Next Article in Journal
A Water-Based Fire-Extinguishing Agent of Lithium Iron Phosphate Battery Fire via an Analytic Hierarchy Process-Fuzzy TOPSIS Decision-Marking Method
Previous Article in Journal
A Multi-Encoder BHTP Autoencoder for Robust Lithium Battery SOH Prediction Under Small-Sample Scenarios
Previous Article in Special Issue
Thermally Stable Carbon Materials from Polybenzoxazines: Structure, Properties, and Supercapacitor Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Application of MOFs in Supercapacitors

by
Christos Argirusis
1,2,*,
Maria-Eleni Katsanou
2,
Niyaz Alizadeh
3,
Nikolaos Argirusis
3 and
Georgia Sourkouni
2
1
Laboratory of Inorganic Materials Technology, School of Chemical Engineering, National Technical University of Athens, 15773 Athens, Greece
2
Clausthal Centre of Materials Technology (CZM), Clausthal University of Technology, 38678 Clausthal-Zellerfeld, Germany
3
MAT4NRG—Gesellschaft für Materialien und Energieanwendungen mbH, Burgstätter Str. 42, 38678 Clausthal-Zellerfeld, Germany
*
Author to whom correspondence should be addressed.
Batteries 2025, 11(5), 181; https://doi.org/10.3390/batteries11050181
Submission received: 5 March 2025 / Revised: 22 April 2025 / Accepted: 23 April 2025 / Published: 2 May 2025
(This article belongs to the Special Issue High-Performance Supercapacitors: Advancements & Challenges)

Abstract

:
As the need for energy is constantly increasing and in the long term fossil fuels are not an option because of global overheating due to the greenhouse effect, alternative energy production concepts such as photovoltaics, wind energy, IR energy harvesters etc., have been developed. The problem is that renewable energy sources are stochastic, and therefore there is a need for electrical energy storage either in rechargeable batteries or in high-performance supercapacitors. In this respect, novel materials are needed to meet the challenges that are related to these technologies. Metal–organic frameworks (MOFs) represent highly promising materials for energy storage applications in supercapacitors (SCs) and thus in recent years have become essential for clean and efficient energy conversion and storage. Metal–organic frameworks (MOFs) present numerous benefits as electrocatalysts, electrolyte membranes, and fuel storage materials; they exhibit exceptional design versatility, extensive surface-to-volume ratios, and permit functionalization with multivalent ligands and metal centers. Here we present an overview of MOF-based materials for electrical energy storage using high-performance supercapacitors. This review deals with recent advances in MOF-based materials for supercapacitors. Finally, an outlook on the future use and restrictions of MOFs in electrochemical applications, with focus on supercapacitors, is given.

1. Introduction

It is estimated that the world will need up to 500 EJ/year by 2050. The need for clean and renewable sources of energy is and will continue to be the most compelling task of 21st-century science [1]. The worldwide urbanization and increasing population have led to a substantial rise in the production of greenhouse gases, particularly CO2. There is agreement about the need to reduce greenhouse gases and to shift energy production to green and renewable, so-called alternative sources, like photovoltaics and wind energy. As these sources are not producing continuous electrical energy, there is a need for storage of the produced and not immediately used excess electrical energy. This can be realized by using advanced batteries and supercapacitors [2].
As civilization increasingly relies on electricity-based technologies, energy demand will inevitably rise. The sustainable utilization of green (renewable) energy sources, like solar, wind, and tidal energy, are becoming more common. These energy forms are dependent on natural fluctuations and must be converted into electricity and stored. On the other hand, the decrease in fossil fuel sources and increase of carbon footprint of these sources resulted in the requirement of new and renewable energy.
In recent years, supercapacitors (SCs) have attracted significant interest as an innovative energy storage solution [3,4]. One distinguishes two main types of supercapacitors: electric double layer capacitors (EDLCs) and pseudocapacitors [5]. The materials used in electric double layer capacitors (EDLC) exhibit a high specific surface area, and the charge storage mechanism is based on surface phenomena. This leads to a rapid process that enables high power density and cycling stability, albeit at the cost of lower energy density [6].
Among the two principal storage processes for supercapacitors, the electrochemical double-layer capacitor (EDLC) depends on charge adsorption and desorption at the carbon electrode/aqueous electrolyte interface, without demonstrating a Faradaic reaction within its operational potential range. In the mechanism utilizing the pseudocapacitor principle, energy storage transpires via surface redox processes or Faradaic charge transfer reactions involving transition metal oxides or hydroxides and an aqueous electrolyte [7,8]. Pseudo-capacitors (PCs) were developed by integrating electric double layer (EDL) and Faradaic processes within the electrode materials. The electrochemical properties of PCs encompass Faradaic processes and have a capacitive signature; thus, they do not contain exclusively capacitive or Faradaic characteristics. Pseudocapacitance arises from surface-bound redox reactions and enhances total capacitance [9,10]. The pseudocapacitor is a type of capacitor capable of storing larger charges than EDLCs, providing superior energy density due to redox reactions taking place at the material’s surface or within its interior [11]. Electric double-layer capacitors (EDLCs) store charge by accumulating ions at the interface of the electrode and electrolyte, resulting in a high-power density. Pseudocapacitors (PCs) predominantly depend on electron transfer at the electrode/electrolyte interface for charge storage, resulting in a higher energy density than electric double-layer capacitors (EDLCs) [12]. In Figure 1, charge storage processes in EDLCs, PCs, and battery resembling electrodes are presented. The quasi-rectangular cyclic voltammetry curve of PCs resembles capacitive electric double layer feedback; nevertheless, it arises from many Faradaic redox processes with a potential distribution. The redox apex must accurately coincide without any peak-to-peak gap. Given that PCs are not physical storage devices yet exhibit linear GCD curves and nearly rectangular CV profiles akin to EDLC, they may be regarded as a complementary variant of EDLC [9,13].
Metal–organic frameworks (MOFs) represent a novel category of materials that have attracted considerable interest over the past 20 years [14,15]. MOFs are open networks composed of metal-centered secondary building units (SBUs) interconnected by organic linkers, resulting in extensive one-dimensional (1-D), two-dimensional (2-D), or three-dimensional (3-D) structures [16,17]. The impact of incorporating ligands on the chemical stability of MOFs can be demonstrated through the application of density functional theory (DFT) [18]. The structures possess a crystalline character, exhibiting long-range organization. Exceptionally uniform pores or channels are inherently present within the framework, typically accommodating guest entities such as solvent molecules (introduced during the synthesis process) or counter ions that neutralize the overall charges on the framework (resulting from the charged metal nodes) [14,15,19]. In contrast to other porous materials like zeolites and carbons, the essential and distinguishing characteristic of MOFs that makes them unique and highly functional is the ability to achieve a “directed” structure by meticulous selection of metals and organic linkers [20,21,22]. In comparison to traditional supercapacitor materials, the distinctive characteristics of metal–organic frameworks (MOFs), including their extraordinarily high surface area, adjustable porosity, uniformly structured nano and microscale cavities, potential for in-pore functionality and external surface modification, excellent thermal stability, and the variety of available metal and functional groups, make them highly appealing for electrical energy storage. Furthermore, an endless array of potential combinations of metal ions and organic linkers enables the synthesis of MOFs with diverse characteristics [23].
Metal–organic frameworks (MOFs) have the benefit of tunable morphology, enabling the exact synthesis of carbon-based materials with customized structures that exhibit enhanced electrical conductivity, variable porosity, extensive surface area, and robust stability [24]. Consequently, bimetallic MOFs have attracted significant scientific attention due to their remarkable electrical conductivity, strong structural stability, and straightforward production. Liang and coworkers [25] employed a hydrothermal technique to synthesize NinComMOFs characterized by a layered and channelized structure. The capacitance of Ni1Co1MOF is 1333 F g−1 at 2 A g−1, with Ni1Co1MOF//AC demonstrating a high energy density of 28 Wh/kg at 444 W kg−1. This value exceeds the capacitance of the individual metal MOFs (Ni-MOF and Co-MOF). The bimetallic oxide ZnCo2O4 synthesized from bimetallic cobalt-based metal–organic frameworks (MOFs) offers a 300% higher specific capacity compared to Co3O4 generated from MOF along with improved stability [26]. Zhao and coworkers [27] have prepared iron-cobalt oxide nanowires strongly bound on gold-modified porous carbon nanosheets (CPCN) derived from the bimetallic MOFs. A three-electrode design demonstrated exceptional electrochemical performance. The FCO/Au/CPCN@CC electrode exhibits a consistent cyclic voltammetry profile at 4800 mV s−1, achieving a peak energy density of 1090.8 W h kg−1 at 571.43 W kg−1, and an energy density of 344.89 W h kg−1 at the maximum power density of 4000 W kg−1. After 10,000 cycles of GCD, the capacitance ratio is 84.27%.
Optimizing the architectures, compositions, and morphologies of MOFs is very important for enhancing the supercapacitor performance. Otun et al. [28] employed a ligand-engineering technique to synthesize ZnCo-bimetallic MOFs with distinct characteristics utilizing three different ligands (2-methylimidazole, terephthalic acid, and 2-amino terephthalic acid) under a uniform synthesis process. The disparity in the electron-donating capacity of the three ligands resulted in alterations to their structural, morphological, and electrochemical characteristics. In comparison to other metal–organic frameworks (MOFs), the imidazole-based ZnCo-MOF (ZnCo-MOF-HMIM), characterized by its dodecahedral morphology, substantial specific surface area, and mild pore attributes, offers significant electron transport pathways for ion migration at the electrode surface, hence ensuring enhanced charge storage capacity.
Besides the classical solvothermal preparation method [29], several green synthesis methods [30] became in recent years more and more attractive [31]. Methods like electrochemical [32], sonochemical [33,34], microwaves [35], and mechanochemical [36,37] often resulting to different properties and micro-structure of the same MOF [31,38], have been used for the preparation of MOF and descendant products [39].
While the solvothermal method is considered the easiest to use in the lab, as it needs only a furnace, the duration of the solvothermal preparation method is very long (often more than 1 day). Therefore, microwave reactors are often used as they reduce the reaction time from days down to hours. The fastest MOF preparation method is the sonochemical method (Figure 2), as the reaction time is sometimes reduced to less than 1 h [40].
The sonochemical method produces differently shaped powders that are mostly finer structured than the ones made using the conventional solvothermal method (Figure 3).
Metal–organic frameworks (MOFs) offer suitable space for electrochemical reactions and ion storage/transfer due to their extensive surface area, making them highly attractive candidates for electrochemical catalysis and energy storage (Figure 4). Compared to other porous materials such as activated carbon and zeolites, MOFs exhibit enhanced architecture and functionalities. Nonetheless, the low conductivity (less than 10−10 S cm−1) of most MOFs impedes their practical electrochemical applications, necessitating enhancement. Materials having a high density of charge carriers can function as metallic conductors or semiconductors [41,42]. MOF materials offer significant potential for enhancing electronic and ion conductivity due to their functional pore surface, high surface areas, and extensive structural tunability [43,44]. A review on the development of conductive MOFs can be found in [45]. A review on comprehensive enhancement strategies for electronically and ionically conductive MOFs, as well as recent research progress, is given by Zhang and coworkers [41].
The extensive research on metal–organic frameworks (MOFs) in the last 10 years has resulted in the publication of over 20,000 papers [46]. Theoretical approaches are often employed alongside experimental characterization techniques to investigate newly synthesized materials and elucidate their properties at the microscopic level. Computational methods, alongside various experimental techniques, have been widely employed for the characterization and elucidation of the properties of MOFs. The mathematical simulation of functionalized metal–organic frameworks (MOFs) for energy and catalytic applications encompasses four primary objectives: (i) elucidating the structure and potential post-modification structural alterations of these materials and their correlation to reactivity; (ii) comprehending the energetics and intricacies of chemical dynamics at catalytic sites, including the modeling of potential energy surfaces for competing mechanisms; (iii) uncovering and elucidating fundamental structure−function relationships that may facilitate further catalyst discovery; and (iv) forecasting novel materials with enhanced catalytic properties [47]. Coudert and Fuchs [48] examine methods for predicting crystal structures, geometrical properties, and conducting large-scale screenings of theoretical MOFs, in addition to their thermal and mechanical properties. A comprehensive analysis of numerous MOFs may yield insights into structure–function relationships. These relationships can be utilized by synthetic chemists for the targeted synthesis of novel materials, or by computational chemists for the optimization of specific properties through extensive computational screening studies, potentially resulting in the discovery of new functional MOFs [49].
The quantum mechanics method (QMM), namely Kohn–Sham (KS) density functional theory (DFT) and different post-Hartree–Fock (Post-HF) approaches, is too expensive for large systems [50]. The computational modeling of MOFs, particularly for large-scale high-throughput screening, is predominantly performed using the molecular mechanics method (MMM), which relies on forcefield parameters. This approach utilizes fitted experimental data or high-level quantum mechanical molecular (QMM) results as the energy expression for the potential energy surface (PES) [51]. This forcefield-based approach eliminates the need for time-intensive self-consistent iterations, resulting in a significant increase in speed. Hai and Wang [52] assert that the theoretical examination of electron motion phenomena must be grounded on Quantum Mechanical Models, such as Density Functional Theory (DFT) or Hartree–Fock (HF).
Despite the increasing computational power, the vast MOF material space necessitates more efficient methodologies for the computational identification of potential MOFs in terms of time and effort. Utilizing data-driven scientific methodologies can provide significant advantages, including expedited design and discovery processes for metal–organic frameworks (MOFs) through the development of machine learning (ML) models and the analysis of intricate structure-performance correlations that surpass expert intuition [53]. ML could also help to identify materials compositions for the improvement of SC electrodes and their combination to more efficient supercapacitors.
The scope of the presented manuscript is to give an overview of recent advances in the use of MOFs in electrochemical devices for electrical energy storage in high-performance supercapacitors. Significant obstacles, limitations for large-scale MOF applications, and prospective developments of MOF-based conductors are also examined.

2. MOF-Based Materials for Supercapacitors

Supercapacitors are nowadays accepted as a viable energy storage solution due to their distinctive characteristics, including rapid charge and discharge rates, high power density, and enhanced cycle efficiency in comparison to rechargeable batteries. Furthermore, SCs are environmentally sustainable devices, rendering them suitable for the demands of sustainable energy storage systems. A supercapacitor, akin to other electrochemical systems, comprises two electrodes, an inorganic (aqueous) or organic electrolyte, and a separator. Electrodes are essential for regulating the performance of the supercapacitor and may be uniform in symmetric cells or distinct in asymmetric cells. Furthermore, high conductivity is essential for high-performance supercapacitor electrode materials. Temperature stability, large specific surface area, pore size distribution optimization, manufacturing ease, corrosion resistance, and cost-effectiveness are all factors to be considered. The choice of suitable materials and the refinement of electrode design are essential approaches to enhance the effectiveness of supercapacitors as energy storage devices [54].
Metal–organic frameworks (MOFs), as highly porous and crystalline solids, have demonstrated incredible value as models for synthesizing functional materials suitable for the fabrication of positive electrodes in supercapacitors (SCs) [55,56]. This is primarily attributable to their extensive surface area, adjustable pore size, diverse synthetic methodologies, potential for post-synthetic modification, and the feasibility of scalable production for specific types [57,58]. Nonetheless, their inadequate electrical conductivity and hindrance to ion insertion remain significant obstacles to the direct utilization of MOFs as supercapacitor electrodes [59]. Structural stability is a further challenge, as metal–organic frameworks (MOFs) typically lack stability in conventional electrolyte solutions used in supercapacitors (SCs). Consequently, MOF composites compared to single metal or bimetallic MOFs, typically reduce disparity by providing adjustable electrochemical activity, enhanced charge capacity, and superior electrical conductivity [60]. However, the research on MOF composites incorporating metal oxides or sulfides/phosphides for such purposes remains sparse. Conversely, composites generated from MOFs have gained significantly greater popularity by utilizing their initial captivating morphologies for templated synthesis [61,62,63].
Li and colleagues [64] offer significant insights into the development and preparation of high-performance HSC-positive electrode materials for practical applications. They have successfully created self-assembled carbon-coated multinuclear Ni9S8@C-5 composites utilizing spherical Ni–MOF as a structural template through a straightforward high-temperature sulfuration process, demonstrating excellent battery-type energy storage characteristics. (1) The oxidation-rich multinuclear Ni9S8 exhibits significant redox activity, establishing a solid foundation for redox reactions in electrode materials; (2) The development of the spherical outer carbon layer effectively prevents structural collapse and agglomeration of Ni9S8 crystals, thereby improving its structural stability; (3) The Ni–MOF-derived carbon skeleton, characterized by a specific surface area of 132.66 m2g−1 and a mesoporous structure of 3–6 nm, is well-suited for electrochemical reactions and substance transport diffusion. Ni9S8@C-5 demonstrates impressive specific capacity (278.06 mAh g−1 at 1 Ag−1) and remarkable cycling stability (capacity retention of 88.97% over 5000 cycles). The constructed Ni9S8@C-5//CTP-AC HSC exhibits an energy density of 69.32 Whkg−1 at a power density of 800.06 Wkg−1, surpassing the values of previously documented MOF-derived nickel sulfide compounds [65,66,67].

2.1. MOFs and Metal Oxides

The synthesis of MO@MOF composites is intricate, necessitating precise conditions to attain uniform metal oxide distribution and robust interfacial contact. Moreover, the addition of metal oxides may diminish the entire porosity of the MOF by obstructing pores, potentially constraining its efficacy in adsorption-based applications. Furthermore, synthesis approaches such as hydrothermal or microwave-assisted procedures might be expensive for energy consumption and equipment, presenting obstacles for large-scale manufacturing. Nonetheless, Metal–oxide@MOF (MO@MOF) composites provide considerable benefits in diverse applications but accompanied by certain obstacles. The use of metal oxides improves the stability of MOFs, rendering the composites appropriate for elevated temperatures and more rigorous environments. This combination enhances catalytic activity by increasing the number of active sites, which is beneficial for pollutant degradation, energy storage, and photocatalysis. Moreover, MOFs provide substantial surface area and porosity, whereas metal oxides impart distinctive capabilities, including electrical, optical, and magnetic properties, hence expanding their applicability in domains such as sensors, batteries, and supercapacitors [68].
The first attempt to create a MOF–oxide composite for supercapacitor applications utilized SnO2 quantum dots, which were deposited on ZIF-8 by an in situ epoxide precipitation technique [69]. SnO2 exhibits excellent chemical stability, transparency, and both optical and electrical properties, with prior studies demonstrating that quantum dots demonstrate pronounced oxidation and reduction peaks, signifying that Faradaic processes are involved [70]. The capacitance properties of the composite electrode (931 F g−1) were better than those of pristine ZIF-8 (99 F g−1) and SnO2 quantum dots (241 F g−1), achieving about 500 F g−1 after 500 cycles.
Subsequently, attempts were undertaken utilizing manganese oxide, an interesting metal oxide with a high theoretical specific capacity (1370 F g−1) and widely recognized in electrochemistry [71], to develop NiO-Ni-r-Graphene Oxide composites [72], as well as Prussian blue (PB) analog composites [56,73]. MnOx flower-like structures were synthesized in situ on manganese hexacyanoferrate nanocubes through the incorporation of NH4F, resulting in a fourfold increase in specific capacitance (1200 F g−1) relative to the pure MOF [74]. The composite had promising outcomes in a flexible supercapacitor, achieving an areal capacitance of 175 mF cm−2 at 0.5 mA cm−2, surpassing the performance of other comparable materials available at that time. In a similar manner, oxide nanolayers were produced on the Prussian Blue microcubes [73] using chemical immersion deposition with KMnO4 and PEG (polyethylene glycol) for the reduction reaction. The ASC utilizing a hybrid positive electrode and polyaniline (PANI)/graphene nanoplatelets demonstrated an energy density of 16.5 Wh kg−1 at 550 W kg−1 (1 A g−1). Furthermore, EIS study demonstrated favorable and stable performance for at least 4000 cycles.
Carbon fabric is often used in electrochemical applications, facilitating either the deposition of the catalyst or its in situ growth. Co3O4 was directly synthesized on carbon cloth, which was subsequently immersed in the MOF precursor solution to fabricate core-shell Co3O4/Ni-BDC [75]. The hybrid battery-SC with the layered structure-based positive electrode demonstrated superior specific capacity (209 mAh g−1) with a retention of 90% over 3000 cycles, in contrast to the sample lacking the oxide (154 mAh g−1) at 1 A g−1. Recently, Zhang et al. [76] altered the cobalt oxide synthesis by calcining ZIF-67, resulting in ribbon-like Co3O4 nanoarrays (Figure 5).
Binary metal oxides have been intensively investigated for energy storage applications due to their structural stability and commendable reversible capacity [77,78]. NiCo2O4 is environmentally friendly and has proven effective as an electrode option owing to its ability to facilitate battery-type Faradaic redox processes [79], while exhibiting superior electronic conductivity compared to both cobalt and nickel oxide [80]. A straightforward solvothermal method was employed to synthesize core-shell NiCo2O4@Ni-MOF electrodes, utilizing carbon cloth as the substrate for the growth of Ni-BDC nanosheets on NiCo2O4 nanowires (NWs) [81]. The composite exhibited a substantial areal capacitance of 7.2 and 5 F cm−2 at current densities of 5 and 30 mA cm−2. Additionally, the bimetallic oxide was employed to synthesize NiCo2O4@polypyrrole as the negative electrode for a hybrid supercapacitor, demonstrating encouraging outcomes with approximately 87% retention after 5000 cycles.
Xiong et al. [82] synthesized a nickel-based composite to mitigate disorderly arrangements during MOF growth by considering other upgrades of nickel-based MOFs, such as metal doping [83] and the incorporation of MWCNTs [84]. They employed NiO/NF as a sacrificial template and precursor to fabricate a well-aligned porous NiO@Ni-BTC/NF cylindrical cage structure. The nickel foam (NF) substrate facilitated electron conductivity and surface area, while the exposed nickel ions’ active sites acted as connecting joints for further MOF growth. The composite exhibited an exceptional specific capacity of 1853 C cm−2 at 1 mA cm−2, much surpassing that of NiO/NF, which is 242 C cm−2. The hybrid supercapacitor with carbon nanotubes as the negative electrode exhibited a specific capacitance of 144 F g−1 and maintained 94% capacity retention after 3000 cycles. Li et al. [85] similarly produced NiO nanoparticles on hexagonal Ni-BTC using the calcination of the metal–organic framework (Figure 6). The TG analysis indicated the temperature at which weight loss commenced; hence, by regulating the calcination temperature at 350 °C, just a portion of the MOF converted to oxide. The distribution of NiO nanoparticles within the metal–organic framework not only enhanced electrolyte ion movement but also augmented the redox active sites. The composite electrode exhibited a better specific capacity (approximately 1100 F g−1) for 5000 cycles at 0.5 A g−1, in contrast to the non-modified MOF (approximately 900 F g−1) and the calcined variants at 450 °C (approximately 250 F g−1) and 550 °C (approximately 110 F g−1). Furthermore, the Ni–MOF@NiO//AC supercapacitor demonstrated a peak specific capacitance of 228.6 F g−1 and an energy density of 62.2 Wh kg−1.
Efforts have been made to employ iron oxides among the many metal oxides utilized in SC composite electrodes. Notwithstanding their low costs and low toxicity, nanoparticles experience aggregation and elevated resistance, which constrains their electrochemical uses [86]. Chameh et al. [87] have synthesized iron oxide composites utilizing ZIF-8 and/or ZIF–67 through decoration using Fe2O3 or applying a layer of ZIF onto Fe3O4 to create a core-shell structure [88]. The ZIF–8 composite exhibited superior specific capacitance at 2.7 A g−1 (1160 F g−1) compared to the ZIF–67 counterpart (573 F g−1), including at elevated current densities. Additionally, the ZIF–8/Fe2O3//AC supercapacitor achieved 28.5 Wh kg−1 at 2.4 kW kg−1, with a 97% retention after 1500 cycles. Conversely, the synthesis of core-shell structures enhanced performance, resulting in ZIF–67 achieving 1334 F g−1, surpassing ZIF-8’s 870 F g−1 at 1 A g−1. The Fe3O4@ZIF-67//AC device demonstrated 27.9 Wh kg−1 at 5.49 kW kg−1, with an 87% retention over 3000 cycles. A magnetic Fe2O3@SiO2 core-shell structure was also synthesized using a solvothermal technique, with a Ni–BTC shell [89]. Upon evaluation as an electrode, it demonstrated a capacitance of 600 F g−1 at a current density of 1.5 A g−1 and exhibited a degradation of 5% following 1000 charge-discharge cycles.
ZnO has proven to be a compelling oxide in the fields of photocatalysis [90,91] and electrocatalysis [92,93]. The ZnO carbon composites have potential outcomes as semiconductor materials due to their thermal and chemical resilience, as well as their electronic conductivity [94]. Liu and coworkers [95] reported an electrode based on PANI/ZnO/ZIF–8/graphene/polyester-textile concerning MOF composites. ZnO was originally coated onto the graphene/polyester substrate and then served as a precursor for the in situ coating of ZIF–8, while PANI was electrochemically deposited. The electrode exhibited remarkable capacity (1378 F cm−2 at 1 mA cm−1) by utilizing the internal ZIF surface area (Electric Double Layer Capacitance (EDLC)) and the PANI electron bridge (pseudocapacitance). The capacitance retention decreased to 73% after 400 cycles, likely due to the volume change of PANI during the charge-discharge process. Furthermore, the built flexible supercapacitor exhibited a commendable energy density of 235 mWh cm−3 at 1542 mW cm−3. Zhu and colleagues have similarly utilized a carbon fabric substrate [96]. The alteration in CV cycles between 5 and 13 influenced the electrodeposition of PANI directly by increasing its thickness and was crucial to the end-product characteristics.
Εlectrochemical investigation confirmed the optimal performance achieved (340.7 F g−1 at 1 A g−1 and 82.5% retention after 5000 cycles at 2 A g−1). Without ZIF–8 in the electrode, the retention decreased to 61.4%, underscoring its essential function in the composite. Zuhri et al. [97] enhanced the supercapacitive performance of a ZnO–FC (functionalized carbon) electrode by including a Ni–Co MOF derived from phthalic acid and pyrazine by a straightforward stirring-blending technique. The MOF implementation indeed enhanced the performance from 1.15 to 2.56 (10 wt%) and 6.62 F g−1 (20 wt%), although the results remained subpar in comparison to analogous materials.
TiO2 has predominantly been investigated for solar and photocatalysis applications owing to its stability, low costs, and non-toxicity [98]. Ramasubbu et al. [99] synthesized a porous TiO2 aerogel/Co-MOF composite with 3D hierarchical structure by modifying a sol-gel method. Aerogels, characterized as porous materials, exhibit substantial specific surface areas, minimal density, and superior thermal conductivity [100,101]. The introduction of the MOF into the aerogel enhanced the electrode’s specific capacity from 66.4 to 111.2 C g−1 at 0.8 A g−1 due to an increase in electroactive sites, a reduction in diffusion route length, and improved electron/ion transport. The constructed so-called supercapattery utilizing a negative electrode based on activated carbon achieved a specific power of 1875 W kg−1 at a specific energy of 0.4 Wh kg−1, with a remaining capacity of 93% after 5000 cycles.
Kitchametti et al. [102] recently presented a study on the application of a core–shell architecture comprising hierarchical 2D Manganese Dioxide (MnO2) nanoflakes and 1-D Nickel Titanate (NiTiO3) (NTO) mesoporous rods as a highly effective supercapacitor electrode, which offers a substantial surface area and enhanced pathways for OH ion diffusion. The hybrid porous core–shell hetero-architecture of MnO2@NTO, processed through a two-step chemical method, achieves a specific capacitance of 1054.7 F/g, a specific power of 11879.87 W/kg, and a specific energy of 36.23 Wh/kg. Additionally, a retention of 85.3% in capacitance is observed after 5000 cycles, along with no degradation in the surface morphological characteristics.

2.2. MOFs@Sulfides

Metal sulfides are another intriguing material for supercapacitive performance. Numerous studies indicate that transition metal sulfides (TMS) are materials characterized by high conductivity and significant redox reactivity [103,104,105,106].
The synergistic interaction between MOFs and sulfides has been investigated in multiple facets of photocatalysis and electro-catalysis [107,108]. Lee et al. [109] reported the in situ synthesis of Ni–BDC on NF utilizing two distinct surfactants. The use of urea enhanced the structural density and thickness due to its functional groups interacting with the BDC linker. Conversely, PVA induced voids in the MOF nanoblocks, resulting in a 2D nano-sized harmonica-like morphology, likely because of the PVA-Ni2+ complexation, which functioned as a structure-directing reagent. XRD data indicated that upon sulfurization, additional peaks emerged corresponding to NiS2 phases, but the primary MOF peaks persisted. The boundary section of the PVA sample was transformed into the NiS2@C core, as confirmed by different analytical methods (XPS, SEM/EDS, TEM). The electrode exhibited the maximum specific capacitance of 2780 F g−1 at 2 A g−1, maintaining 88% of its starting capacity after 10,000 cycles. The built flexible ASC, utilizing an activated carbon-supported negative electrode and PVA-KOH gel electrolyte, had a capacitance of 283.5 F g−1 with an energy density of 77.2 Wh kg−1 at 7000 W kg−1. Zhang et al. [110] have synthesized Ni–BDC@NiS2 nanosheet arrays by initially growing Ni-BDC on a carbon cloth substrate and subsequently partially converting it to NiS2. The ultrathin surface offered an extensive active area, resulting in excellent rate performance and specific capacitance (1128 F g−1 at 2 A g−1), with the constructed Ni–MOF@NiS2//AC supercapacitor retaining 95.2% capacity after 10,000 cycles.
Yue et al. [111] have documented a technique in which MoS2 nanoflowers were incorporated into the accordion-like Ni–BDC. In summary, MoO42− infiltrated the MOF sheets, and sulfur facilitated the production of the composite, with the development of NiS corroborated by XRD and XPS analyses. The incorporation of nonmetals established a framework that enhanced the underpotential deposition sites, resulting in a specific capacitance three times greater (1590 F g−1) than that of the pure MOF. The ASC MoS2@Ni-BDC//Activated-Carbon exhibited an energy density of 72.93 Wh kg−1 at a power density of 375 W kg−1, with a capacitance retention of 94.61% after 10,000 cycles.
Transition metal sulfides (TMSs) are currently being considered as appropriate materials for HSC positive electrodes because of the theoretically expected elevated capacity and electrochemical reactivity [103,112,113]. However, during participation in redox reactions, their crystalline phase structures experience irreversible alterations, resulting in structural collapse [114], which diminishes material stability, reduces electrical conductivity, and hinders advancements in electrode materials. A proficient strategy to tackle this issue involves the integration of TMSs with materials that possess structural integrity, extensive pore architectures, and elevated electrical conductivity, such as carbon, employing particular methodologies [115].

2.3. MOF-Based Composites

Due to their intrinsic features, metal–organic frameworks (MOFs) have been extensively utilized as templates for the synthesis of carbons, oxides, chalcogenides, or hybrids [116]. Cobalt oxide is highly esteemed in the electrochemical field due to its catalytic activity, substantial theoretical capacity, various crystal dimensions, valence states, and low price [117]. ZIF–67, being one of the most extensively researched metal–organic frameworks (MOFs), has frequently been employed by numerous researchers as a cobalt-based template for the synthesis of Co3O4 [118,119].
Recently, Kitchamsetti and Kim [120] have prepared a hierarchical FeCo-MIL-88 (FC-MOF) derived CoFe2O4 and NiMn2O4 (CFO@NMO) composite through a hydrothermal method. The hybrid CFO@NMO electrode demonstrated a remarkable specific capacity of 353.6 mAhg−1, along with an impressive capacity retention of 86.1% after 5000 cycles. A hybrid supercapacitor (HSC) device was constructed to utilize the advantages of extremely high cycling stability and specific capacity, employing AC/NF (Nickel foam (NF), activated carbon (AC)) for the negative electrode and CFO@NMO//NF for the positive electrode, both utilizing an aqueous electrolyte. The constructed hybrid supercapacitor device presented a specific capacitance of 312.8 Fg−1, demonstrating an impressive 88.4% retention in capacitance after 10,000 GCD cycles, alongside a notable energy density of 90.3 Whkg−1 and power density of 12.9 kWkg−1.
Recent works combine metal–organic frameworks with transition metal nitrides like chromium nitride interfacial layers for the preparation of hybrid supercapacitors with high efficiency. A hybrid supercapacitor based on a two-cell mode was developed utilizing activated carbon and chromium nitride/copper–trimesic acid. The results showed an impressive energy density of 94.5 Whkg−1 and a maximum power density of 7750 Wkg−1, operating at 1.6 V, with a cycling stability of 97.9% after 3000 galvanostatic charge-discharge cycles [121].

2.4. MOFs@Oxides/Sulfides

Wang and colleagues [122] synthesized a Co3O4/C@MoS2 core-shell structure using a double thermal treatment of ZIF–67 followed by solvothermal addition of MoS2. The comparative analysis of cycling stability for the composite (specific capacitance of 1076 F g−1 at 1 A g−1), Co3O4/C, and MoS2 demonstrated superior performance, with retention rates of 64.5%, 60.6%, and 50.5% over 5000 cycles at 10 A g−1, respectively. The MoS2 content significantly influences electrochemical behavior; a low quantity (10 mg of Mo precursor) restricts electron and ion transfer, while a higher quantity (30 mg) impedes diffusion, making an average of 20 mg the optimal loading. A further treatment of ZIF–67 involved a chemical conversion to Co3S4 using thioacetamide at 120 °C, followed by the NiO deposition on it to produce Co3S4@NiO hollow dodecahedrons [123]. The resulting architecture facilitates a more concise electron pathway, while the NiO outer layer enhances surface area and active sites. The augmentation of sulfide is demonstrated by the specific capacitance, which rises from 1416 to 1878 F g−1 at 1 A g−1 upon the addition of NiO. The composite electrode was assessed within an asymmetric supercapacitor (ASC) utilizing activated carbon as the negative electrode, achieving a capacitance of 164.8 F g−1 at 1 A g−1 and 115 F g−1 over 10,000 cycles at 5 A g−1, with a retention rate of 86.1%. Zhao et al. [124] have synthesized ZIF materials in order to use them as templates for both negative and positive electrodes of an ASC developed on flexible CNT fibers. The CNTFs@ZnCo2O4@Zn-Co-S positive electrode was synthesized through the processes of annealing and sulfurization of Zn/Co-ZIF, whereas the CNTFs@HCo3O4@CoNC negative electrode was obtained from Co-ZIF (ZIF–67) via oxidation and pyrolytic processes. SEM micrographs and characterization data of the prepared CNTFs are illustrated in Figure 7.
The integration of the ASC in triboelectric nanogenerators for energy harvesting and self-charging, presenting a prospective use in wearable electronics, is very interesting. The investigation of both flexibility and mechanical stability involved the selection of diverse bending angles, resulting in a retention rate of 93.9% after 5000 cycles (Figure 8). The constructed gadget demonstrated a significant areal energy density of 32 mWh cm−2 at a power density of 698 mW cm−2.
Recently, ZIF–67 was cultivated on CuxO nanowires, which were subsequently annealed to form CuxONWs@Co3O4 and then sulfurized to produce CuxONWs@CoS2 [125]. TEM images demonstrate that the NWs penetrate the hollow Co3O4 rhombic dodecahedrons, ensuring mechanical integrity and facilitating electron transport channels. The oxide and sulfide composites were analyzed over 10,000 cycles, achieving retention rates of 112.3% and 100.2%, respectively, attributed to a substantial number of electroactive sites. An uncommon metal selection was utilized with Ce–BTC as the template for CeO2/C/MoS2 [126]. The hybrid electrode is more efficient compared to both MoS2 (300 F g−1) and CeO2/C (727 F g−1), achieving a specific capacitance of 1326 F g−1 at 1 A g−1 and maintaining 92.8% capacity after 1000 cycles. The performance is ascribed to carbon conductivity, 2D MoS2 nanosheets, and the elevated surface area of CeO2. Furthermore, the ASC exhibited commendable stability across an equivalent number of cycles, alongside a high energy density (34.5 Wh kg−1) at 667 W kg−1.

2.5. MOFs@Oxides/Phosphides

Recently, CoPx/CoO composites using the calcination of 2D ZIF–67, followed by partial phosphidation using varying phosphorus concentrations, have been synthesized according to the procedure illustrated in Figure 9 [127]. Hybrids are also documented in OER catalysis [128] and Na–O2 batteries [129]. The combination of phosphides’ redox activity and oxides’ stability led to improved supercapacitor performance [127]. XRD examination and Rietveld refinement demonstrated that increasing the phosphorus content results in the predominance of phosphide phases (CoP, Co2P) over CoO in the composite. Despite utilizing a substantial quantity of phosphorus, the oxide did not achieve complete transformation; the augmentation of phosphorus content from 40 wt% to 100 wt% resulted in merely a 4% improvement from (CoPx)0.90/CoO0.10 to (CoPx)0.94/CoO0.06, attributed to the initial stability of ZIF, which was maintained in the Co3O4 phase, inhibiting the diffusion of the produced PH3 gas into the bulk. Nonetheless, the Co2O3 could not be retained in the final composite due to the potent reducing nature of PH3. (CoPx)0.90/CoO0.10 exhibited a specific capacitance of 467 F g−1 at 5 A g−1 and maintained 91% retention over 10,000 cycles at 30 A g−1, while the equivalent material with rGO demonstrated comparable specific capacitance and 107% retention.

2.6. Other Composite Combinations

Guo and coworkers [130] have prepared metal nitride-based 3-D stacked hierarchical Ni3N-CoN/NC nanosheets for flexible and freestanding supercapacitor electrodes with remarkable energy density of 0.2144 mWh cm−2 and a peak power density of 80 mW cm−2, demonstrating remarkable cycling stability of 92.3% after 15,000 cycles.
Zhao et al. [131] have recently synthesized amorphous spherical carbon materials featuring hierarchical pores (DHTAC-2) through a one-step carbonization process of Al-MOF. The carbon exhibits a significant specific surface area of 2666 m2 g⁻1 and a considerable pore volume of 2.35 cm3 g⁻1. The carbon obtained (DHTAC-2) shows a remarkable specific capacitance of 298.8 F g−1 at a current density of 1 A g−1, along with impressive capacitance retentions of 97.3% and 98.0% after 120,000 cycles at current densities of 50 A g−1 and 100 A g−1, respectively. Furthermore, the Nitrogen-doped carbon (N-DHTAC) exhibits an elevated presence of pyridine-N, pyrrole-N, and graphite-N content. The three types of N led to notable enhancements in wettability, reversible pseudocapacitance, and conductivity, culminating in an impressive capacitance of 355.0 F g−1 at 1 A g−1 and an increased energy density of 17.7 Wh kg−1 at 350 W kg−1 for the assembled electric double-layer capacitor (EDLC) device.
Zhang et al. [132] produced a series of chalcogenides generated from ZIF–67 on cotton fabric, which was further subjected to thermal treatment under various conditions along with chalcogenide powders to obtain CF@CoX (X = O, S, Se, Te). The flakes were uniformly deposited onto the carbon fibers, creating a three-dimensional hierarchical structure. Of the four synthesized electrodes, the sulfide exhibited the highest areal specific capacity (3.58 F cm−2) at 5 mA cm−2, followed by the selenide (2.43 F cm−2), the telluride (1.8 F cm−2), and the oxide (0.67 F cm−2). CF@CoS demonstrated superior performance relative to other components in the ASC device, utilizing an AC cloth (made through a process akin to cotton pyrolysis) as the negative electrode, achieving 0.77 F cm−2 and an energy density of 149 mWh cm−2 at 4.3 mW cm−2.
Zhang et al. [133] summarize, in a recent work, advancements in the field of multidimensional MOF-based materials and their derivatives across various dimensionalities, encompassing 1-D nanopowders, 2-D nanofilms, 3-D aerogels, and 4-D self-supporting electrodes for supercapacitors. Additionally, the primary challenges and viewpoints regarding MOFs and their derivative materials for practical and sustainable electrochemical energy conversion and storage applications are succinctly addressed, potentially serving as a framework for the design of next-generation electrode materials across various dimensionalities.

2.6.1. MOFs/Graphene

Graphene and graphene-like materials have gained significant interest because of their high electrical conductivity, superior power density, and stable chemical properties. The exploration of nanostructures has been extensive in various fields, including energy storage [134,135]. Metal–organic frameworks have been utilized in energy systems; however, they exhibit low conductivity and stability challenges in practical applications [136].
A recent study of Saxena et al. [137] presents the preparation of a rGO/Ppy/Zn-MOF ternary composite using graphene oxide, pyrrole, and a zinc metal–organic framework precursor, specifically Zn(NO3)2·6H2O and imidazole, through a straightforward hydrothermal one-pot method. The synthesized rGO/Ppy/Zn-MOF composite demonstrated exceptional electrochemical properties. In energy storage applications, the synthesized rGO/Ppy/Zn-MOF demonstrated remarkable specific capacitance of 175 F g−1 at a current density of 1 A g−1, a specific energy of 19.68 Wh kg−1, and a specific power of 1792 W kg−1. Exhibiting cyclic stability at 10 A g−1, it maintains up to 82% of its starting capacitance even after more than 7000 repetitive cycles.
In order to further increase the active surface graphene, aerogels have also been employed. Graphene has been utilized in methodological energy strategies owing to its remarkable structure and characteristics; however, its application is constrained by issues related to nanosheet aggregation [136]. The development of nitrogen-doped graphene aerogel (N-GA) featuring a three-dimensional structure can significantly inhibit the stacking and aggregation of graphene flakes while preserving the benefits of electrical conductivity [138]. In this respect, MoS2 derived from MOFs with a mixed 1T/2H phase was synthesized using a straightforward one-step hydrothermal method, and the assembly of MOFs-derived MoS2 with nitrogen-doped graphene aerogel (N-GA) resulted in the formation of MoS2/N-GA. The findings indicated that employing N-GA as the carrier for MOFs-derived MoS2 not only improved the electrical conductivity of the composites but also mitigated the volume expansion and contraction of MoS2 during the charging and discharging processes, attributed to its distinctive three-dimensional structure. The capacitive performance of MNGA20 (20% N-GA) outperformed that of MNGA10 (10% N-GA) and MNGA30 (30% N-GA). MNGA20 composites demonstrated the lowest charge transfer resistance and achieved an optimal specific capacitance of 530 F g−1 at 1 A g−1, maintaining 80.2% capacitance after 1000 charge-discharge cycles at 10 A g−1 [139].
Trimesic acid-based MOFs are often the choice for the preparation of composites with graphene materials [140]. Other MOFs used in similar composites with graphene are based on imidazolate [141,142], zirconium [143], as well as terephthalic acid [144].

2.6.2. MOFs/Phosphides

Transition metal phosphides (TMPs) exhibit high electrical conductivity and remarkable physicochemical properties, demonstrating significant potential for enhancing the capacitance and energy density of supercapacitors [145]. MOF-derived dual phosphide composites are rare as supercapacitor electrodes in the development of phosphide materials. Conversely, MOF-derived carbon/phosphide hybrid materials have been utilized in several applications in recent years [146].
Zhou et al. [147] synthesized NiCoP/C nanohybrids with varying ratios, that involve monometallic phosphides, contingent upon the initial concentration during the production of the metal–organic framework utilizing a phenanthroline linker. The annealing of the MOF occurred in a nitrogen environment, leading to the anchoring of NiCoP nanoparticles onto the carbon. The composite exhibited a notable specific capacity of 775.7 C g−1 at 1 A g−1 and 582.4 C g−1 at 20 A g−1, while the completed supercapacitor provided 47.6 Wh kg−1 at 799 W kg−1. Liu and coworkers [148] have prepared nickel phosphide and nickel cobalt phosphide heterostructure nanocomposites (NP@NCP) synthesized through the phosphidization of metal–organic framework (MOF) precursors leading to a morphology with nanosheets hierarchically arranged to form microflower-like particles. The NP@NCP exhibit a morphology like that of MOF, yet they possess a significantly higher porosity. The NP@NCP31 material on carbon cloth, when utilized as the electrode material, demonstrates a specific capacity of 868.1 C g−1 at 1 A g−1, with 82.05% of its capacity retained after 5000 cycles in 4 M KOH. An asymmetric supercapacitor made with NP@NCP31 and activated carbon demonstrates an energy density of 61.9 Wh kg−1 at a power density of 800.1 W kg−1, along with a capacity retention rate of 94.17% after 5000 cycles. The impressive charge-storage performance of the NP@NCP is attributed by the authors to the development of a heterointerface, which induces charge redistribution, enhances the number of electrochemically active sites, and promotes ion diffusion.
A different viable synthetic method has been investigated involving the use of an ionic liquid (IL) in the synthesis of MOFs. Although ionic liquids (ILs) are primarily utilized as substitutes for organic solvents, their incorporation into DMF was essential for nitrogen and phosphorus doping, as well as for the introduction of a different morphology with wrinkles and folds resulting from partial collapse during the following carbonization and phosphorization processes under argon/hydrogen flow, ultimately yielding the NiCoP/NPC composite [149]. Additionally, the scientists noted applying this procedure results to a five times higher surface area, while the constructed ASC demonstrated 40.2 Wh kg−1 at a power density of 800.2 W kg−1. Similar electrode materials based on C/NiCoP have been recently developed and tested. The prepared C/NiCoP material exhibits a flower-like structure with specific capacitance of 773 C g−1 at 1 A g−1 [150].
Bimetallic phosphides based on nickel and manganese have been synthesized through the phosphidization of bimetallic hydroxides derived from metal–organic frameworks. The bimetallic phosphides consist of nanosheets and display a crystalline Ni2P phase, with a high specific surface area. When utilized on carbon cloth (CC) and employed as the electrode material in 4 M KOH, the (Ni0.93Mn0.07)2P-18/CC exhibits a specific capacitance of 851.1 C g−1 at 1 A g−1 and retains 84.87% capacity following 5000 cycles. The investigation explores the origin of the high specific capacity, pointing out the electron interactions between Mn and Ni sites, the elevated specific surface area, and the ease of ion transport, all of which contribute to the remarkable specific capacity and stability observed [151]. For Cu/Co phosphides density-functional theory (DFT) computations demonstrate that the Cu3P/CoP heterostructures enhance the metallic conductivity of the composite, exhibiting an increased density of electron-occupied states near the Fermi level and improving the charge-transfer efficiency of the heterostructure. The findings indicate that the Cu3P/CoP-1 composite, possessing a 1:1 Cu/Co molar ratio, has superior electrochemical performance. At a current density of 1 Ag−1, a remarkable specific capacitance of 804.3 F g−1 has been demonstrated, and when the current density is elevated to 10 Ag−1, it maintains 92.8% of its capacitance over 10,000 cycles [152].

2.6.3. MOFs-Sulfides

The facile agglomeration of transition metal sulfides leads to a diminished specific surface area and sluggish reaction kinetics, resulting in volumetric expansion that compromises structural integrity and diminishes performance during prolonged electrochemical processes. To address these deficiencies, a preferred strategy is to develop a stable composite structure with rapid charge transfer capabilities. Numerous researchers have integrated metal sulfides generated from MOFs with preferably one-dimensional conductive materials to establish a stable structure. A new composite of nanoscale one-dimensional (1D) Mxene fibers infused with metal sulfides produced from MOFs (Ni(Co)3S4@MXene-fibers) has been created (Figure 10). The electrochemical investigation indicates that the specific capacitance of Ni(Co)3S4@MXene-fibers attains 829.4 C g−1 at a current density of 1 A g−1. Furthermore, the asymmetric device (Ni(Co)3S4@MXene-fibers//AC) exhibits an energy density of 63.3 Wh kg−1 at a power density of 850 W kg−1 [153].
Mono as well as bimetallic sulfides have proven to be viable materials due to simple fabrication techniques for energy applications [154,155,156]. Notably, NiCo2S4 has been thoroughly investigated due to its exceptional electrochemical performance, resulting to the emergence of MOF-derived variants in literature as well [154,157,158,159]. Regarding sulfide composites, Li et al. [160] synthesized amorphous Ni/Zn-BDC spheres, which were then converted into NiS2/ZnS. The reaction duration in the autoclave at 160 °C influences the diffusion of S2− ions, resulting in the formation of core-shell structures (1 h), yolk-shell structures (3 h), and hollow spheres (6 h), as confirmed by SEM/TEM examination. The hollow spheres demonstrated commendable capacitance (1198 F g−1 at 1 A g−1) relative to other analogous nickel sulfide materials [161,162], which can be ascribed to the synergy between NiS2 and ZnS, enhancing ion/electron transfer, as well as the reduction in diffusion path owing to the mesoporous architecture. The NiS2/ZnS//AC ASC exhibited a specific capacitance of 89.7 F g−1 at 1 A g−1, with a retention rate of 88.7% after 1500 cycles.
Shrivastav and colleagues synthesized Co3S4/WS2 utilizing a ZIF-67 template [163]. The hollow Co3S4 microspheres were solvothermally modified with 2D WS2 nanorods to improve energy density and facilitate ion/electron transfer. The electrode exhibited a capacitance of 412.7 F g−1 at a current density of 1 A g−1 and was utilized in a symmetrical supercapacitor with a broad potential window of 2 V, demonstrating 92% stability over 2000 cycles. Zhao et al. [164] created a flower-like trimetallic sulfide composite (Figure 11), referred to as Cu(NiCo)2S4/Ni3S4. The combination exhibited a specific capacitance of 1320 F g−1 (1 A g−1), in direct comparison to the CuNiCo-OH intermediate product, which measured 670 F g−1. Furthermore, the Cu(NiCo)2S4/Ni3S4//AC supercapacitor exhibited satisfactory cycling stability (85% retention) and achieved a maximum energy density of 40.8 Wh kg−1, with a power density of 7859 W kg−1.
Recently, Li et al. [165] have prepared flower like Co/Ni-MOFs (ZIF–67@Ni–salicylate) for supecapacitor electrodes exhibiting exceptional performance such as high specific capacitance of 1493 F g−1 at 1 A g−1 and at the same time excellent cycle stability. Xiong et al. [166] also synthesized bimetallic MOFs on Ni foam with two distinct techniques. The one-pot method utilized two metal precursors (Ni–Fe or Ni–Co) concurrently, whereas the two-step synthesis (MOF-on-MOF) introduced the second metal precursor in the subsequent phase, following the formation of the initial MOF. The synergistic interaction between the two metals imparts enhanced efficiency to bimetallic catalysts in electrocatalysis, attributed to the enrichment of redox sites. Notably, FeMOF/NiMOF/NF exhibited exceptional performance, demonstrating modest overpotentials of 293 mV at 50 mA cm−2 and 359 mV at 100 mA cm−2, along with commendable cycling stability, positioning it as an optimal candidate for OER catalysis. Additionally, it exhibited an areal capacitance of 1166 mF cm−2 at a current density of 1 mA cm−2. A sulfurization method was conducted to investigate the sulfides in an ASC (with AC as the negative electrode), revealing that NiCoS/NF demonstrated a high specific capacitance of 2815.4 F g−1 at 1 mA cm−2.

2.6.4. MOFs–Selenides

In addition to sulfides, selenides are effective alternatives in hybrid semiconductor devices, exhibiting superior conductivity (1 × 10−3 S m−1) and metallic characteristics compared to sulfur (5 × 10−28 S m−1) [167,168,169,170]. Zhang et al. [171] synthesized nitrogen-doped cobalt selenides-carbon composites utilizing ZIF–67. By altering the treatment temperature, they generated distinct structures, e.g., a solid structure at 300 °C, a yolk-shell at 450 °C, and a double-shell at 600 °C. Mei et al. [172] also emphasized the importance of the structure during the synthesis of the MOF (Ni-BTC). As the reaction time extended (T = 150 °C), Ostwald ripening let the inner sphere progressively dissolve while the outer shell expanded, ultimately resulting in a hollow structure. After 10 h, a yolk-shell (730 C g−1 at 2 A g−1) was obtained, whereas after 15 h, a hollow sphere (~600 C g−1 at 2 A g−1) was ultimately formed. Additional carbon composites, including Co [173], Cu-Co [174], or Ni-Co [175], have also been examined as supercapacitor electrodes. Hollow and porous nickel–zinc–cobalt selenide (Ni–Zn–Co–Se) nanosheet configurations were, for the first time, synthesized on chemically reduced rGO-coated nickel foam (rGO/NF), utilizing a Zn–Co-organic framework as a template. The Ni–Zn–Co–Se@rGO/NF electrode, benefiting from its distinctive porous and hollow nanoarchitecture, offers numerous electroactive sites, extensive accessible areas for efficient electrolyte insertion, and reduced ion diffusion pathways, demonstrating outstanding electrochemical properties, including a specific capacity of 464.4 mA h g−1, a rate performance of 68.9%, and a remarkable cyclic lifespan of 93.24% after 7000 cycles in a three-electrode assembly for supercapacitors. An asymmetric supercapacitor (ASC) device was constructed (Figure 12) utilizing Ni–Zn–Co–Se@rGO/NF as the positive electrode and MOF-derived hollow porous carbon (MDHPC) as the negative electrode to illustrate its practical applicability. The obtained Ni–Zn–Co–Se@rGO/NF//MDHPC device demonstrated an energy density of 74.6 Wh kg−1 at a power density of 796.91 W kg−1, along with an impressive durability of 96.4% after 7000 consecutive charge–discharge cycles [176].
A recent study by Dong and coworkers [177] centers on the advancement of multi-metal selenides (MMSes) that utilize synergistic effects to improve electrochemical performance by exploiting plentiful active sites and redox reactions. The metal–organic framework ZIF–67 is used as a morphological template by cultivating ZIF–67 on a NiMoO4·xH2O substrate. Subsequently it was selenized, leading to synthesis of a multi-metal selenide electrode material abundant in Ni, Co, and Mo, comprising NiSe, Ni3Se2, CoSe, and MoSe2 phases. This electrode demonstrated a specific capacitance of 15.49 F cm−2 at a current density of 5 mA cm−2 and retains 96.5% of its capacitance after 10,000 cycles. The constructed asymmetric supercapacitor demonstrated an areal energy density of 0.301 mWh cm−2 and an areal power density of 4 mW cm−2, with a capacity retention of 95.6% after 10,000 cycles at a current density of 50 mA cm−2.

2.6.5. MOFs–Sulfoselenides

Metal selenides are attractive materials for electrodes in advanced supercapacitors and energy harvesting applications [178]. Nonetheless, their performance remains constrained by inadequate utilization of active materials and the poor porosity of the electrodes. A potential solution to address these shortcomings is to manufacture composites using materials with inherent high porosity, such as MOFs.
A straightforward strategy is presented for the in situ synthesis of hierarchical mesoporous selenium@bimetallic selenide quadrilateral nanosheet arrays on diverse conductive substrates, utilizing two-dimensional bimetal porphyrin paddlewheel framework (MCo–TCPP, where TCPP denotes tetrakis(4-carboxyphenyl)porphyrin and M represents Ni or Zn) nanosheets as templates [179]. Yi et al. [180] produced a bimetallic metal–organic framework (NiCo-BTC) for the production of a sulfoselenide. The MOF microspheres were annealed in an Ar/H2 environment to obtain NiCo2 on graphitic carbon (GC), which was further converted into NiCo2(SxSe1-x)5 nanoparticles confined within GC hollow spheres. Among the investigated ratios, NiCo2(S0.78Se0.22)5 exhibited enhanced specific capacity of 560.7 C g−1 at 1 A g−1 and 440.1 C g−1 at 20 A g−1, alongside pure sulfide and selenide. The constructed ASC utilizing the specified cathode and Bi2O2.33/rGO anode achieved an energy density of 47.2 Wh kg−1 at 801 W kg−1 and a capacity of 212 C g−1 at 1 A g−1.
Recently, Lu et al. [181] presented an eco-friendly, safe, and straightforward approach to synthesize a bimetallic sulfoselenide–selenite heterojunction, (Ni,Co)(Se,S)2/(Ni,Co)SeO3, via a one-pot hydrothermal process with NiCo–LDH as a template. The coarse texture of the nanosheet array increases the number of active sites, while the multi-interface enhances the electronic structure for electrochemical processes. The optimized (Ni,Co)(Se,S)2/(Ni,Co)SeO3−1 serves as a battery-type cathode for supercapacitors, demonstrating a high specific capacitance of 7.61 F cm−2 at a current of 2 mA cm−2. The completed HSC device exhibits a high energy density of 0.59 mW h cm−2 at a power density of 1.44 mW cm−2 and demonstrates a capacitance retention rate of 84.38% after 5000 charge/discharge cycles.
In the following, Table 1 indicative performance metrics of representative MOF-based materials for supercapacitors are tabulated.

3. Limitations and Outlook

Traditional MOFs possess fundamental drawbacks, particularly their limited conductivity and structural integrity, which constrain their application in electrochemical applications. However, specific challenges related to MOFs must be tackled to enable their broader commercialization. The inadequate electrochemical conductivity significantly restricts their potential applications, prompting the design and proposal of various nanocomposites with diverse structures across different dimensionalities.
The functionalization of MOF derivatives is still somewhat restricted; however, there exists significant potential to improve their electrochemical properties. This can be achieved by encapsulating functional groups and various species within their frameworks, aiming to develop high-performance energy storage devices through the identification of appropriate MOF derivatives for both anodes and cathodes in electrochemical devices.
Another possibility to improve MOF-based hybrid materials for electrochemical applications is the use of MXenes as hybrid components. MXenes have gained significant research interest in energy-storage applications arising from their exceptional metallic conductivity, superior hydrophilicity, unique layered structure, and abundant surface functional groups [194]. Nonetheless, MXenes have comparatively inferior thermodynamic and environmental stability in relation to MOFs [195]. As MXenes exhibit superior conductivity and satisfactory catalytic properties for electrochemical applications [196], consequent engineering the characteristics of the MOFs with MXenes can improve the electrical conductivity of the MOFs and at the same time increase the stability of the MXenes [197]. MXene/MOF-based composites have emerged also as excellent hydrophilic materials, facilitating immersion in aqueous electrolytes and quick charge transfer. As an example, the catalytic efficacy of CoBDC/MXene was assessed in alkaline, acidic, and neutral electrolytes, demonstrating low overpotentials of 29, 41, and 76 mV, respectively. To enhance comprehension of the operational mechanisms of these composites in energy storage and conversion, it is essential to refine and complete theoretical calculations and in situ characterizations. Nonetheless, numerous published studies exhibit a deficiency in comprehensive characteristic assessments and theoretical simulations of MXene/MOF composites and their derivatives [198]. MOFs are also an interesting alternative for the development of water-based electrolytes. The stability of MOFs in moisture as well as water is acknowledged as a critical characteristic concerning their prospective applications as the majority of MOFs exhibit a certain degree of water-lability. Aqueous energy devices necessitate metal–organic frameworks that exhibit exceptional stability in the presence of moisture and water. However, most MOFs exhibit low structural tolerance in aqueous systems. While MOFs and their derivatives have the potential to fulfill almost all essential criteria for use in aqueous energy devices, there are still challenges that must be addressed to achieve greater success. While several MOFs have been developed exhibiting remarkable chemical and thermal stability, a need for further enhancement to withstand more extreme acidic or basic environments is required. The incorporation of guest molecules of appropriate dimensions could serve as a viable approach to improve the structural resilience of MOFs while preventing pore obstruction [199]. Another potential approach to address the degradation of MOFs in the presence of water is through surface hydrophobic modification [200].
MOF derivatives have surfaced as potential materials to tackle these limitations and discover applications across diverse electrochemical domains. These materials act as efficient templates, allowing for meticulous control over morphology, chemical modification, heteroatom doping, increased surface area, and structural alterations. There is huge progress potential for future applications that needs more attention by the research community.
In the future, the emphasis of exploration in MOFs might transition from traditional domains like storage or separation to more pressing and vital issues, such as the generation of clean energy. Notable progress has been made in improving the performance of MOF-based components to align with industrial and commercial standards. However, as mentioned above, further progress will require that MOFs surpass the standards set by conventional materials.
Transferring MOFs production technology from laboratory environments to commercial applications necessitates the capability to produce them at the appropriate quality, scale, and cost. Consequently, the principles of sustainable development and the circular economy are crucial, including the minimization of energy consumption, raw materials, reagent toxicity, and waste, as well as the design of materials for reuse and recyclability [201,202]. Transitioning from a batch to a continuous production process is significantly advantageous, as it facilitates expedited manufacturing (attributable to improved heat and mass transfer), enhanced reproducibility, diminished solvent and energy usage, increased space-time yield (STY) and reactor scalability, as well as decreased downtime and labor expenses (due to fewer steps between batches) [203]. The main issue in the scalability of MOFs production remains the price of the precursors. While some of them are not expensive (e.g., terephthalic acid or many metal precursors), more complex organic molecules can substantially increase the price of the final product.
In summary, the materials derived from MOFs demonstrate considerable promise and numerous advantages for application in energy production and storage systems. Considering the various existing challenges that are complex to tackle in the short term, a strong belief that a bright future awaits these innovative materials based on MOFs in real-world applications is maintained.

4. Summary

This review concludes with the latest developments in MOFs-based materials for high-performance supercapacitor applications. It highlights how different synthesis strategies influence the morphologies and sizes of these materials, showcasing the extensive potential of MOFs to enhance electrochemical devices for energy storage. The connections between the electrochemical performances and the morphologies of the electrode active materials are highlighted. Additionally, the main challenges and viewpoints regarding materials based on MOFs and their derivatives for practical and sustainable electrochemical energy storage applications are briefly addressed.
MOFs exhibit a remarkable versatility, allowing for rational modifications tailored to meet various emerging needs. In this context, MOFs can play a significant role in advancing the hydrogen economy through their applications in electrocatalysis for hydrogen production and oxygen reduction, PEM fuel cells, and energy storage (rechargeable batteries and supercapacitors). Their high surface areas (up to 5.0 × 103 m2 g⁻1), highly uniform pores, adjustable pore geometries, adaptable structures, and diverse chemistry make MOFs compelling options for supercapacitor technologies. Furthermore, the use of multifunctional bridging ligands and post-synthetic modifications has enabled the careful design and functionalization of MOFs, facilitating the rational tuning of their properties. The wide range of possible combinations of metal centers and organic linkers allowing to design pores and increase active surface indicates a promising future for the application of MOFs in supercapacitors. The ability to manipulate pore structures alongside surface engineering could pave the way for a systematic investigation into how structural parameters and chemical compositions influence the performance of MOF-based electrochemical devices, potentially yielding important fundamental insights.
A significant body of work on MOFs and their derivatives has demonstrated and validated the consistency of their structures and chemical properties for current electrochemical applications at the laboratory scale. Nevertheless, additional advancements are essential, as numerous challenges in the realm of MOF-based electrochemical devices are currently being encountered and must be addressed promptly.

Author Contributions

Conceptualization, C.A. and G.S.; Validation, C.A., N.A. (Nikolaos Argirusis), N.A. (Niyaz Alizadeh) and G.S.; Investigation, C.A. and M.-E.K.; Data curation, C.A.; Writing—original draft preparation, C.A. and N.A. (Nikolaos Argirusis); Writing—review and editing, C.A., N.A. (Nikolaos Argirusis), N.A. (Niyaz Alizadeh), and M.-E.K.; Supervision, C.A. and G.S.; Project administration, G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data has been produced for this work.

Conflicts of Interest

Authors Niyaz Alizadeh and Nikolaos Argirusis were affiliated with the company MAT4NRG. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. DNV As, Norway, Energy Transition Outlook 2024—A Global and Regional Forecast to 2050. Available online: www.dnv.com (accessed on 22 April 2025).
  2. Wang, Y.; Wang, H.; Qu, G. Molybdenum-Based Electrode Materials Applied in High-Performance Supercapacitors. Batteries 2023, 9, 479. [Google Scholar] [CrossRef]
  3. Qi, P.; Wang, H.; Lu, Y.; Chen, M.; Liu, G.; Li, W.; Huang, C.; Tang, Y. Ammonia-induced N-doped NiCoO2 nanosheet array on Ni foam as a cathode of supercapacitor with excellent rate performance. J. Alloys Compd. 2022, 895, 162535. [Google Scholar] [CrossRef]
  4. El-Shamy, O.A.A.; Siddiqui, M.R.; Mele, G.; Mohsen, Q.; Alhajeri, M.M.; Deyab, M.A. Synthesis, characterization and application of CNTs@Co-MOF and FCNTs@Co-MOF as superior supercapacitor: Experimental and theoretical studies. J. Mol. Struct. 2025, 1321, 140161. [Google Scholar] [CrossRef]
  5. Kulkarni, O.; Bhosale, R.; Narale, D.; Pise, S.; Shaikh, T.; Kolekar, S. Dual redox center-based copper-cobalt metal–organic framework as pseudocapacitive electrode material for supercapacitor. Inorg. Chem. Commun. 2025, 172, 113711. [Google Scholar] [CrossRef]
  6. Bojarajan, A.K.; Gunasekaran, S.S.; Kalluri, S.; Al Omari, S.A.B.; Bakenov, Z.; Sangaraju, S. Tuning capacitance of bimetallic ZnCo2O4 using anionic, cationic and non-ionic surfactants by hydrothermal synthesis for high-performance asymmetric supercapacitor. Inorg. Chem. Commun. 2024, 169, 113035. [Google Scholar] [CrossRef]
  7. Minakshi, M.; Samayamanthry, A.; Whale, J.; Aughterson, R.; Shinde, P.A.; Ariga, K.; Shrestha, L.K. Phosphorous—Containing Activated Carbon Derived from Natural Honeydew Peel Powers Aqueous Supercapacitors. Chem. Asian J. 2024, 19, e202400622. [Google Scholar] [CrossRef] [PubMed]
  8. Minakshi, M.; Mujeeb, A.; Whale, J.; Evans, R.; Aughterson, R.; Shinde, P.A.; Ariga, K.; Shrestha Kumar, L. Synthesis of Porous Carbon Honeycomb Structures Derived from Hemp for Hybrid Supercapacitors with Improved Electrochemistry. ChemPlusChem 2024, 89, e202400408. [Google Scholar] [CrossRef]
  9. Minakshi, M.; Wickramaarachchi, K. Electrochemical aspects of supercapacitors in perspective: From electrochemical configurations to electrode materials processing. Prog. Solid State Chem. 2023, 69, 100390. [Google Scholar] [CrossRef]
  10. Alam, S.; Urooj, A.; Rehman, S.; Iqbal, M.Z.; Hegazy, H.H. Investigation of metal organic frameworks and their derivatives as electrode materials for hybrid energy storage devices. Mater. Chem. Phys. 2023, 304, 127877. [Google Scholar] [CrossRef]
  11. Fleischmann, S.; Mitchell, J.B.; Wang, R.; Zhan, C.; Jiang, D.; Presser, V.; Augustyn, V. Pseudocapacitance: From fundamental understanding to high power energy storage materials. Chem. Rev. 2020, 120, 6738–6782. [Google Scholar] [CrossRef]
  12. Brousse, T.; Belangér, D.; Long, J.W. To Be or Not To Be Pseudocapacitive? J. Electrochem. Soc. 2015, 162, A5185–A5189. [Google Scholar] [CrossRef]
  13. Costentin, C.; Porter, T.R.; Savéant, J.-M. How do pseudocapacitors store energy? Theoretical analysis and experimental illustration. ACS Appl. Mater. Interfaces 2017, 9, 8649–8658. [Google Scholar] [CrossRef] [PubMed]
  14. Kitagawa, S.; Kitaura, R.; Noro, S. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef] [PubMed]
  15. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef]
  16. Li, J.X.; Bao, T.; Zhang, C.; Song, H.; Zou, Y.Y.; Yuan, I.; Xi, Y.; Yu, C.Z.; Liu, C. A general strategy for direct growth of yolk-shell MOF-on-MOF hybrids. Chem. Eng. J. 2023, 472, 144926. [Google Scholar] [CrossRef]
  17. Arbulu, R.C.; Jiang, Y.B.; Peterson, E.J.; Qin, Y. Metal-organic framework (MOF) nanorods, nanotubes, and nanowires. Angew. Chem. Int. Ed. 2018, 57, 5813–5817. [Google Scholar] [CrossRef]
  18. Mohammadi, T.; Hosseini, M.G.; Rezvani Jalal, N.; Alavipour, E.; Pastor, E. Binder-free supercapacitors based on nanosheets of Bi-ligand Co metal-organic frameworks: Density functional theory validation. J. Power Sources 2025, 632, 236385. [Google Scholar] [CrossRef]
  19. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 974. [Google Scholar] [CrossRef]
  20. Lei, L.; Cheng, Y.; Chen, C.; Kosari, M.; Jiang, Z.; He, C. Taming structure and modulating carbon dioxide (CO2) adsorption isosteric heat of nickel-based metal organic framework (MOF-74(Ni)) for remarkable CO2 capture. J. Colloid Interface Sci. 2022, 612, 132–145. [Google Scholar] [CrossRef]
  21. Zhang, W.; Bojdys, M.J.; Pinna, N. A universal synthesis strategy for tunable metal-organic framework nanohybrids. Angew. Chem. Int. Ed. 2023, 62, e202301021. [Google Scholar] [CrossRef]
  22. Della Roca, J.; Liu, D.; Lin, W. Nanoscale Metal–Organic Frameworks for Biomedical Imaging and Drug Delivery. Acc. Chem. Res. 2011, 44, 957–968. [Google Scholar] [CrossRef] [PubMed]
  23. Ren, Y.; Chia, G.H.; Gao, Z. Metal–organic frameworks in fuel cell technologies. Nano Today 2013, 8, 577–597. [Google Scholar] [CrossRef]
  24. Zhang, H.; Lu, X.; Yang, L.; Hu, Y.; Yuan, M.; Wang, C.; Liu, Q.; Yue, F.; Zhou, D.; Xia, Q. Efficient air epoxidation of cycloalkenes over bimetal-organic framework ZnCoMOF materials. Mol. Catal. 2021, 499, 111300. [Google Scholar] [CrossRef]
  25. Liang, Y.N.; Yao, W.G.; Duan, J.X.; Chu, M.; Sun, S.Z.; Li, X. Nickel cobalt bimetallic metal-organic frameworks with a layer-and-channel structure for high performance supercapacitors. J. Energy Storage 2021, 33, 102149. [Google Scholar] [CrossRef]
  26. Lim, G.J.H.; Liu, X.M.; Guan, C.; Wang, J. Co/Zn bimetallic oxides derived from metal organic frameworks for high performance electrochemical energy storage. Electrochim. Acta 2018, 291, 177–187. [Google Scholar] [CrossRef]
  27. Zhao, H.; Xu, W.; Li, M.; Meng, Z.; Ullah, I.; Nawaz, M.Z.; Wang, J.; Wang, C.; Chu, P.K. (Fe, Co) oxide nanowires on gold nanoparticles modified MOF-derived carbon nanoflakes for high-efficiency sodium-ion batteries and supercapacitors across electrolytes. J. Power Sources 2025, 626, 235793. [Google Scholar] [CrossRef]
  28. Otun, K.O.; Diop, N.F.; Fasakin, O.; Adam, R.A.M.; Rutavi, G.; Manyala, N. Engineering the structures of ZnCo-MOFs via a ligand effect for enhanced supercapacitor performance. RSC Adv. 2025, 15, 4120–4136. [Google Scholar] [CrossRef]
  29. Li, H.I.; Eddaoudi, M.; Keeffe, M.O.; Yaghi, O.M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402, 276–279. [Google Scholar] [CrossRef]
  30. Hao, X.; Song, W.; Wang, Y.; Qin, J.; Jiang, J. Recent Advancements in Electrochemical Sensors Based on MOFs and Their Derivatives. Small 2025, 21, 2408624. [Google Scholar] [CrossRef]
  31. Xie, C.; Li, H.; Niu, B.; Guo, H.; Lin, X. Comparison of ultrasonic vs mechanochemistry methods for fabrication of mixed-ligand Zn-based MOFs for electrochemical determination of luteolin. J. Alloys Compd. 2024, 989, 174363. [Google Scholar] [CrossRef]
  32. Kumar, R.S.; Kumar, S.S.; Kulandainathan, M.A. Efficient electrosynthesis of highly active Cu3(BTC)2-MOF and its catalytic application to chemical reduction. Microporous Mesoporous Mater. 2013, 168, 57–64. [Google Scholar] [CrossRef]
  33. Kaur, G.; Kandwal Komal, P.; Sud, D. Sonochemically synthesized Zn (II) and Cd (II) based metal-organic frameworks as fluoroprobes for sensing of 2,6-dichlorophenol. J. Solid State Chem. 2023, 319, 123833. [Google Scholar] [CrossRef]
  34. Glowniak, S.; Szczesniak, B.; Choma, J.; Jaroniec, M. Recent developments in sonochemical synthesis of nanoporous materials. Molecules 2023, 28, 2639. [Google Scholar] [CrossRef]
  35. Mendes, R.F.; Rocha, J.; Almeida Paz, F.A. Chapter 8—Microwave Synthesis of Metal-Organic Frameworks. In Metal-Organic Frameworks for Biomedical Applications; Mozafari, M.B.T., Ed.; Woodhead Publishing: Cambridge, UK, 2020; pp. 159–176. [Google Scholar] [CrossRef]
  36. Kumari, P.; Kareem, A.; Jhariat, P.; Kumar, S.S.; Panda, T. Phase purity regulated by mechano-chemical synthesis of metal-organic frameworks for the electrocatalytic oxygen evolution reaction. Inorg. Chem. 2023, 62, 3457–3463. [Google Scholar] [CrossRef]
  37. Wang, W.P.; Chai, M.; Zulkifli, M.Y.B.; Xu, K.J.; Chen, Y.L.; Wang, L.Z.; Chen, V.; Hou, J.W. Metal-organic framework composites from a mechanochemical process. Mol. Syst. Des. Eng. 2023, 8, 560–579. [Google Scholar] [CrossRef]
  38. Vaitsis, C.; Kanellou, E.; Pandis, P.K.; Papamichael, I.; Sourkouni, G.; Zorpas, A.A.; Argirusis, C. Sonochemical synthesis of zinc adipate Metal-Organic Framework (MOF) for the electrochemical reduction of CO2: MOF and circular economy potential. Sustain. Chem. Pharm. 2022, 29, 100786. [Google Scholar] [CrossRef]
  39. Shakeel, N.; Khan, J.; Al-Kahtani, A.A. Morphology-driven electrochemical attributes of Cu-MOF: A high-performance anodic material for battery supercapacitor hybrids. RSC Adv. 2024, 14, 33941–33951. [Google Scholar] [CrossRef]
  40. Vaitsis, C.; Sourkouni, G.; Argirusis, C. Metal Organic Frameworks (MOFs) and ultrasound: A review. Ultrason. Sonochemistry 2019, 52C, 106–119. [Google Scholar] [CrossRef]
  41. Muzaffar, N.; Afzal, A.M.; Hegazy, H.H.; Iqbal, M.W. Recent advances in two-dimensional metal-organic frameworks as an exotic candidate for the evaluation of redox-active sites in energy storage devices. J. Energy Storage 2023, 64, 107142. [Google Scholar] [CrossRef]
  42. Zhang, G.; Jin, L.; Zhang, R.; Bai, Y.; Zhu, R.; Pang, P. Recent advances in the development of electronically and ionically conductive metal-organic frameworks. Coord. Chem. Rev. 2021, 439, 213915. [Google Scholar] [CrossRef]
  43. Lee, S.J.; Telfer, S.G. Multicomponent metal-organic frameworks. Angew. Chem. Int. Ed. 2023, 135, e202306341. [Google Scholar] [CrossRef]
  44. Lin, Z.X.; Richardson, J.J.; Zhou, J.J.; Caruso, F. Direct synthesis of amorphous coordination polymers and metal-organic frameworks. Nat. Rev. Chem. 2023, 7, 273–286. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, H.N.; Meng, X.; Dong, L.Z.; Chen, Y.; Li, S.L.; Lan, Y.Q. Coordination polymer-based conductive materials: Ionic conductivity vs. electronic conductivity. J. Mater. Chem. A 2019, 7, 24059–24091. [Google Scholar] [CrossRef]
  46. Kumar, P.; Vellingiri, K.; Kim, K.-H.; Brown, R.J.C.; Manos, M.J. Modern progress in metal-organic frameworks and their composites for diverse applications. Microporous Mesoporous Mater. 2017, 253, 251–265. [Google Scholar] [CrossRef]
  47. Bernales, V.; Ortuño, M.A.; Truhlar, D.G.; Cramer, C.J.; Gagliardi, L. Computational Design of Functionalized Metal−Organic Framework Nodes for Catalysis. ACS Cent. Sci. 2018, 4, 5–19. [Google Scholar] [CrossRef]
  48. Coudert, F.X.; Fuchs, A.H. Computational characterization and prediction of metal–organic framework properties. Coord. Chem. Rev. 2016, 307, 211–236. [Google Scholar] [CrossRef]
  49. McCarver, G.A.; Rajeshkumar, T.; Vogiatzis, K.D. Computational catalysis for metal-organic frameworks: An overview. Coord. Chem. Rev. 2021, 436, 213777. [Google Scholar] [CrossRef]
  50. Hai, G.; Gao, H.; Huang, X.; Tan, L.; Xue, X.; Feng, S.; Wang, G. An efficient factor for fast screening of high-performance two-dimensional metal-organic frameworks towards catalyzing the oxygen evolution reaction. Chem. Sci. 2022, 13, 4397–4405. [Google Scholar] [CrossRef]
  51. Chatterjee, P.; Sengul, M.Y.; Kumar, A.; MacKerell, A.D., Jr. Harnessing deep learning for optimization of Lennard-Jones parameters for the polarizable classical Drude oscillator force field. J. Chem. Theory Comput. 2022, 18, 2388–2407. [Google Scholar] [CrossRef]
  52. Hai, G.; Wang, H. Theoretical studies of metal-organic frameworks: Calculation methods and applications in catalysis, gas separation, and energy storage. Coord. Chem. Rev. 2022, 469, 214670. [Google Scholar] [CrossRef]
  53. Demir, H.; Daglar, H.; Gulbalkan, H.C.; Aksu, G.O.; Keskin, S. Recent advances in computational modeling of MOFs: From molecular simulations to machine learning. Coord. Chem. Rev. 2023, 484, 215112. [Google Scholar] [CrossRef]
  54. Zhao, J.; Burke, A.F. Electrochemical capacitors: Materials, technologies and performance. Energy Storage Mater. 2021, 36, 31–55. [Google Scholar] [CrossRef]
  55. Song, J.L.; Chai, L.; Kumar, A.; Zhao, M.; Sun, Y.Z.; Liu, X.G.; Pan, J.Q. Precise tuning of hollow and pore size of bimetallic MOFs derivate to construct high-performance nanoscale materials for supercapacitors and sodium-ion batteries. Small 2023, 20, 2306272. [Google Scholar] [CrossRef]
  56. Kitchamsetti, N. A review on recent advances in Prussian blue, its analogues, and their derived materials as electrodes for high performance supercapacitors. J. Energy Storage 2023, 73, 108958. [Google Scholar] [CrossRef]
  57. Xu, B.; Zhang, H.; Mei, H.; Sun, D. Recent progress in metal-organic framework-based supercapacitor electrode materials. Coord. Chem. Rev. 2020, 420, 213438. [Google Scholar] [CrossRef]
  58. Yin, Z.; Wan, S.; Yang, J.; Kurmoo, M.; Zeng, M.-H. Recent advances in post-synthetic modification of metaleorganic frameworks: New types and tandem reactions. Coord. Chem. Rev. 2019, 378, 500–512. [Google Scholar] [CrossRef]
  59. Li, C.; Sun, X.; Yao, Y.; Hong, G. Recent advances of electrically conductive metal-organic frameworks in electrochemical applications. Mater. Today Nano 2021, 13, 100105. [Google Scholar] [CrossRef]
  60. Sundriyal, S.; Kaur, H.; Bhardwaj, S.K.; Mishra, S.; Kim, K.-H.; Deep, A. Metal-organic frameworks and their composites as efficient electrodes for supercapacitor applications. Coord. Chem. Rev. 2018, 369, 15–38. [Google Scholar] [CrossRef]
  61. Chen, D.; Wei, L.; Li, J.; Wu, Q. Nanoporous materials derived from metal-organic framework for super-capacitor application. J. Energy Storage 2020, 30, 101525. [Google Scholar] [CrossRef]
  62. Liu, Y.; Xu, X.; Shao, Z.; Jiang, S.P. Metal-organic frameworks derived porous carbon, metal oxides and metal sulfides-based compounds for supercapacitors application. Energy Storage Mater. 2020, 26, 1–22. [Google Scholar] [CrossRef]
  63. Wang, Y.; Li, B.; Zhang, B.; Tian, S.; Yang, X.; Ye, H.; Xia, Z.; Zheng, G. Application of MOFs-derived mixed metal oxides in energy storage. J. Electroanal. Chem. 2020, 878, 114576. [Google Scholar] [CrossRef]
  64. Li, S.; Luo, J.; Wang, J.; Zhu, Y.; Feng, J.; Fu, N.; Wang, H.; Guo, Y.; Tian, D.; Zheng, Y.; et al. Hybrid supercapacitors using metal-organic framework derived nickel-sulfur compounds. J. Colloid Interface Sci. 2024, 669, 265–274. [Google Scholar] [CrossRef]
  65. Hao, Y.; Guo, H.; Yang, F.; Zhang, J.; Wu, N.; Wang, M.; Li, C.; Yang, W. Hydrothermal synthesis of MWCNT/Ni-Mn-S composite derived from bimetallic MOF for high-performance electrochemical energy storage. J. Alloy. Compd. 2022, 911, 164726. [Google Scholar] [CrossRef]
  66. Qu, Y.; Sun, L.; Xie, F.; Hu, J.; Tan, H.; Zhang, Y. Metal-organic framework derived r-Ni3S2 nanoparticles with enriched sulfur vacancies for supercapacitor application. Appl. Surf. Sci. 2023, 623, 157037. [Google Scholar] [CrossRef]
  67. Lee, P.Y.; Lin, L.Y.; Yougbaré, S. Sulfurization of nickel-cobalt fluoride decorating ammonia ions as efficient active material of supercapacitor. J. Solid State Chem. 2022, 313, 123345. [Google Scholar] [CrossRef]
  68. Naseer, M.; Siyal, S.H.; Najam, T.; Afzal, S.; Iqbl, R.; Ismail, M.A.; Rauf, A.; Shah, S.S.A.; Nazir, M.A. Engineering of metal oxide integrated metal organic frameworks (MO@ MOF) composites for energy and environment sector. Mater. Sci. Eng. B 2025, 313, 117909. [Google Scholar] [CrossRef]
  69. Gao, Y.; Wu, J.; Zhang, W.; Tan, Y.; Zhao, J.; Tang, B. The electrochemical performance of SnO2 quantum dots@zeolitic imidazolate frameworks-8 (ZIF-8) composite material for supercapacitors. Mater. Lett. 2014, 128, 208–211. [Google Scholar] [CrossRef]
  70. Cui, H.; Liu, Y.; Ren, W.; Wang, M.; Zhao, Y. Large scale synthesis of highly crystallized SnO2 quantum dots at room temperature and their high electrochemical performance. Nanotechnology 2013, 24, 345602. [Google Scholar] [CrossRef]
  71. Yu, G.; Xie, X.; Pan, L.; Bao, Z.; Cui, Y. Hybrid nanostructured materials for high-performance electrochemical capacitors. Nano Energy 2013, 2, 213–234. [Google Scholar] [CrossRef]
  72. Gowdhaman, A.; Kumar, S.A.; Elumalai, D.; Balaji, C.; Sabarinathan, M.; Ramesh, R.; Navaneethan, M. Ni-MOF derived NiO/Ni/r-GO nanocomposite as a novel electrode material for high-performance asymmetric supercapacitor. J. Energy Storage 2023, 60, 106769. [Google Scholar] [CrossRef]
  73. Das, A.K.; Bera, R.; Maitra, A.; Karan, S.K.; Paria, S.; Halder, L.; Si, S.K.; Bera, A.; Khatua, B.B. Fabrication of an advanced asymmetric supercapacitor based on a microcubical PB@MnO2 hybrid and PANI/GNP composite with excellent electrochemical behaviour. J. Mater. Chem. A 2017, 5, 22242–22254. [Google Scholar] [CrossRef]
  74. Zhang, Y.Z.; Cheng, T.; Wang, Y.; Lai, W.Y.; Pang, H.; Huang, W. A simple approach to boost capacitance: Flexible supercapacitors based on manganese Oxides@MOFs via chemically induced in situ self-transformation. Adv. Mater. 2016, 28, 5242–5248. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, L.; Zhang, Y.; Huang, S.; Yuan, Y.; Li, H.; Jin, Z.; Wu, J.; Liao, Q.; Hu, L.; Lu, J.; et al. Co3O4/Ni-based MOFs on carbon cloth for flexible alkaline battery-supercapacitor hybrid devices and near-infrared photocatalytic hydrogen evolution. Electrochim. Acta 2018, 281, 189–197. [Google Scholar] [CrossRef]
  76. Zhang, F.; Zhang, J.; Song, J.; You, Y.; Jin, X.; Ma, J. Anchoring Ni-MOF nanosheet on carbon cloth by zeolite imidazole framework derived ribbonlike Co3O4 as integrated composite cathodes for advanced hybrid supercapacitors. Ceram. Int. 2021, 47, 14001–14008. [Google Scholar] [CrossRef]
  77. Zheng, J.H.; Zhang, R.M.; Yu, P.F.; Wang, X.G. Binary transition metal oxides (BTMO) (Co-Zn, Co-Cu) synthesis and high supercapacitor performance. J. Alloys Compd. 2019, 772, 359–365. [Google Scholar] [CrossRef]
  78. Vaitsis, C.; Mechili, M.; Argirusis, N.; Kanellou, E.; Pandis, P.K.; Sourkouni, G.; Zorpas, A.; Argirusis, C. Ultrasound-Assisted preparation methods of nanoparticles for energy-related applications. In Nanotechnology and the Environment; Sen, M., Ed.; IntechOpen: London, UK, 2020; pp. 77–103. [Google Scholar] [CrossRef]
  79. Shen, L.; Che, Q.; Li, H.; Zhang, X. Mesoporous NiCO2O4 nanowire arrays grown on carbon textiles as binder-free flexible electrodes for energy storage. Adv. Funct. Mater. 2014, 24, 2630–2637. [Google Scholar] [CrossRef]
  80. Wu, P.; Cheng, S.; Yao, M.; Yang, L.; Zhu, Y.; Liu, P.; Xing, O.; Zhou, J.; Wang, M.; Luo, H.; et al. A low-cost, self-standing NiCO2O4@CNT/CNT multilayer electrode for flexible asymmetric solid-state supercapacitors. Adv. Funct. Mater. 2017, 27, 1702160. [Google Scholar] [CrossRef]
  81. Zhang, F.; Ma, J.; Yao, H. Ultrathin Ni-MOF nanosheet coated NiCO2O4 nanowire arrays as a high-performance binder-free electrode for flexible hybrid supercapacitors. Ceram. Int. 2019, 45, 24279–24287. [Google Scholar] [CrossRef]
  82. Xiong, S.; Jiang, S.; Wang, J.; Lin, H.; Lin, M.; Weng, S.; Liu, S.; Jiao, Y.; Xu, Y.; Chen, J. A high-performance hybrid supercapacitor with NiO derived NiO@Ni-MOF composite electrodes. Electrochim. Acta 2020, 340, 135956. [Google Scholar] [CrossRef]
  83. Wang, J.; Zhong, Q.; Xiong, Y.; Cheng, D.; Zeng, Y.; Bu, Y. Fabrication of 3D Co-doped Ni-based MOF hierarchical micro-flowers as a high-performance electrode material for supercapacitors. Appl. Surf. Sci. 2019, 483, 1158–1165. [Google Scholar] [CrossRef]
  84. Wang, Q.; Wang, Q.; Xu, B.; Gao, F.; Gao, F.; Zhao, C. Flower-shaped multiwalled carbon nanotubes@nickel-trimesic acid MOF composite as a high-performance cathode material for energy storage. Electrochim. Acta 2018, 281, 69–77. [Google Scholar] [CrossRef]
  85. Li, N.; Li, Y.; Li, Q.; Zhao, Y.; Liu, C.S.; Pang, H. NiO nanoparticles decorated hexagonal Nickel-based metal-organic framework: Self-template synthesis and its application in electrochemical energy storage. J. Colloid Interface Sci. 2021, 581 Pt B, 709–718. [Google Scholar] [CrossRef]
  86. Chen, J.; Xu, J.; Zhou, S.; Zhao, N.; Wong, C.-P. Template-grown graphene/porous Fe2O3 nanocomposite: A high-performance anode material for pseudocapacitors. Nano Energy 2015, 15, 719–728. [Google Scholar] [CrossRef]
  87. Chameh, B.; Moradi, M.; Hessari, F.A. Decoration of metal organic frameworks with FeO3 for enhancing electrochemical performance of ZIF-(67 and 8) in energy storage application. Synth. Met. 2020, 269, 116540. [Google Scholar] [CrossRef]
  88. Chameh, B.; Moradi, M.; Hajati, S.; Hessari, F.A. Design and construction of ZIF(8 and 67) supported Fe3O4 composite as advanced materials of high performance supercapacitor. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 126, 114442. [Google Scholar] [CrossRef]
  89. Shabani Shayeh, J.; Salari, H. Dendritic fibrous nano metal organic framework: A magnetic core-shell structure as high performance material for electrochemical capacitors. J. Energy Storage 2020, 32, 101734. [Google Scholar] [CrossRef]
  90. Albiter, E.; Merlano, A.S.; Rojas, E.; Barrera-Andrade, J.M.; Salazar, Á.; Valenzuela, M.A. Synthesis, characterization, and photocatalytic performance of ZnO-graphene nanocomposites: A review. J. Compos. Sci. 2020, 5, 4. [Google Scholar] [CrossRef]
  91. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent developments of zinc oxide based photocatalyst in water treatment technology: A review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef]
  92. Sulciute, A.; Nishimura, K.; Gilshtein, E.; Cesano, F.; Viscardi, G.; Nasibulin, A.G.; Ohno, Y.; Rackauskas, S. ZnO nanostructures application in electrochemistry: Influence of morphology. J. Phys. Chem. C 2021, 125, 1472–1482. [Google Scholar] [CrossRef]
  93. Luo, W.; Zhang, Q.; Zhang, J.; Moioli, E.; Zhao, K.; Züttel, A. Electrochemical reconstruction of ZnO for selective reduction of CO2 to CO. Appl. Catal. B Environ. 2020, 273, 119060. [Google Scholar] [CrossRef]
  94. Saranya, M.; Ramachandran, R.; Wang, F. Graphene-zinc oxide (G-ZnO) nanocomposite for electrochemical supercapacitor applications. J. Sci. Adv. Mater. Dev. 2016, 1, 454–460. [Google Scholar] [CrossRef]
  95. Liu, Y.N.; Jin, L.N.; Wang, H.T.; Kang, X.H.; Bian, S.W. Fabrication of three-dimensional composite textile electrodes by metal-organic framework, zinc oxide, graphene and polyaniline for all-solid-state supercapacitors. J. Colloid Interface Sci. 2018, 530, 29–36. [Google Scholar] [CrossRef] [PubMed]
  96. Zhu, C.; He, Y.; Liu, Y.; Kazantseva, N.; Saha, P.; Cheng, Q. ZnO@MOF@PANI core-shell nanoarrays on carbon cloth for high-performance supercapacitor electrodes. J. Energy Chem. 2019, 35, 124–131. [Google Scholar] [CrossRef]
  97. Zuhri, F.U.; Diantoro, M.; Suryanti, L.; Suprayogi, T.; Nasikhudin, S.; Meevasana, W. ZnO-FC-NiCo MOF for prospective supercapacitor materials. Mater. Today Proc. 2021, 44, 3385–3389. [Google Scholar] [CrossRef]
  98. Reghunath, S.; Pinheiro, D.; Kr, S.D. A review of hierarchical nanostructures of TiO2: Advances and applications. Appl. Surf. Sci. Adv. 2021, 3, 100063. [Google Scholar] [CrossRef]
  99. Ramasubbu, V.; Omar, F.S.; Ramesh, K.; Ramesh, S.; Shajan, X.S. Three-dimensional hierarchical nano-structured porous TiO2 aerogel/Cobalt based metal-organic framework (MOF) composite as an electrode material for supercapattery. J. Energy Storage 2020, 32, 101750. [Google Scholar] [CrossRef]
  100. Feng, J.; Feng, J.; Zhang, C. Thermal conductivity of low density carbon aerogels. J. Porous Mater. 2011, 19, 551–556. [Google Scholar] [CrossRef]
  101. Keshavarz, L.; Ghaani, M.R.; MacElroy, J.M.D.; English, N.J. A comprehensive review on the application of aerogels in CO2-adsorption: Materials and characterization. Chem. Eng. J. 2021, 412, 128604. [Google Scholar] [CrossRef]
  102. Kitchamsetti, N.; Samtham, M.; Singh, D.; Choudhary, E.; Rondiya, S.R.; Ma, Y.R.; Cross, R.W.; Dzade, N.Y.; Devan, R.S. Hierarchical 2D MnO2@1D mesoporous NiTiO3 core-shell hybrid structures for high-performance supercapattery electrodes: Theoretical and experimental investigations. J. Electroanal. Chem. 2023, 936, 117359. [Google Scholar] [CrossRef]
  103. Khan, M.W.; Khan, A.S.; Rohma; Pai, S.H.S.; Mohapatra, G.; Kumar, A.; Ali, R.B.; Sial, Q.A.; Seo, H. Electrodeposition of redox active nickel-manganese metal organic framework and manganese sulfide composite for supercapacitors. Electrochim. Acta 2025, 513, 145554. [Google Scholar] [CrossRef]
  104. Feng, G.; Zhao, T.; Zhou, L.; Wang, X.; Ding, H.; Jiang, F.; Li, H.; Liu, Y.; Yu, Q.; Cao, H.; et al. Nanocubic transition metal sulfides (FeMn)S2@NC synthesized by MOFs facily for supercapacitors. Vacuum 2025, 231, 113819. [Google Scholar] [CrossRef]
  105. Kim, M.K.; Kim, W.J.; Kim, M.L.; Ahn, G.H.; Hong, J.P.; Ryu, G.H.; Hong, J. Hierarchically structured transition metal (Cu, Ni) sulfide core–shell electrode for high-performance supercapacitor. J. Alloys Compd. 2025, 1014, 178717. [Google Scholar] [CrossRef]
  106. Lu, P.; Jiang, X.; Guo, W.; Wang, L.; Zhang, T.; Boyjoo, Y.; Si, W.; Hou, F.; Liu, J.; Dou, S.X.; et al. A Ni–Co sulfide nanosheet/carbon nanotube hybrid film for high-energy and high-power flexible supercapacitors. Carbon 2021, 178, 355–362. [Google Scholar] [CrossRef]
  107. Liu, Y.; Huang, D.; Cheng, M.; Liu, Z.; Lai, C.; Zhang, C.; Zhou, C.; Xiong, W.; Qin, L.; Shao, B.; et al. Metal sulfide/MOF-based composites as visible-light-driven photocatalysts for enhanced hydrogen production from water splitting. Coord. Chem. Rev. 2020, 409, 213220. [Google Scholar] [CrossRef]
  108. Song, X.Z.; Zhang, N.; Wang, X.F.; Tan, Z. Recent advances of metal-organic frameworks and their composites toward oxygen evolution electrocatalysis. Mater. Today Energy 2021, 19, 100597. [Google Scholar] [CrossRef]
  109. Lee, C.S.; Lim, J.M.; Park, J.T.; Kim, J.H. Direct growth of highly organized, 2D ultra-thin nano-accordion Ni-MOF@NiS2@C core-shell for high performance energy storage device. Chem. Eng. J. 2021, 406, 126810. [Google Scholar] [CrossRef]
  110. Zhang, J.; Li, Y.; Han, M.; Xia, Q.; Chen, Q.; Chen, M. Constructing ultra-thin Ni-MOF@NiS2 nanosheets arrays derived from metal organic frameworks for advanced all-solid-state asymmetric supercapacitor. Mater. Res. Bull. 2021, 137, 111186. [Google Scholar] [CrossRef]
  111. Yue, L.; Wang, X.; Sun, T.; Liu, H.; Li, Q.; Wu, N.; Guo, H.; Yang, W. Ni-MOF coating MoS2 structures by hydrothermal intercalation as high-performance electrodes for asymmetric supercapacitors. Chem. Eng. J. 2019, 375, 121959. [Google Scholar] [CrossRef]
  112. Wei, J.; Hu, F.; Shen, X.; Chen, B.; Chen, L.; Wang, Z.; Lv, C.; Ouyang, Q. Defective core–shell NiCO2S4/MnO2 nanocomposites for high performance solid-state hybrid supercapacitors. J. Colloid Interface Sci. 2023, 649, 665–674. [Google Scholar] [CrossRef]
  113. Ali, M.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Imran, M.; Ashraf, F.; Rehman, A.U.; Muhammad, F. 2D-TMDs based electrode material for supercapacitor applications. Int. J. Energy Res. 2022, 46, 22336–22364. [Google Scholar] [CrossRef]
  114. Wang, J.; Li, S.; Zhu, Y.; Zhai, S.; Liu, C.; Fu, N.; Hou, S.; Niu, Y.; Luo, J.; Mu, S.; et al. Metal-organic frameworks-derived NiSe@RGO composites for high-performance hybrid supercapacitors. J. Electroanal. Chem. 2022, 919, 116548. [Google Scholar] [CrossRef]
  115. Zhang, N.; Meng, Q.; Wu, H.; Hu, X.; Zhang, M.; Zhou, A.; Li, Y.; Huang, Y.; Li, L.; Wu, F.; et al. Co-MOF as Stress-Buffered Architecture: An engineering for improving the performance of NiS/SnO2 heterojunction in lithium storage. Adv. Energy Mater. 2023, 13, 2300413. [Google Scholar] [CrossRef]
  116. Hu, M.-L.; Masoomi, M.Y.; Morsali, A. Template strategies with MOFs. Coord. Chem. Rev. 2019, 387, 415–435. [Google Scholar] [CrossRef]
  117. Mei, J.; Liao, T.; Ayoko, G.A.; Bell, J.; Sun, Z. Cobalt oxide-based nanoarchitectures for electrochemical energy applications. Prog. Mater. Sci. 2019, 103, 596–677. [Google Scholar] [CrossRef]
  118. Chen, T.-Y.; Lin, L.-Y.; Geng, D.-S.; Lee, P.-Y. Systematic synthesis of ZIF-67 derived Co3O4 and N-doped carbon composite for supercapacitors via successive oxidation and carbonization. Electrochim. Acta 2021, 376, 137986. [Google Scholar] [CrossRef]
  119. Xu, J.; Liu, S.; Liu, Y. Co3O4/ZnO nanoheterostructure derived from core-shell ZIF-8@ZIF-67 for supercapacitors. RSC Adv. 2016, 6, 52137–52142. [Google Scholar] [CrossRef]
  120. Kitchamsetti, N.; Kim, D. High performance hybrid supercapacitor based on hierarchical MOF derived CoFe2O4 and NiMn2O4 composite for efficient energy storage. J. Alloys Compd. 2023, 959, 170483. [Google Scholar] [CrossRef]
  121. Iqbal, M.Z.; Zakir, A.; Shoukat, W.; Badi, N.; Afzal, A.M.; Fouda, A.M.; Hegazy, H.H. Enhancing the electrochemical efficiency of metal-organic frameworks through integration of chromium nitride interfacial layer for efficient hybrid supercapacitors. J. Energy Storage 2025, 105, 114606. [Google Scholar] [CrossRef]
  122. Wang, B.; Tan, W.; Fu, R.; Mao, H.; Kong, Y.; Qin, Y.; Tao, Y. Hierarchical mesoporous Co3O4/C@MoS2 core-shell structured materials for electrochemical energy storage with high supercapacitive performance. Synth. Met. 2017, 233, 101–110. [Google Scholar] [CrossRef]
  123. Hou, S.; Lian, Y.; Bai, Y.; Zhou, Q.; Ban, C.; Wang, Z.; Zhao, J.; Zhang, H. Hollow dodecahedral Co3S4@NiO derived from ZIF-67 for supercapacitor. Electrochim. Acta 2020, 341, 136053. [Google Scholar] [CrossRef]
  124. Zhao, J.; Li, H.; Li, C.; Zhang, Q.; Sun, J.; Wang, X.; Guo, J.; Xie, L.; Xie, J.; He, B.; et al. MOF for template-directed growth of well-oriented nanowire hybrid arrays on carbon nanotube fibers for wearable electronics integrated with triboelectric nanogenerators. Nano Energy 2018, 45, 420–431. [Google Scholar] [CrossRef]
  125. Prasad Ojha, G.; Muthurasu, A.; Prasad Tiwari, A.; Pant, B.; Chhetri, K.; Mukhiya, T.; Dahal, B.; Lee, M.; Park, M.; Kim, H.-Y. Vapor solid phase grown hierarchical CuxO NWs integrated MOFs-derived CoS2 electrode for high-performance asymmetric supercapacitors and the oxygen evolution reaction. Chem. Eng. J. 2020, 399, 125532. [Google Scholar] [CrossRef]
  126. Govindan, R.; Hong, X.-J.; Sathishkumar, P.; Cai, Y.-P.; Gu, F.L. Construction of metal-organic framework-derived CeO2/C integrated MoS2 hybrid for high-performance asymmetric supercapacitor. Electrochim. Acta 2020, 353, 136502. [Google Scholar] [CrossRef]
  127. Gayathri, S.; Arunkumar, P.; Saha, D.; Han, J.H. Composition engineering of ZIF-derived cobalt phosphide/cobalt monoxide heterostructures for high-performance asymmetric supercapacitors. J. Colloid Interface Sci. 2021, 588, 557–570. [Google Scholar] [CrossRef]
  128. Hu, X.; Zhang, S.; Sun, J.; Yu, L.; Qian, X.; Hu, R.; Wang, Y.; Zhao, H.; Zhu, J. 2D Fe-containing cobalt phosphide/cobalt oxide lateral heterostructure with enhanced activity for oxygen evolution reaction. Nano Energy 2019, 56, 109–117. [Google Scholar] [CrossRef]
  129. Wang, J.; Gao, R.; Zheng, L.; Chen, Z.; Wu, Z.; Sun, L.; Hu, Z.; Liu, X. CoO/CoP heterostructured nanosheets with an O-P interpenetrated Interface as a bifunctional electrocatalyst for Na-O2 battery. ACS Catal. 2018, 8, 8953–8960. [Google Scholar] [CrossRef]
  130. Guo, J.W.; Zhao, H.B.; Yang, Z.W.; Wang, Y.W.; Liu, X.L.; Wang, L.F.; Zhao, Z.H.; Wang, A.Z.; Ding, L.H.; Liu, H.; et al. Hierarchical porous 3D Ni3N-CoN/NC heterojunction nanosheets with nitrogen vacancies for high-performance flexible supercapacitor. Nano Energy 2023, 116, 108763. [Google Scholar] [CrossRef]
  131. Zhao, J.J.; Liu, N.; Sun, Y.Z.; Xu, Q.H.; Pan, J.Q. Nitrogen-modified spherical porous carbon derived from aluminum-based metal-organic frameworks as activation-free materials for supercapacitors. J. Energy Storage 2023, 73, 109070. [Google Scholar] [CrossRef]
  132. Zhang, S.; Dai, P.; Liu, H.; Yan, L.; Song, H.; Liu, D.; Zhao, X. Metal-organic framework derived porous flakes of cobalt chalcogenides (CoX, X = O, S, Se and Te) rooted in carbon fibers as flexible electrode materials for pseudocapacitive energy storage. Electrochim. Acta 2021, 369, 137681. [Google Scholar] [CrossRef]
  133. Zhang, A.; Zhang, Q.; Fu, H.C.; Zong, H.W.; Guo, H.W. Metal-organic frameworks and their derivatives based nanostructure with different dimensionalities for supercapacitors. Small 2023, 19, 2303911. [Google Scholar] [CrossRef]
  134. Kumar, S.; Kumar, V.; Bulla, M.; Devi, R.; Dahiya, R.; Sisodiya, A.K.; Singh, R.B.; Mishra, A.K. Hydrothermally reduced graphene oxide based electrodes for high-performance symmetric supercapacitor. Mater. Lett. 2024, 364, 136364. [Google Scholar] [CrossRef]
  135. Feng, W.; Liu, C.; Liu, Z.; Pang, H. In-situ growth of N-doped graphene-like carbon/MOF nanocomposites for high-performance supercapacitor. Chin. Chem. Lett. 2024, 35, 109552. [Google Scholar] [CrossRef]
  136. Kausar, A.; Ahmad, I. Graphene-MOF hybrids in high-tech energy devices—Present and future advances. Hybrid Adv. 2024, 5, 100150. [Google Scholar] [CrossRef]
  137. Saxena, N.; Bondarde, M.P.; Lokhande, K.D.; Bhakare, M.A.; Dhumal, P.S.; Some, S. One-pot synthesis of rGO/Ppy/Zn-MOF, ternary composite for High-Performance supercapacitor application. Chem. Phys. Lett. 2024, 856, 141605. [Google Scholar] [CrossRef]
  138. Lian, P.C.; Liu, H.H.; Li, H.; Zhang, Y.; Mei, Y. A facile and mild route to synthesize ultralight and flexible 3D functionalized graphene. J. Porous. Mater. 2018, 25, 905–911. [Google Scholar] [CrossRef]
  139. Tang, Z.; Lei, Y.; Deng, Y.; Chen, J.; Liu, K.; Zhang, C.; Wang, T.; Lei, Y. MOF-derived MoS2/nitrogen-doped graphene aerogel for supercapacitor electrodes. Diam. Relat. Mater. 2025, 153, 112098. [Google Scholar] [CrossRef]
  140. Siddiqui, R.; Rani, M.; Shah, A.A.; Razaq, A.; Iqbal, R.; Neffati, R.; Arshad, M. Fabrication of tricarboxylate-neodymium metal organic frameworks and its nanocomposite with graphene oxide by hydrothermal synthesis for a symmetric supercapacitor electrode material. Mater. Sci. Eng. B 2023, 295, 116530. [Google Scholar] [CrossRef]
  141. Ashraf, M.; Umar, E.; Sunny, M.A.; Iqbal, M.W.; Hassan, H.; Alrobei, H.; Ismayilova, N.A.; Alawaideh, Y.M.; Al-Buriahi, M.S.; Alomairy, S. Reduced graphene oxides embedded zeolitic imidazolate framework-8 and copper sulfide as electrodes for asymmetric supercapacitors and hydrogen evolution reactions. J. Energy Storage 2024, 104, 114479. [Google Scholar] [CrossRef]
  142. Kumar, R.V.; Vickraman, P.; Raja, T.A.; Bharathi, M.S.R. Exploratory mass dispersoid rGO in-situ influence on nickel phosphate-trimesic acid metal-organic framework for higher energy density hybrid supercapacitors. J. Energy Storage 2024, 102, 114140. [Google Scholar] [CrossRef]
  143. Hassan, H.; Shoaib, M.; Khan, K.; Ghanem, M.A.; Osman, M. Graphene quantum dots decorated on chromium oxide and zirconium metal-organic framework composite (GQDs@Zr-MOF/Cr2O3) for asymmetric supercapacitors and hydrogen production. Mater. Chem. Phys. 2025, 332, 130225. [Google Scholar] [CrossRef]
  144. Xiong, C.; Zhang, Y.; Zheng, C.; Yin, Y.; Xiong, Q.; Zhao, M.; Wang, B. Fabrication of metal-organic framework@cellulose nanofibers/reduced graphene oxide-Vitrimer composite electrode materials with shape memory for supercapacitors. Electrochim. Acta 2024, 493, 144373. [Google Scholar] [CrossRef]
  145. Wang, H.; Wang, L.; Zhao, P.; Zhang, X.; Lu, X.; Qiu, Z.; Qi, B.; Yao, R.; Huang, Y.; Wang, L.; et al. Metal-organic framework-mediated construction of confined ultrafine nickel phosphide immobilized in reduced graphene oxide with excellent cycle stability for asymmetric supercapacitors. J. Colloid Interface Sci. 2023, 649, 616–625. [Google Scholar] [CrossRef]
  146. Tang, X.R.; Li, N.; Pang, H. Metal-organic frameworks-derived metal phosphides for electrochemistry application. Green Energy Environ. 2022, 7, 636–661. [Google Scholar] [CrossRef]
  147. Zhou, Q.; Gong, Y.; Tao, K. Calcination/phosphorization of dual Ni/Co-MOF into NiCoP/C nanohybrid with enhanced electrochemical property for high energy density asymmetric supercapacitor. Electrochim. Acta 2019, 320, 134582. [Google Scholar] [CrossRef]
  148. Liu, S.; Xu, W.; Feng, K.; Shi, X.; Xu, Z.; Teng, R.; Fan, X.; Wang, C. Three-dimensional heterostructured nickel phosphide @ nickel cobalt phosphide nanocomposites with highly porous surface for high-performance supercapacitor. Colloids Surf. A Physicochem. Eng. Asp. 2024, 686, 133342. [Google Scholar] [CrossRef]
  149. Yi, M.; Lu, B.; Zhang, X.; Tan, Y.; Zhu, Z.; Pan, Z.; Zhang, J. Ionic liquid-assisted synthesis of nickel cobalt phosphide embedded in N, P codoped-carbon with hollow and folded structures for efficient hydrogen evolution reaction and supercapacitor. Appl. Catal. B Environ. 2021, 283, 119635. [Google Scholar] [CrossRef]
  150. Zhu, Y.; Lao, J.; Xu, F.; Sun, L.; Shao, Q.; Luo, Y.; Fang, S.; Chen, Y.; Yu, C.; Zou, Y. A structurally controllable flower-shaped phosphide derived from metal-organic frameworks for high-performance supercapacitors. J. Electroanal. Chem. 2024, 966, 118376. [Google Scholar] [CrossRef]
  151. Liu, S.; Xu, W.; Feng, K.; Shi, X.; Wang, C. Bimetallic MOF derived NiMn phosphide for high-performance supercapacitor electrode material. J. Energy Storage 2024, 96, 112684. [Google Scholar] [CrossRef]
  152. Zhou, H.; Nasser, R.; Zhang, L.; Li, Z.; Li, F.; Song, J.-M. Copper/Cobalt based metal phosphide composites as positive material for supercapacitors. Chem. Eng. J. 2024, 496, 154102. [Google Scholar] [CrossRef]
  153. Lao, J.; Zhu, Y.; Xu, F.; Sun, L.; Fang, S.; Shao, Q.; Liao, L.; Guan, Y.; Zou, Y. A novel 1D MXene fibers-loaded metal-organic framework derived bimetallic sulfide for high-performance supercapacitors. Fuel 2024, 362, 130904. [Google Scholar] [CrossRef]
  154. Ramesh, S.; Rabani, I.; Sivasamy, A.; Kakani, V.; Haldorai, Y.; Seo, Y.-S.; Kim, J.-H.; Kim, H.S. Fabrication of nickel-copper sulfide nanoparticles decorated on metal-organic framework composite for supercapacitor application by hydrothermal process. J. Alloys Compd. 2024, 977, 173375. [Google Scholar] [CrossRef]
  155. Guo, H.; Zhang, H.; Wu, N.; Pan, Z.; Li, C.L.; Chen, Y.; Cao, Y.J.; Yang, W. Trimesic acid-modified 2D NiCo-MOF for high-capacity supercapacitors. J. Alloys Compd. 2023, 934, 167779. [Google Scholar] [CrossRef]
  156. Liu, Y.; Li, Y.; Kang, H.; Jin, T.; Jiao, L. Design, synthesis, and energy-related applications of metal sulfides. Mater. Horiz. 2016, 3, 402–421. [Google Scholar] [CrossRef]
  157. Ouyang, Y.; Zhang, B.; Wang, C.; Xia, X.; Lei, W.; Hao, Q. Bimetallic metal-organic framework derived porous NiCo2S4 nanosheets arrays as binder-free electrode for hybrid supercapacitor. Appl. Surf. Sci. 2021, 542, 148621. [Google Scholar] [CrossRef]
  158. Wang, D.; Tian, L.; Huang, J.; Li, D.; Liu, J.; Xu, Y.; Ke, H.; Wei, Q. “One for two” strategy to prepare MOF- derived NiCo2S4 nanorods grown on carbon cloth for high-performance asymmetric supercapacitors and efficient oxygen evolution reaction. Electrochim. Acta 2020, 334, 135636. [Google Scholar] [CrossRef]
  159. Cai, P.; Liu, T.; Zhang, L.; Cheng, B.; Yu, J. ZIF-67 derived nickel cobalt sulfide hollow cages for high-performance supercapacitors. Appl. Surf. Sci. 2020, 504, 144501. [Google Scholar] [CrossRef]
  160. Li, G.-C.; Liu, M.; Wu, M.-K.; Liu, P.-F.; Zhou, Z.; Zhu, S.-R.; Liu, R.; Han, L. MOF-derived self-sacrificing route to hollow NiS2/ZnS nanospheres for high performance supercapacitors. RSC Adv. 2016, 6, 103517–103522. [Google Scholar] [CrossRef]
  161. Cai, F.; Sun, R.; Kang, Y.; Chen, H.; Chen, M.; Li, Q. One-step strategy to a three-dimensional NiS-reduced graphene oxide hybrid nanostructure for high performance supercapacitors. RSC Adv. 2015, 5, 23073–23079. [Google Scholar] [CrossRef]
  162. Ruan, Y.; Jiang, J.; Wan, H.; Ji, X.; Miao, L.; Peng, L.; Zhang, B.; Lv, L.; Liu, J. Rapid self-assembly of porous square rod-like nickel persulfide via a facile solution method for high-performance supercapacitors. J. Power Sources 2016, 301, 122–130. [Google Scholar] [CrossRef]
  163. Shrivastav, V.; Sundriyal, S.; Goel, P.; Shrivastav, V.; Tiwari, U.K.; Deep, A. ZIF-67 derived Co3S4 hollow microspheres and WS2 nanorods as a hybrid electrode material for flexible 2V solid-state supercapacitor. Electrochim. Acta 2020, 345, 136194. [Google Scholar] [CrossRef]
  164. Zhao, W.; Yan, G.; Zheng, Y.; Liu, B.; Jia, D.; Liu, T.; Cui, L.; Zheng, R.; Wei, D.; Liu, J. Bimetal-organic framework derived Cu(NiCo)2S4/Ni3S4 electrode material with hierarchical hollow heterostructure for high performance energy storage. J. Colloid Interface Sci. 2020, 565, 295–304. [Google Scholar] [CrossRef] [PubMed]
  165. Li, M.D.; Jiang, X.Y.; Liu, J.J.; Liu, Q.; Lv, N.J.; Qi, N.; Chen, Z.Q. A flower-like Co/Ni bimetallic metal-organic framework based electrode material with superior performance in supercapacitors. J. Alloys Compd. 2023, 930, 167354. [Google Scholar] [CrossRef]
  166. Xiong, D.; Gu, M.; Chen, C.; Lu, C.; Yi, F.-Y.; Ma, X. Rational design of bimetallic metal-organic framework composites and their derived sulfides with superior electrochemical performance to remarkably boost oxygen evolution and supercapacitors. Chem. Eng. J. 2021, 404, 127111. [Google Scholar] [CrossRef]
  167. Nataraj, N.; Dash, P.; Sakthivel, R.; Lin, Y.-C.; Fang, H.-W.; Chung, R.-J. Simultaneous electrochemical and colorimetric detection of tri-heavy metal ions in environmental water samples employing 3D-MOF/nickel selenide as a synergistic catalyst. Chem. Eng. J. 2024, 485, 149965. [Google Scholar] [CrossRef]
  168. Al-Shaibah, F.N.; Ibrahim, M.A.A.; Abu Hatab, A.S.; Abotaleb, A.; Sinopoli, A.; Zekri, A.; Ahmad, Y.H.; Al-Qaradawi, S.Y. Rational synthesis of MOF-Derived Cobalt-Based binary selenides nanocrystals for electrochemical oxygen evolution reaction. Appl. Surf. Sci. 2025, 688, 162479. [Google Scholar] [CrossRef]
  169. Sheikhi, S.; Jalali, F. Copper selenide—Porous carbon derived from metal-organic frameworks as an efficient electrocatalyst for methanol oxidation. Int. J. Hydrogen Energy 2024, 55, 864–874. [Google Scholar] [CrossRef]
  170. Zhai, Z.; Yan, W.; Dong, L.; Wang, J.; Chen, C.; Lian, J.; Wang, X.; Xia, D.; Zhang, J. Multi-dimensional materials with layered structures for supercapacitors: Advanced synthesis, supercapacitor performance and functional mechanism. Nano Energy 2020, 78, 105193. [Google Scholar] [CrossRef]
  171. Zhang, Y.; Pan, A.; Wang, Y.; Cao, X.; Zhou, Z.; Zhu, T.; Liang, S.; Cao, G. Self-templated synthesis of N-doped CoSe2/C double-shelled dodecahedra for high-performance supercapacitors. Energy Storage Mater. 2017, 8, 28–34. [Google Scholar] [CrossRef]
  172. Mei, H.; Zhang, L.; Zhang, K.; Gao, J.; Zhang, H.; Huang, Z.; Xu, B.; Sun, D. Conversion of MOF into carbon-coated NiSe2 yolk-shell microspheres as advanced battery-type electrodes. Electrochim. Acta 2020, 357, 136866. [Google Scholar] [CrossRef]
  173. Wang, Q.; Ran, X.; Shao, W.; Miao, M.; Zhang, D. High performance flexible supercapacitor based on metal-organic-framework derived CoSe2 nanosheets on carbon nanotube film. J. Power Sources 2021, 490, 229517. [Google Scholar] [CrossRef]
  174. Sun, P.; Zhang, J.; Huang, J.; Wang, L.; Wang, P.; Cai, C.; Lu, M.; Yao, Z.; Yang, Y. Bimetallic MOF-derived (CuCo)Se nanoparticles embedded in nitrogen-doped carbon framework with boosted electrochemical performance for hybrid supercapacitor. Mater. Res. Bull. 2021, 137, 111196. [Google Scholar] [CrossRef]
  175. Li, Z.; Tian, M.; Chen, Y.; Liu, Y.; Cai, Y.; Wei, W. MOFs derived (Ni0.75Co0.25)Se2 nanoparticles embedded in N-doped nanocarbon for hybrid supercapacitors. Ceram. Int. 2021, 47, 12623–12630. [Google Scholar] [CrossRef]
  176. Acharya, J.; Ojha, G.P.; Pant, B.; Park, M. Construction of self-supported bimetallic MOF-mediated hollow and porous tri-metallic selenide nanosheet arrays as battery-type electrodes for high-performance asymmetric supercapacitors. J. Mater. Chem. A 2021, 9, 23977–23993. [Google Scholar] [CrossRef]
  177. Dong, W.; Li, X.; Ye, F.; Xu, X.; Zhang, X.; Yang, F.; Shen, D.; Hong, X.; Yang, S. Synergistically enhanced multimetallic selenide electrode materials derived from ZIF-67 templates for high-performance supercapacitors. J. Energy Storage 2025, 114, 115870. [Google Scholar] [CrossRef]
  178. Charis Caroline, S.; Das, B.; Pramana, S.S.; Batabyal, S.K. Nickel sulfide-nickel sulfoselenide nanosheets as a potential electrode material for high performance supercapacitor with extended shelf life. J. Energy Storage 2023, 68, 107812. [Google Scholar] [CrossRef]
  179. Zhao, Y.; Wang, S.; Ye, F.; Liu, W.; Lian, J.; Li, G.; Wang, H.; Hu, L.; Wu, L. Hierarchical mesoporous selenium@bimetallic selenide quadrilateral nanosheet arrays for advanced flexible asymmetric supercapacitors. J. Mater. Chem. A 2022, 10, 16212–16223. [Google Scholar] [CrossRef]
  180. Yi, M.; Wu, A.; Chen, Q.; Cai, D.; Zhan, H. In situ confined conductive nickel cobalt sulfoselenide with tailored composition in graphitic carbon hollow structure for energy storage. Chem. Eng. J. 2018, 351, 678–687. [Google Scholar] [CrossRef]
  181. Lu, X.; Chen, Q.; Wu, L.; Cui, S.; Li, G.; Zhao, W.; Han, L. A multi-interface bimetallic sulfoselenide–selenite heterojunction as a battery-type cathode for high-performance supercapacitors. New J. Chem. 2024, 48, 15227–15239. [Google Scholar] [CrossRef]
  182. Huang, S.; Shi, X.-R.; Sun, C.; Duan, Z.; Ma, P.; Xu, S. The Application of Metal–Organic Frameworks and Their Derivatives for Supercapacitors. Nanomaterials 2020, 10, 2268. [Google Scholar] [CrossRef]
  183. Lan, M.; Wang, X.; Zhao, R.; Dong, M.; Fang, L.; Wang, L. Metal-organic framework-derived porous MnNi2O4 microflower as an advanced electrode material for high-performance supercapacitors. J. Alloys Compd. 2020, 821, 153546. [Google Scholar] [CrossRef]
  184. Javed, M.S.; Aslam, M.K.; Asim, S.; Batool, S.; Idrees, M.; Hussain, S.; Shah, S.S.A.; Saleem, M.; Mai, W.; Hu, C. High-performance flexible hybrid-supercapacitor enabled by pairing binder-free ultrathin Ni–Co–O nanosheets and metal-organic framework derived N-doped carbon nanosheets. Electrochim. Acta 2020, 349, 136384. [Google Scholar] [CrossRef]
  185. Liu, J.; Wang, Z.; Bi, R.; Mao, F.; Wang, K.; Wu, H.; Wang, X. A polythreaded MnII-MOF and its super-performances for dye adsorption and supercapacitors. Inorg. Chem. Front. 2020, 7, 718–730. [Google Scholar] [CrossRef]
  186. Deng, T.; Shi, X.; Zhang, W.; Wang, Z.; Zheng, W. In-plane Assembly of Distinctive 2D MOFs with Optimum Supercapacitive Performance. iScience 2020, 23, 101220. [Google Scholar] [CrossRef] [PubMed]
  187. Shi, X.; Yu, J.; Liu, Q.; Shao, L.; Zhang, Y.; Sun, Z.; Huang, H. Metal-Organic-Framework-Derived N-, P-, and O-Codoped Nickel/Carbon Composites Homogeneously Decorated on Reduced Graphene Oxide for Energy Storage. ACS Appl. Nano Mater. 2020, 3, 5625–5636. [Google Scholar] [CrossRef]
  188. Maliha, Z.; Rani, M.; Neffati, R.; Mahmood, A.; Iqbal, M.Z.; Shah, A. Investigation of copper/cobalt MOFs nanocomposite as an electrode material in supercapacitors. Int. J. Energy Res. 2022, 46, 17404–17415. [Google Scholar] [CrossRef]
  189. Singh, M.K.; Krishnan, S.; Singh, K.; Rai, D.K. CNT Interwoven Cu-MOF: A Synergistic Electrochemical Approach for Solid-State Supercapacitor and Hydrogen Evolution Reaction. Energy Fuels 2024, 38, 12098–12110. [Google Scholar] [CrossRef]
  190. Anwer, A.H.; Ansari, M.Z.; Mashkoor, F.; Zhu, S.; Shoeb, M.; Jeong, C. Synergistic effect of carbon nanotube and tri-metallic MOF nanoarchitecture for electrochemical high-performance asymmetric supercapacitor applications and their charge storage mechanism. J. Alloys Compd. 2023, 955, 170038. [Google Scholar] [CrossRef]
  191. Zaka, A.; Iqbal, M.W.; Afzal, A.M.; Hassan, H.; Alharthi, S.; Amin, M.A.; Saeedi, A.M.; Albargi, H.B.; Alhadrami, A.; Alqarni, N.D.; et al. Synergistic innovations in energy Storage: Cu-MOF infused with CNT for supercapattery devices and hydrogen evolution reaction. Inorg. Chem. Commun. 2024, 159, 111739. [Google Scholar] [CrossRef]
  192. Ji, Y.; Li, W.; You, Y.; Xu, G. In situ synthesis of M (Fe, Cu, Co and Ni)-MOF@MXene composites for enhanced specific capacitance and cyclic stability in supercapacitor electrodes. Chem. Eng. J. 2024, 496, 154009. [Google Scholar] [CrossRef]
  193. Iqbal, M.Z.; Aziz, U.; Aftab, S.; Wabaidur, S.M.; Siddique, S.; Iqbal, M.J. A Hydrothermally Prepared Lithium and Copper MOF Composite as Anode Material for Hybrid Supercapacitor Applications. Angew. Chem. 2023, 8, e202204554. [Google Scholar] [CrossRef]
  194. Suganthi, S.; Ahmad, K.; Oh, T.H. Progress in MOFs and MOFs-Integrated MXenes as Electrode Modifiers for Energy Storage and Electrochemical Sensing Applications. Molecules 2024, 29, 5373. [Google Scholar] [CrossRef] [PubMed]
  195. Do, H.H.; Cho, J.H.; Han, S.M.; Ahn, S.H.; Kim, S.Y. Metal–Organic-Framework- and MXene-Based Taste Sensors and Glucose Detection. Sensors 2021, 21, 7423. [Google Scholar] [CrossRef]
  196. Su, H.; Jin, C.; Zhang, X.; Yu, Z.; Zeng, X. Recent progress in the synthesis and electrocatalytic application of MXene-based metal phosphide composites. Carbon Neutralization 2024, 3, 1009–1035. [Google Scholar] [CrossRef]
  197. Basha, S.I.; Shah, S.S.; Helal, A.; Aziz, M.A.; Yoo, D.-Y. Unveiling the limitless potential: Exploring metal–organic frameworks (MOFs)/MXene based construction materials. Case Stud. Constr. Mater. 2024, 21, e03586. [Google Scholar] [CrossRef]
  198. Yang, H.; Zhang, G.-X.; Zhou, H.-J.; Sun, Y.-Y.; Pang, H. Metal–Organic Frameworks Meet MXene: New Opportunities for Electrochemical Application. Energy Mater. Adv. 2023, 4, 0033. [Google Scholar] [CrossRef]
  199. Tan, H.; Zhou, Y.; Qiao, S.-Z.; Fan, H.J. Metal organic framework (MOF) in aqueous energy devices. Mater. Today 2021, 48, 270–284. [Google Scholar] [CrossRef]
  200. Ding, M.; Jiang, H.-L. Improving Water Stability of Metal–Organic Frameworks by a General Surface Hydrophobic Polymerization. CCS Chem. 2021, 3, 2740–2748. [Google Scholar] [CrossRef]
  201. Woodliffe, J.L.; Molinar-Díaz, J.; Clowes, R.; Hussein, O.H.; Lester, E.; Ferrari, R.; Ahmed, I.; Laybourn, A. Continuous flow synthesis of MOF UTSA-16(Zn), mixed-metal and magnetic composites for CO2 capture—Toward scalable manufacture. J. Environ. Chem. Eng. 2024, 12, 114167. [Google Scholar] [CrossRef]
  202. Woodliffe, J.L.; Ferrari, R.S.; Ahmed, I.; Laybourn, A. Evaluating the purification and activation of metal-organic frameworks from a technical and circular economy perspective. Coord. Chem. Rev. 2021, 428, 213578. [Google Scholar] [CrossRef]
  203. Laybourn, A.; Robertson, K.; Slater, A.G. Quid pro flow. J. Am. Chem. Soc. 2023, 145, 4355–4365. [Google Scholar] [CrossRef]
Figure 1. Charge storage process in electric double layer capacitors, pseudo capacitors, and battery resembling electrodes. (Reproduced from [9] with permission).
Figure 1. Charge storage process in electric double layer capacitors, pseudo capacitors, and battery resembling electrodes. (Reproduced from [9] with permission).
Batteries 11 00181 g001
Figure 2. Synthesis of Cu-MOF through the hydrothermal and sonochemical approach. (Reproduced from [39] with permission).
Figure 2. Synthesis of Cu-MOF through the hydrothermal and sonochemical approach. (Reproduced from [39] with permission).
Batteries 11 00181 g002
Figure 3. SEM images of Zn-adp prepared via (ad) conventional and (e,f) sonochemical synthesis. (Reproduced from [38] with permission).
Figure 3. SEM images of Zn-adp prepared via (ad) conventional and (e,f) sonochemical synthesis. (Reproduced from [38] with permission).
Batteries 11 00181 g003
Figure 4. Schematic illustrations of the MOF materials as a versatile platform for electronic conduction and ionic conduction. (Reproduced from [41] with permission).
Figure 4. Schematic illustrations of the MOF materials as a versatile platform for electronic conduction and ionic conduction. (Reproduced from [41] with permission).
Batteries 11 00181 g004
Figure 5. SEM images of ribbonlike Co3O4 nanoarrays (ac), Co3O4@Ni-MOF composites (df), Ni-MOF directly grown on carbon cloth (gi), and EDX element mappings of Co3O4@Ni-MOF composites (j). (Reproduced from [76] with permission).
Figure 5. SEM images of ribbonlike Co3O4 nanoarrays (ac), Co3O4@Ni-MOF composites (df), Ni-MOF directly grown on carbon cloth (gi), and EDX element mappings of Co3O4@Ni-MOF composites (j). (Reproduced from [76] with permission).
Batteries 11 00181 g005
Figure 6. Schematic illustration of the preparation of Ni–MOF@NiO. (Reproduced from [85] with permission).
Figure 6. Schematic illustration of the preparation of Ni–MOF@NiO. (Reproduced from [85] with permission).
Batteries 11 00181 g006
Figure 7. (a) SEM images of CNTFs. (b) Low- and (c) high-magnification SEM images of CNTFs@H-Co3O4 NA. (d) SEM images of as-prepared CNTFs@H-Co3O4@ZIF-67 HA. Inset: enlarged SEM image. (e) SEM images of as-prepared CNTFs@H-Co3O4@CoNC HA. Inset: enlarged SEM image. (f) Powder XRD patterns of as-prepared CNTFs@H-Co3O4@CoNC HA. (g) Raman spectra of as-prepared CNTFs@H-Co3O4@CoNC HA. (h) EDS of the as-prepared samples. (i) N2 adsorption/desorption isotherms of the as-prepared CNTFs@H-Co3O4@CoNC HA. Inset image shows the pore size distribution. (Reproduced from [124] with permission.
Figure 7. (a) SEM images of CNTFs. (b) Low- and (c) high-magnification SEM images of CNTFs@H-Co3O4 NA. (d) SEM images of as-prepared CNTFs@H-Co3O4@ZIF-67 HA. Inset: enlarged SEM image. (e) SEM images of as-prepared CNTFs@H-Co3O4@CoNC HA. Inset: enlarged SEM image. (f) Powder XRD patterns of as-prepared CNTFs@H-Co3O4@CoNC HA. (g) Raman spectra of as-prepared CNTFs@H-Co3O4@CoNC HA. (h) EDS of the as-prepared samples. (i) N2 adsorption/desorption isotherms of the as-prepared CNTFs@H-Co3O4@CoNC HA. Inset image shows the pore size distribution. (Reproduced from [124] with permission.
Batteries 11 00181 g007
Figure 8. The assembled FASC device. (a) Schematic of the assembled structure of the FASC. (b) CV of the assembled device measured for different voltage windows. (c) GCD curves of the assembled flexible asymmetric supercapacitor (FASC) at different current densities. (d) Ragone plot (power density vs. energy density) of the FASC device. (e) Cyclic stability. (f) Capacitance retention after 5000 cycles of bending. (Reproduced from [124] with permission).
Figure 8. The assembled FASC device. (a) Schematic of the assembled structure of the FASC. (b) CV of the assembled device measured for different voltage windows. (c) GCD curves of the assembled flexible asymmetric supercapacitor (FASC) at different current densities. (d) Ragone plot (power density vs. energy density) of the FASC device. (e) Cyclic stability. (f) Capacitance retention after 5000 cycles of bending. (Reproduced from [124] with permission).
Batteries 11 00181 g008
Figure 9. Synthesis scheme of leaf-shaped CoPx/CoO heterostructure and structural analysis of precursor. (a) Scheme for the synthesis of CoPx/CoO heterostructure; Powder XRD of (b) ZIF-Co-L and (c) Co3O4-L precursors; (d) SEM and (e) TEM of ZIF-Co-L precursor; (f) SEM and (g) TEM images of Co3O4-L precursor. (Reproduced from [127] with permission).
Figure 9. Synthesis scheme of leaf-shaped CoPx/CoO heterostructure and structural analysis of precursor. (a) Scheme for the synthesis of CoPx/CoO heterostructure; Powder XRD of (b) ZIF-Co-L and (c) Co3O4-L precursors; (d) SEM and (e) TEM of ZIF-Co-L precursor; (f) SEM and (g) TEM images of Co3O4-L precursor. (Reproduced from [127] with permission).
Batteries 11 00181 g009
Figure 10. Schematic diagram of the synthesis of Ni(Co)3S4@MXene-fibers. (Reproduced from [153] with permission).
Figure 10. Schematic diagram of the synthesis of Ni(Co)3S4@MXene-fibers. (Reproduced from [153] with permission).
Batteries 11 00181 g010
Figure 11. (ac) SEM images of CoCuNi-OH and (d) the corresponding elemental mapping; (eg) SEM images of Cu(NiCo)2S4/Ni3S4 and (h) the corresponding elemental mapping images. (Reproduced from [164] with permission).
Figure 11. (ac) SEM images of CoCuNi-OH and (d) the corresponding elemental mapping; (eg) SEM images of Cu(NiCo)2S4/Ni3S4 and (h) the corresponding elemental mapping images. (Reproduced from [164] with permission).
Batteries 11 00181 g011
Figure 12. (a) Schematic diagram of the as-assembled ASC device, (b) CV curves of the cathode and anode within the respective potential of −0.2 to 0.7 and −1 to 0 V at a scan rate of 20 mV s1, (c) CV curve of the ASC device at different scan rates of 5–200 mV s1, (d) GCD curves of the ASC device at different current densities of 1 to 10 A g1, (e) specific capacity and coulombic efficiency at different current densities, (f) cyclic performance of the ASC device (inset figure shows the first and last five cycles), and (g) Ragone plot (inset shows the digital photograph of a yellow and blue LED bulb ignited with the help of the ASC device when connected in a series). (Reproduced from [176] with permission).
Figure 12. (a) Schematic diagram of the as-assembled ASC device, (b) CV curves of the cathode and anode within the respective potential of −0.2 to 0.7 and −1 to 0 V at a scan rate of 20 mV s1, (c) CV curve of the ASC device at different scan rates of 5–200 mV s1, (d) GCD curves of the ASC device at different current densities of 1 to 10 A g1, (e) specific capacity and coulombic efficiency at different current densities, (f) cyclic performance of the ASC device (inset figure shows the first and last five cycles), and (g) Ragone plot (inset shows the digital photograph of a yellow and blue LED bulb ignited with the help of the ASC device when connected in a series). (Reproduced from [176] with permission).
Batteries 11 00181 g012
Table 1. Electrochemical performance indicators for MOF-based supercapacitor materials and devices.
Table 1. Electrochemical performance indicators for MOF-based supercapacitor materials and devices.
Material(s)/DevicesSpecific Capacitance or Capacity *
[F/g] or [C/g] *
Energy
Density
[Wh/kg]
Power
Density
[W/kg]
Cycling
Stability
[% @ Cycles]
Ref.
Cu-MOF (sonochemically)594.234.413,76597.95 @ -[39]
Ni-MOF derived NiO/Ni/r-GO//r-GO172.2 *39.641,36080 @ 10,000[72]
NiCo-PTA@PNTs110941.237579.1 @ 10,000[182]
NiO@Ni-BTC/NF-121853 *29.1700094 @ 3000[82]
Ni/Mn-PTA2848142.880093.25 @ 5000[183]
ZIF-8/Fe2O3//AC116028.5239897 @ 1500[87]
Fe3O4@ZIF-67//AC1334.027.9548887 @ 3000[88]
ZIF-8/ZIF-67: 50/502016984072 @ 5000[184]
MnO2@NTO1054.736.211,879.985.3 @ 5000[102]
Mn-PTA/NF10.256644181.2 @ 10,000[185]
CoFe2O4@NiMn2O4/NF//AC312.890.312,90088.4 @ 10,000[120]
rGO/Ppy/Zn-MOF175.019.7179282 @ 7000[137]
NiCoP/C nanohybrids//AC775.747.6798.978.1 @ 10,000[147]
(Ni0.93Mn0.07)2P-18/CC851.1 *59.875084.87 @ 5000[151]
NiCo2(S0.78Se0.22)5/GC//Bi2O2.33/rGO476.247.2 93.7 @ 100[161]
Ni-MOF-24/Cu3 (HITP)2/CFP142457150094.3 @ 7000[186]
Ni-CTP-COOH/GO1258.779.71275110 @ 5000[187]
Cu-Co-MOF/rGO935.845.22495.5-[188]
Cu-MOF/CNT Composite348.5627.7164090.15 @ 10,000[189]
Cu-MOF/CNT166.423.6501.579.2 @ 10,000[190]
Cu-MOF/CNT//AC1875 *4792098.3 @ 15,000[191]
NiMOF@MX 2//AC (ASC)1160.548.215,00094 @ 10,000[192]
Li-Cu-MOF//AC171.136.1510082.1 @ 1000[193]
* indicates specific capacity values.
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

Argirusis, C.; Katsanou, M.-E.; Alizadeh, N.; Argirusis, N.; Sourkouni, G. Recent Advances in the Application of MOFs in Supercapacitors. Batteries 2025, 11, 181. https://doi.org/10.3390/batteries11050181

AMA Style

Argirusis C, Katsanou M-E, Alizadeh N, Argirusis N, Sourkouni G. Recent Advances in the Application of MOFs in Supercapacitors. Batteries. 2025; 11(5):181. https://doi.org/10.3390/batteries11050181

Chicago/Turabian Style

Argirusis, Christos, Maria-Eleni Katsanou, Niyaz Alizadeh, Nikolaos Argirusis, and Georgia Sourkouni. 2025. "Recent Advances in the Application of MOFs in Supercapacitors" Batteries 11, no. 5: 181. https://doi.org/10.3390/batteries11050181

APA Style

Argirusis, C., Katsanou, M.-E., Alizadeh, N., Argirusis, N., & Sourkouni, G. (2025). Recent Advances in the Application of MOFs in Supercapacitors. Batteries, 11(5), 181. https://doi.org/10.3390/batteries11050181

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