratory, Centre for Nanoscience and Technology, Pondicherry University, Puducherry 605 014, India. E-mail: a.subramania@gmail.com
cCentre of Excellence in Advanced Materials and Green Technologies, Amrita School of Engineering, Amrita Vishwa Vidyapeetham University, Coimbatore-641112, India
dDepartment of IT-Energy Convergence, Korea National University of Transportation, Chungju 27469, Republic of Korea. E-mail: sjlee@ut.ac.kr
First published on 30th June 2025
Climate change, global warming, and other adverse environmental impacts are largely driven by carbon dioxide (CO2) emissions. One promising pathway to mitigate these issues is the growing eco-friendly hydrogen production technologies. Hydrogen, as a clean energy carrier, has the potential to transition industries toward decarbonization. Amongst the numerous hydrogen production approaches, water splitting via electrocatalysis presents a sustainable route. However, achieving huge productivity in the hydrogen evolution reaction (HER) requires advanced catalytic agents with enhanced active sites, huge porosity, and robust adaptability. Recently, materials based on metal–organic frameworks (MOFs) have received more consideration in electrocatalysis for environmental remediation and energy. The metal component of MOFs typically consists of metal ions (often transition metals) or metal clusters. These metal ions act as the nodes in the framework, coordinating with the organic ligands. The choice of metal determines the chemical properties, stability, and reactivity of the MOF. Numerous MOF-based materials were effectively established for the applications of the hydrogen evolution process. To produce hydrogen, this review article examines various MOF-related electrocatalysts, which include MOF-derived metals, metal oxides, metal phosphides, metal nitrides, metal chalcogenides, dichalcogenides, and their composites. Furthermore, the pros and cons of various MOF-based materials as water-splitting catalysts are discussed. Lastly, the present challenges and future prospects of these materials as electrocatalysts are also discussed.
Sustainability spotlightMetal–organic framework-based electrocatalysts can be used in electrolyzers to produce hydrogen from water, offering a promising route for sustainable hydrogen production due to their individual properties like high specific area, high porosity, wide open structure, and tunability. Recently, metal–organic frameworks (MOFs) demonstrated high efficiency in the hydrogen evolution reaction (HER), with some reports showing overpotentials as low as 10 mV. |
Consequently, several techniques, such as the reformation of steam methane, the combustion of coal, and the splitting of water, were investigated to produce hydrogen for industrial purposes.6,7 Though coal gasification and steam methane reforming can produce a significantly larger amount of hydrogen than water electrolysis, further issues, including CO2 emissions and precursors made from fossil fuels, must be considered. Thus, the electrocatalytic water-splitting method of producing hydrogen is highly regarded. It is much more affordable, environmentally friendly, and more practical than other approaches because it employs abundant water as the base material and limitless solar energy as the driving force.7–9
The procedure for water splitting in an electrolytic cell comprises two essential half-reactions: the cathodic hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER). In essence, compared to the HER, the OER process needs a high overpotential to achieve the necessary current density. As a prerequisite for the thermodynamically uphill process to occur in the electrolyzer, water splitting involves a thermodynamic Gibbs free energy (DG) of about 237.2 kJ mol−1, which is comparable to a standard potential (DE) of 1.23 V versus the reversible hydrogen electrode (RHE). Homogeneous and heterogeneous catalysts are the two different kinds of water-splitting catalysts; of these, semiconductor catalysts, as heterogeneous catalysts, have attracted a lot of interest owing to their robust anti-toxicity, huge thermal endurance, and adaptable catalytic characteristics.10–12
Omar Yaghi and his colleagues introduced the concept of metal–organic frameworks (MOFs) in the 1990s by demonstrating that metal ions and organic linkers could self-assemble into highly porous crystalline structures. Since then, MOFs have attracted significant attention as electrocatalysts for water splitting due to their high surface area, tunable porosity, and diverse chemical compositions. Initially developed for gas storage and separation applications, MOFs have since been adapted for catalytic purposes, including electrocatalytic water splitting, owing to their ability to be engineered with specific active sites and tailored pore architectures. Crystalline permeable polymers composed of organic intermediates and metal ions, which are identified as metal–organic frameworks (MOFs), were found to be a promising candidate in numerous fields, because of their characteristics like excellent specific area, wide open structures, high porosity, tunability, etc.13
The following are the unique properties of MOFs as an ideal candidate for application in electrocatalytic reactions. Because of their high porosity, MOFs have a wide surface area and make it easier for ions and electrons to move freely during electrochemical reactions.14,15 During synthesis, MOF characteristics like pore size, metal sites, and ligand functionalization can be adjusted to optimize them for particular electrocatalytic reactions. MOFs may be more effective than conventional catalysts due to their enormous surface area, which frequently exceeds 7000 m2 g−1, and offer more active sites for electrochemical processes. During electrochemical reactions, the metal centers in MOFs can experience reversible coordination changes that increase their catalytic activity. MOFs with conductive metal centers or functionalized linkers can help enhance charge transfer during electrocatalytic reactions.16,17 Hybridization with conductive materials like graphene or carbon nanotubes can improve the overall conductivity of MOFs, which is critical for efficient electrocatalysts. MOFs have a well-organized donor–acceptor interface that allows for efficient electron–hole separation and improved porosity, making them excellent candidates for use in electrocatalysis.18 However, in their pristine form, the majority of them have limited stability in water, weak photocatalytic activity, and chemical instability. These constraints make it difficult to employ them for environmental remediation. Nonetheless, MOFs have a special feature that allows for easy tuning of their chemical compositions, allowing for the modification of both their structural and chemical characteristics to improve performance. Numerous changes, such as manufacturing heterostructures and defect engineering of heterojunctions, have been recently documented. Researchers can improve the reliability, conductivity, and reactivity of MOFs by altering the metal centers or organic ligands. By integrating MOFs with conductive substances like graphene or carbon nanotubes, their electrical conductivity and stability can be increased, increasing their usefulness for electrocatalytic processes.19,20 By adding dopants like sulfur or nitrogen to the structure of MOFs, their stability and electrochemical characteristics can be improved. Core–shell MOFs can enhance stability and performance in electrocatalytic reactions, and this is achieved by coating the MOF surface with a more stable metal or substance. In the electrocatalytic hydrogen evolution reaction, MOF-based materials are highly recommended as superior building blocks for the creation of extremely effective electrocatalysts. To date, various electrocatalysts for HER, including pristine MOFs, MOFs with nanoparticles, MOF nitrides, MOF phosphides, and metal chalcogenides, result in excellent catalytic activity and stability.
Herein, we highlight an analysis of recent developments in MOFs and their composites with engineering features for hydrogen production. This article provides a brief history of MOFs and their characteristics, merits, limitations, applications, and enhancement techniques to boost electrocatalytic performance. In addition, for hydrogen generation applications, a thorough analysis of the morphological development and optimization techniques was conducted. Lastly, conclusions and suggestions for the future are offered to give guidance for further study in this area.
Generally, two electrochemical processes occur on two electrodes in electrocatalytic water splitting. Specifically, the anode electrode experiences the oxidation of water, whereas the cathode electrode experiences the reduction reaction of H2O.21,22 The HER is the primary reaction for producing hydrogen gas at the cathode. The process comprises the degradation of protons (H+) to yield hydrogen gas (H2).23 The overall reaction for the HER is:
2H+ + 2e− → H2; |
In a typical HER process, the protons are reduced at the cathode, where they gain electrons from the peripheral power source, forming hydrogen gas (H2). In an acidic medium, the HER occurs in the subsequent three basic phases (Fig. 1);
![]() | ||
Fig. 1 Reaction involved in the HER process. Image is reproduced with permission from ref. 19 under Creative Commons Attribution License CC BY, Copyright © CrystEngComm 2020. |
H+ + e− → Hads |
Hads + H+ + e− → H2 |
This is often termed the Heyrovsky step, named after the scientist who first proposed it.
2Hads → H2 |
![]() | ||
Fig. 2 Structure of a MOF. Image is reproduced with permission from ref. 25 under Creative Commons Attribution License CC BY, Copyright © Nano-Micro Letters 2020. |
(i) Transition metals: zinc (Zn), iron (Fe), cobalt (Co), etc., are the most commonly used metals due to their ability to form stable coordination bonds with organic ligands.33
(ii) Lanthanides: some MOFs use rare-earth metals like cerium (Ce) or yttrium (Y) for specialized applications like luminescence or catalysis.34
(iii) Alkaline earth metals: magnesium (Mg) and calcium (Ca) can be used in MOFs, particularly for their role in creating highly porous materials.35
(iv) Metal clusters: in some MOFs, metal clusters like copper tetrahedra or iron-oxide clusters are used, which enhance stability and connectivity.36,37
(i) Carboxylate-based ligands: ligands with carboxyl groups (–COOH) are the most common, where the oxygen atoms coordinate to the metal centers. Some examples of carboxylate-based ligands are terephthalic acid, phthalic acid, isophthalic acid, and fumaric acid.40,41
(ii) Bipyridine ligands: ligands like 2,2′-bipyridine, where nitrogen atoms coordinate to metals, are often used in combination with other ligands.42,43
(iii) Phosphonate ligands: these contain a phosphonic acid group (–PO3H2) and can form stable bonds with metals.44
(iv) Imidazole-based ligands: organic molecules like 2-methylimidazole, where nitrogen atoms coordinate with metals like Zn to form robust frameworks.45,46
(v) Aromatic or heterocyclic ligands: these ligands contain aromatic rings or heteroatoms like nitrogen, which act as coordination sites for the metals.47
(i) One-dimensional (1D) structures: these consist of linear chains where metal ions are linked by organic ligands.
(ii) Two-dimensional (2D) structures: the framework comprises sheets or layers, with ligands attaching to metal nodes to create a two-dimensional network.
(iii) Three-dimensional (3D) structures: these are the most common and involve the development of a three-dimensional network where metal ions are associated by organic linkers to create porous structures. This is the classic form of MOFs and is often referred to as a “Framework”.
Also, MOFs are grouped according to the topology of their framework.50,51 The framework can be highly diverse, and several common structures are:
(i) Cubic frameworks: some of the most common MOFs exhibit a cubic or octahedral symmetry, where metal centers are linked by organic ligands in a symmetrical pattern. Hexagonal frameworks: hexagonal MOFs consisting of organic linkers and metal centers are also frequently seen. These structures often have a high degree of symmetry and can form 2D or 3D networks.
(ii) Layered frameworks: some MOFs have 2D layered structures, where the metal centers and ligands form planar sheets stacked together. These materials are useful for applications requiring interlayer porosity or for the intercalation of guest molecules.
(iii) Open-framework structures: some MOFs are characterized by open frameworks where the network is highly porous, allowing for the inclusion of a wide range of guest molecules like gases or liquids.
![]() | ||
Fig. 3 (a) Diagrammatic illustration of MOF structures. (b) Techniques for MOF synthesis. The image is reproduced with permission from ref. 46 under Creative Commons Attribution License CC BY, Copyright © Nano-Micro Letters 2021. |
MOFs are synthesized using innumerable techniques, some of which are covered in this segment. In the solvothermal method, metal salts and organic ligands undergo dissolution in an appropriate solvent and are then heated for a while at elevated temperatures and pressurized in a Teflon-lined autoclave. This reaction can be referred to as solvothermal if water is employed. MOFs are created when the ensuing slurry exhibits nucleation and growth. Using the solvothermal method, MOFs like MOF-5,53 MOF-74,54 MOF-177,55 ZIF-8,56 and HKUST-1 (ref. 57) are synthesized under ambient conditions. Next, the hydrothermal method uses high pressure and temperature to start the formation of MOFs, just like the solvothermal method does. However, water was employed as the solvent instead of an organic one. The reaction mixture is also placed in a Teflon-lined autoclave for the controlled synthesis of MOFs. MOFs like Materials of Institute Lavoisier-101 (MIL-101) were synthesized through the hydrothermal method.58 Other than solvothermal and hydrothermal methods of synthesis, utilizing microwave irradiation to expedite the synthesis process, which is also known as microwave-assisted synthesis, was a quick and effective process to produce MOFs like the University of Oslo-66 (UiO-66) by mixing the organic ligands and metal salts in a microwave reactor.59 Rather than using water or organic solvents, MOFs were produced in an ionic solution. Although this approach uses specialized equipment, it is environmentally beneficial. The ionic liquid method is used to produce an MOF-like zeolitic imidazolate framework-7 (ZIF-7) by combining Zn(NO3)2 with 2-methylimidazole in an ionic solution, which is then heated to create a highly porous structure with adjustable characteristics.60 Another quick and easy synthesis method is rapid liquid phase synthesis, which produces MOFs in the liquid phase. The collected liquid phase of the MOF product is then separated through filtration. Although it is a quick and easy process, hazardous solvents must be used. For instance, the Hong Kong University of Science and Technology-1 (HKUST-1) was synthesized through a rapid liquid phase method by combining Cu(CH3COO)2 and trimesic acid in a solution of acetic acid and water, and then filtering the mixture to get an extremely permeable and robust structure.61
![]() | ||
Fig. 4 Illustration of various MOFs, MOF-derived materials and their composites used as electrocatalysts for hydrogen production. |
Applying the Langmuir–Blodgett technique, Dong et al. prepared a huge and independent one-layer sheet of two-dimensional supramolecular polymer (2DSP) that contains triphenylene-fused nickel bis(dithiolene) complexes.70 With a Tafel slope of 80.5 mV dec−1 and an overpotential of 333 mV at a current density of 10 mA cm−2, these 2DSPs demonstrate exceptional catalytic sites for hydrogen evolution from water, outperforming carbon nanotube-supported molecular catalysts and heteroatom-doped graphene catalysts. Li et al. produced 2D polyhalogenated Co(II) coordination isostructural polymers using a simple and facile hydrothermal method.71 He observed that the Co-MOF and acetylene black combination enhanced the hydrogen production reactivity in acidic media, and the Co–Cl4-MOF's HER electrocatalytic activity was noticeably better than that of Co–Br4-MOF and Co–F4-MOF. Among all composites, the AB and Co–Cl4-MOF (3:
4) surprisingly show the highest HER action because of their long-term stability, lower Tafel slope of 86 mV dec−1, and smaller overpotential of 283 mV at a current density of 10 mA cm−2. The electrochemical methodology used catalysts based on nickel metal to produce hydrogen. But their poor activity limited their use in the catalysis industry. There have been several methods reported for improving the HER's electrocatalytic performance.25 Initially, using a pyrolysis technique, Wang et al. produced nickel NPs over the carbon form Ni-MOF for deployment in the HER.72 The surface of Ni metal was modified using NH3 gas. According to the findings, Ni nanoparticles produced at 0.4 bar of pressure had a low overvoltage of 88 mV at 20 mA cm−2.
![]() | ||
Fig. 5 The decoupling process and the linked evolution of metallic and single Co sites. The image is reproduced with permission from ref. 68 under Creative Commons Attribution License CC BY, Copyright © ChemSusChem 2018. |
ZIF-67/Cu(OH)2 was calcined to produce bimetallic Cu, Co nanoparticles over a carbonaceous framework (CuCo@NC). The outcome demonstrates that CuCo@NC has outstanding catalytic capabilities, exhibiting an overpotential of 145 mV at 10 mA cm−2, which is given in Fig. 6(a–d). This CuCo@NC electrocatalyst greatly improves electrocatalytic performances in the HER by offering a large number of catalytic sites, substantial nitrogen mixing, robust synergistic interaction, and enhanced mass transfer.74
![]() | ||
Fig. 6 (a) LSV curves for hydrogen evolution in 0.5 M of H2SO4 for several samples, (b) corresponding Tafel curves, (c) LSV curves displaying hydrogen evolution in 0.1 M of KOH for different samples, and (d) CuCo@NC's HER stability tests in 0.5 M of H2SO4 and 0.1 M of KOH. The image is reproduced with permission from ref. 69 under Creative Commons Attribution License CC BY, Copyright © Advanced Energy Materials 2017. |
![]() | ||
Fig. 7 (a) Photograph of the two-electrode water splitting and gas collection device. (b) Polarization curves of Co@Co3O4-NC‖Co@Co3O4-NC, Pt/C‖IrO2, Pt/C‖Pt/C and IrO2‖IrO2. (c) Stability test of Co@Co3O4-NC before and after 3000 cycles at 50 mV s−1 (d) Volume of generated H2 and O2 at a constant potential of 2 V. Image is reproduced with permission from ref. 72 under Creative Commons Attribution License CC BY, Copyright © Journal of Materials Chemistry A 2013. |
In addition, Ji et al. investigated TMP-based catalysts for the HER.80 In this study, he synthesized CoP nanoframes (CoP NFs) through precipitation and chemical etching, followed by a lower-temperature phosphidation method. The SEM, TEM, HRTEM, and EDAX images of CoP NFs are displayed in Fig. 8(A–G). With a current density of 10 mA cm−2 and a cell voltage of just 1.65 V, CoP NFs produce exceptional bifunctional catalysts for the HER and OER, hence providing a very effective water electrolyzer. He also extended the method to make Co dichalcogenide (CoX2) nanoframes by considering X = S, Se, and Te, respectively. HER activity is according to the following sequence, according to the results of electrochemical tests confirmed by DFT calculations: CoP NFs > CoSe2 NFs > CoS2 > CoTe2 NFs. Long Jiao et al. studied a layered CoP/reduced graphene oxide (rGO) composite that has been effectively generated by pyrolysis and a subsequent phosphating procedure using a logically constructed sandwich-type metal–organic framework/graphene oxide as a template and precursor.81 In acid solution, the resulting CoP/rGO-400 shows outstanding HER activity.
![]() | ||
Fig. 8 The SEM images (A and B), TEM images (C and D), HRTEM images (E), SAED pattern (F), and TEM EDAX (G) for CoP NFs. The image is reproduced with permission from ref. 75 under Creative Commons Attribution License CC BY, Copyright © ACS Catalysis 2020. |
Wang et al. reported the production of controlled spongy Ni2P nanosheets with NiO-MOF-74 as intermediates and a traditional phosphorization technique.82 In the HER, the spongy Ni2P NSs demonstrated outstanding catalytic behavior with quite a low Tafel slope of 63 mV dec−1 and a low overpotential of 168 mV at a current density of 10 mA cm−2 in 1.0 M KOH. This result is due to the electron conduction channels made possible by the porous architectures of Ni2P nanosheets, which speed up the dispersion of H2 and O2 bubbles on the electrode and enable electron transfer. In another study, Suqi He and his co-workers used a modest and rapid microwave-assisted process for the synthesis of MOF-74-Ni. They utilized it as a substrate to produce a permeable hybrid Ni2P/C.83 In the HER process, the prepared Ni2P/C demonstrated outstanding electrocatalytic performance with a poor starting potential of −94 mV, stability (>24 h), and a Tafel slope of 113.2 mV dec−1. The outstanding performance has been ascribed to the huge superficial area due to the porous framework, inherently increased electrocatalytic behavior of Ni2P, and elevated electrical conduction from the carbonaceous elements. Tian et al. synthesized nickel phosphide nanoparticles through a solid chemical transformation method, using Ni-BTC MOF as a precursor under moderate conditions.84 The prepared sample was examined through XRD studies to confirm the formation of Ni2P nanoparticles, which are shown in Fig. 9(a). The surface morphology, the spectrum of elemental mapping, the HRTEM image, equivalent line characteristics, and FFT patterning of the Ni2P NPs are given in Fig. 9(b–g). With a reduced potential of 75 mV and an improved current density, these affordable, earth-abundant MOF-derived Ni2P NPs showed excellent electrocatalytic behavior for the process of HER. Their performance is constructively comparable to that of metal phosphides, indicating a cheap substitute for expensive Pt catalysts in real-world practice. Furthermore, this preparation approach is straightforward, adjustable, affordable, and versatile, making it extremely promising for large-scale manufacturing. This approach was expanded to include the broad synthesis of more TMPs from MOFs. In addition, Tian and his co-workers used a solid-state interaction involving Ni-BTC and NaH2PO2 to produce Ni2P and Ni12P5 nanopowder. In terms of hydrogen generation, Ni2P performed better than Ni12P5.85
![]() | ||
Fig. 9 (a) XRD, (b) SEM, (c) EDS spectrum, (d) SEM-mapping, (e) HRTEM, and (f and g) equivalent line profile and FFT pattern of the Ni2P NPs. Image is reproduced with permission from ref. 78 under Creative Commons Attribution License CC BY, Copyright © RSC Advances 2015. |
Feng et al. used a practical and simple synthesis method for developing fibrous rod-like cobalt–nickel bimetal nitride using a bimetal–organic framework precursor.86 The prepared CoxNiyN exhibits a high surface area, abundant mesoporous structure, and uniformly distributed metal active sites. Furthermore, the presence of Ni results in abundant active sites and changes the Co atom's electrical surroundings. From Fig. 10(A–F), it was shown that, in an all-pH environment, porous Co2Ni1N exhibits outstanding stability and strong catalytic activity, with weak overpotentials of 102.6, 92.0, and 152.8 mV at a current density of 10 mA cm−2 in various solutions, respectively.
![]() | ||
Fig. 10 (A) LSV polarization graph of Pt/C, bare carbon cloth, Co3N, Co2Ni1N, Co1Ni1N, Co1Ni2N, and Ni3N in 1.0 M KOH solution at 5 mV−1. (B) Tafel plots of Pt/C and CoxNiyN materials. (C) EIS Nyquist plot of Co3N, Co2Ni1N, Co1Ni1N, Co1Ni2N, and Ni3N from 10 mHz to 100 kHz (inset: the equivalent circuit. Rs and Rct correspond to the electrolyte and charge transfer resistance, respectively). (D) EIS Nyquist plots of Co2Ni1N from 10 mHz to 100 kHz at various potential levels. (E) Capacitive currents against the scan rate for CoxNiyN materials. (F) Polarization curves of Co2Ni1N before and after 1000 CV sweeps (inset: time dependence on current density for Co2Ni1N). Image is reproduced with permission from ref. 79 under Creative Commons Attribution License CC BY, Copyright © Applied Materials 2019. |
![]() | ||
Fig. 11 (a) Polarization curves of DC, CoSe2, CoSe2/DC, CoSe2@DC, and 20 wt% Pt/C without IR correction. (b) Polarization curves of Co@C, Co3O4@DC, CoSe2@DC and Co@C–Se. (c) Tafel plots from (a and b). (d) Nyquist plots of different samples using modified electrodes that have an overpotential of 200 mV. The equivalent circuit and Nyquist plots of CoSe2@DC modified electrodes with varying overpotentials are shown in the inset figure, (e) the capacitive currents as a function of scan rates, and (f) the chronoamperometric response for CoSe2@DC. The inset figure shows the polarization plots of CoSe2@DC before and after 1000 cycles. The image is reproduced with permission from ref. 81 under Creative Commons Attribution License CC BY, Copyright © Nano Energy 2016. |
Pyrite NiSe2 has recently been regarded as a promising HER catalyst because of its superior conductivity, robust corrosion resistance, and affordability. NiSe2's HER performance in alkaline electrolytes remains unsatisfactory, which may be due only to activated water dissociation in alkaline media.89 Using this perception, Liu et al. described the fabrication of new electrocatalysts of Ni(OH)2/NiSe2 NS on carbon cloth, which has exceptional catalytic activity with an inferior potential of 82 mV to exhibit a current density of 10 mA cm−2 and maintains stability for 12 h in 1.0 M KOH.90 This is higher than that of the most recently developed NiSe2/CC electrocatalysts. For the first time, Zhou et al. established a strategy for growing NiSe2 on conductive nickel foam to produce a three-dimensional fibrous NiSe2/Ni material by thermal selenization of commercially existing Ni foams.91 This approach for material fabrication is inexpensive and time-efficient, with the minimal cost of nickel foam. Despite the straightforward procedure, the grown NiSe2/Ni catalysts exhibit magnificent electrochemical stability, significant cathodic current densities of 100 mA cm−2 at −143 mV, and lower Tafel slopes of about 49 mV dec−1. On considering the primary transition metal dichalcogenide-based catalysts, Zhou et al. described permeable NiSe2 catalysts that are better for water electrolysis.92 NiSe2 displays HER performance comparable to the most advanced platinum catalysts. They developed NiSe2 directly from commercial nickel foam using surface roughness engineering aided by acetic acid. To enhance the TMD performance, Wang et al. utilized first-principles estimation for performing efficient studies of the relationship between transition metal doping and NiSe2 catalytic activity.93 He confirms the formation of Ni1−xFexSe2 porous NSs grown on CFC through XRD and Raman spectroscopy. The structural morphology of as-prepared various combinations of Ni1−xFexSe2 porous NSs grown on CFC was examined by SEM and HRTEM, and their corresponding elemental mapping images showed the formation of Ni1−xFexSe2 porous NSs grown on CFC. From the results, it is shown that Fe is the ideal element to modify NiSe2's electrocatalytic activity with reduced ΔGH* values and enhanced electrical conductivity. Fe/NiSe2 NSs are successfully produced on carbon cloth to offer additional experimental support. When comparing these nanosheets to their undoped counterparts, the efficiency of the hydrogen evolution reaction is noticeably higher. At a reduced overpotential of 64 mV, the optimized Ni0.8Fe0.2Se2 electrocatalysts produce a current density of 10 mA cm−2 with extraordinary stability, as shown in Fig. 12(a–f).
![]() | ||
Fig. 12 (a) Linear sweep voltammetry curves and (b) Tafel plots for NiSe2/CFC, Ni0.9Fe0.1Se2/CFC, Ni0.8Fe0.2Se2/CFC, and Pt/C measured in 0.5 M of H2SO4, (c) current density curve with time (inset) and LSV graphs before and after a stability test of 10![]() |
Other than nickel and cobalt-based catalysts, MoSe2 nanosheets are thought to be potential electrocatalysts for the HER. However, due to significant aggregation or restacking of MoSe2 nanosheets, the manufacturing of MoSe2 electrodes for extensive industrial use remains complex. Recently, it was found that 3D hierarchical constructions worked well to overcome this obstacle. Carbon fiber paper (CFP) was used to construct independent hierarchical structures for electrocatalysts because of its porous structure, extreme electrical conduction, strong mechanical stability, and exceptional corrosion impedance in acidic and alkaline solutions. Based on this idea, MoSe2 NSs transversally oriented on SnO2 nanotubes were constructed by Huang et al., to improve the amount of unveiled active edges of MoSe2.94 This resulted in a reduction of the Tafel slope of MoSe2@SnO2 NTs to 51 mV dec−1 from 70 mV dec−1 of MoSe2 and reduction of the overpotential at a current density of 10 mA cm−2 to 0.174 V (vs. RHE) from 0.247 V. Despite the increased performance, the electron transport from the electrode to the catalysts is limited by weak conductivity of SnO2 NTs. Therefore, using templates with higher conductivity could enhance the activity even further. On highly conductive NiSe2 porous foam, Zhou et al. produced MoS2(1−x)Se2x particles, which showed significantly improved electrochemical characteristics as depicted in Fig. 13(a–e). According to DFT calculations, the ΔGH on MoS2(1−x)Se2x/NiSe2 in (100) and MoS2(1−x)Se2x/NiSe2 in (110) planes were reduced from 8.4 kcal mol−1 on MoS2(1−x)Se2x to 2.7 kcal mol−1 and 2.1 kcal mol−1, respectively.95 Hence, MoS2(1−x)Se2x and NiSe2 work in concert to significantly increase electrocatalytic activity. This work inspired Zhang et al. to believe that the catalytic performance of highly conductive NiSe2 and 2D MoSe2 nanosheets may be significantly enhanced by integrating them into a 3D self-standing hierarchical framework. He synthesized 3D hierarchical MoSe2/NiSe2 NWs arrays through a two-step hydrothermal technique, which exhibits a Tafel slope of 46.9 mV dec−1 with a small overpotential of 249 mV respectively.96
![]() | ||
Fig. 13 (a) Procedure for developing MoS2(1−x)Se2x on porous NiSe2 foam. (b and c) SEM images of NiSe2 foam expanded at a temperature of 600 °C from viable nickel foam. (d and e) SEM images of MoS2(1−x)Se2x particles spread over the surface of NiSe2 foam at 500 °C (b and d) scale bar, 50 μm, and (c and e) scale bar, 1 μm. Image is reproduced with permission from ref. 88 under Creative Commons Attribution License CC BY, Copyright © Nature Communications 2016. |
MOF-driven metal sulfides were assessed as competitors for modifying the platinum in the water-splitting process, much like the metal selenides. Guan et al. proposed a logical design for new CoS2 nanotube arrays that are put together on a flexible support and may be used right away as a very dynamic bifunctional electrocatalyst for the water-splitting process.97 At the very first moment, uniform wire-like MOF nanoarrays were developed, and then the MOF arrays were converted into CoS2 nanotube arrays by a sulfidation procedure using heat treatment. With a cell voltage of 1.67 V and the use of CoS2 nanotube arrays as catalysts, a water-splitting current density of 10 mA cm−2 in an alkaline solution is attained. The steady current was sustained for 20 hours even when the electrode was bent.
Graphene has been employed in many scientific applications owing to its remarkable surface, conductivity, and exceptional stability, and is still reasonably priced. Because of these features, graphene and graphene-based nanomaterials are now widely used in electroanalytical applications. In graphene/MOF composites, higher rates of electron transfer are seen because graphene functions as a conducting bridge in these composites. Based on this concept, firstly, Jayaramulu et al. explained the construction of a nanofibrous NGO and a nickel-doped MOF (MOF-74) and showed that the MGO/MOF composite exhibits micro–mesoporous behavior, where the mesopores come from the nickel clusters of MOF associating with the O2 and N2 functional groups in the NGO structure, while the micropores come from the pristine MOF.98 Secondly, with thiourea as a sulfur source, he successfully transforms the NGO/MOF hybrid into a multi-sheet 2D nanocomposite (NGO/Ni7S6). In an alkaline solution, the NGO/Ni7S6 composite functions as a perfect bifunctional catalyst for the HER with extreme stability and efficiency. The synergistic effect between Ni7S6 and NGO could be liable for the excellent catalytic activity in mass transfer, durability against corrosion over the HER, the quantity of nitrogen, attainable Ni active sites for superior electrical conductivity, and hierarchical porous construction that facilitates rapid mass movement. In another study, Yan et al. established the preparation of a AuPd-MnOx nanocomposite anchored on ZIF-8-rGO bi-supported through wet-chemical techniques99 as shown in Fig. 14. For the production of hydrogen from FA, the resulting AuPd-MnOx/ZIF-8-rGO exhibits outstanding catalytic behavior, and at 298 K, the turnover frequency (TOF) achieves its maximum value of 382.1 mol H2 catalytst per h devoid of any additives. The tuned Pd structure in the AuPd-MnOx/ZIF-8-rGO hybrids, the reduced dimensions and large diffusion of the AuPd-MnOx composite, and the durable metal–support interrelation among the AuPd-MnOx and ZIF-8-rGO bi-support are the motives for this satisfactory catalytic performance of this composite. Through a two-step hydro and solvothermal procedure, Zhu et al. loaded MoS2 onto the Cu-centered MOF and graphene oxide composite to create effective electrocatalysts for the HER.100 The MoS2/rGO-MOF composite catalysts have a reduced Tafel slope of 36 mV dec−1, a tiny overpotential of 60 mV, and good electrocatalytic HER activity. The increased surface area resulting from the GO-MOF hybrid's mesoporous structure and the synergistic relation among the GO-MOF matrix and MoS2 nanosheets were associated with the improvement in catalytic activity. Yun Liu et al. successfully used a simple in situ space-confined growing approach to create a unique type of MoS2/3D-NPC composite. High electrocatalytic activity for the HER can be achieved by using MoS2 nanosheets grown in the pores of highly conductive 3D-NPC. This is because the 3D hierarchical structure can expose the maximum number of active edge sites, provide a strong binding with the conductive carbon host, and facilitate charge transfer during the electrochemical reaction. With a tiny initial overpotential of about 0.16 V, strong cathodic currents, a small Tafel slope of 5 mV per decade, and good stability in acidic media, this 3D architectural composite shows excellent HER activity.101 Hao Bin Wu et al. provided a new method for creating nanostructured MoCx nano-octahedra as a very effective electrocatalyst for the HER that is aided by MOFs.102 To achieve uniform formation of metal carbide nanocrystallites without coalescence and excessive growth, this strategy depends on the confined and in situ carburization reaction that takes place in a unique MOF-based compound (NENU-5) that consists of a Cu-based MOF (HKUST-1) host and guest Mo-based Keggin POMs residing in pores. These porous MoCx nano-octahedra, which benefit from the desired nanostructure, show excellent electrocatalytic activity for the HER in basic and acidic solutions with good stability.
![]() | ||
Fig. 14 Diagrammatic representation of the synthesis process of the AuPd-MnOx/ZIF-8-rGO composite. Image is reproduced with permission from ref. 93 under Creative Commons Attribution License CC BY, Copyright © Advanced Energy Materials 2015. |
Using a “killing three birds with one stone” approach, Li et al. developed Fe3C/Mo2C-containing N, P, co-doped graphitic carbon from POM@MOF-100 (Fe) (referred to as Fe3C/Mo2C@NPGC).103 One of the best non-noble metal HER catalysts in acidic media reported to date is the Fe3C/Mo2C@NPGC catalyst, which exhibits excellent electrocatalytic activity and stability towards the HER with a low onset overpotential of 18 mV (vs.RHE), a small Tafel slope of 45.2 mV dec−1, and long-term durability for 10 h. Jia Lu et al. used MOFs as functional precursors in a simple pyrolysis process to create multilevel core–shell Au@Zn–Fe–C hybrids.104 A uniform Zn–Fe-MOFs shell coated on an Au nanoparticle was directly pyrolyzed to develop the nanocomposite hybrids, which demonstrated outstanding electrocatalytic performance for the HER with a low onset overpotential of −0.08 V and a stable current density of 10 mA cm−2 at −0.123 V in 0.5 M H2SO4. It was determined that the electrocatalytic activity varied with the loading of the Au nanoparticle cores and the thickness of the metal–carbon shells. This was explained by the conducting shells of Zn–Fe–C for quick electron transport and the creation of active sites for electrocatalytic reactions facilitated by the encapsulated Au nanoparticle.
Pure MOFs' low electroconductivity limits their use in electrochemical domains. To get over this restriction, MXene has become a viable 2D material that can be coupled with MOFs to produce electrocatalysts with excellent surface area, rapid charge transport, and exceptional mechanical stability. Still, the evolution of exceptionally good electrodes is severely constrained in the case where key catalysis mechanisms are poorly understood, and the advancements are still at the synthesis stage. Thus, to expand the applications of electrocatalytic water splitting, a few problems still need to be recognized and resolved. Recently, two-dimensional (2D) MOF nanosheets have been considered ideal electrocatalysts because of their huge number of exposed active metal centers, quick mass and ion transport across their thickness, and porous structure. It is expected that their combination with electrically conductive 2D nanosheets will result in even better electrocatalysis. Regarding this, Zhao and co-workers hybridized two-dimensional cobalt 1,4-benzene dicarboxylate (CoBDC) with Ti3C2Tx (the MXene phase) through an in situ method.105 The resultant hybrid material exhibits a Tafel slope of 48.2 mV per decade in 0.1 M KOH and a current density of 10 mA cm−2 at a voltage of 1.64 V vs. the reversible hydrogen electrode. These outcomes are on par with those attained by the earlier documented transition metal-based catalysts and surpass those achieved by the conventional IrO2-based catalysts. Very recently, Gothandapani et al. synthesized nickel using MXenes (Ni–Ti3C2) via heating of Ni-MOF at a temperature of 650 °C.106 The prepared material was examined through XRD, FTIR, FESEM, and BET analysis. Comparable to Ti3C2, the resultant Ni–Ti3C2 displayed high porosity and a large surface area after calcination. The resulting material was then employed as an electrode for the HER process and examined using several electrochemical techniques, including linear sweep voltammetry (LSV), electrochemical impedance analysis (EIS), and cyclic voltammetry (CV) in an alkaline medium. Due to its surface area and easily obtained active sites, the driven Ni–Ti3C2 showed a small Tafel value of 56.15 mV dec−1 with a voltage of 181.15 mV, delivering superior electron transport to the Ni–Ti3C2 hybrid.
Zong et al. proposed MXene-covered MOF-derived cobalt phosphide (Ti2NTx@MOF-CoP), a bifunctional catalyst with high activity and low cost.107 The Ti2NTx@MOF-CoP material is developed by phosphating ZIF-67, a substrate that assembles with MXene nanosheets. In a broad pH range, the Ti2NTx@MOF-CoP electrode exhibits modest hydrogen evolution capacity at a current density of 10 mA cm−2 and has an inferior overpotential of 112 mV, particularly at pH = 14. They also constructed a dual-electrode for the HER based on the bifunctional activity of Ti2NTx@MOF-CoP in alkaline media. Consequently, this study revealed that CoP generated by MOFs offers a large surface area and strong active sites. As a new representative of the MXene, Ti2NTx could enhance the heterojunction catalyst stability to expand the active site region.
Incorporating the substrate with ZIF-8 for the HER in an acidic medium, Hao et al. reported the production of a Ti3C2Tx NS integrated MOF catalyst (Ti3C2Tx@ZIF-8) with extreme catalytic performance and inexpensiveness.108 The catalyst has a lower Tafel slope value of 77 mV dec−1 and an overpotential of only 507 mV at 20 mA cm−2, which is shown in Fig. 15 (a–d). The active surface area (ECSA) of 122.5 cm2 is also indicated by CV and chronopotentiometry, which shows that the catalysts are stable over 20 hours with no appreciable variations in the overpotential value.
![]() | ||
Fig. 15 (a) LSV analysis of ZIF-8, Ti3C2Tx, and Ti3C2Tx@ZIF-8 for the HER process. (b) Tafel slope values of ZIF-8, Ti3C2Tx, and Ti3C2Tx@ZIF-8. (c) Stability and durability of the Ti3C2Tx@ZIF-8 electrocatalysts. (d) Histogram for the corresponding overpotential values. Image is reproduced with permission from ref. 97 under Creative Commons Attribution License CC BY, Copyright © Catalysts 2023. |
The combination of CNTs and MOFs results in CNT@MOF hybrids, which have a porous framework that facilitates electrolyte transport through conductive pathways. The addition of CNTs to the CNT@MOF composite, enhanced its conductivity and imparted long-range electron transport capability to the hybrid material. Furthermore, under certain circumstances, by increasing the ratio of surface area to active sites, CNTs help improve the composite's catalytic process, which expands the electrocatalytic activity of MOFs. When electrically conductive CNTs are combined with extremely porous, huge-surface-area MOFs, active CNT@MOF hybrids are produced that function in electrochemical methods. According to Wu et al., cobalt phosphide polyhedral NPs and CNTs were coupled using a MOF templated approach.109 CoP-CNT hybrids as synthesized exhibit outstanding HER performance because of the synergic catalytic effect among CoP polyhedra and CNTs. In 0.5 M of H2SO4, the resulting CoP-CNTs exhibit superior catalytic activity, with an inferior overpotential of ∼139 mV at 10 mA cm−2, a tiny onset overpotential of ∼64 mV, and a Tafel slope of 52 mV dec−1. The catalysts also exhibit exceptional endurance in acidic solutions.
Transition metal oxides (TMOs) were highly important because of their strong structural stability and superior internal affinity for hydrogen-containing intermediates. Also, by altering the three-dimensional electronic structure of the core metal atom, heterostructure construction could produce an active surface site/relative to the two stages, to enhance electrocatalyst activity. Based on this concept, in recent times, a nanoparticle enriched MOF has been produced by Zhang et al., to synthesize CoFeOx nanoparticles to build a single-layer benzimidazole-based Co-MOF (M-PCBN)110 With overpotentials of 232, 316, and 348 mV at 10 mA cm−2, M-PCBN showed higher electrocatalytic performance than the massive and pristine MOF, highlighting the importance of including highly catalytically active NPS. A reduced Tafel slope of 32 mV dec−1, and a 60-hour chronoamperometric experiment at 1.48 V vs. RHE revealed that M-PCBN exhibited significant stability.
Transition metal sulfides (TMSs) have received attention in electrocatalysts in recent times, due to their excellent redox properties, distinct stoichiometric proportions, high crystallinity, and extremely stable structure. To sustain the reaction intermediates and improve HER capabilities, the extremely electronegative sulfide in TMSs removes electrons from the metal and active sites. Still, several documented TMS-based electrocatalysts have issues such as low conductivity, inferior active sites, and unstable working efficiency. A heterostructure is generally a sensible option for increasing the active area to improve the interfusion effect. Furthermore, by significantly changing the localized ion and electron transmission behavior, the electron or atomic arrangement at the hetero-unique interface might improve HER catalytic activities. MoSx on UiO-66-NH2 was demonstrated by Dai and colleagues via a solvothermal approach.111 At a current density of 10 mA cm−2, the unique construction of the UiO-66-NH2- stabilized MoSx-based material showed exceptional HER activity with a Tafel slope of 59 mV dec−1, a potential of 125 mV, and an overpotential of 200 mV. Additionally, it showed exceptional stability in an acidic medium. Its numerous active regions, wide surface area, and rapid electron and proton transport in MOF were the causes of its remarkable HER performance. Qiao and Colleagues recently developed unique 2D hybrid NSs combining MOF-Co-BDC and MoS2 nanosheets with the electrochemical activity of manufacturing catalysts by a sonication-assisted process.112 The Co-BDC/MoS2 electrocatalysts showed significant reaction activity with a Tafel slope of 86 mV dec−1 and an overpotential of 248 mV. Additionally, pristine MoS2 and Co-BDC overpotentials were 349 mV and 529 mV. Furthermore, the nanosheets showed a high degree of dependability for 15 hours at a constant current density of 10 mA cm−2. Co-BDC decorates the semiconducting 2H MoS2 phase, resulting in a notable phase change to the metallic 1T-MoS2, which is more suited for the HER process.
Zhen-Feng Huang et al. used a simple self-template strategy to develop hollow Co-based bimetallic polyhedra for application in the HER. In this work, solvothermal sulfidation and thermal annealing convert homogeneous bimetallic metal–organic frameworks into hollow bimetallic sulfides. Co3S4's HER activity is greatly increased by the combination of the hollow structure and homo-incorporation of a second metal, according to electrochemical experiments and density functional theory calculations. In particular, the Co3S4 lattice's homogeneous doping enhances electrical conductivity and maximizes the Gibbs free energy for H* adsorption. Over a broad pH range, hollow Zn0.30Co2.70S4 demonstrates electrocatalytic HER activity better than that in the majority of the reports of noble-metal-free electrocatalysts. In 0.5 M H2SO4, 0.1 M phosphate buffer, and 1 M KOH, the overpotentials are 80 mV, 90 mV, and 85 mV at 10 mA cm−2, and 129 mV, 144 mV, and 136 mV at 100 mA cm−2 respectively (Table 1).113
Electrocatalyst | Electrolyte | Tafel slope (mV dec−1) | η @ j (mV@mA−2) | Ref. |
---|---|---|---|---|
AB&CTGU-5 | 0.5 M H2SO4 | 45 | 44 | 67 |
Ni-MOF | 0.5 M H2SO4 | 60 | 350 | 68 |
Co-MOF | 0.5 M H2SO4 | 121 | 101 | 69 |
2DSP | 0.5 M H2SO4 | 80.5 | 333 | 70 |
AB & Co–Cl4-MOF (3![]() ![]() |
0.5 M H2SO4 | 86 | 283 | 71 |
Py-ZIF | 0.5 M H2SO4 | 84 | 260 | 73 |
CuCo@NC | 0.5 M H2SO4 | 79 | 115 | 74 |
NiOx@BCNTs | 2 M KOH | 119 | 79 | 76 |
Co@Co3O4-NC | 1 M KOH | 77.3 | 221 | 77 |
CoP@NC/CF | 1 M KOH | 64.8 | 151.3 | 78 |
CoP/NCNHP | 0.5 M H2SO4 | 70 | 310 | 79 |
CoP NFs | 0.5 M H2SO4 | 49.6 | 122 | 80 |
Ni2P NS | 1 M KOH | 63 | 168 | 82 |
Ni2P/C | 0.5 M H2SO4 | 113.2 | −94 | 83 |
Ni2P NP | 0.5 M H2SO4 | 62 | 270 | 84 |
Co2Ni1N | 0.5 M H2SO4 | 60.17 | 102.6 | 86 |
Fe–CoSe2@NC | 0.5 M H2SO4 | 40 | −143 | 87 |
CoSe2@DC | 0.5 M H2SO4 | 82 | −40 | 88 |
NiW-CNT/PC/CC | 1 M KOH | 112.9 | 45 | 89 |
Ni(OH)2/NiSe2/CC | 1 M KOH | 60 | 82 | 90 |
NiSe2/Ni | 0.5 M H2SO4 | 49 | 143 | 91 |
H-NiSe2 | 0.5 M H2SO4 | 42.6 | −107 | 92 |
HP-NiSe2 | 0.5 M H2SO4 | 43 | −57 | 92 |
Ni0.8Fe0.2Se2/CFC | 0.5 M H2SO4 | 43 | 64 | 93 |
SnO2@MoSe2 | 0.5 M H2SO4 | 51 | 174 | 94 |
MoS2(1−X)Se2X/NiSe2 | 0.5 M H2SO4 | 42.1 | −69 | 95 |
MoSe2/NiSe2 NWs | 0.5 M H2SO4 | 46.9 | 249 | 96 |
MoSe2 | 0.5 M H2SO4 | 69.2 | 305 | 96 |
NiSe2 | 0.5 M H2SO4 | 68.8 | 345 | 96 |
CoS2 NTA/CC | 1 M KOH | 88 | 276 | 97 |
Co3O4 NWA/CC | 1 M KOH | 90 | 350 | 97 |
NGO/Ni7S6 | 0.1 M KOH | 45.4 | 161 | 98 |
MoS2/rGO-MOF | 0.5 M H2SO4 | 36 | 60 | 100 |
Ni–Ti3C2 | 0.5 M H2SO4 | 56.15 | 181.1 | 106 |
Ti2NTx@MOF-CoP | 0.5 M H2SO4 | 96.7 | 129 | 107 |
MOF-CoP | 0.5 M H2SO4 | 109.4 | 203 | 107 |
ZIF-67 | 0.5 M H2SO4 | 127.2 | 256 | 107 |
Ti2NTx | 0.5 M H2SO4 | 172.8 | 401 | 107 |
Ti2AlN | 0.5 M H2SO4 | 245.1 | 545 | 107 |
Ti3C2Tx@ZIF-8 | 1 M KOH | 77 | 507 | 108 |
CoP-CNTs | 0.5 M H2SO4 | 52 | 139 | 109 |
M-PCBN/CC | 1 M KOH | 32 | 232 | 110 |
Zr-MOF | 0.5 M H2SO4 | 59 | 125 | 111 |
Co-BDC/MoS2 | 1 M KOH | 86 | 155 | 112 |
Zn0.30Co2.70S4 | 0.5 M H2SO4 | 47.5 | 80 | 113 |
CoP–FeP | 0.5 M H2SO4 | 45.1 | 163 | 113 |
MoS2/3D-NPC | 0.5 M H2SO4 | 51 | 160 | 101 |
MoCx nano-octahedrons | 0.5 M H2SO4 | 53 | 142 | 102 |
CoP/GO-400 | 1 M KOH | 135 | 470 | 81 |
Fe3C/Mo2C@NPGC | 0.5 M H2SO4 | 45.2 | 18 | 103 |
Au@Zn–Fe–C | 0.5 M H2SO4 | 130 | 80 | 104 |
To address these challenges, researchers are focusing on several strategies. These include developing MOFs with chemically robust linkers, enhancing the stability of MOF-derived materials through protective coatings or encapsulation strategies, and optimizing interfaces in composite materials to withstand operational stress. Additional efforts involve creating conductive MOFs by integrating highly conductive constituents like graphene, carbon nanotubes, or metallic nanoparticles with redox-active or π-conjugated linkers. Engineering MOFs with hierarchical porosity can improve mass transport, while controlled pyrolysis techniques help to preserve or expose active sites in derived materials. Designing composites with well-distributed and accessible catalytic centres is also critical. Addressing these challenges will not only unlock their potential in electrocatalysis but also expand their applicability to other areas like energy storage, gas separation, and sensing.
This journal is © The Royal Society of Chemistry 2025 |