Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Metal–organic framework/MXene-based materials: preparations and applications

Shasha Zheng*, Keke Zhou, Xiaoxue Zhang and Ning Ren
College of Energy Engineering, Huanghuai University, Zhumadian, 463000, Henan, P. R. China. E-mail: zhengshasha@huanghuai.edu.cn

Received 22nd June 2025 , Accepted 11th August 2025

First published on 29th August 2025


Abstract

Metal–organic framework (MOF) materials have attracted significant attention due to their high surface area and adjustable pore structure, which enable potential applications across various fields. However, their practical application is often hindered by poor electrical conductivity and limited structural stability. Integrating MOF with two-dimensional transition metal carbides/nitrides (MXene) offers a powerful strategy to overcome these limitations, synergistically combining the porous architecture of MOF with the exceptional conductivity and mechanical robustness of MXenes. Notably, the performance of MOF/MXene-based materials is governed by the diversification of MOF components. This review provides a systematic and comprehensive analysis of MOF/MXene-based materials, emphasizing a component-driven design approach. These materials are classified into three different types: pure MOF/MXene, MOF composite/MXene, and MOF derivative/MXene, and their synthetic strategies are detailed, including in situ growth, solvothermal/hydrothermal methods, self-assembly, and co-precipitation method. Further exploration was conducted on their advanced performance in multifunctional applications such as batteries, supercapacitors, sensors, catalysts, and electromagnetic wave absorbing materials.


1. Introduction

The relentless progress in scientific innovation has driven the development of novel nanomaterials, whose outstanding physicochemical characteristics enable extensive applicability across diverse fields.1–4 Among these, metal–organic framework (MOF) nanomaterials, a class of compounds constructed through the coordination of metal ions or clusters with organic ligands to form one-, two-, and three-dimensional (1D, 2D, and 3D) structures, have garnered significant attention.5–8 The exceptionally high specific surface area and tunable porosity of MOF materials endow them with superior adsorption and separation performance.9–11 Furthermore, the structural and functional versatility of MOF materials has enabled their application in fields such as catalysis,12–14 energy storage,15–17 and so on.18–20 Despite these advantages, MOF materials also present certain limitations, including insufficient chemical stability, inadequate mechanical strength, poor electrical conductivity, and restricted adsorption capacity and selectivity.21–23 To address these challenges and fully exploit the potential of MOF materials, the design of MOF-based composites has emerged as a promising strategy.24–27 By integrating MOF materials with other functional materials, synergistic effects can be achieved, effectively combining the strengths of each component while compensating for their individual deficiencies.28–31

MXene, a novel class of 2D transition metal carbides and nitrides with the general formula Mn+1XnTx (M = early transition metals; X = carbon and/or nitrogen; Tx = surface functional groups), has garnered significant attention in advanced materials science.32–36 Its unique 2D layered structure imparts exceptional electrical conductivity, good mechanical flexibility, intrinsic hydrophilicity, and abundant surface functional groups (e.g., –OH, –F, –O).37–41 These properties render MXene a highly promising material for diverse applications, including batteries,42,43 supercapacitors (SCs),44,45 catalysis,46,47 water purification,48–50 adsorption,51,52 sensors,53–55 and various other fields.56–60 Nevertheless, practical implementation of MXene is hindered by several intrinsic limitations, such as oxidative instability, restacking of the 2D nanosheets, and a relatively limited number of accessible active sites.61–63 A particularly effective approach to overcome these challenges involves the integration of MXene with MOF porous materials. This hybridization generates significant synergistic effects: the porous structure of MOF materials can serve as a structural scaffold to prevent the restacking of MXene nanosheets, thereby preserving their high surface area and excellent electrical conductivity.64–68 Conversely, the highly conductive MXene network facilitates efficient electron transport, serving as a rapid charge transfer channel that markedly enhances the typically sluggish charge transfer kinetics of the inherently less conductive MOF component.69–71 The distinctiveness of MOF/MXene-based materials lies in the unique synergy offered by MXene, which sets them apart from other common MOF-based hybrids. Compared to MOF/carbon nanotubes (CNTs), the 2D structure of MXene provides a larger contact interface and better dispersion.63,72 Unlike MOF/graphene, MXene's rich surface functional groups enable strong chemical bonding with MOF materials, forming stable heterojunctions.73,74 In contrast to MOF/metal nanoparticles (NPs), MXene enhances both conductivity and stability cost-effectively.75,76 Finally, compared to MOF/polymer composites, MXene integration directly addresses the core issue of poor electrical conductivity.77–79 This multifaceted advantage positions MOF/MXene-based materials as a highly attractive class of materials.

Numerous comprehensive reviews have been conducted on MOF/carbon-based materials,80–84 MOF/polymers,85–90 MOF/quantum dots (QDs),91–93 MOF/metal NPs,94–96 and other MOF composites.97–103 However, in the field of MOF/MXene-based materials, despite a series of valuable reviews have been reported, they generally have limited scope. Many reviews focus on specific applications (such as energy storage,72,77,104–109 catalysis,75,110,111 biomedical,78 wastewater treatment,112,113 electromagnetic interference shielding,114 etc.) or specific types of materials (such as MOF derivatives/MXene115), resulting in a fragmented understanding of this field. Therefore, there is currently a lack of a comprehensive, systematic, and up-to-date review that covers the complete system of MOF/MXene-based materials from basic classification, synthesis strategies to diverse applications. Compared with these existing reviews, the main purpose of this review is to provide a comprehensive and detailed analysis of MOF/MXene-based materials, providing a reference for researchers in materials science, chemistry, and engineering fields to fully understand this dynamic field.

It is worth noting that the diversity of MOF components (including pure MOF, MOF composite and MOF derivative) in MOF/MXene-based materials has a decisive influence on the material properties. To our knowledge, there are currently no comprehensive reports on the component classification, synthesis strategies and multiple application potentials of MOF/MXene-based materials. Therefore, this review focuses on the classification of MOF/MXene-based materials, with emphasis on their preparation techniques and application fields. As shown in Scheme 1, MOF/MXene-based materials are systematically classified into three primary categories based on their MOF constituents: pure MOF/MXene, MOF composite/MXene, and MOF derivative/MXene. Their applications span a wide range of fields, including batteries, SCs, sensors, catalysis, and absorption, among others. Furthermore, the future development trends and challenges that need to be addressed with MOF/MXene-based materials are also discussed.


image file: d5ra04436g-s1.tif
Scheme 1 Overview of MOF/MXene-based materials and their application fields.

2. Synthesis of MOF/MXene-based materials

MXene materials, featuring abundant surface functional groups and excellent electrical conductivity, are considered promising substrates for enhancing the stability and electrical conductivity of MOF materials.116–119 The integration of MOF and MXene is driven by interfacial interactions between the two components. The surface functional groups of MXene serve as active sites for the heterogeneous nucleation of MOF. Metal ions from the MOF precursor solution are attracted to the negatively charged oxygen-containing groups on the MXene surface via electrostatic interactions. This initial adsorption facilitates the coordination of organic ligands, initiating MOF growth directly on the MXene substrate.65,120 Crucially, beyond physical interactions, strong chemical bonding often occurs at the interface, forming stable heterojunctions. Coordination bonds can form between the metal atoms (Ti) in the MXene layer and the donor atoms (O, N) from the organic ligands of MOF, such as Ti–O–C or Ti–N.113,121

This synergistic interaction has attracted significant research attention toward developing MOF/MXene-based materials through various synthetic strategies. These composites not only promote advancements in material design but also expand the application potential of MOF-based materials.73,122–125 As shown in Scheme 2, these composites are systematically classified into three main types based on their MOF constituents: pure MOF/MXene, MOF composites/MXene, and MOF derivatives/MXene. The primary synthesis methods for MOF/MXene-based materials encompass in situ synthesis, hydrothermal synthesis, electrostatic self-assembly, solvothermal method, and co-precipitation method. This section provides a comprehensive overview of the preparation of various MOF/MXene-based materials employing these different methods.


image file: d5ra04436g-s2.tif
Scheme 2 Classification and synthesis methods of MOF/MXene-based materials.

2.1. Pure MOF/MXene

Pure MOF/MXene composite materials are categorized into seven distinct categories according to the type of metal ions present in the MOF structure: (1) Fe-MOF/MXene, exemplified by Ti3C2Tx/MIL-53(Fe),126 synthesized via an in situ growth method; (2) Co-MOF/MXene, represented by Ti3C2-ZIF-67,127 fabricated through self-assembly; (3) Ni-MOF/MXene, such as Ni-MOF@Ti3C2Tx,128 constructed using a solvothermal approach; (4) Cu-MOF/MXene, including Cu-MOF/MXene/glassy carbon electrode (GCE),129 directly synthesized without intermediate steps; (5) Zn-MOF/MXene, typified by Zn-MOF-Ti3C2,130 prepared via a one-pot electrostatic self-assembly strategy in aqueous solution; (6) Zr-MOF/MXene, as demonstrated by MOF-801@MXene nanosheets,131 synthesized via an in situ strategy under ambient conditions; (7) bimetallic MOF/MXene, such as CoNi-ZIF-67@Ti3C2Tx,132 synthesized through coprecipitation process.
2.1.1. Fe-MOF/MXene. Liu et al.126 fabricated a novel Ti3C2Tx/MIL-53(Fe) composite (TiCFe) through in situ integration of MIL-53(Fe) with layered Ti3C2Tx MXene. Initially, Ti3AlC2 MAX phase was selectively etched using hydrofluoric acid (HF) to generate accordion-like Ti3C2Tx nanosheets. The negatively charged surface of the exfoliated MXene facilitated the adsorption of metal ions, promoting their coordination with MIL-53(Fe). This synergistic interaction resulted in a hierarchical TiCFe architecture, which exhibited an enhanced specific surface area and abundant accessible active sites. Furthermore, a Fe-MOF/MXene hybrid material was developed by Xiao et al.133 through a straightforward solvothermal approach, enabling the direct growth of Fe-MOF crystals on Ti3C2Tx MXene substrates. This composite retained the characteristic octahedral morphology of pristine Fe-MOF while significantly enhancing charge transfer efficiency through the incorporation of MXene. In the same year, Jia et al.134 designed a 2D hybrid electrode by utilizing Mnn+1XnTx as the electrode of the SCs, integrating electrochemically active sites from Fe-MOF (MIL-100(Fe)) to construct an interfacial columnar structure on the Ti3C2Tx MXene layer. Through chemical bond coupling, the Fe-MOF nanospheres and MXene nanosheets were transformed into a 3D multilayer porous nanostructure, which not only suppressed MXene aggregation but also preserved the laminar flow of reaction sites within the composite electrode. Similarly, Li et al.135 synthesized butylimidazolium chloride (BIC)-modified MIL-101(Fe), denoted as BIC@MIL-101(Fe), by the solvothermal synthesis method. Subsequently, by employing vacuum-assisted self-assembly technology, they successfully prepared MXene-BIC@MIL-101(Fe) composite membrane exhibiting excellent structural stability, high mechanical strength, and good separation performance. Recently, Wang et al.136 fabricated a hydroxyl-rich MXene@MOF composite by the in situ growth of Fe-based MOF on MXene, inspired by the structure of bacterial cell walls, for the efficient isolation of lysozyme. In contrast, Han et al.137 developed an innovative solar-driven evaporation system through synergistically integrating MXene with Fe-MOF@cellulose acetate/polyvinyl pyrrolidone. Their approach combined electrospinning and vacuum-assisted filtration to effectively couple the photothermal conversion properties of MXene with the enhanced hydraulic conductivity of electrospun cellulose acetate substrates. This design optimized both solar energy utilization and fluid transport, leading to significantly improved evaporation efficiency under solar irradiation. Meanwhile, Liu et al.138 synthesized self-reporting molecularly imprinted polymers as internal reference probes and in situ grew Fe-MOF on MXene as signal nanoprobes, which were applied in a ratiometric electrochemical biosensor for rapid and effective detection of Staphylococcus aureus.
2.1.2. Co-MOF/MXene. In 2018, Ramachandran and colleagues fabricated a Co-MOF/Ti3C2Tx composite directly on Ni foam (NF) through in situ hydrothermal deposition. Fig. 1a schematically illustrates the fabrication procedure for the Co-MOF/Ti3C2Tx@NF composite. The porous structure of Co-MOF matrices, coupled with the high electrical conductivity of 2D Ti3C2Tx nanosheets, synergistically enhances the availability of catalytic sites and facilitates efficient ionic diffusion. As illustrated in Fig. 1c and d, scanning electron microscopy (SEM) analysis confirms the uniform coating and strong interfacial adhesion between the 2D Ti3C2Tx nansheets and Co-MOF on the NF substrate.139 In 2021, Liang et al. utilized in situ growth to prepare ZIF-67@Ti3C2.140 Transmission electron microscopy (TEM) images reveal the hybrid structure, with ZIF-67 particles, averaging approximately 800 nm in diameter, uniformly anchored on the surface of Ti3C2. Subsequently, Wan et al. developed an innovative ZIF-67 functionalized MXene composite using sequential precipitation synthesis. This hybrid material was subsequently integrated into thermoplastic polyurethane (TPU) matrices through melt compounding technology. The synthesis process involved dissolving 2.9 g (0.01 mol) of cobalt nitrate hexahydrate in 250 mL of methanol under continuous magnetic stirring for 10 min, followed by a 24 h aging period. Post-synthesis purification included sequential washing, vacuum filtration, and thermal drying at 80 °C for 24 h to obtain the final MXene/Co-ZIF composite. As illustrated in Fig. 1e and f, the resulting hybrid structure demonstrates unique interfacial features. MXene nanosheets with reactive surface groups (–O, –OH, and/or –F) facilitate effective chemical grafting, while ZIF-67 NPs provide abundant unsaturated cobalt coordination sites, promoting strong heterogeneous interfacial bonding. This mutual interaction promotes uniform dispersion of ZIF-67 crystallites across MXene lamellae, simultaneously achieving particle size refinement through surface confinement effects. The engineered TPU nanocomposites exhibited remarkable thermal performance enhancements, particularly demonstrating significant reduction in heat release rate alongside improved thermostability.141
image file: d5ra04436g-f1.tif
Fig. 1 (a) The synthesis process of Co-MOF/Ti3C2Tx@NF composite. (b) Synthesis mechanism of the MXene-ZIF67. SEM images of Co-MOF/Ti3C2Tx@NF (c and d), ZIF-67@MXene hybrid (e and f), and MXene-ZIF67 (g). (a, c and d) Reproduced with permission.139 Copyright 2018, Elsevier. (e and f) Reproduced with permission.141 Copyright 2022, Multidisciplinary Digital Publishing Institute. (b and g) Reproduced with permission.127 Copyright 2022, Elsevier.

Concurrently, Yang et al. developed a “sandwich-like” Ti3C2-ZIF-67 composite through cubic-structured intercalation within MXene layer. Fig. 1b illustrates the fabrication process, in which MXene nanosheets were derived through selective HF etching of Al atoms from Ti3AlC2 precursors. The surface functional groups, which acquired negative charges during the etching process, facilitated the immobilization of ZIF-67 on the MXene substrate. Structural characterization, as evidenced by SEM image (Fig. 1g), confirmed the formation of ZIF-67 crystals, exhibiting a distinct crystalline morphology. The ZIF-67 NPs were uniformly modified on both sides of the MXene nanosheets. This composite material harnessed the combined advantages of its constituent components, namely, the hydrophilicity and conductivity of MXene and the biocompatibility of MOF. The composite exhibited exceptional colonization potential for microorganisms and maintained remarkable species diversity, while the synergistic interactions within the microbial consortium significantly improved bioelectricity production efficiency.127 In addition to the above preparation strategies, there are some synthesis strategies about Ti3C2Tx/ZIF-67 have been reported simultaneously.142,143

In addition, Roy et al.144 constructed a hierarchical Co-MOF-MXene@carbon black electrode architecture for electrochemical sensor electrodes and point-of-care testing chips by synthesizing MXene-supported Co-MOF-based nanocomposites and integrating them with carbon black ink through 3D printing technology. Ding et al.145 adopted an interdiffusion reaction strategy to construct a 0D/2D heterostructure via self-assembling Co-MOF onto highly conductive MXene nanosheets, enabling the design of highly reliable room-temperature gas sensing devices. In this architecture, the MXene nanosheets served as highly conductive core substrates, promoting efficient electron and mass transport while suppressing the agglomeration of MOF NPs during solution self-assembly. Concurrently, the Co-MOF functional sensing layer effectively separated the MXene nanosheets, mitigating their restacking and enhancing the overall thermodynamic stability of the composite.

2.1.3. Ni-MOF/MXene. Qin et al.146 utilized a straightforward self-assembly strategy to construct MXene QDs integrated with 2D Ni-MOF architectures in tunable ratios, thereby forming type II heterojunction structures. The composites demonstrated enhanced absorption and superior interfacial charge transfer ability, as evidenced by their optical and optoelectronic properties. The following year, Cheng et al. used layered Ti3C2 MXene as a catalyst and employed a facile electrostatic self-assembly process to synthesize 2D composites incorporating ultrathin Ni-MOF nanosheets. The similar layered structure of Ti3C2 MXene and Ni-MOF facilitated the formation of a conductive intimate interface, thereby promoting charge transfer and accelerating the separation of photogenerated charge carriers (electrons and holes).147 Concurrently, Zhang and co-workers reported an innovative MXene-directed synthesis strategy for high-performance SCs using Ti3C2Tx MXene/Ni-based MOF composite.148 Simultaneously, Zhang et al. developed an advanced battery-type nickel phosphates-MXene electrode material that is highly suitable for SCs. This material effectively inhibited the aggregation of MXene nanosheets while maintaining a laminar flow structure within the electrode, thereby facilitating rapid ion migration.149 In 2022, Ganiyat Olatoye et al. resolved the challenge of low specific capacitance by depositing Ni-ZIF-67 onto MXene substrates. The researchers first synthesized Ni-ZIF-67 hybrid materials under a controlled temperature gradient (150–450 °C), followed by the fabrication of Ni-ZIF-67-200/K-Ar-MXene composites through modification of intercalated 2D MXene layers (K-Ar-MXene) serving as a structural scaffold.150

In addition, to address the inherent limitations of MOF, such as poor conductivity, unsatisfactory stability, and easy aggregation, Zheng et al. successfully developed Ti3C2 MXene-supported pillared-layer MOF with enhanced cyclability (MXene@Ni-MOF). The hybrid material features a 3D architecture stabilized by Ni–N coordination networks, which confers exceptional structural integrity, as evidenced by sustained performance during cyclic testing. Furthermore, the immobilization of Ni-MOF nanosheets on the MXene substrate facilitates rapid charge transfer between the two components, effectively mitigating the intrinsic low conductivity of Ni-MOF and preventing nanosheet agglomeration (Fig. 2a and b).70 In Fig. 2c, Wan et al. successfully fabricated 3D flower-like Ni-MOFs on multiterminal lamellar MXene using a solvothermal approach and investigated its flame retardant properties in TPU.128 Concurrently, Yang and colleagues prepared Ni-MOF/V2CTx-MXene composites through precisely controlled thermal annealing, resulting in heterostructured materials with hierarchically porous nanorod architectures. This engineered composite demonstrated significant improvements in both charge transport efficiency and electrochemically active surface characteristics.151 In 2023, Li et al. leveraged the conductive properties of Ni-MOF and the high strength of Ti3C2Tx to enhance the stability of Ni-MOF. The prepared an innovative hybrid material featuring vertically aligned Ni-MOF nanorods on MXene substrates via controlled hydrothermal synthesis (Fig. 2d).152 Additionally, Nanda et al. introduced a nickel-containing MXene composite with a nanosheet morphology, synthesized via solvothermal treatment in an autoclave. The resulting material demonstrated a high specific energy capacity and improved power delivery characteristics.153


image file: d5ra04436g-f2.tif
Fig. 2 (a) Schematic illustration of the synthetic process for MXene@Ni-MOF. (b) SEM and optical images of MXene@Ni-MOF. (c) SEM image for Ni-MOF@MXene hybrids. (d) Schematic diagram for the preparation of Ni-MOF/MXene. (a and b) Reproduced with permission.70 Copyright 2022, Elsevier. (c) Reproduced with permission.128 Copyright 2022, Elsevier. (d) Reproduced with permission.152 Copyright 2023, Multidisciplinary Digital Publishing Institute.

The composite of highly conductive MXene and porous conductive MOF (Ni3(HITP)2, HITP = 2,3,6,7,10,11-hexaaminotriphenylene) has also been widely reported, Lin et al.154 fabricated electrophysiological Ni3(HITP)2 MOF/MXene electrodes. Xu et al.155 fabricated a temperature sensor with high thermal sensitivity and excellent flexibility by loading conductive Ni3(HHTP)2 MOF (HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene) onto MXene via in situ polymerization. Moreover, the electrochemical performance of MXene was significantly improved by compounding it with Ni-HHTP156 and (Fe-, Cu-, Co-, Ni)-1,3,5-benzene tricarboxylate (BTC) MOFs,157 which showed excellent performance for SCs. Shivade et al. used an innovative strategy to synthesize Ni-MOF/MXene composites assisted by biomolecule folic acid, which effectively prevented the oxidation and restacking of MXene.158

2.1.4. Cu-MOF/MXene. In 2021, a novel 3D flower-like Cu-MOF combined with ultrathin MXene nanosheets was synthesized and studied by Cheng et al.129 The Cu-MOF was synthesized via a straightforward, one-step process utilizing copper nitrate and 2-aminoterephthalic acid as the starting materials (Fig. 3a–c). Subsequent progress in MOF engineering was demonstrated by Gu et al. through strategic modification of the conventional HKUST-1 framework.159 Their methodology involved the incorporation of 2D Ti3C2Tx MXene components, resulting in a novel composite material exhibiting enhanced structural integrity under hydrothermal conditions (Fig. 3d). Hassan et al.160 prevented the restacking of MXene sheets by integrating them with Cu-BTC MOF, thereby improving the electrochemical performance. However, Liu et al.161 adopted an innovative approach combining thermal induction with solvent-assisted oxygen anion etching, enabling precise modulation of pores in rigid microporous Cu-based MOFs. These engineered MOFs were subsequently assembled with MXene-derived aerogels through self-assembly to construct flexible NO2 sensing platforms. Sharma et al.162 designed MXene@Cu-TCPP (TCPP = tetrakis (4-carboxyphenyl) porphyrin) hybrid architectures that formed 2D heterogeneous nanostructures by van der Waals forces, exhibiting excellent stability and conductivity. Zhu et al.163 constructed a dimensional-matched organic–inorganic nanohybrid with high light absorption capacity and strong interfacial attraction by combining MXene/TiO2 with a small amount of Cu-TCPP through a one-step solvothermal method. In addition, Lin et al.164 reported a novel MXene/Cu-TCPP lamellar membrane that achieved efficient pervaporation energy generation by synergistically combining 2D MXene and Cu-TCPP MOF. Wang et al.165 constructed MXene@MOF-based electrochemical biosensors via heterogeneous interface design: the TCPP organic ligand was first employed to functionalize MXene nanosheets, followed by surfactant-assisted solvothermal synthesis of MOF, yielding a MXene@MOF architecture with superior stability, enhanced conductivity, and a multilayered folded morphology. This architecture significantly increased electrochemical active sites, enabling label-free, ultrasensitive detection of glycoprotein nonmetastatic melanoma protein B in serum samples from patients with Parkinson's disease.
image file: d5ra04436g-f3.tif
Fig. 3 (a) Preparation of Cu-MOF and MXene nanosheets. TEM images of Cu-MOF (b) and MXene (c). (d) Schematic diagram for the synthesis of HKUST-1/Ti3C2Tx MXene hybrid materials. (a–c) Reproduced with permission.129 Copyright 2021, Wiley. (d) Reproduced with permission.159 Copyright 2021, American Chemical Society.

Recently, Zhou et al.166 drew inspiration from the 3D interlocking hierarchical structure of the skin to construct a multifunctional artificial epidermal device through in situ growth of Cu3(HHTP)2 NPs on the surface of hollow spherical MXene, resulting in a biomimetic MXene@Cu3(HHTP)2 composite sensor. In this design, the gas-sensitive Cu3(HHTP)2 is heterogeneously assembled in situ at the MXene interface, mimicking the hollow spherical morphology found in skin, thereby structurally emulating the stratum spinosum and stratum granulosum of the epidermis. However, Ge et al.167 proposed an all-solution process method to fabricate a photodetector by sequentially spraying conductive Cu3(HHTP)2 and MXene onto a Si substrate, forming a van der Waals double heterojunction.127 The heterojunction structure promotes unidirectional electron hole separation with type I band alignment within the Cu3(HHTP)2/Si interface, while it also takes advantage of the potential barrier difference between the Cu3(HHTP)2/Si and MXene/Cu3(HHTP)2 Schottky junctions. The MXene/Cu3(HHTP)2/Si photodetector exhibits excellent photoelectric properties.

2.1.5. Zn-MOF/MXene. In 2020, Li et al. employed a gentle water freezing-and-thawing approach to exfoliate multilayer MXene, followed by the integration of ZIF-8 NPs into the MXene layers through a straightforward layer-by-layer assembly strategy. The MXene/ZIF-8 composite exhibits unique architectural characteristics, featuring flexible MXene nanosheets with surface corrugations that serve as effective nucleation sites for the controlled growth of ZIF-8. This methodology provides an efficient solution to the persistent challenge of MXene layer restacking commonly encountered in conventional synthesis methods.168 Concurrently, Yao and co-workers utilized the MXene/ZIF-8 nanocomposite as an electrode material through an in situ growth method.169 At almost the same time, Wang and co-workers prepared MXene/ZIF-8 nanocomposites by mixing the Zn(NO3)2·6H2O solution with the 2-methylimidazole solution, followed by stirring for approximately 24 h, and subsequent centrifugation, washing, and drying steps.170 Rabiee et al. constructed an innovative inorganic MXene/MOF-5 nanostructure using a vacuum-assisted filtration process.171 In 2022, Guo et al. successfully synthesized a bioactive Zn-MOF over the Ti3C2Tx nanosheets through in situ growth.172 The fabrication pathway of the Zn-MOF@Ti3C2Tx composite is schematically depicted in Fig. 4a. Expanding on earlier work, Yao et al. systematically investigated ZIF-8@Ti3C2Tx hybrid materials through an optimized single-step microemulsion synthesis.173 The experimental scheme involved the gradual incorporation of microemulsion phases A and B into Ti3C2Tx suspensions, followed by vigorous stirring under ambient conditions. The product was then isolated via centrifugation, purified through repeated methanol washes, and finally dehydrated under vacuum at 60 °C. More recently, Jose Paul et al. synthesized [Zn4(1,2,4,5-benzene-tetra carboxylate)2(H2O)6]n·3nH2O MOF using pyromellitic acid-based linkers and Zn(NO3)2·6H2O as precursors via a reticular synthesis approach. Multilayer MXene was prepared through conventional HF etching, and the MOF-MXene hybrid composite was subsequently constructed through the one-step electrostatic self-assembly process in aqueous medium. As demonstrated in Fig. 4b, this integration process successfully occurred under ambient conditions within distilled water through reticular synthesis principles. SEM and TEM images of the MOF-MXene microstructure (Fig. 4c–e) indicate that the MOF envelops MXene with a casually oriented, block-structured configuration.130 Deng et al. developed polyurethane (PU)-MXene-MOF sponge using an in situ growth technique.174 As shown in Fig. 4f and g, the ZIF-8 NPs were uniformly dispersed on the surface of the PU-MXene sponge, resulting in a significantly roughened surface morphology of the PU-MXene-MOF sponge skeleton. Building upon previous research, Hao et al. successfully synthesized MXene@ZIF-8 to a mitigate the inherently low electrical conductivity of MOF.175 In addition, Shahriyari Far and co-workers fabricated MXOF composites via solvothermal growth, demonstrating the synergistic effect between Zn-MOF and MXene nanosheets.176
image file: d5ra04436g-f4.tif
Fig. 4 (a) Schematic diagram for the preparation of Zn-MOF@MXene hybrid. (b) Schematic illustration of the synthetic process for MOF-MXene composite. (c) SEM image of MOF-MXene. (d and e) TEM images of MOF-MXene. (f and g) SEM images of PU-MXene-MOF sponge. (a) Reproduced with permission.172 Copyright 2022, Wiley. (c–e) Reproduced with permission.130 Copyright 2023, Elsevier. (f and g) Reproduced with permission.174 Copyright 2023, Elsevier.
2.1.6. Zr-MOF/MXene. Li et al.131 constructed a significantly improved MOF@MXene membrane for gas separation by a room temperature in situ synthesis strategy. Initially, the negatively charged MXene nanosheets immobilized adjacent Zr metal ions through electrostatic interactions. Subsequently, under ambient conditions, these metal ions coordinated with organic ligands, promoting the homogeneous nucleation and growth of MOF-801 crystalline structures on the MXene surfaces, thereby yielding MOF-801@MXene nanosheets. The as-prepared nanocomposites were then uniformly deposited onto a porous organic support via vacuum-assisted filtration, ultimately forming the MOF@MXene composite membrane. Wen et al.177 synthesized the 3D hierarchical Zr-MOF/Ti3C2Tx nanocomposite through a facile method that involved electrostatic interaction between the Ti3C2Tx MXene nanosheet and Zr4+, followed by the solvothermal method for the in situ synthesis of Zr-MOF on the Ti3C2Tx nanosheets. In addition, Kashif et al.178 fabricated UiO-66/MXene hybrids via solvothermal method. UiO-66 and MXene were heat treated in solvent to make them closely combined to form composite materials. Xiang et al.179 prepared a polyethylene glycol/MXene@MOF membrane with stable interlayer spacing and excellent antifouling performance by a simple interface self-assembly strategy. In this approach, UiO-66-NH2 MOF NPs were used as structural supports to induce intercalation within layered MXene nanosheets, effectively tuning the interlayer spacing and enhancing the permeability. Subsequently, hydrophilic polyethylene glycol was introduced through hydrogen bonding, further stabilizing the intercalated architecture. However, Maleki et al.180 focused on the preparation of a novel MXene modified polyether sulfone membrane. By compounding UiO-66-NH2 NPs with MXene, UiO-66-NH2@MXene 2D nanocomposites were created to modify polyethersulfone membrane. During the membrane modification process, 0.25 wt% UiO-66-NH2@MXene nanocomposite was used as modifier to successfully synthesize the modified MXene based polyethersulfone membrane. Recently, Meng et al.181 employed electrostatic self-assembly to rationally integrate NH2-UiO-66 with MXene nanosheets, resulting in an innovative hybrid material. This process involved the solution-phase mixing of oppositely charged Zr-MOF and MXene components, which spontaneously assembled into an ordered multilayer structure driven by electrostatic interactions.
2.1.7. Bimetallic MOF/MXene. In 2019, Wen et al.132 synthesized CoNi-ZIF-67@Ti3C2Tx by in situ coprecipitation growth of bimetallic CoNi-ZIF-67 rhombic dodecahedrons on Ti3C2Tx. The resulting hybrid structure is clearly depicted in the SEM image (Fig. 5a), where the layered structure of Ti3C2Tx is well preserved, with uniformly dispersed CoNi-ZIF-67 particles embedded within the interlayer spaces. For comparison, SEM analysis (Fig. 5b) of pristine CoNi-ZIF-67 highlights its distinct morphological features in the absence of Ti3C2Tx support. Liu et al.182 engineered 3D MXene/NiCo-MOF composite films by integrating 2D nanosheets with NiCo-MOF nanoflakes through hydrogen bonds, employing vacuum-assisted filtration. This approach initiated with the synthesis of 2D NiCo-MOF nanosheets under ambient conditions, followed by multilayered MXene production through HF etching of Ti3AlC2 and subsequent alkalization (Fig. 5f). In 2021, Du et al.183 obtained Ti3C2Tx/CoNi-MOF nanosheets by in situ growth process. Additionally, Wu and co-workers developed the 1D heterostructure via integrating MXene fibers with CoNi-MOF.184 In recent years, Xiao et al. demonstrated a novel 2D heterostructure synthesized by vertically growing NiCo-MOF ultrathin nanosheets on exfoliated Ti3C2Tx MXene substrates. This architecture effectively integrates the synergistic merits of conductive MXene with the enhanced electrochemical properties inherent to bimetallic MOF nanomaterials.185 The superior electrical conductivity and effective interfacial compatibility of aminated MXene with electrolytes enables homogeneous in situ anchoring of bimetallic NiCo-MOF on its surface. Yue et al. successfully prepared highly redox-active NiCo-MOF that were in situ stabilized using 2D aminated Ti3C2Tx.186 Recently, Tan et al.187 fabricated multiple 2D MOF/Ti3C2Tx hybrid materials through electrostatically directed assembly (Fig. 5g). As illustrated in Fig. 5c–e, the atomic force microscope (AFM) characterization of Co2Ni-MOF@Ti3C2Tx reveals a uniform thickness distribution, in contrast to its individual components (Co2Ni-MOF and Ti3C2Tx). Beyond these strategies, recent studies188–191 have reported analogous in situ growth methods for fabricating MXene-supported bimetallic NiCo-MOF composites.
image file: d5ra04436g-f5.tif
Fig. 5 SEM images of CoNi-ZIF-67@Ti3C2Tx (a) and pure CoNi-ZIF-67 (b). AFM images of Co2Ni-MOF (c), Ti3C2Tx (d), and Co2Ni-MOF@Ti3C2Tx (e). (f) Fabrication process of MXene/NiCo-MOF. (g) A scheme for the preparation of MXene/NiCo-MOF. (a and b) Reproduced with permission.132 Copyright 2019, Molecular Diversity Preservation International. (c–e and g) Reproduced with permission.187 Copyright 2022, Elsevier. (f) Reproduced with permission.182 Copyright 2020, Molecular Diversity Preservation International.

In addition to the above-mentioned NiCo-MOF/MXene, other bimetallic MOF/MXene composite materials such as VCo-MOF/MXene, FeCu-MOF/MXene, FeNi-MOF/MXene, FeCo-MOF/MXene, CuZn-MOF/MXene and NiMn-MOF/MXene have also been reported. Li and co-workers synthesized a composite catalyst by in situ growth of VCo-MOF on the surface of MXene using a solvothermal method.192 Furthermore, M. Adil and co-workers synthesized the FeCu MOF/MXene composite via in situ hydrothermal growth on NF to form a binder-free electrode.193 A hybrid electrocatalyst, FeNi-MOF@MXene, composed of Fe, Ni-based MOF (FeNi-BTC MOF) and MXene, was prepared via a solvothermal reaction.194 Meanwhile, Thai et al.195 developed a novel bimetallic FeCo-MOF composite (MIL-100@ZIF-67) supported on MXene nanosheets, resulting in the hybrid material MIL-100@ZIF-67@MXene. To optimize the ion transport pathway and improve the ion permeability, Li et al.196 used block copolymers to generate mesopores within 2D MOF nanosheets, thereby constructing hierarchically porous Cu–Zn-MOF@MXene hybrids. In addition, both Ali et al.197 and Siva Shalini et al.198 synthesized NiMn-MOF nanocomposites through a hydrothermal method, and similarly achieved MXene incorporation into NiMn-MOF to improve electrochemical performance through the same synthetic route.

2.1.8. Others. As mentioned above, a variety of single-metal (Fe, Co, Ni, Cu, Zn, and Zr) and bimetallic (NiCo, VCo, FeCo, FeNi, FeCu, ZnCu, and NiMn) MOF-based materials have been reported for the construction of composites with MXene. In addition to these, several other representative MOF/MXene-based materials incorporating Bi-, Ce-, Al-, Ti-, La-, and Mo-based MOF have also been recently developed. For example, Gong et al. fabricated leaf-like MXene-based flame retardants (MXene@Bi-MOF) through the in situ growth of shuttle-shaped Bi-MOF on the surface of MXene.199 Chen et al. successfully prepared a novel Ce-MOF/MXene composite via employing the 3D straw-sheaf-like Ce-MOF and MXene nanosheets.200 Meanwhile, Li et al.201 prepared MXene@Ce-MOF composites featuring abundant oxygen vacancies. Zhang et al.202 synthesized MXene@MOF-303(Al) composites through in situ growth of MOF-303(Al) on MXene, effectively combining the advantageous properties of both components. Subsequently, a composite membrane composed of MXene@MOF-303(Al), polyvinyl alcohol, and polyacrylonitrile was prepared using the composite as a filler and a polyacrylonitrile ultrafiltration membrane as substrate. In addition, Li et al.203 utilized MXene as a Ti source to synthesize layered Ti-MOF nanosheets via organic ligand coordination, followed by strategic integration of mesopores to create multiscale porous MXene-MOF composites with superior charge storage characteristics. Zhao et al.204 achieved in situ synthesis of a Ti-based MOF (NH2-MIL-125) using 2-aminoterephthalic acid and tetrabutyl titanate, concurrently hybridizing it with MXene to form a Ti-MOF@MXene composite material, which was further functionalized with polyethyleneimine and covalently connected immobilized onto a gold electrode to construct a sensing platform. Yu et al.205 developed lanthanide-based La-MOF on MXene nanosheets via an in situ growth strategy, producing La-MOF@MXene hybrid flame retardants. In addition, Lin et al.206 proposed an interface engineering strategy to construct defect-engineered La-MOF@MXene nanohybrids with oxygen vacancies by synergistic defect and in situ growth. Kaur et al.207 fabricated a composite by combining Mo-MOF with MXene (NH2–Mo-MOF/MXene), which was subsequently deposited onto screen-printed electrodes.

To address structural limitations inherent in conventional MOF synthesis, Wang et al.208 proposed an innovative metal-ion-assisted conversion methodology employing 2D MXene as a precursors. This method enabled the successful fabrication of 3D catechol-functionalized TiCu-HHTP MOFs with a non-interpenetrated SrSi2 architecture. The transformation mechanism involves three synergistic processes: (1) electron transfer from MXene substrates to adsorbed Cu2+, initiating oxidation–reduction reactions, (2) cleavage of Ti–C bonds releasing Ti4+ ions, and (3) subsequent coordination of Ti4+/Cu2+ with the HHTP ligand to form a stable MOF structure. Notably, this approach exhibits remarkable versatility, V2CTx can substitute conventional MXene precursors, and various Mn+ (Ni2+, Co2+, Mn2+, and Zn2+) can effectively replace the original metal ions, highlighting the broad adaptability of the methodology.

2.1.9 Summary. The synthesis of pure MOF/MXene composites, as detailed in the previous sections, relies on a variety of strategies, among which the selection of metal ions plays a crucial role in determining the best approach and the properties of the composites. A critical comparison reveals the different patterns in the synthesis approaches employed for different metal ions.

(1) In situ growth is the most versatile method, successfully applied for Fe-MOF/MXene (e.g., Ti3C2Tx/MIL-53(Fe)126), Zr-MOF/MXene (e.g., MOF-801@MXene131), Zn-MOF/MXene (e.g., ZIF-8@Ti3C2Tx173), and bimetallic MOF/MXene (e.g., CoNi-ZIF-67@Ti3C2Tx132). This method generally promotes strong interfacial interaction and good dispersion of MOF on the MXene surface, making it a robust choice for achieving intimate contact.

(2) Solvothermal and hydrothermal methods are frequently used for Co-MOF/MXene (e.g., Co-MOF/Ti3C2Tx@NF139) and Ni-MOF/MXene (e.g., Ni-ZIF-67/MXene150). These techniques offer good control over crystal morphology but require high temperatures, which can be detrimental to the stability of MXene if the temperature cannot be precisely controlled.

(3) Self-assembly strategies, including electrostatic self-assembly and vacuum-assisted filtration, are effective for Zn-MOF/MXene (e.g., Zn-MOF-Ti3C2Tx via one-pot strategy130), Ni-MOF/MXene (e.g., MXene QDs coupled with 2D Ni-MOF146), and Cu-MOF/MXene (e.g., Cu-BTC MOF/MXene160). These methods usually carried out under mild conditions, which is beneficial for preserving the integrity of MXene and is particularly suitable for creating layered or membranous structures.

(4) Coprecipitation is a straightforward method used for the synthesis of bimetallic MOF/MXene composites, such as CoNi-ZIF-67@Ti3C2Tx,132 offering a rapid route to bimetallic integration.

In conclusion, while in situ growth emerges as the most widely adopted and effective strategy for achieving strong interfacial bonding across various metal ions, the selection of a synthesis method must be carefully tailored to the specific MOF and the desired composite morphology. The diversity of approaches highlighted in this section underscores the flexibility in designing pure MOF/MXene materials, with each method offering a unique balance of interfacial strength, morphological control, and compatibility with the MXene substrate.

2.2. MOF composite/MXene

In recent years, significant research efforts have been devoted to the fabrication of MOF composite/MXene-based materials. These MOF composites typically involve the integration of MOF materials with various functional components, including metal NPs, metal oxides, carbon-based materials, organic compounds, metal hydroxides, and other materials.

With respect to MOF@metal NPs/MXene composites, Dai and co-workers209 successfully synthesized Zr-Fc MOF/AuNPs/4-mercaptophenylboronic acid (4-MPBA) nanocomposites, which were subsequently integrated with 2D Ti3C2Tx MXene. Huang et al.210 introduced a bottom-up synthesis of porous heterostructure substrates by stepwise solvothermal reactions with Zn-based MOFs and MXene nanosheets, and spatially constructed stable ultrafine Pt nanocrystals on them (Pt/ZIF-Ti3C2Tx). In addition, Wen et al.211 developed a hierarchical assembly approach involving the direct in situ growth of porous ZIF-8 nanocrystals on multilayer MXene substrates, followed by the precise immobilization of abundant Pt NPs on the MXene@ZIF-8 hybrid substrate. The resulting MXene@ZIF-8@Pt NPs composite material demonstrates enhanced electrochemical durability and catalytic efficiency, attributed to the complementary characteristics of its constituent materials: the conductive MXene framework, surface-enhanced ZIF-8 matrix, and catalytically active Pt nanostructures. Recently, Zhang et al.212 successfully prepared a Ru-doped Co-MOF@MXene composite catalyst with high conductivity and high catalytic activity, achieved by modulating the electronic structure of Co-pyridinedicarboxylic MOF through Ru doping and room temperature hybridization with MXene nanosheets.

The development of MOF@metal oxide/MXene composites has also advanced significantly. Lu et al.213 successfully fabricated an innovative Fe3O4@MXene@MOF-74 composite through a straightforward synthesis strategy. Structural characterization revealed the formation of well-defined crystalline NPs after Fe3O4 modification, with Fig. 6a illustrating the homogeneous magnetic integration across the MXene substrate. Elemental mapping analysis quantitatively verified the coexistence of Zn, Fe, C, Ti, and O, thereby verifying the compositional homogeneity and structural integrity of the composite. Liu et al.120 reported a universal strategy for constructing 3D MXene/MOF composites. Through in situ growth via a coprecipitation reaction, a series of hollow 3D MXene/MOF composites (Ti3C2Tx/Cu-BTC, Ti3C2Tx/ZIF-8, Ti3C2Tx/ZIF-67, Ti3C2Tx/FeCo-PBA, and Ti3C2Tx/ZIF-67/CoV2O) were successfully prepared (Fig. 6e). Furthermore, Chi et al.214 prepared a MXene@SnO2–Ce-MOF Z-scheme heterojunction with abundant oxygen vacancies as a photoactive material by hydrothermally combining the etched MXene with SnO2, and then combining it with Ce-MOF prepared by co-precipitation method.


image file: d5ra04436g-f6.tif
Fig. 6 SEM images of Fe3O4@MXene@MOF-74 (a), MXene/CNHs/β-CD-MOFs (b), MOF-525 (c), and MXene@Co-NPC (d). Synthesis procedure of 3D MXene/MOF (e), ZIF-8@MXene (f) and PE-ZIF-8@MXene (g). (h) Schematic illustration of the intermolecular interaction between ZIF-8@MXene, polymer molecules and lithium salt. (a) Reproduced with permission.213 Copyright 2022, Elsevier. (b) Reproduced with permission.215 Copyright 2020, Elsevier. (c and d) Reproduced with permission.216 Copyright 2022, Elsevier. (e) Reproduced with permission.120 Copyright 2022, Wiley. (f–h) Reproduced with permission.221 Copyright 2022, Elsevier.

Recently, MOF@carbon-based material/MXene composites have attracted considerable attention. Tu et al.215 synthesized a novel composite material through the integration of MXene with carbon nanohorns and β-cyclodextrin MOFs (MXene/CNHs/β-CD-MOFs). The layered MXene component maintained its characteristic accordion-like configuration, while the CNHs preserved their quasi-spherical nanostructure. Fig. 6b demonstrates the successful encapsulation of MXene/CNHs assemblies within elongated β-CD-MOF crystalline matrices. Rana et al.216 employed a MOF-525@Ecoffex nanocomposite featuring a layer of Ecoffex@MXene@Co-nanoporous carbon (NPC) to construct stretchable and humidity-resistant multifunctional triboelectric nanogenerator (TENG). In this configuration, MOF-525 (Fig. 6c) served as the charge generation layer, while the Co-NPC@MXene nanocomposite (Fig. 6d) functioned as the conductive intermediate layer. However, Rajan et al.217 constructed a ternary heterojunction photocatalyst (Ni-MOF/graphitized carbon nitride (g-C3N4)/MXene) through integrated solvothermal synthesis and wet impregnation techniques. This innovative architecture combined Ni-MOF, g-C3N4, and MXene, establishing dual charge-transfer mechanisms through an engineered Type II heterojunction configuration coupled with a noble-metal-free Schottky interface. Subsequently, Qi et al.218 engineered a Fe-MOF-NH2/CNT-NH2/MXene composite through electrostatic self-assembly, strategically combining positively charged amino-functionalized components (Fe-MOF-NH2 and CNT-NH2) with negatively charged MXene, and further prepared molecularly imprinted polymer modified resistive type electrochemical sensor using electropolymerization technology. In addition, Hassan et al.219 synthesized carbon aerogel induced chromium MOF (CA@MIL-101) with Ti3C2Tx MXene nanocomposite material (CA@MIL-101-(Cr)/Ti3C2Tx) by hydrothermal method, which had interconnected porous structure that provided abundant divalent ionic active sites and significantly enhanced charge transfer kinetics. Recently, Lin et al.220 prepared g-C3N4 nanosheets using thermal polymerization and peeling method under air conditions. Subsequently, MIL-101(Fe)/g-C3N4 composites were constructed using hydrothermal method at 120 °C. Subsequently, MXene nanosheets were added, and MXene@MIL-101(Fe)/g-C3N4 composite materials were prepared by solvothermal method. The results showed that MIL-101(Fe)/g-C3N4 successfully deposited on the MXene nanosheets.

Building upon advancements in MOF@carbon-based materials/MXene composites, investigators have extended their efforts to engineer MOF@organic compound/MXene composites, aiming to achieve richer functions and performance. Zhao and colleagues221 engineered a polymer composite electrolyte reinforced with ZIF-8@MXene nanosheet (PE-ZIF-8@MXene) via an in situ growth process. Fig. 6f–h demonstrate the fabrication process for both ZIF-8@MXene and PE-ZIF-8@MXene, accompanied by molecular-level interaction schematics among the hybrid nanofillers, polymer matrix, and lithium salt components. Guo et al.222 successfully synthesized ZIF-67@MXene/polyimide (PI) hybrid materials with ordered lamellar structures through thermal compression of 3D networked aerogels, where PI was synergistically integrated with MXene-supported ZIF-67 crystalline frameworks. This synthesis method took advantage of the compressibility of the aerogel, which kept the aerogel walls always parallel to the sliding direction and thus provided significant antifriction effect. In addition, Liang et al.223 employed a ternary mixed ligand system comprising Eu3+, 1,3,5-benzenetricarboxylic acid, and terephthalic acid with polydopamine-functionalized MXene (PDA@MXene) to fabricate multilayered sandwich-structured Eu-MOF/PDA@MXene composites through an in situ assembly strategy. This synthesis method not only endowed the material with a multilayer porous structure but also realized the magnetic functionalization of the material by uniformly immobilizing Fe3O4 NPs in Eu-MOF/PDA@MXene through self-assembly. However, Hou et al.224 reported the successful synthesis of MXene@PDA/MOF composites with a unique honeycomb-like morphology, achieved by in situ coordination of 2-methylimidazole and Zn2+ on the surface of PDA-modified MXene. This synthesis method endowed the material with extensive surface area and superior electrical conductivity, establishing fundamental advantages for electrochemical sensing applications. Recently, Zhu et al.225 successfully intercalated tannic acid (TA) etched ZIF-8 (TZIF-8) nanocrystals with hollow structures into the interlayer spacing of MXene, thus forming MXene/TZIF-8 membranes with high permeability and good separation effects on various oil-water emulsions. Moreover, Yang et al.226 reported a novel photocatalyst created through in situ growth and self-assembly, where MXene served as a support for ZIF-8 MOF immobilization. The MXene@ZIF-8 was further modified with TA and polyphenol compounds to form a cross-linked network that enhanced the heterogeneous interface, which is essential for efficient photocatalysis. The performance of the photocatalyst was optimized by exploring different concentrations of TA.

In addition to the previously described metal NPs, metal oxide, carbon-based materials, organic compound@MOF composites combined with MXene, Sun et al.227 adopted a novel kirkendall effect assisted electrostatic self-assembly method to construct a MOF (MIL-88A) substrate surface decorated with Ni–Fe layered double hydroxide (LDH) and formed multilayer nanocages, which were coated with MXene on the outer surface of the nanocage. By adjusting the adsorption of MXene on the LDH surface, the heterogeneous interface in the MOF-LDH-MXene ternary composite exhibited excellent interfacial polarization loss. Recently, hybrid systems incorporating MnCO3@MOF,228 polydimethylsiloxane-MOF@Ag,229 and the reduced graphene oxide (rGO)/PDA/FeCu-MOF230 composites with MXene have also been investigated.

The synthesis of MOF composite/MXene materials, as detailed in the preceding sections, is characterized by a remarkable diversity in both the composition of the MOF composite and the strategies used to integrate it with MXene. A critical analysis reveals that the choice of synthesis method often depends on the properties of the MOF composite and the desired final architecture.

(1) In situ growth is a powerful strategy for constructing multi-layered or core–shell architectures. This approach was used to prepare Pt/ZIF-Ti3C2Tx by sequentially growing ZIF-67 on MXene and then depositing Pt nanocrystals,210 and to construct the hierarchical MXene@ZIF-8@Pt NPs composite.211 This method allows for precise, step-by-step engineering of complex heterostructures with strong interfacial contact.

(2) Solvothermal and hydrothermal methods are particularly effective for the synthesis of MOF@metal oxide/MXene and MOF@carbon-based material/MXene composites. The synthesis of Fe3O4@MXene@MOF-74 (ref. 213) and the MXene@SnO2-Ce-MOF Z-scheme heterojunction214 relied on these techniques. Similarly, the integration of MXene with pre-formed MIL-101(Fe)/g-C3N4 composites was achieved via a solvothermal process.220 These methods are highly suitable for reactions that require controlled temperature and pressure to form crystalline composites.

(3) Self-assembly strategies, including electrostatic self-assembly, vacuum-assisted filtration, and thermal compression, are highly versatile for creating layered, membrane-like, or 3D aerogel-based composites. This approach was used to synthesize MOF-525@Ecoffex nanocomposites with a MXene@Co-NPC hybrid,216 to construct ZIF-67@MXene/PI hybrid materials via thermal compression of aerogels,222 and to fabricate MXene/TZIF-8 membranes with high permeability.225 The self-assembly of MOF@organic compound/MXene composites is also common, as demonstrated by the electrostatic self-assembly of Fe-MOF-NH2/CNT-NH2/MXene218 and the in situ growth of ZIF-8@MXene for polymer electrolytes.221 These methods are often conducted under mild conditions, preserving the integrity of MXene.

(4) Coprecipitation is a direct and scalable method for the in situ formation of MOF composites on MXene, as demonstrated by the universal strategy for fabricating hollow 3D MXene/MOF composites.120

In conclusion, the synthesis of MOF composite/MXene materials is highly flexible, with the choice of method enabling the design of a wide range of morphologies and architectures. In situ growth excels in building complex, multi-component systems, solvothermal/hydrothermal methods are ideal for crystalline composite integration, and self-assembly offers a powerful route to layered and functionalized materials.

2.3. MOF derivative/MXene

MOF derivatives, which exhibit high specific surface area and porosity, structural flexibility, tunable functionality, and excellent chemical stability, have been effectively integrated with MXene, known for its ultra-thin nature, robust mechanical toughness, and superior electrical conductivity. This synergistic combination has garnered significant attention from research groups worldwide. Several successful syntheses of MOF derivative/MXene composites are documented herein, categorized into seven distinct groups based on the variations in MOF derivatives: (1) MOF-derived metal/MXene; (2) MOF-derived metal oxide/MXene; (3) MOF-derived metal phosphide/MXene; (4) MOF-derived metal selenide/MXene; (5) MOF-derived metal sulfide/MXene; (6) MOF-derived metal hydroxide/MXene; (7) carbon-based material derived from MOF/MXene-based materials. Each category presents unique challenges and opportunities, highlighting the diverse potential applications of these composite materials.
2.3.1. MOF-derived metal/MXene. With the increasing demand for high-performance composite materials, the combination of MOF-derived metal-based materials with MXene has emerged as a prominent research focus. Wang et al.231 developed a hierarchical composite through electrostatic-mediated assembly. By leveraging cationic attraction between Co and Fe ions and anionic MXene substrates functionalized with surface-terminating groups, they achieved uniform adsorption of metallic precursors. Subsequent introduction of 1,4-benzenedicarboxylic acid as bridging ligand induced spontaneous coordination assembly, generating interlamellar FeCo-MOF spacers that effectively mitigated MXene restacking through structural confinement. The carbonization protocol under controlled pyrolysis enabled in situ formation of bimetallic nanocatalysts, where the spatial restriction from the MOF-derived matrix ensured homogeneous dispersion of FeCo NPs across the MXene scaffolds. The resulting engineered nanohybrids exhibited dual functional advantages: (1) simultaneous high polarity and ionic-electronic conductivity facilitating polysulfide redox mediation, and (2) optimized sulfur electrochemistry through shortened charge-transfer pathways. Notably, the 0D–2D architecture demonstrated synergistic catalytic functionality by integrating metallic active sites with conductive MXene platforms, significantly enhancing sulfur utilization efficiency in energy storage systems.

Furthermore, considering the suboptimal electrochemical performance of MXene layered 2D structures, Haridas B. Parse et al.232 reported the preparation and electrocatalytic ability of a composite material comprising MXene and Co NPs derived from Co-MOF, encapsulated within N-doped carbon (NC). Han and co-workers233 explored the construction of accordion-like MXene@Co-CZIF and MXene@Ni-CZIF composites (derived from the pyrolysis of Co-ZIF and Ni-ZIF, denoted as Co-CZIF and Ni-CZIF, respectively) via an electrostatic self-assembly strategy coupled with subsequent pyrolysis in H2/Ar atmosphere. As depicted in Fig. 7b, the Co-CZIF component maintained well-defined polyhedral geometry throughout MXene's surface and internal matrix. Fig. 7c demonstrates the effective incorporation of spherical Ni-ZIF nanostructures with hierarchical architecture into the MXene framework through this assembly process, with NPs distributed across both the exterior and interior regions. Xiang et al.234 prepared laminated Ti3C2Tx/CNTs/Co through combining microwave-assisted synthesis, in situ carbonization, and electrostatic assembly methods. A schematic illustration of the nanocomposite formation mechanism is presented in Fig. 7a, while Fig. 7d specifically illustrates the anchoring of sea urchin-shaped CNTs/Co nanostructures onto Ti3C2Tx MXene substrates. In a subsequent advancement, Liu and co-workers235 fabricated Ti3C2Tx/carbon nanofibers (CNFs)/TiO2/CoNi nanocomposites via electrostatic self-assembly combined with thermal treatment. Fig. 7e–g provide a comprehensive overview of the synthetic pathway and morphological characterization, as revealed by SEM analysis.


image file: d5ra04436g-f7.tif
Fig. 7 (a) Synthetic route for Ti3C2Tx/CNTs/Co nanocomposites. SEM images of MXene@Co-CZIF (b), MXene@Ni-CZIF (c), and Ti3C2Tx/CNTs/Co (d). (e) Scheme for the preparation of Ti3C2Tx/CNFs/TiO2/CoNi nanocomposites. SEM images of CoNi-MOF (f) and Ti3C2Tx/CNFs/TiO2/CoNi nanocomposites (g). (a and d) Reproduced with permission.234 Copyright 2021, Springer. (b and c) Reproduced with permission.233 Copyright 2020, American Chemical Society. (e–g) Reproduced with permission.235 Copyright 2022, Elsevier.
2.3.2. MOF-derived metal oxide/MXene. MOF-derived metal oxide-based materials/MXene composites have garnered significant research interest owing to their distinctive structural characteristics and synergistically enhanced functional performance.236 These hybrid materials effectively combine the intrinsic high catalytic activity of MOF-derived oxides with the excellent electrical conductivity and robust mechanical stability of MXene. This section provides a systematic overview of the prevailing synthetic strategies employed in the fabrication of MOF-derived metal oxide/MXene.

Ji et al.237 prepared uniform-sized MOF structures via hydrothermal synthesis, which were subsequently subjected to thermal decomposition to yield hollow Fe2O3. Concurrently, MXene was synthesized via in situ etching and processed under controlled conditions to enable directional alignment of MXene nanosheets driven by ice-crystal compression, resulting in highly ordered MXene aerogel frameworks. The hollow Fe2O3 structures were then dispersed in a silica sol and infiltrated into the porous MXene structure. Ultimately, the processed composite underwent lyophilization, yielding MXene-Fe2O3–SiO2 aerogels. Xie and co-workers238 successfully synthesized a flexible electrode film through vacuum filtration integration of Ti3C2Tx and MOF-derived Co–Fe oxide porous nanorods. Fig. 8a illustrates the composite fabrication process, which involves three principal stages. Firstly, the precursor Ti3AlC2 undergoes chemical etching using HCl/LiF solution, yielding multilayer MXene sediment that is subsequently delaminated into colloidal suspension via manual agitation. Concurrently, hydrothermal synthesis produces Co–Fe-MOF nanorod precursors that are subsequently calcined in oxidative atmosphere to generate porous mixed oxide phases. The resulting dispersions are combined and subjected to vacuum-assisted filtration, yielding a free-standing Co–Fe oxide/MXene composite film. In 2022, Lei et al.239 prepared a layered Ti3C2Tx hybrid material designated as Ti3C2Tx@CoFe@TiO2 through microwave-assisted thermal treatment of CoFe-MOF@Ti3C2Tx precursor at multiple temperature gradients, employing an integrated in situ synthesis strategy. Shi et al.240 also demonstrated the construction of MOF-derived porous CoFe2O4@carbon@alkalized MXene composite architecture using electrospinning, in situ growth, and subsequent calcination. Shingte et al.241 embedded MOF-derived rhombic nickel ferrite NPs (NFO NPs) into Ti3C2Tx MXene layers by a straightforward blending technique. The embedded NFO NPs served as effective interlayer spacers, mitigating restacking of MXene nanosheets and thereby increasing the accessible surface area of the composite.


image file: d5ra04436g-f8.tif
Fig. 8 Schematic fabrication process of Co–Fe oxide/MXene composite film (a), ZnCo2O4@MXene/S composite (b), and In2O3/Ti3C2Tx MXene composite (c). (d and e) SEM images of In2O3/Ti3C2Tx MXene composite. (f) Schematic illustration of the synthesis approach of MXene@Co3O4/ZnO composite. (g and h) SEM images of MXene@Co3O4/ZnO and Co3O4/MXene nanocomposite. (a) Reproduced with permission.238 Copyright 2020, Elsevier. (b) Reproduced with permission.244 Copyright 2021, Elsevier. (c–e) Reproduced with permission.242 Copyright 2022, Elsevier. (f and g) Reproduced with permission.246 Copyright 2022, Elsevier. (h) Reproduced with permission.247 Copyright 2022, Elsevier.

Liu et al.242 fabricated a MOF-derived In2O3/Ti3C2Tx MXene composite using a simple oil-bath method. The synthetic process is illustrated in Fig. 8c. Morphological characterization (Fig. 8d and e) demonstrates the successful growth of pristine In2O3 microtubes on delaminated MXene substrates, confirming effective heterostructure formation through this solution-phase approach. Subsequently, Zhang et al.243 successfully synthesized In2O3/ZnO/MXene by integrating MOF-derived hollow In2O3/ZnO nanotubes with MXene. The process began with the hydrothermal synthesis of MIL-68(In), followed by the introduction of Zn(NO3)2·6H2O to form In2O3/ZnO hollow nanotubes after calcination. These nanotubes, surface-modified with (3-aminopropyl) triethoxysilane, were dispersed in deionized water and stoichiometrically combined with MXene to yield the final ternary nanocomposite.

Wei and colleagues244 synthesized the ZnCo2O4@MXene/S composite. The process involves coating the MXene layer onto hollow ZnCo2O4 polyhedron through electrostatic assembly, as shown in Fig. 8b. The synthesis commenced with the production of faceted ZIF-8 nanocrystals through Zn2+ coordination with 2-methylimidazole in methanol, which subsequently acted as templates for the epitaxial deposition of ZIF-67 shells. A sequential growth approach was then implemented in Co2+-containing methanol solution to construct the core–shell ZIF-8@ZIF-67 architecture. Hollow porous ZnCo2O4 polyhedrons were obtained after annealing treatment. Afterwards, MXene layers were coated on the surface of hollow ZnCo2O4 polyhedrons through electrostatic assembly to form ZnCo2O4@MXene composites. Finally, sulfur was introduced into ZnCo2O4@MXene at 155 °C to easily synthesize ZnCo2O4@MXene/S composites. In 2022, Kshetri et al.245 engineered two specialized electrode architectures: a capacitor-type Co-porous carbon (PC)@MXene-CNF and a battery-type MnO2@Co3O4-PC@MXene-CNF assembly, both strategically constructed through hierarchical integration of Co-MOF@MXene-CNF. By strategically combining MOF-derived components with MXene substrates, Wu et al.246 developed a MXene@Co3O4/ZnO hybrid material using an optimized fabrication protocol. The architectural evolution of this multi-level MXene@MOF derivative system is schematically presented in Fig. 8f, revealing its sequential assembly mechanism. The morphological characteristics of the synthesized MXene@Co3O4/ZnO material are further evidenced by SEM analysis, with representative microstructural features displayed in Fig. 8g. Recently, Bu et al. synthesized the novel mesoporous Co3O4/MXene nanocomposites by direct calcination of ZIF-67 grown in situ on MXene sheets. The SEM image of Co3O4/Ti3C2Tx nanocomposite (Fig. 8h) reveals that the surfaces of the Ti3C2Tx sheets are entirely covered by Co3O4 NPs, demonstrating strong interfacial coupling between constituent phases.247 Then, Oh et al. constructed a simple strategy involving the construction of a 2D structural hybrid material consisting of Co3O4 and MXene nanosheets coated with N-doped graphite C (NGC), represented as Co3O4@NGC. The presence of MXene nanosheets can impart high electronic conductivity and structural stability to the composite material.248 Given that the restacking of 2D layered materials such as MXene limits their practical scalability, Zhou et al. prepared an electrocatalyst consisting of MOF-derived hollow CoV2O6 nanocubes (TS-V2CTx/CoV2O6 HN) anchored on lattice-stretched V2CTx MXene, achieved through an ion-exchange process followed by liquid nitrogen quenching.249 To enhance the electronegativity of nanocomposites, Rahman et al. engineered a dual-layer triboelectric device with enhanced flexibility and multifunctionality by incorporating MOF-derived nanoporous cobalt oxide (NPCO) blended with silicone polymer alongside MXene-based nanocomposites.250 Wang et al. conducted a series of studies on the construction of a hollow Co9S8 core@multi-shell MXene@Bi2O3 structure using MOF as a precursor, with optimization of the flexible substrate by varying the concentration of Co9S8.251

In addition, Wang et al.252 designed a Ti3C2Tx MXene composite integrated with MOF-derived CuO components. This architecturally layered structure, composed of alternating MXene sheets and CuO NPs derived from MOF precursors, significantly augmented the effective surface area of the resultant MXene-CuO hybrid material. Song et al.253 introduces the strategic combination of MnO2/Mn3O4 derived from MOFs with MXene, leading to the construction of a multifunctional MnO2/Mn3O4/MXene/Au NPs composite. The hierarchical architecture features 3D MnO2/Mn3O4 microcuboids embedded with vertically oriented nanosheets, which, in synergistic combination with MXene-supported Au NPs, substantially improves both electrochemical characteristics and interfacial accessibility. This design achieves dual enhancement in signal transduction efficiency and biocompatibility through complementary material properties.

2.3.3. MOF-derived metal phosphide/MXene. Zong et al.254 synthesized a novel MOFs-CoP@MXene composite via the self-assembly method. The MOFs-CoP component is easily deformable, and its heterostructure forms a hollow polyhedral structure that enhances the stability of the MOFs-CoP@MXene composite. Similarly, Liu and co-workers255 successfully synthesized the hierarchical sandwich-like CoP-NC@Ti3C2Tx composite through argon-protected in situ intercalation and phosphorylation of MOF@MXene precursors (Fig. 9a). Their approach involved the deposition of geometrically distorted ZIF-67 rhombic dodecahedrons onto Ti3C2Tx substrates, creating a ZIF-67@Ti3C2Tx architecture. Subsequent thermal treatment triggered ZIF-67 decomposition into 100–200 nm CoP-NC NPs. In a related methodology, Zong et al.256 engineered Ti2NTx@MOF-CoP heterostructures by phosphate-functionalizing Ti2NTx@ZIF-67 assemblies. This protocol commenced with Ti2NTx MXene synthesis, followed by electrostatic-driven ZIF-67 integration and final phosphidation. Li and colleagues reported MXene@MOF-Co2P nanocomposites through ZIF-67 templating, where nitrogen/carbon-doped Co2P architectures were immobilized on bimetallic MXene matrices.257 Building on their previous study, Wang and co-workers developed a simple procedure to prepare CoPC/CNTs@MXene absorbers by combining 0D/1D stratified CoPC/CNTs with MXene via electrostatic self-assembly.258
image file: d5ra04436g-f9.tif
Fig. 9 (a) Formation approach of CoP-NC@Ti3C2Tx. SEM images of NiSe2–CoSe2@C (b) and NiSe2–CoSe2@C/Ti3C2Tx (c). (d) Schematic view of formation for the hierarchical 2D MXene@CoSe2/Ni3Se4 nanosheets. (e) SEM and (f) TEM images of MXene@CoSe2/Ni3Se4 nanosheets. (g) Synthetic route of NCMS@C/MX. (h) SEM image of NCMS@C/MX. (a) Reproduced with permission.255 Copyright 2021, Elsevier. (b and c) Reproduced with permission.262 Copyright 2021, Elsevier. (d–f) Reproduced with permission.263 Copyright 2022, Elsevier. (g and h) Reproduced with permission.267 Copyright 2023, Wiley.

Ma et al.259 employed a self-assembly strategy combined with in situ phosphorization to construct a porous core–shell CoP@NC nano-polyhedral composite system (CoP@NC/MXene), using 2D MXene nanosheets as a conductive substrate. The precursor framework was first formed via self-assembly, followed by a phosphorization process that simultaneously generated the CoP core and a NC shell. The incorporation of MXene enhanced charge transfer efficiency through its 2D conductive network and suppressed the pulverization of active materials. Cui et al.260 focused on the integration of MOFs with 2D materials, synthesizing a layered porous material (graphene oxide (GO)/MXene@NiZrP) by in situ compositing UIO-66 with GO and MXene via hydrothermal and calcination methods. This approach leveraged the dual-substrate synergistic effects of GO and MXene to offer plentiful active sites and modulate interlayer spacing. During calcination, the MOF structure was transformed into metal phosphides while retaining high specific surface area and surface defect characteristics. Zhang et al.261 further expanded the MOF-derived strategy for MXene composites by embedding Co-doped Cu3P/NC octahedra (derived from MOF precursors) into MXene interlayers through an in situ growth method, forming a 3D@2D heterostructure (Co–Cu3P/NC@MXene). This synthetic route not only modulated the electronic structure of Cu3P via heterometal doping but also leveraged the 2D interlayer confinement effect of MXene to facilitate mass transport, thereby achieving dual optimization of hierarchical architecture and electronic properties.

2.3.4. MOF-derived metal selenide/MXene. Through a practical two-stage synthesis approach involving hydrothermal treatment and subsequent selenization at 400 °C in an argon environment, Hou et al. successfully fabricated multilayer NiSe2–CoSe2@C/Ti3C2Tx composites. As depicted in Fig. 9b and c, the petal-like structure was disrupted following high-temperature selenization, resulting in the appearance of carbon-coated particles adhering to each other.262 In 2022, Yang and co-workers achieved the fabrication of distinctive 2D hierarchical MXene@CoSe2/Ni3Se4 nanosheets through an innovative template-assisted approach. This method encompassed the in situ growth of MOF on MXene substrates, followed by sequential ion exchange and controlled selenization processes. Fig. 9d presents the synthetic pathway for constructing these multilayered heterostructures. Morphological characterization (Fig. 9e and f) reveals a well-defined architecture in which CoSe2/Ni3Se4 nanolayers uniformly adhere to both sides of the MXene substrates, forming a stable sandwich-like configuration.263 More recently, Li et al. reported on the recent research development regarding the VSe2–ZrO2/C/MXene composite by a practical solvothermal method and in situ selenization process.264 Additionally, Shi et al. synthesized the CoSe2@C/MXene composite material by annealing a pre-prepared Co-MOF/MXene (mass ratio 10[thin space (1/6-em)]:[thin space (1/6-em)]1) with selenium powder at 650 °C in an Ar atmosphere for 3 h. The resulting CoSe2 NPs were uniformly dispersed both on the surface and within the multi-heteroatom-doped carbon matrix, while the MXene nanosheets maintained their sheet structure within the composite material.265 Wang et al. reported a series of hollow-structured ZnSe–CoSe composites (45@ZnSe–CoSe/S, 45@ZnSe–CoSe/P/S, and 45@ZnSe–CoSe/MXene/S) with distinct hollow architectures, fabricated through a simple two-step synthesis method. This approach combined TA-assisted etching with subsequent selenization treatment of ZIF-8@ZIF-67.266 Fig. 9g presents a schematic diagram outlining the synthesis procedure of MOFs/MXene-derived multi-metal selenides@C/MXene (NCMS@C/MX). Liang et al. achieved the preparation of NCMS@C/MX by first synthesizing MOF and exfoliating MXene, followed by further selenization of the selenides@C/MXene composite. The morphology of NCMS@C/MX composite is depicted in the SEM image (Fig. 9h), where Ni3Se4 and CoSe2 NPs, as well as Mo3Se4 nanosheets, are uniformly distributed within the multi-heteroatom-doped carbon matrix.267 Furthermore, the advancement of K-ion batteries (KIBs) faces challenges stemming from insufficient anode materials possessing optimal theoretical capacity, rapid K+ transport characteristics, and environmental sustainability. To address this challenge, Hu et al. developed a MOF-templated synthetic route to fabricate ZnTe@C composites. Their strategy enabled homogeneous dispersion of ZnTe NPs within a conductive carbon framework, which was subsequently immobilized on MXene substrates. Crucially, electronic structure modulation of ZnTe enhanced K+ adsorption kinetics, while their pioneering work elucidated the conversion mechanism: yZnTe + xK+ + xeyZn + KxTey, establishing ZnTe as a promising KIB anode.268 Jeong Ho Na and colleagues successfully prepared 3D MXene microspheres via spray-drying assembly. This configuration effectively mitigates MXene sheet restacking while encapsulating MOF-derived ZnSe@NC heterostructures, demonstrating enhanced electrode stability.269 Vallem et al.270 proposed a rational design of a high-efficiency selenium host material, Fe-CNT/TiO2, based on a tailored composite nanostructure. The synthetic protocol involved the initial construction of a MOF-MXene composite, followed by CNT surface decoration, and final high-temperature sintering. The key innovation lies in utilizing the MOF-derived Fe-CNT structure to provide both conductive networks and catalytic active sites, while the MXene transforms into a TiO2 phase through sintering, forming a hierarchical porous architecture.
2.3.5. MOF-derived metal sulfide/MXene. Metal sulfides are regarded as highly promising nanomaterials. However, their practical applications are hindered by intrinsic limitations such as poor conductivity and significant volume expansion during electrochemical processes. To address these issues, metal sulfides derived from MOF offer a viable solution by enhancing electrical conductivity and mitigating volume expansion. Moreover, integrating MOF-derived metal sulfides to MXene conductive substrates can further improve their conductivity.271 In 2018, Zou et al. synthesized a unique porous Ni–Co mixed metal sulfide (NiCoS), which was grown on MXene derived from ZIF-67/MXene. The preparation process for NiCoS/MXene is depicted in Fig. 10a.272 Subsequently, Wei et al. designed and constructed a MOF-derived NiCo2S4 attached to MXene to form a NiCo2S4@MXene composite. The MOF-derived NiCo2S4 promotes the surface hydroxylation of MXene, facilitating the adsorption of metal ions and inducing sulfur vacancy formation. In addition, MXene provides a high specific surface area, enabling rapid charge carrier transport. Through synergistic interactions between components, the resulting composites have significantly enhanced nonlinear optical and electrochemical properties.273 In Fig. 10b, Li et al. addressed the primary limitations of metal sulfide materials, namely slow kinetics and severe volume expansion, via utilizing poly(methyl methacrylate) spheres as sacrificial template to form MXene nanosheets. Subsequently, through a multi-step procedure involving in situ growth, carbonization, and sulfurization processes, they successfully synthesized MXene@CoS2/NC composite materials comprising MOF-derived CoS2/NC integrated with MXene nanosheets.274 More recently, Adil et al. developed an inventive method for creating the MXene cobalt sulfide (MXene-CoS) composite material. A cost-effective Co-MOF was sulfided and directly integrated with an MXene film via a binder-free in situ growth strategy on NF, resulting in the formation of MXene-CoS/NF. This architecture exhibits favorable structural characteristics, including increased porosity, abundant active sites, enlarged interfacial area, and superior electrocatalytic activity.275 Additionally, Farooq et al.276 established a synthetic strategy for preparing bifunctional electrocatalyst CoS@C/MXene through high-temperature vulcanization. The process involves first combining ZIF-67 with MXene to form a precursor hybrid material, followed by high-temperature vulcanization treatment in a H2S atmosphere. For instance, the CoS@C/MXene underwent vulcanization at 800 °C, which transformed the cobalt species in ZIF-67 into CoS while simultaneously carbonizing them into a carbon layer structure encapsulating CoS. This structure tightly integrated with the MXene, ultimately forming the CoS@C/MXene nanocomposite.
image file: d5ra04436g-f10.tif
Fig. 10 (a) The preparation process of NiCoS/MXene. (b) Schematic representation of the synthesis of MXene@CoS2/NC. (a) Reproduced with permission.272 Copyright 2018, American Chemical Society. (b) Reproduced with permission.274 Copyright 2022, Elsevier.
2.3.6. MOF-derived metal hydroxide/MXene. In recent years, the synthesis strategies of MOF-derived metal hydroxides/MXene composite have demonstrated unique advantages in structural engineering. Hu et al.277 innovatively constructed CoFe LDH/MXene composite structure through an “etching-coprecipitation” post-treatment method: by utilizing MXene as a multifunctional support, they anchored MOF precursors and achieved controlled etching to release metal ions, thereby inducing the in situ coprecipitation growth of LDH. This interfacial engineering not only optimized the charge transport pathways but also regulated the intermediate adsorption strength at active sites through strong electronic interactions between CoFe LDH and MXene. In contrast, Kasinathan et al.278 adopted a “charge-directed self-assembly” strategy, leveraging electrostatic interactions to integrate chitosan (CS)-modified MXene nanosheets with cobalt-tungsten LDH, resulting in the formation of a CS/MXene@LDH hybrid structure. The self-assembly process, driven by the complementary surface charges of negatively charged MXene and positively charged CoW LDH, enabled uniform dispersion and robust integration of the nanoscale components. Although these two studies differ in application fields, both highlight the shared philosophy of precise structural regulation and functional integration in composite system fabrication through interfacial synergistic design with MXene, employing distinct pathways of chemical etching reconstruction and physical electrostatic adsorption respectively.
2.3.7. Carbon-based material from MOF/MXene-based materials. Synthesis methods of MOF-derived carbon/MXene-based materials have demonstrated diversified innovative pathways in recent years.279–281 To advance the development of high-performance microwave absorption materials, Zhou et al.282 developed an innovative approach by integrating 1D CNTs derived from Ni-MOFs (exhibiting hollow solid microsphere architecture) with 2D MXene to establish an interconnected 3D conductive framework (Fig. 11a–c). In addition, Gu et al. reported a MXene-supported MOF-derived Fe–N–C catalyst (Fe–Nx/N/MXene), which exhibits good oxygen reduction reaction (ORR) activity and long-term stability.283 To achieve optimal electromagnetic wave (EMW) absorption capability, Wu et al. synthesized a Co–Ni-MOF-derived waxberry-like magnetic porous carbon (CoNi-MPC)@CNTs/MXene hybrid (CoNi-MPC@CNTs/MXene) through a simple solution precipitation, heat treatment, and electrostatic self-assembly strategy (Fig. 11d).284
image file: d5ra04436g-f11.tif
Fig. 11 (a) Schematic illustration of the synthesis of MXene/Ni/N-CNTs. (b and c) TEM images different magnifications of MXene/Ni/N-CNTs nanocomposites. (d) Schematic illustration of the fabrication of CoNi-MPC@CNTs/MXene. (a–c) Reproduced with permission.282 Copyright 2021, Elsevier. (d) Reproduced with permission.284 Copyright 2024, Elsevier.

In 2018, in order to further enhance the electromagnetic attenuation ability, Liao and co-workers285 prepared laminated Co/TiO2–C hybrids through thermal treatment of Ti3C2Tx/Co-MOF, integrating them with MXene-derived TiO2–C laminates (Fig. 12a–d). Building upon this foundation, Deng et al.286 subsequently engineered innovative sandwich-structured 2D nanocomposites containing Fe & TiO2 NPs embedded in carbon matrices, utilizing MXene-MOF hybrid architectures as synthetic platforms. As shown in Fig. 12e–h, Fe-MOFs were initially transformed into Fe3O4/PC, and subsequently formed carbon-coated Fe NPs (Fe@C) under a reducing atmosphere. With increasing pyrolysis temperature, both TiO2 and Fe particles exhibited progressive growth and pronounced aggregation. Recently, Wang and colleagues287 successfully prepared 3D spherical Ni/TiO2/C microspheres alongside hierarchically layered Ni/TiO2/C nanocomposites through an innovative synthesis approach. Utilizing Ni-BTC and Ti3C2Tx MXene as precursor materials, the researchers implemented an in situ solvothermal self-assembly strategy coupled with controlled thermal annealing. The resulting nanostructures incorporate precisely arranged 2D carbon matrices hosting uniformly dispersed TiO2 NPs (Fig. 12i). Cui et al.288 employed biomass-derived bamboo as carbon source to construct a 3D interconnected network through porous structure engineering strategy. Briefly, bamboo cellulose was assembled with few-layer Nb2CTx MXene to form a 3D aerogel framework. Subsequently, ZIF-67 precursors were uniformly anchored via chemical deposition, and ultimately Co NPs/Nb2CTx/CA composite was obtained through high-temperature pyrolysis (Fig. 12j and k). This architecture establishes a conductive network through synergistic interactions between the biomass-derived carbon matrix and MXene components, while simultaneously optimizing impedance matching via interfacial polarization and magnetic loss contributions from MOF-derived magnetic nanoparticles. Zheng et al.289 proposed a biomimetic structural design strategy. Their methodology employed solvothermal synthesis to realize controlled in situ crystallization of bimetallic CoNi-MOF within MXene layers, which subsequently underwent pyrolysis conversion into magnetically responsive carbon hybrids (CoNi@C). This transformation process yielded a hierarchically organized structure mimicking marine coral morphology. This strategy effectively alleviates MXene stacking through interlayer-embedded MOF-derived carbon nanorods, while introducing porous heterogeneous interfaces to enhance dielectric-magnetic coupling effects. Furthermore, Guo et al.290 innovatively addressed limitations associated with single-layer architectures through a sequential vacuum-assisted filtration strategy. Their Janus architecture combined two functionally distinct components: a magnetic CNF composite derived from CoNi-MOF-74 served as the electromagnetic absorption layer, and MXene/CNF acting as the reflective layer. The integrated system achieved robust interfacial bonding via hydrogen bonding interactions between constituent layers. This dual-layer architecture not only optimizes electromagnetic dissipation pathways through “absorption-reflection-reabsorption” mechanism, but also enhances multifunctionality by harnessing MXene's plasmonic effects and the mechanical reinforcement properties of CNF.


image file: d5ra04436g-f12.tif
Fig. 12 (a) Illustration of the preparation of Co/TiO2–C. (b–d) SEM images of Ti3C2Tx, Co-MOF, and Co/TiO2–C composites. (e) Preparation of Fe & TiO2@C. (f–h) TEM images of Ti3C2Tx, Fe & TiO2@C (700 °C), and Fe & TiO2@C (800 °C). (i) Illustration of the fabrication process of Ni/TiO2/C nanocomposites. (j and k) SEM images of ZIF-67/Nb2CTx/cellulose aerogel and Co/Nb2CTx/cellulose CA. (a–d) Reproduced with permission.285 Copyright 2018, American Chemical Society. (e–h) Reproduced with permission.286 Copyright 2020, Springer. (i) Reproduced with permission.287 Copyright 2022, Elsevier. (j and k) Reproduced with permission.288 Copyright 2023, American Chemical Society.
2.3.8 Summary. The synthesis of MOF derivative/MXene composites, as detailed in the previous sections, is fundamentally different from that of pure MOF/MXene or MOF composite/MXene materials. The defining characteristic of this type of material is the essential role of post-synthetic thermal treatment (e.g., calcination, pyrolysis, sintering) in transforming the MOF structure into the final derivative while preserving the structural integrity of the MXene substrate.

A critical analysis reveals that the synthesis strategy is a multi-step process: firstly, a precursor composite is formed by integrating the MOF with MXene, and secondly, this composite undergoes a controlled thermal transformation to generate the derivative. The choice of the initial integration method for the precursor is diverse. In situ growth is a common approach, used to construct MOF@MXene that are subsequently converted into metal phosphides261 or metal sulfides.275 Electrostatic self-assembly is also highly effective for creating the initial hybrid, as demonstrated by the formation of MXene@Co-CZIF and MXene@Ni-CZIF composites, which were then pyrolyzed to form metal/carbon hybrids.233 Solvothermal methods are employed to create layered porous materials, such as GO/MXene@NiZrP, which are then transformed into phosphides via calcination.260

The thermal treatment step is the most critical and challenging aspect. The temperature, atmosphere (e.g., N2, Ar), and heating rate must be meticulously controlled. The objective is to completely decompose the MOF into the desired derivative (e.g., metal, oxide, phosphide, sulfide), while minimizing the oxidation and degradation of the MXene. For instance, the successful synthesis of MOF-derived metal phosphides,260,261 sulfides,275,276 and hydroxides277 depends on finding this delicate balance. The use of MXene as a conductive and mechanically robust substrate not only provides a support for MOF precursors but also plays a crucial role in stabilizing the derivative phases and preventing its aggregation during the high-temperature process.

In conclusion, the synthesis of MOF derivative/MXene composites is a complex process, with the core being the integration of MOF precursors with MXene, followed by precisely controlled thermal transformation. The success of this strategy depends on the synergy between a well-designed precursor architecture and optimized thermal conditions to achieve high-performance hybrid materials.

3. Applications

The versatility of MOF/MXene-based materials is illustrated in Scheme 3, highlighting their broad applicability in batteries,291–293 SCs,120,245 sensors,294–296 catalysts,297–299 and absorption.300–302 In advanced battery systems, MOF compounds serve bifunctional roles. These porous materials serve not only as high-performance anode components in lithium-ion batteries (LIBs) due to their exceptional specific capacity and outstanding energy storage properties, but also show promise as cathode materials by maintaining stable redox activity and effectively mitigating detrimental shuttling effects.303 For SC applications, MOF/MXene composites demonstrate ultra-high specific capacitance and excellent reversibility, rendering them strong candidates as binder-free electrode materials.105,196 Furthermore, MOF/MXene composites are extensively utilized in multiple sensing applications, showcasing remarkable adaptability. For example, efficient sensors for arsenate detection in electrochemistry, self-powered biomotion and tactile sensing applications, electrochemically sensitive sensors for H2O2 detection, and as organoid gas sensors.129,133,216 In summary, these sensors demonstrate outstanding selectivity, consistent reproducibility, and durable operational stability, attributes that collectively contribute to technological advancements in societal infrastructure. The unique combination of adjustable surface properties, architectural versatility, and expansive active interfaces has positioned MOF/MXene hybrid systems as a focal point in advanced materials research. These attributes make it a promising catalyst material with high electrical conductivity, including oxygen evolution reaction (OER),304 hydrogen evolution reaction (HER),305 ORR,306 and photocatalysts.307,308
image file: d5ra04436g-s3.tif
Scheme 3 Application of MOF/MXene-based materials.

In the field of EMW absorption, MOF/MXene-based materials demonstrate exceptional EMW attenuation capabilities, significantly enhancing absorption efficiency and exhibiting great potential for next-generation intelligent electromagnetic shielding systems.290 Due to their lightweight nature and broad absorption bandwidth, these materials can address a range of electromagnetic pollution issues in daily life. High-performance absorbing materials can markedly enhance the absorption performance of EMW and positively impact microwave absorption performance, showcasing an excellent absorption bandwidth.286 This suggests that MOF/MXene-based materials play a significant role in societal development. MOF/MXene-based materials can serve as highly effective flame retardants, enhancing fire safety while also possessing adsorption and sterilization effects.141,172 These hybrid systems show particular promise for dual applications in agricultural engineering (particularly paddy field management) and biomedical wound care, with documented therapeutic implementations already contributing to enhanced healthcare outcomes and quality-of-life improvements.309

3.1. Batteries

MOF materials demonstrate significant potential as battery electrode materials, offering advantages including numerous active sites, hierarchically porous architectures, minimized ion transport distances, and tunable chemical compositions.310 However, their relatively low electronic conductivity remains a limitation. Therefore, developing innovative electrode architectures with enhanced energy storage performance and mechanical robustness represents a vital research direction. The research findings on MOF/MXene-based materials for battery over the past few years are presented in Table 1.
Table 1 MOF/MXene-based materials for batteries, including LIBs, LSBs and others
MOF/MXene-based materials Type of batteries Reversible capacity/current density Cycling performance (capacitance retention, cycle, current density) Ref.
Ti3C2 MXene@NiCo-MOF LIBs 402 mA h g−1/0.1 A g−1 85.7%, 400, 1 A g−1 182
MXene@NiCo-MOF LIBs 637 mA h g−1/0.2 A g−1 304 mA h g−1, 800, 1 A g−1 185
Ni-MOF@MXene LIBs 980 mA h g−1/0.2 A g−1 80%, 5000, 5 A g−1 311
PE-ZIF-8@MXene LIBs 121 mA h g−1/5C 89.6%, 500, 1C 221
MOFs-CoP@MXene LIBs 706.5 mA h g−1/0.2 A g−1 585.8 mA h g−1, 1000, 0.5 A g−1 254
VSe2–ZrO2@C@MXene LIBs 612.2 mA h g−1/0.1 A g−1 430 mA h g−1, 1000, 1 A g−1 264
CoSe2/ZnSe@MXene LIBs 536.7 mA h g−1/1 A g−1 63.1%, 2000, 1 A g−1 310
ZnxInyS/MXene LIBs 1300 mA h g−1/0.5 A g−1 1097 mA h g−1, 1000, 1 A g−1 271
Co3O4@NGC/MXene LIBs 327 mA h g−1/50 A g−1 830 mA h g−1, 500, 1 A g−1 248
Zr-MOF@Ti3C2Tx LSBs 1017 mA h g−1/0.1C 46%, 1000, 1C 177
Ti3C2Tx-UIO-66-NH2 LSBs 840 mA h g−1/0.5C 38.5%, 1500, 1C 293
ZnCo2O4@Ti3C2/S LSBs 547 mA h g−1/0.2C 26.8%, 400, 0.5C 244
45@ZnSe–CoSe/MXene LSBs 478 mA h g−1/0.2C 387.2 mA h g−1, 400, 0.5C 266
FeCo@MXene LSBs 703 mA h g−1/4C 58%, 1000, 1C 231
ZnTe@C/Ti3C2Tx KIBs 408 mA h g−1/0.1 A g−1 230.2 mA h g−1, 3500, 1 A g−1 268
MXene/ZnSe@NC KIBs 180 mA h g−1/0.5 A g−1 238 mA h g−1, 1000, 0.5 A g−1 269
Co3C/MXene@C KIBs 236 mA h g−1/0.2 A g-1 129.3 mA h g−1, 500, 0.5 A g−1 312
CoP-NC@Ti3C2Tx SIBs 142.3 mA h g−1/0.05 A g−1 101.6 mA h g−1, 500, 0.5 A g−1 255
CoP@N–C/MXene SIBs 658.6 mA h g−1/0.05 A g−1 61 mA h g−1, 500, 2 A g−1 259
MXene@CoS2/NC SIBs 620 mA h g−1/0.2 A g−1 87.5%, 5000, 5 A g−1 274
CoSe2@C/MXene SIBs 402.9 mA h g−1/0.1 A g−1 259.8 mA h g−1, 1800, 1 A g−1 265
CoFe2O4@carbon@MXene SIBs 416 mA h g−1/0.1 A g−1 130, 1000, 1 A g−1 240


Within the realm of MOF/MXene composites, both 3D Ti3C2 MXene/NiCo-MOF composite films182 and 2D MXene/NiCo MOF heterostructures185 have been identified as highly effective anodes for high-performance LIBs. The optimized MXene/NiCo-MOF electrode showcases remarkable electrochemical characteristics, delivering a reversible capacity of 402 mA h g−1 at 0.1 A g−1 with sustained rate capability (256 mA h g−1 at 1 A g−1) through 300 cycles. These enhancements stem from synergistic effects including three-dimensional porous networks, expanded surface area, accelerated electron transport kinetics, and optimized Li+ diffusion pathways. Long-term cycling stability data (Fig. 13a) reveal 85.7% capacity retention after 400 cycles at 1 A g−1, confirming structural resilience under high current operation. Complementary cyclic voltammetry (CV) analysis conducted at various scan rates (Fig. 13b) further confirms the superior electrochemical behavior of these composites.182 Electrochemical characterization of the Ti3C2Tx/NiCo MOF heterostructure displays progressive capacity enhancement during cycling (Fig. 13c), a phenomenon associated with gradual electrode activation and dynamic evolution of a functional gel-like interfacial layer. The material's exceptional rate adaptability, as shown in Fig. 13d, combined with its distinctive voltage profiles across cycling stages, highlights the advantages of combining conductive MXene substrates with redox-active bimetallic MOF structure (Fig. 13e).185 Notably, the MOF@MXene heterostructure anode developed by Sun et al. exhibits remarkable electrochemical performance, maintaining 80% capacity retention over 5000 cycles at 5A g−1, a significant improvement compared to individual MXene and MOF components. Li+ kinetics were investigated via CV measurements across scan rates ranging from 0.1 to 1 mV s−1 (Fig. 13f), with long-term cycling stability at high current density is illustrated in Fig. 13g. Structural analysis in Fig. 13h identifies two distinct Li+ adsorption sites (labeled 1 and 2) in the pure NF-MOF. Overall, this hybrid architecture demonstrates enhanced Li storage capabilities and superior rate performance as LIB anodes, exceeding the electrochemical characteristics of both constituent materials when used separately.311


image file: d5ra04436g-f13.tif
Fig. 13 (a) Cycle performance of MXene/NiCo-MOF at 1 A g−1. (b) CV curves of the MXene/NiCo-MOF at different scan rates. (c) The discharging/charging curves of the Ti3C2Tx/NiCo-MOF electrode at 0.2 A g−1. (d) Rate performance of the Ti3C2Tx/NiCo-MOF. (e) Illustration of the Li+ intercalation of the MXene/NiCo-MOF heterostructure. (f) CV curves of NF-MOF@MXene at different scan rates. (g) Cycling performance at 5 A g−1. (h) Lithium adsorbed at different sites. (a and b) Reproduced with permission.182 Copyright 2020, Molecular Diversity Preservation International. (c–e) Reproduced with permission.185 Copyright 2022, Elsevier. (f–h) Reproduced with permission.311 Copyright 2022, Elsevier.

Furthermore, a novel polymer composite electrolyte reinforced with ZIF-8@MXene nanosheets (PE-ZIF-8@MXene) was developed by Zhao et al. When implemented in symmetric lithium battery configurations, the PE-ZIF-8@MXene enabled stable operation for 2000 h at 0.5 mA cm−2. Electrochemical performance evaluation revealed that Li|PE-ZIF-8@MXene|Li batteries maintained 89.6% of their initial capacity through 500 cycles (Fig. 14c). Comparative analysis with conventional PE demonstrated the modified electrolyte's superior voltage hysteresis characteristics across varying current densities (Fig. 14a). Current-dependent polarization effects were observed in the voltage–capacity profiles, with minimal increases noted when scaling current density from 0.5C to 2C (Fig. 14b).221


image file: d5ra04436g-f14.tif
Fig. 14 (a) Rate performances of symmetric batteries with PE-ZIF-8@MXene and PE. (b) Concentrated voltage–capacity curves of PE-ZIF-8@MXene. (c) Cycling performance of PE-ZIF-8@MXene and PE. (d) CV curves of MOFs-CoP/S and MOFs-CoP@MXene/S. (e) Discharge–charge curves of MOFs-CoP@MXene. (f) Rate capability of MXene/S, MOFs-CoP/S and MOFs-CoP@MXene/S. (g) Schematic illustration of the inhibition mechanism of MOFs-CoP/S and MOFs-CoP@MXene/S. (h) Schematic of VSe2–ZrO2/C/MXene for the Li-ion storage mechanism. (i) Potential vs. time curve and (j) graph for the change of Rtot values at specific potential levels of Co3O4@NGC/MXene electrode for LIBs during the initial cycling. (k) The initial profile curves and (l) rate performance of Co3O4@NGC and Co3O4@NGC/MXene electrodes for LIBs. (a–c) Reproduced with permission.221 Copyright 2022, Elsevier. (d–g) Reproduced with permission.254 Copyright 2021, Elsevier. (h) Reproduced with permission.264 Copyright 2023, Royal Society of Chemistry. (i–l) Reproduced with permission.248 Copyright 2021, Elsevier.

In the development of MOF derivative@MXene composites, Zong et al. synthesized an innovative MOFs-CoP@MXene hybrid through structural engineering. When employed as a LIB anode, this composite demonstrated exceptional cycling durability. As a cathode material for lithium–sulfur batteries (LSBs), it exhibited dual functionality: stable sulfur redox kinetics and significant mitigation of polysulfide shuttle phenomena through chemical confinement. Electrochemical characterization revealed distinct redox peaks in CV profiles (1.5–3.0 V, 0.1 mV s−1), with the MOFs-CoP@MXene/S configuration showing enhanced cathodic activity evidenced by dual reduction peaks at 2.02 and 2.37 V, compared to the MOFs-CoP/S counterpart (Fig. 14d). The initial charge and discharge capacities of the MOFs-CoP@MXene/S were 1138.8 mA h g−1 at 0.2C (Fig. 14f). Fig. 14g presents the inherent mechanism by which MOFs-CoP@MXene/S inhibits the sulfide shuttle effect, thereby exerting its catalytic potential.254 A MXene-based composite integrating VSe2–ZrO2 and carbon components (VSe2–ZrO2/C/MXene) was developed by Li and co-workers as advanced LIB anode material. This engineered material exhibited enhanced electrochemical characteristics, initially displaying capacity fluctuation before achieving remarkable improvement from 461.2 to 1238.5 mA h g−1 at 100 mA g−1. Notably, the composite exhibited exceptional durability with 430 mA h g−1 retention after 1000 cycles under1.0 A g−1, confirming its structural resilience. The underlying charge storage mechanism is visually elucidated in the corresponding schematic diagram (Fig. 14h).264 Recent studies by Oh et al. have identified the Co3O4@NGC/MXene hybrid as a promising LIBs anode material, demonstrating remarkable energy storage characteristics. The composite maintained 830 mA h g−1 after 500 cycles at 1.0 A g−1, along with superior rate performance delivering 327 mA h g−1 at 50.0 A g−1. The electrochemical behavior analysis reveals: Fig. 14i quantifies cycle endurance across different voltage windows, while Fig. 14j tracks the evolution of total resistance (Rtot) during initial battery operation. Comparative charge–discharge curves of Co3O4@NGC and Co3O4@NGC/MXene electrodes at 1.0 A g−1 are presented in Fig. 14k. Both configurations showed stable long-term cyclability, sustaining performance over 500 cycles as evidenced in Fig. 14l.248

LSBs have garnered significant attention as a promising rechargeable battery.313 Wen et al. initially synthesized the 3D hierarchical Zr-MOF/Ti3C2Tx nanocomposite, which displayed a substantial discharge/charge capacity of 801 mA h g−1 (Fig. 15b) at 1C and excellent cycling properties. Fig. 15c illustrates the long-term cycling performance of the Zr-MOF/Ti3C2Tx composite over 1000 cycles at 1C rate. Fig. 15a depicts the working mechanism of the Zr-MOF/Ti3C2Tx/S composite in LSBs.177 The superior architecture of the Zr-MOF/Ti3C2Tx nanocomposites significantly enhances the electronic conductivity and charge transfer kinetics of the entire cathode, accelerates the redox conversion rate of chemisorbed polysulfide, and effectively protects the lithium anode from contamination. Building upon previous advancements, Wei and co-workers further engineered a ZnCo2O4@Ti3C2/S composite serving as an effective sulfur host material for LSBs. This innovative configuration achieved remarkable electrochemical metrics, including a primary discharge capacity of 1283.9 mA h g−1 and an exceptional initial coulombic efficiency of 98.7% under 0.1C. The system's electrochemical durability was further validated through extended cycling evaluations at 0.5C, with corresponding data visualization provided in Fig. 15e. The CV curves analysis across scanning rates (0.1–0.4 mV s−1), as depicted in Fig. 15d, confirmed the material's stable redox behavior.244 Recent developments introduced the 45@ZnSe–CoSe/MXene/S composite as a high-performance cathode candidate for LSB systems. Comprehensive electrochemical characterization demonstrated multiple superior attributes: a substantial initial capacity of 1198.5 mA h g−1 at 0.1C, robust rate capability maintaining 489.8 mA h g−1 at 2C, along with exceptional cycling stability and electrochemical reversibility. These collective properties significantly enhance sulfur redox kinetics, positioning this material as a promising solution for durable LSB applications.266


image file: d5ra04436g-f15.tif
Fig. 15 (a) The working mechanism of the Zr-MOF/Ti3C2Tx/S in LSBs. (b) Charge/discharge voltage profiles of the Zr-MOF/Ti3C2Tx. (c) Cycle performance of the Zr-MOF/Ti3C2Tx at 1C. (d) CV curves of ZnCo2O4@Ti3C2/S. (e) Cycling performance for ZnCo2O4@Ti3C2/S cathodes at 0.5C. (f) Schematic illustration of KPB‖ZCT full cells. (a–c) Reproduced with permission.177 Copyright 2021, American Chemical Society. (d and e) Reproduced with permission.244 Copyright 2021, Elsevier. (f) Reproduced with permission.268 Copyright 2022, Wiley.

In addition to the previously discussed LIBs and LSBs, several other representative battery systems have been explored. For instance, Hu and co-workers demonstrated ZnTe an effective anode candidate for K+ storage, subsequently synthesizing the ZnTe@C/Ti3C2Tx (ZCT) nanohybrids.268 The ZCT exhibited a high capacity of 408.0 mA h g−1 at 0.1 A g−1 and superior long-term cycle ability (230.2 mA h g−1 at 1.0 A g−1 after 3500 cycles), making it a promising candidate for practical applications. Fig. 15f provides a schematic illustration of potassium Prussian blue (KPB)‖ZCT full cells. Furthermore, Na et al. have highlighted the application of 3D MXene/ZnSe@NC as anode materials in KIBs. This innovative architecture exhibited enhanced electrochemical performance, maintaining a capacity retention of 238 mA h g−1 at 0.5 A g−1 through 1000 charge–discharge cycles, while delivering 110 mA h g−1 under 2.0 A g−1.269 Concurrently, sodium-ion batteries (SIBs) have garnered as a cost-effective energy storage solution, benefiting from the abundant availability and versatile substitution characteristics of Na+-related materials, establishing it as a leading candidate to replace Li+-based systems in contemporary battery applications. Yang and co-workers successfully prepared Ti3C2-ZIF-67, which was utilized as microbial fuel cell anode material. This engineered interface significantly enhanced extracellular electron transport efficiency through optimized microorganism colonization and enhanced biodiversity preservation at the bio-electrode interface. These findings establish new paradigms for MXene-based nanomaterials in bioelectrochemical energy conversion systems.127 Shi and co-workers reported the CoSe2@C/MXene composite, which, when used as an electrode for SIBs, exhibited high cyclic stability and rate performance.265 Subsequently, the CoFe2O4@carbon@alkalized MXene composite was found to reduce the diffusion pathways for both Li+ and Na+. As a negative electrode material for batteries, it demonstrated superior rate capability and high capacity retention, achieving a satisfactory rate performance of 149 mA h g−1 at 5 A g−1 and a discharge capacity of 130 mA h g−1 at 1 A g−1 even after 1000 cycles for SIBs.240 In parallel, the Co3C/MXene@C electrodes were constructed, exhibiting exceptional energy storage characteristics including enhanced specific capacity and remarkable electrochemical activity. The optimized architecture enabled stable charge storage with capacity retention of 236 mA h g−1 after 100 cycles for KIB, while SIB systems demonstrated 275.9 mA h g−1 under identical cycling conditions. Particularly noteworthy were the high-rate capabilities, with capacities maintaining 80.6 mA h g−1 for KIBs and 105 mA h g−1 for SIBs at 2000 mA g−1.312 In complementary research, Liu et al. developed a hierarchically structured CoP-NC@Ti3C2Tx composite. Electrochemical evaluation revealed that this sandwich-type architecture achieved outstanding rate performance (147.8 mA h g−1 at 2000 mA g−1) along with unprecedented cycling stability, sustaining capacity retention exceeding 100% through 2000 continuous charge–discharge cycles.255 Recently, Sun et al.314 developed zinc ion batteries (ZIBs) based on 3D cross-linked VOX/V2CTX-MXene-x composite, the device delivers specific capacities of 336.39 mA h g−1 and 213.06 mA h g−1 at 1 A g−1 and 10 A g−1 with excellent rate performance, and an energy density of 356.27 Wh kg−1 at 1280 W kg−1.

3.2. Supercapacitors

SCs have emerged as promising candidates for next-generation energy storage devices due to their exceptional cycle longevity, fast charge/discharge capabilities, and sustainable characteristics. These distinctive advantages enable SCs to outperform conventional energy storage technologies including dielectric capacitors, fuel cells, and rechargeable batteries.139 The implementation of MOF materials leverages their unparalleled structural versatility, adjustable nanoscale porosity, and expansive interfacial areas,70 while MXene demonstraties particular efficacy as an electrochemical substrate through its remarkable mechanical resilience and densely populated reactive sites.152 Therefore, the unique properties of MOF/MXene composites are crucial for SCs.315,316 A comprehensive compilation of recent advancements in MOF/MXene composite research specific to SCs has been systematically organized in Table 2.
Table 2 MOF/MXene-based materials for SCs
MOF/MXene-based materials Surface area Current density Capacitance Cycling performance (capacitance retention, cycle, current density) Electrolyte Ref.
a Electrochemical properties measured in a two-electrode configuration, others in three-electrode configuration.
Pure MOF/MXene
MIL-100(Fe)/Ti3C2Tx 529.49 m2 g−1 0.5 A g−1 962.17 F g−1 93%, 10[thin space (1/6-em)]000, 0.5 A g−1 6 M KOH 134
Co-MOF/Ti3C2Tx 14.012 m2 g−1 3 mA cm−2 18.7 F cm−2 92.1%, 3000, 6 mA cm−2 3 M KOH 139
Ti3C2Tx/Ni-MOF 96.6 m2 g−1 1 A g −1 1124 F g−1 83.2%, 4000, 10 A g−1 6 M KOH 148
MXene/Ni-MOF 184 m2 g−1 0.5 A g−1 979 F g−1 98%, 5000, 3 mA cm−2a 3 M KOH 70
Ni-ZIF-67-200/K-Ar-MXene 549.78 m2 g−1 0.5 A g−1 557C g−1 66%, 5000, 2 A g−1 6 M KOH 150
Ni-MOF/Ti3C2Tx 167.74 m2 g−1 0.5 A g−1 497.6 F g−1 85%, 1000, 0.5 A g−1 6 M KOH 152
Ni-BDC/V2CTx/NF 14.34 m2 g−1 1 A g−1 2453 F g−1 83.84%, 3000, 1 A g−1 1 M KOH 151
MXene@Ni-HHTP 0.5 A g−1 416.6 F g−1 102.13%, 3000, 3 A g−1 3 M KOH 156
Ni-MOF@MXene 1 A g−1 1160.5 F g−1 52.2%, 10[thin space (1/6-em)]000, 10 A g−1 3 M KOH 157
Ni-MOF/MXene 1 A g−1 716.19 F g−1 74.22%, 2000, 5 A g−1a 1 M KOH 158
MXene/ZIF-8 0.5 A g−1 1279.5 F g−1 90.88%, 10[thin space (1/6-em)]000, 5 A g−1 7 M KOH 168
Ti-MOF@Ti3C2Tx 85 m2 g−1 0.2 A g−1 154 F g−1 84%, 5000, 1 A g−1 1 M KOH 203
Ni/Co-MOF@Ti3C2Tx-NH2 129.22 m2 g−1 0.5 A g−1 1924 F g−1 58.13%, 10[thin space (1/6-em)]000, 10 A g−1 3 M KOH 186
NiCo-MOF/MXene 38.2 m2 g−1 1 A g−1 1176.8 F g−1 72.5%, 5000, 10 A g−1 2 M KOH 188
MXene/Ni,Co-MOF 11 m2 g−1 1 A g−1 58 mA h g−1 89%, 10[thin space (1/6-em)]000, 1 A g−1 1 M KOH 189
MXene-FeCu-MOF 3 mA cm−2 440 mA h g−1 93%, 5000, 15 mA cm−2 2 M KOH 193
Meso-CuZn-MOFs@Ti3C2Tx 72 m2 g−1 0.2 A g−1 404 F g−1 88.7%, 5000, 1 A g−1a 3 M KOH 196
MnNi-MOF-MXene 251.77 m2 g−1 1 A g−1 1028C g−1 68.9%, 10[thin space (1/6-em)]000, 10 A g−1 2 M KOH 198
[thin space (1/6-em)]
MOF composite/MXene
Ti3C2Tx/ZIF-67/CoV2O6 49.466 m2 g−1 1 A g−1 285.5 F g−1 94.4%, 4000, 3 A g−1 3 M KOH 120
CA@MIL-101(Cr)/MXene 208 m2 g−1 1.5 A g−1 2720 F g−1 84%, 10[thin space (1/6-em)]000, — 219
[thin space (1/6-em)]
MOF derivative/MXene
Co3O4@MXene 76.14 m2 g−1 1 A g−1 732.5 F g−1 83%, 5000, 7 A g−1 6 M KOH 317
CoFe2O4/Ti3C2Tx 0.2 mA cm−2 2467.6 F cm−3 88.2%, 10[thin space (1/6-em)]000, 1.5 mA cm−2a 1 M LiCl 238
MnO2@Co3O4-PC@MXene-CNF 230.24 m2 g−1 1 A g−1 475.4 mA h g−1 92.08%, 10[thin space (1/6-em)]000, 2 A g−1 2 M KOH 245
Co9S8@MXene@Bi2O3 73.57 m2 g−1 1 A g−1 646.1 F g−1 86.8%, 7000, 2 A g−1 6 M KOH 251
MXene@NiCo2S4 53.92 m2 g−1 1 A g−1 451.12 mA h g−1 90.68%, 10[thin space (1/6-em)]000, 5 A g−1a 2 M KOH 273
CoNi2S4/MXene/NF 1 A g−1 933C g−1 80.5%, 10[thin space (1/6-em)]000, 10 A g−1a 3 M KOH 318
MXene-CoS 399 m2 g−1 3 mA cm−2 447 mA h g−1 94%, 5000, 30 mA cm−2 2 M KOH 275
MXene@CoSe2/Ni3Se4 740.93 m2 g−1 1 A g−1 283 mA h g−1 75%, 10[thin space (1/6-em)]000, 5 A g−1 3 M KOH 263
GO/MXene@NiZrP 48.89 m2 g−1 1 A g−1 2358 F g−1 87.5%, 10[thin space (1/6-em)]000, 10 A g−1 6 M KOH 260
MXene@PCNF 405.59 m2 g−1 1 A g−1 572.7 F g−1 96.4%, 10[thin space (1/6-em)]000, 2 A g−1 3 M KOH 319
MXene@cobalt hydroxide 85.10 m2 g−1 1 A g−1 348.55 F g−1 3 M KOH 320


Early studies by Jia et al. engineered MIL-100(Fe)/Ti3C2Tx MXene electrodes for SCs. The nanospheres, embedded within MXene nanosheets, were able to store more ionic charge. The composite electrode exhibited a maximum energy density of 85.53 Wh kg−1 coupled with 962.17 F g−1 specific capacitance at 0.5 A g−1.134 More recently, Ramachandran et al. successfully synthesized Co-MOF/Ti3C2Tx, which exhibited an ultra-high areal capacitance of 18.7 F cm−2 and demonstrated good reversibility. This material was proposed as a novel binder-free electrode material for SCs. Fig. 16a presents the CV response of Co-MOF/Ti3C2Tx@Ni at various scan rate. The cycling stability of Co-MOF@Ni and Co-MOF/Ti3C2Tx@Ni was comparatively analyzed through galvanostatic charge–discharge (GCD) measurements at 6 mA cm−2, as illustrated in Fig. 16b. Notably, the Co-MOF/Ti3C2Tx@Ni hybrid demonstrated enhanced electrochemical durability with 92.1% capacitance preservation after 3000 cycles.139 Yang et al. fabricated MOF/MXene/NF-300 as electrode, which demonstrated superior performance (1 A g−1, 2453 F g−1). Notably, the prepared MOF/MXene/NF-300 electrode material showed exceptional electrochemical durability, maintaining 83.84% capacitance retention through 3000 cycles.151 Zhang et al. constructed a Ti3C2Tx/Ni-MOF-based SC achieving remarkable energy storage performance, with specific capacitance reaching 1124 F g−1 at 1 A g−1(Fig. 16c). This configuration maintained 62% capacity retention at 20 A g−1.148 Further advancements were reported by Zheng et al. through MXene@Ni-MOF composites synthesis, attaining 979 F g−1 specific capacitance at 0.5 A g−1 with merely 2% capacity degradation after 5000 cycles. The electrochemical charge storage mechanism of this hybrid material is visually detailed in Fig. 16e.70 Olatoye and co-workers introduced a novel Ni-ZIF-67 (NZ) composite architecture, followed by the development of NZ-R-2-200/K-Ar-MXene through surface modification of K-Ar-MXene with NZ-R-2-200 NPs. The fabricated asymmetric supercapacitor, employing NZ-R-2-200/K-Ar-MXene as positive electrode in combination with NPC/rGO negative electrode, demonstrated exceptional electrochemical performance. This configuration achieved specific energy and power values of 27.48 Wh Kg−1 and 400 W kg−1 respectively, while maintaining operational stability through 2000 cycles.150 Fig. 16f depicts an all-solid-state asymmetric flexible SC assembled using a Ni-MOF/Ti3C2Tx nanocomposite as the cathode and activated carbon (AC) as the anode.152 The Ni-MOF//MXAC composite was electrochemically characterized in a 2 M KOH electrolyte under a two-electrode configuration. CV analysis at 30 mV s−1 revealed capacitive performance of 536 F g−1 (1 A g−1, 0–0.4 V) and 298 F g−1 (−1–0 V) for this electrode system.153 In comparative studies, MXene/ZIF-8-4 hydrogel-based asymmetric supercapacitors achieved notable energy storage characteristics with 354.7 F g−1 at 0.5 A g−1 and 57.84 Wh kg−1 energy density at 0.628 kW kg−1.168 Another system, Ni/Co-MOF@MXene-NH2//AC configuration demonstrated enhanced energy storage metrics in three-electrode testing, delivering 1924 F g−1 at 0.5 A g−1 with sustained 98.1 Wh kg−1 at 600 W kg−1, alongside exceptional cycling stability through 15[thin space (1/6-em)]600 cycles.186 The NiCo-MOF/MXene system exhibited superior charge storage capacity (1176.8 F g−1 at 1 A g−1) combined with 72.5% capacity retention after extended cycling, confirming its structural durability.188 MXene-FeCu-MOF//AC hybrids displayed remarkable electrochemical persistence, maintaining 89% initial capacity (440 mA h g−1) through 10[thin space (1/6-em)]000 operational cycles.193


image file: d5ra04436g-f16.tif
Fig. 16 (a) CV curves of Co-MOF/Ti3C2Tx@NF at different scan rates. (b) Capacitance retention of Co-MOF@NF and Co-MOF/Ti3C2Tx@NF. (c) Rate performance of Ti3C2Tx/Ni-MOF. (d) GCD curves of MXene@Ni-MOF. (e) The charge/discharge mechanism of MXene@Ni-MOF. (f) Schematic illustration of flexible SC and the CV curves of Ni-MOF/Ti3C2Tx//AC device. (g) Schematic of electrolyte/electron transport in Ti3C2Tx, Ti3C2Tx/ZIF-67 and hollow Ti3C2Tx/ZIF-67/CoV2O6 electrodes. (a and b) Reproduced with permission.139 Copyright 2018, Elsevier. (c) Reproduced with permission.148 Copyright 2021, Elsevier. (d and e) Reproduced with permission.70 Copyright 2022, Elsevier. (f) Reproduced with permission.152 Copyright 2023, Multidisciplinary Digital Publishing Institute. (g) Reproduced with permission.120 Copyright 2023, Wiley.

In the realm of MOF composites/MXene, Liu et al. strategically engineered 3D hollow-structured Ti3C2Tx/ZIF-67/CoV2O6 hybrids through the synergistic integration of ZIF-67 and Ti3C2Tx/ZIF-67. Electrochemical evaluations revealed that the optimized composite electrode exhibited a specific capacitance of 285.5 F g−1, a significant increase compared to that of the ZIF-67 and Ti3C2Tx/ZIF-67 electrode. The schematic representation in Fig. 16g illustrates electron/electrolyte transport pathways within Ti3C2Tx-based electrodes.120

In the field of MOF derivatives/MXene composites, Xie and colleagues fabricated a porous Co–Fe oxide nanorod derived from Ti3C2Tx-supported MOF. This innovative design enabled the fabrication of fully pseudocapacitive flexible supercapacitors with enhanced energy storage characteristics. The optimized composite exhibited outstanding electrochemical performance in 1 M LiCl electrolyte, achieving a high volumetric capacitance of 2467.6 F cm−3. The flexible energy storage device maintained 88.2% initial capacitance after 10[thin space (1/6-em)]000 cyclic operations, demonstrating remarkable durability.238 Kshetri et al. engineered multifunctional composite electrodes through an innovative integration strategy: capacitive-type Co-PC@MXene-CNF paired with battery-type MnO2@Co3O4-PC@MXene-CNF. This configuration enabled their flexible wearable hybrid SC to achieve 72.5 Wh kg−1 energy density at 832.4 W kg−1, while demonstrating exceptional cyclic stability with >90.36% capacity preservation.245 In a parallel advancement, Wu et al. synthesized a Ti3C2@Co3O4/ZnO composite. When paired with AC cathodes in Li-ion hybrid capacitors, the system exhibited remarkable energy storage characteristics: 196.8 Wh kg−1 at 174.9 W kg−1, sustaining 87.5Wh kg−1 even under high-power conditions (3500 W kg−1). The device maintained 75% capacitance retention after 6000 cycles at 2 A g−1, confirming its electrochemical durability.246 Wang et al. synthesized Co9S8-2@MXene@Bi2O3 composite, which exhibited long-term cyclic stability and high discharge capacitance (646.1 F g−1 at 1 A g−1), these enhanced electrochemical properties were attributed to the unique core–shell architecture and optimized interfacial interactions between MXene and Bi2O3.251 In contrast, Yang et al. fabricated a hierarchically structured MXene@CoSe2/Ni3Se4 with honeycomb morphology, delivering 283 m Ah g−1 capacity at 1 A g−1 while maintaining 80% initial capacity after 5000 cycles, thereby establishing new benchmarks for SC energy storage systems.263 Meanwhile, MXene-CoS hybrids derived from MOFs have emerged as promising alternatives, leveraging their mesoporous structure (447 mA h g−1 at 3 mA cm−2) with enhanced ion transport pathways and surface reactivity.275 Most recently, Hussain et al.321 reported a novel NH4+ hybrid SC, assembled by Co-MOF@MXene carbonized nanofibers (MXCNF) cathode constructed by Co-MOF nanoflowers hydrothermally anchored on MXCNF through integrated electrospinning-carbonization synthesis, and a MXCNF anode. The architecture leveraged vertically aligned MXene cores for enhanced electrical conduction, complemented by homogeneously dispersed Co-MOF surface modifications that optimized redox-active surface utilization. This configuration achieved remarkable electrochemical performance, delivering 41.5 mW h kg−1 at 800 mW kg−1 power output while maintaining 87% cyclic stability over 16[thin space (1/6-em)]000 cycles.

3.3. Sensors

MOF/MXene-based materials are extensively utilized in the realm of sensing.322,323 Table 3 provides a comprehensive summary of MOF/MXene-based materials for sensors.
Table 3 MOF/MXene-based materials for sensorsa
MOF/MXene-based materials Detection Electrode Linear ranges Detection limits Selectivity Sensitivity Ref.
a AA: ascorbic acid; DA: dopamine; UA: uric acid; Glu: glucose; 5-AVA: 5-aminovaleric acid; L-Trp: L-tryptophan; L-Cys: L-cysteine; GPNMB: glycoprotein nonmetastatic melanoma protein B; GSH: glutathione; IgG: human serum immunoglobulin G; L-His: L-histidine; TMAO: trimethylamine oxide; L-Arg: L-Arginine; L-Leu: L-leucine; L-Phe: L-phenylalanine; L-Met: L-methionine; L-Asp: L-aspartate; HIV: human immunodeficiency virus; DMF: N,N-dimethylformamide; BSA: bovine serum albumin; PSA: prostate-specific antigen; AFP: alpha-fetoprotein; CEA: carcinoembryonic antigen.
Pure MOF/MXene
Cu-MOF/MXene H2O2 GCE 1–6120 μM 0.35 μM AA, DA, UA, Glu 129
MOF-Ti3C2 DA GCE 90–300 nM 110 nM AA, 5-AVA 130
Fe-MOF/MXene As(III) GCE 1–10 ng L−1 0.58 ng L−1 8.94 μA ng−1 L1 cm−2 133
MXene@MOF GPNMB 180.33 pg mL−1 AA, GSH, IgG, Glu, L-His 165
MXene/ZIF-8 N2H4 GCE 10–7700 μM 5.1 μM NO3, SO42−, K+, Cl, DMF, ethanol, acrylic acid, 2-propanol 169
Ti3C2Tx/ZIF-8 HIV protein GCE 1 fM–1 nM 0.3 fM BSA, PSA, AFP, CEA 170
Ce-MOF/Ti3C2Tx L-Trp GCE 0.2–139 μM 0.19 μM AA, DA, UA, L-Cys, Glu, KCl, NaCl 0.1556 μA μM−1 cm−2 200
MXene-NH2@CeFe-MOF-NH2 Pb2+, Hg2+ GCE 5–50 nM, 1–35 nM 0.95 nM, 0.32 nM Na+, K+, Ca2+, Fe3+, Ni2+, Mn7+, Mg2+, Al3+, Cl, SO42− —, — 322
[thin space (1/6-em)]
MOF composite/MXene
Co-MOF-MXene@carbon black Blood creatinine 10–800 μM 0.005 μM UA, urea, DA, GSH 1.1 μA μM−1 cm−2 144
MXene/CNHs/β-CD-MOFs Carbendazim GCE 0.003–10.0 μM 1.0 nM Cl, SO42−, Na+, Cu2+, Al3+, UA, AA, Glu, malathion, thiabendazole, chlorophene, fenitrothion 5.253 A mol−1 L 215
Eu-MOF/PDA@MXene/Fe3O4 TMAO Magnetic GCE 9 × 10−18–9 × 10−15 M 1.25 × 10−18 M Urea, Plasma 63.73 μA M−1 cm2 223
MXene@PDA/MOF L-Cys GCE 0.01–5 (10−6 M) 0.00374 (10−6 M) L-Arg, L-Leu, L-Trp, L-Phe, L-Met, L-Asp, AA, UA 224
rGO/PDA@MXene/FeCu-MOF Ribavirin GCE 0.053 nM Lamivudine, sulfadiazine, inosine, uridine, DA, GSH, Glu 1.43 μA nM−1 cm−2 230
[thin space (1/6-em)]
MOF derivative/MXene
In2O3/Ti3C2Tx MXene NH3 Au 5–100 ppm 5 ppm HCHO, CH3OH, CH3COCH3, C2H5OH, C6H6, CO, NO2 242
In2O3/ZnO/Ti3C2TX MXene Ethanol gas 2.15–10 ppm C2H5OH, CH3OH, C3H6O, C3H9N, HCHO, NH3, C6H15N, NO2 243
Co3O4/Ti3C2Tx MXene Ethanol gas 1 ppm CH3OH, C3H8O, C3H6O, NH3 247
MnO2/Mn3O4@Ti3C2 MXene/Au NPs Methamidophos GCE 10−12–10−6 M 1.34 × 10−13 M Glu, citric acid, NO3, SO42−, PO43−, Cu2+, Fe3+, Zn2+, Pb2+ 253
MXene-Cu/Cu2O/C Glu Au 3–10 mM 1.70 mM Lactose, fructose, sucrose, AA, urea, NaCl, KCl 279
TiO2/Ni-NC Hg2+ GCE 0.001–10 μM 0.79 nM Ca2+, Pb2+, Mn2+, Zn2+, Cd2+, Cu2+ 324


A novel electrochemical approach for As(III) monitoring was developed by Xiao et al., utilizing a Fe-MOF/MXene composite-modified GCE. Fig. 17a illustrates the operational mechanism of adsorption-enhanced As(III) quantification through square wave anodic stripping voltammetry (SWASV). Experimental results indicate that the synergistic interaction between Fe-MOF nanostructures and MXene substrates significantly enhances both the preconcentration efficiency of As(III) and the electrochemical detection sensitivity, outperforming the individual components. This sensor has been successfully applied to the determination of As(III) in real water samples. Fig. 17b presents the SWASV responses of the Fe-MOF/MXene-1.5/GCE in 0.1 M PBS electrolyte containing progressively increasing concentrations of As(III). A linear relationship between electrochemical signals and analyte levels is demonstrated in the inset through current-concentration calibration curves.133 Wang et al. successfully applied TiO2/Ni-NC in detecting Hg2+, demonstrating good catalytic activity for the electrochemical reaction of Hg2+.324 Subsequently, Cheng et al. constructed a H2O2-sensitive electrochemical platform based on Cu-MOF/MXene/GCE. Chronoamperometric analysis at −0.35 V revealed a wide detection range (1 μM–6.12 mM). The sensor displayed ultrahigh sensitivity, rapid response kinetics, and straightforward operational design, highlighting its strong potential for practical diagnostic applications.129 Parallel research efforts addressed DA sensing requirements through innovative nanocomposite design. Paul and co-workers engineered a MOF-Ti3C2 hybrid system combining Zn-based MOF ([Zn4(1,2,4,5-benzene-tetra carboxylate)2(H2O)6]n·3nH2O) with MXene substrates for redox-active DA quantification. Optimized conditions established linear current-concentration correlations (90–300 nM) via differential pulse voltammetry, achieving detection thresholds between 90–130 nM through enhanced anodic peak amplification.130 MOF/MXene-based composites have further expanded biosensing applications: Ti3C2Tx/ZIF-8 composites enabled hydrazine monitoring169 and HIV-1 protein,170 with mechanistic illustrations provided in Fig. 17c. Notably, a Ce-MOF/MXene assembly was developed as a selective L-Trp detector, maintaining stable performance across 10–50 μM concentrations (Fig. 17d).200 Furthermore, Ravipati et al. engineered an innovative MOF-71/V2C MXene hydrogel using solvothermal synthesis followed by lyophilization, establishing this nanocomposite as a precision biosensor for dual-analyte quantification of levothyroxine (LT4) and carbamazepine (CBZ).325 The developed sensor demonstrated exceptional analytical performance with extended linear response ranges spanning 10 nM to 100 μM (LT4) and 10 nM to 500 μM (CBZ), coupled with excellent selectivity and remarkably low detection thresholds of 5.6 nM (LT4) and 6.7 nM (CBZ). Owing to its intrinsic adaptability, the hydrogel showed applicability in flexible electrochemical wearables, creating new opportunities for biomedical sensor development.


image file: d5ra04436g-f17.tif
Fig. 17 (a) The schematic illustration of the adsorption-assisted As(III) detection on Fe-MOF/MXene/GCE by SWASV. (b) SWASV curves of the Fe-MOF/MXene/GCE in 0.1 M PBS solution with different As(III) concentrations. (c) Schematic illustration of N2H4 sensing. (d) differential pulse voltammetry curves of the Ce-MOF/Ti3C2Tx MXene/GCE. (e) Schematic of the single electrode Ecoffex@MXene@Co-NPC nanocomposite based TENG. (f) The long-term stability of self-powered MXene/CuO sensor. (g) Schematic diagram for the possible mechanism of In2O3/Ti3C2Tx MXene composites toward NH3. (h) Schematic diagram of the gas sensing mechanism of In2O3/ZnO/Ti3C2Tx MXene-based sensor. (a and b) Reproduced with permission.133 Copyright 2021, Elsevier. (c) Reproduced with permission.169 Copyright 2021, Wiley. (d) Reproduced with permission.200 Copyright 2022, Elsevier. (e) Reproduced with permission.216 Copyright 2022, Elsevier. (f) Reproduced with permission.252 Copyright 2021, American Chemical Society. (g) Reproduced with permission.242 Copyright 2022, Elsevier. (h) Reproduced with permission.243 Copyright 2023, Elsevier.

Recent advances in MOF composite/MXene architectures have demonstrated remarkable versatility in self-powered biomotion and tactile sensing applications. A prominent example is the Ecoflex@Co-NPC@MXene developed by Rana et al., which has been successful implementation in self-powered biodynamic detection systems.216 This innovative configuration employs a hierarchical nanocomposite as a charge-enhancement layer in TENG, achieving a fourfold amplification in power generation efficiency compared to conventional designs. The structural architecture of this single-electrode TENG system is schematically illustrated in Fig. 17e. In the field of pathogen detection, the Zr-Fc MOF/AuNPs/4-MPBA sensing platform has emerged as a promising candidate for the ultrasensitive detection of E. coli O157:H7.209 The engineered MOF matrix capitalizes on its expansive surface topology to facilitate 4-MPBA through boronic acid-cis-diol interactions with bacterial lipopolysaccharides. This interaction enables signal labels to be fixed on the electrodes, forming a faradaic cage-type structure. The resulting sensor exhibits high sensitivity, excellent selectivity, and robust stability, highlighting its potential for real-world diagnostic applications. Given the significant health risks posed by pesticide residues, there is an urgent need to develop a sensitive and rapid method for pesticide monitoring. Addressing this critical demand for agrochemical monitoring, Tu et al.215 engineered a nanoengineered electrochemical sensing platform based on MXene/CNHs/β-CDMOFs for sensitive detection of carbendazim pesticide. The synergistic action of MXene/CNHs and β-CD-MOFs extends the linear range of the MXene/CNHs/β-CD-MOFs electrode from 3.0 nM to 10.0 μM, with a low detection limit of 1.0 nM (S/N = 3). The sensor demonstrates outstanding selectivity against common interferents, excellent reproducibility, and long-term operational stability, underscoring its practical applicability for on-site and routine monitoring of pesticide contaminants.

In the field of MOF derivatives integrated with MXene architectures, Song et al. developed an innovative biosensing system by combining MnO2/Mn3O4 with MXene/Au NPs hybrids for the detection of organophosphorus pesticides. This electrochemical platform exhibited exceptional analytical performance, demonstrating a broad detection window spanning six orders of magnitude (10−12 M to 10−6 M) and an ultralow detection threshold of 1.34 × 10−13 M. The superior sensing characteristics stem from the collaborative interaction between the MnO2/Mn3O4 and MXene/Au NPs composites, which collectively establish extensive active surfaces for electrochemical processes while preserving acetylcholinesterase functionality through favorable biocompatibility. These attributes position the platform as a highly promising tool for environmental monitoring and agricultural safety evaluation.253 Regarding MXene/CuO, an organ-like gas sensor was developed for the detection of NH3. The TENG-driven self-powered NH3 sensor exhibited a strong response at room temperature (Vg/Va = 24.8@100 ppm) and showed promise in monitoring pork quality. As depicted in Fig. 17f, the MXene/CuO hybrid sensor maintained consistent performance over time, confirming its operational durability.252 Similarly, Liu et al. reported MOF-derived In2O3/Ti3C2Tx MXene composites that achieved efficient NH3 detection with strong linear correlation (R2 = 0.9785) across 5–100 ppm concentrations. The architecture demonstrated rapid response and recovery dynamics (3/2 s), with Fig. 17g visually summarizing the proposed room-temperature sensing mechanism of these engineered nanomaterials.242 More recently, Co3O4/Ti3C2Tx nanocomposites with a mesoporous structure were synthesized by Bu and co-workers for ethanol sensing applications. The resulting sensor displayed excellent sensitivity, fast response and recovery kinetics, a low detection limit, high selectivity, and robust long-term stability, making it suitable for practical gas-sensing applications.247 Additionally, In2O3/ZnO/Ti3C2Tx MXene nanocomposites have been identified as a promising sensing material for ethanol detection, as depicted in Fig. 17h. These nanocomposites demonstrate a strong response to ethanol concentrations ranging from 2.15 to 10 ppm, along with outstanding selectivity towards ethanol gas.243 Furthermore, a novel wearable, stretchable multifunctional double-layer TENG has been developed, which combines the superior nanoporosity of NPCO with the charge capture and transport characteristics of MXene/silicone. This TENG functions as a high-performance self-powered biosensor and delivers a high power density of 10.4 W m−2.250

3.4. Catalysts

In light of rapid societal advancement, environmental pollution and escalating energy consumption have emerged as critical challenges threatening sustainable human development. Consequently, There is a growing demand for renewable energy and green energy sources within society.326,327 In response, an increasing number of researchers are dedicated to identifying viable, sustainable, and pollution-free solutions to address these urgent global issues. Among these efforts, catalysts-serving as the core tool of green chemistry, are becoming the key to break through the technical bottlenecks.297,328 This section systematically examines recent advancements in MXene/MOF composite catalysts and their electrocatalytic applications for energy conversion systems.

OER, HER, and ORR are three crucial reactions in the pursuit of sustainable energy.329,330 Among these, OER serves as a fundamental process in numerous renewable energy systems. Consequently, the search for cost-effective and efficient electrocatalysts to overcome the slow kinetics and high overpotential associated with OER is paramount.331 MOF materials have emerged as particularly attractive catalytic candidates owing to their customizable chemical properties, varied architectural features, and expansive active surfaces, positioning them at the forefront of catalyst innovation. Nevertheless, intrinsic limitations such as poor electrical conductivity and structural instability under operational conditions restrict their standalone performance. To address these challenges, researchers have strategically integrated MOFs with conductive MXene nanostructures, leveraging the excellent electron transport properties of MXenes to significantly enhance the overall electrocatalytic activity of the hybrid systems.332,333 Notable examples include CoNi-ZIF-67@MXene and Co2Ni-MOF@MXene, representative MOF/MXene composites developed for OER applications. Wen et al. fabricated a Ti3C2Tx MXene-supported CoNi-ZIF-67 composite (CoNi-ZIF-67@Ti3C2Tx), demonstrating enhanced OER performance relative to pristine CoNi-ZIF-67, specifically achieving 275 mV onset potential versus RHE and a Tafel slope of 65.1 mV dec−1. Comparative electrochemical metrics at 10 mA cm−2 are visually contrasted in Fig. 18a.132 Furthermore, Tan et al. developed a series of 2D MOF/MXene composites, in which the Co2Ni-MOF layer is tightly anchored onto the MXene nanosheet, and the well-defined interface between the two layers facilitates the accelerated charge transfer kinetics during electrocatalytic operation. The OER performances of the Co2Ni-MOF@MXene are depicted in Fig. 18b, while the Tafel curves of Co2Ni-MOF@MXene are presented in Fig. 18c. Among the series, Co2Ni-MOF@MXene-1 demonstrated superior electrocatalytic activity compared to other evaluated materials, achieving both the smallest Tafel slope (51.7 mV dec−1) and minimal overpotential requirement (265 mV at 10 mA cm−2). These results indicate that electrostatically driven directional assembly is a viable and effective strategy for fabricating high-performance 2D OER catalysts.187 In the realm of MOF derivatives/MXene, Zou et al. synthesized NiCoS/Ti3C2Tx hybrid demonstrating enhanced electrocatalytic performance. This improvement was attributed to the strong synergistic coupling between NiCoS and Ti3C2Tx, which facilitated improved charge carrier mobility and optimized electron transport pathways.272 Conversely, Du et al. investigated the influence of MXene incorporation on the OER activity of CoNi-MOF@MXene nanocomposites and observed an unexpected decrease in catalytic efficiency. This behavior was mechanistically attributed to the electron-donating nature of Ti3C2Tx MXene, which may alter the electronic structure of the active metal sites and suppress oxidative processes.183 Furthermore, Hu's group successfully employed CoFe-LDH/Ti3C2 as potential electrocatalysts for OER, achieving remarkable catalytic efficiency manifested by ultra-low overpotentials of 170 mV (10 mA cm−2) and 238 mV (100 mA cm−2).277 Fig. 18d presents free energy diagrams at 0 V for the OER on the surface of individual CoFe-LDH and CoFe-LDH/MXene composite. Additionally, MXene@MOF-Co2P composite material with engineered porosity demonstrated superior electrocatalytic behavior, requiring only 246 mV at 10 mA−2 with accelerated kinetics evidenced by a 28.18 mV dec−1 Tafel slope. This performance enhancement arises from three synergistic effects: (1) heteroatom (C, N) doping-induced electronic modulation, (2) hierarchical pore networks preserved from MOF precursors, and (3) interfacial charge transfer between Co2P and MXene components.257 HER significantly contributes to carbon mitigation strategies and net-zero emissions objectives. For instance, Zong et al. synthesized the MXene@MOF-CoP composite as a bifunctional catalyst. This innovative electrode configuration achieved notable hydrogen generation efficiency in diverse pH environments, particularly demonstrating 112 mV overpotential at pH = 14. The bifunctional activity based on MXene@MOF-CoP under alkaline conditions had a bias of 1.61 V at 10 mA cm−2. The catalyst prepared using the MXene@MOF-CoP electrode exhibited excellent catalytic stability.256 Similarly, Ti3C2Tx@ZIF-8 catalyst developed by Hao et al. emerges as a viable alternative to expensive Pt-based catalysts, owing to its structural adaptability, enhanced conductivity, and expanded surface characteristics. CV measurements indicate an electrochemical active surface area of 122.5 cm2 for this material, implying sustained operational stability through extended testing periods.175 In energy conversion systems, ORR is a critical electrochemical process. Research by Parse et al. highlights the exceptional ORR catalytic behavior of Co-NC/Ti3C2-800 composite, achieving an onset potential (Eonset) of 1.04 V with current density reaching 4.8 mA cm−2 and half-wave potential (E1/2) at 0.93 V. This performance enhancement originates from synergistic effects between the Co REDOX center and the NC matrices that form conformal coatings on Ti3C2Tx skeleton, establishing optimized electron transport pathways during redox processes. Concurrent surface oxidation generating TiO2 on MXene substrates further improves reaction kinetics by facilitating mass transport, ultimately boosting catalytic efficiency.232 A schematic illustration of the oxygen electroreduction mechanism on the Co-NC/Ti3C2-800 composite is provided in Fig. 18f, while Fig. 18g compares the CV responses of various catalysts in N2-saturated electrolyte.


image file: d5ra04436g-f18.tif
Fig. 18 (a) A comparison of the CoNi-ZIF-67@Ti3C2Tx and CoNi-ZIF-67 catalysts in the onset potential and overpotential at 10 mA cm−2. (b) LSV curves and (c) Tafel plots of Co2Ni-MOF and Co2Ni-MOF@MXene. (d) Free energy diagrams at 0 V for OER on the surface of individual CoFe LDH and CoFe LDH/Ti3C2O2. (e) A schematic illustration of degradation of Ti3C2Tx over Ti3C2Tx/MIL-53(Fe) composite under visible light irradiation. (f) A schematic illustration of electroreduction of oxygen over Co-NC/Ti3C2-800. (g) Comparative CV curves for the Co-NC/Ti3C2 composites. (h) Variation of open circuit potential for MXene-CoO@Co–N–C, ZIF, and MXene catalysts. (a) Reproduced with permission.132 Copyright 2019, Multidisciplinary Digital Publishing Institute. (b and c) Reproduced with permission.187 Copyright 2022, Elsevier. (d) Reproduced with permission.277 Copyright 2021, Elsevier. (e) Reproduced with permission.126 Copyright 2020, Elsevier. (f and g) Reproduced with permission.232 Copyright 2021, American Chemical Society. (h) Reproduced with permission.334 Copyright 2022, Elsevier.

Photocatalysts, which exhibit excellent catalytic abilities under visible light, play a crucial role in the degradation of numerous wastes. Liu et al. developed a novel Ti3C2–Fe composite demonstrating enhanced visible-light photocatalytic performance with exceptional operational stability.126 Fig. 18e schematically illustrates the tetracycline decomposition mechanism over TiCFe-1 under visible light. Subsequently, Qin's group engineered a Ti3C2-QD/Ni-MOF architecture that achieved exceptional N2 reduction efficiency, generating 88.79 μmol gcat−1 h−1 of NH3.146 Addressing the environmental concerns associated with methylene blue (MB), Cheng et al. synthesized Ni-MOF/Ti3C2 composites, which exhibited superior activity for the photocatalytic degradation of MB. When Ti3C2 was loaded with 1.5% wt%, the activity increased nearly fourfold.147 MOF composites with high porosity and tunable bandgap energy (4.99 eV), combined with exceptional surface area characteristics, effectively enhance charge transport kinetics while suppressing photogenerated electron–hole recombination, demonstrating superior dye degradation capabilities in photocatalytic applications.176 Zhu and co-workers developed a MXene/CuPMOF nanohybrid, which achieved efficient hydrogen evolution photocatalysis through optimized interphase connectivity.163 Other examples of the catalytic aspects of MOF/MXene composites include a novel hybrid material synthesized by the hybrid reaction of hust-1 and Ti3C2Tx by Gu's group. The resulting hybrid material exhibited good catalytic performance, with a conversion rate of 76.7% in the ring–opening reaction of styrene's catalytic oxidation, significantly higher than the 23.1% achieved with pure hust-1, with full activity recovery maintained through six catalytic cycles.159 Cheng et al. further engineered 2D/2D Ni-MOF/Ti3C2 heterostructures with conductive interfaces, showing 4-fold photocatalytic activity enhancement compared to pure Ni-MOF at optimal 1.5% wt% Ti3C2 loading.147 In electrocatalysis, Zhan et al. synthesized nitrogen-doped MXene-supported carbon-encapsulated CoP nanoarrays decorated with ultrafine Pt nanoparticles. The synergistic interaction between lattice-incorporated nitrogen, CoP, and Pt effectively mitigated CO poisoning, stabilized Pt nanoparticles, and enhanced the intrinsic activity of Pt active sites, resulting in superior electrocatalytic performance for relevant energy conversion reactions.335

Additionally, Li et al. developed a VCo-MOF@Ti3C2Tx composite catalyst demonstrating remarkable peroxymonosulfate (PMS) activation efficiency, achieving 96.14% ciprofloxacin (CIP, 20 mg L−1) within 30 min using merely 5 mg catalyst. This enhanced performance stems from the material's dual functionality in PMS adsorption and activation processes.192 Furthermore, Guo's group engineered a MXene-CoO@Co–N–C heterostructure with sandwich-like architecture that maintained 100% bisphenol A degradation efficiency under hypersaline conditions (200 mM). The system achieved exceptional catalytic turnover (TOF = 8.64 min−1) at low reagent concentrations (0.05 g per L catalyst + 0.1 g L−1 of PMS), representing a 22.5-fold enhancement over MOF-derived catalyst without MXene. Comparative open-circuit potential variations across catalyst types are illustrated in Fig. 18h.334 This MXene-CoO@Co–N–C catalyst has been successfully employed to remove trace organic contaminants from brackish water using PMS, exhibiting robust decontamination efficacy even under high-salinity environments.334 Additionally, the TS-V2CTx/CoV2O6 HN hybrid demonstrates enhanced electrocatalytic performance in both HER and OER. The MOF-derived CoV2O6 nanocubes possess hollow architectures that optimize electrolyte infiltration while promoting efficient charge transfer kinetics. When configured as a water splitting module, this heterostructure achieves stable current output at 10 mA cm−2 with a low battery voltage of 1.358 V.249

3.5. Electromagnetic/microwave absorption

Growing environmental challenges, particularly pervasive air pollution, have spurred intensive research into advanced adsorbent materials. Parallel demands exist for electromagnetic pollution mitigation, driving the need for lightweight broadband EMW absorbers.336,337 MOF/MXene-based materials play a pivotal role in advancing critical technological applications.338 The results of the latest studies on electromagnetic/microwave absorption are summarized in Table 4.
Table 4 MOF/MXene-based materials for electromagnetic/microwave absorption
MOF/MXene-based materials EABa/GHz Thicknesses/mm RLb/dB Ref.
a EAB: effective absorption bandwidth.b RL: reflection loss.
Electromagnetic absorption
Ti3C2Tx/CNTs/Co 6.10 1.40 −85.80 234
NiSe2–CoSe2@C/Ti3C2Tx 5.68 2.60 −60.46 262
Co/TiO2–C 9.00 2.00 −41.10 285
Fe&TiO2@C 6.50 1.60 −51.80 286
Ti3C2Tx@CoFe@TiO2 5.02 1.40 −62.90 239
Co/Nb2CTx/carbon 4.00 1.67 −60.25 288
CoPC/CNTs@MXene 5.60 1.80 −10.00 258
CNCM-7 6.80 2.70 −65.30 284
[thin space (1/6-em)]
Microwave absorption
MXene/Co-CZIF 50% 10.20 2.70 −60.09 233
MXene/Ni-CZIF 50% 4.56 3.40 −64.11 233
MXene/1D N-CNTs 8.40 1.49 −57.78 282
MXene fibers/CoNi/C and CNTs/CoNi 4.50 1.60 −51.60 184
Ti3C2Tx/CNFs/TiO2/CoNi 4.00 1.76 −54.60 235
Ni/TiO2/C 3.40 1.50 −45.60 287


High-performance EMW absorption materials with multifunctional properties have attracted considerable scientific and technological interest. Nevertheless, significant challenges remain in their development and optimization. Xiang et al. reported a Ti3C2Tx/CNT/Co nanocomposite exhibiting outstanding EMW absorption performance. The enhanced EMW capability originates from synergistic mechanisms, including charge carrier migration, dipolar polarization (both electric and magnetic), interfacial polarization effects, natural resonance phenomena, and optimized multi-path reflection configurations. As predicted, the engineered composite manifests outstanding absorption characteristics with a remarkable reflection loss value reaching −85.8 dB while maintaining an ultrathin profile of merely 1.4 mm.234 Similarly, Hou et al. prepared a device with a maximum RL of −60.46 dB at 2.6 mm using a NiSe2–CoSe2@C Ti3C2Tx composite material with a mass fraction of Ti3C2Tx at 40 wt%. The RL <−10 db can reach 5.68 GHz (10.32–16 GHz). This innovative EMW absorption material derives its enhanced performance from two distinctive attributes: a hierarchically layered architecture and intrinsic metalloid characteristics.262 Additionally, Liao and co-workers synthesized a novel Co/TiO2–C composite material through thermal treatment. This material exhibits superior microwave absorption characteristics, attaining optimal RL of −41.1 dB at 9.0 GHz (3.0 mm thickness) and −31.0 dB at 13.9 GHz (2.0 mm thickness). The operational bandwidth (RL ≤−10 dB) reached 3.04 GHz and 4.04 GHz at these frequencies, respectively. Comparative analysis with pristine MOF-derived carbon composites reveals three primary enhancement mechanisms: (1) multilayered microstructure-induced EMW scattering; (2) interface/dipole polarization effects arising from heterojunctions, structural defects, and carbon matrix interactions; (3) improved dielectric dissipation through optimized electrical conductivity. These attributes render the Co/TiO2–C hybrid composite a promising candidate for electromagnetic absorption applications.285

EMW absorbing materials play a crucial role in military and engineering applications, underscoring the pressing demand for advanced absorption solutions. Lei et al. developed a sandwich-like Ti3C2Tx-based nanocomposite that demonstrated superior EMW absorption properties, attaining −62.9 dB minimum RL at 1.2 mm thickness within 17.95 GHz spectrum. The material demonstrated 5.02 GHz EAB (12.74–17.76 GHz) when fabricated at 1.4 mm depth, highlighting its capability as a thin, broadband absorber for mitigating pervasive electromagnetic interference in modern electronic environments.239 In a related advancement, Deng et al. investigated an innovative 2D hybrid architecture incorporating Fe&TiO2 NPs within carbon matrices. As depicted in Fig. 19a, the composite achieved 6.5 GHz EAB absorption bandwidth at 1.6 mm thickness while demonstrating −51.8 dB peak attenuation at 6.6 GHz with 3 mm configuration, offering new perspectives for MXene-derived nanomaterial synthesis.286 Employing a porous structural engineering approach, Cui et al. synthesized a porous Co/Nb2CTx/CA composite. This composite, characterized by its low density (54.03 mg cm−3), superior reflection loss, and expansive effective absorption bandwidth (reaching −60.25 dB and 4 GHz, respectively) at an ultra-thin thickness of 1.67 mm and an ultra-low filling content of 10 wt%, demonstrated a radar cross section reduction value of 31.24 dB m−2, suggesting promising potential for excellent thermal insulation and flame retardant properties.288 To develop efficient EMW absorbing materials with robust absorption capacity, Wang and colleagues synthesized the CoPC/CNTs@MXene (CCM) absorber through controlled pyrolysis of ZnCo-MOF precursors at 800 °C, followed by electrostatic self-assembly within the Ti3C2Tx matrix. Notably, the CCM-20 composite demonstrated exceptional EMW absorption at 5.56 GHz, achieving a minimum RL of −54.2 dB. More remarkably, the CCM-30 composite variant attained effective microwave attenuation (EAB <−10 dB) at 5.6 GHz with merely 1.8 mm thickness, attributed to its hierarchical 0D/1D/2D structure that enhances electromagnetic energy dissipation through synergistic scattering mechanisms.258


image file: d5ra04436g-f19.tif
Fig. 19 (a) Illustration of EMW absorption mechanisms for Fe&TiO2@C nanocomposites. (b) Schematic illustration of the possible MA mechanism for MXene@Co-CZIF and MXene@Ni-CZIF. (a) Reproduced with permission.286 Copyright 2020, Springer. (b) Reproduced with permission.233 Copyright 2020, American Chemical Society.

High-performance microwave absorbing materials featuring layered and 3D architectures with abundant interfacial structures can significantly enhance EMW absorption efficiency. Such structural designs also contribute to improved microwave absorption performance, particularly by broadening the effective absorption bandwidth. Han et al. fabricated advanced MXene@Co-CZIF and MXene@Ni-CZIF composites (Fig. 19b) through electrostatic self-assembly, achieving exceptional EMW attenuation. The MXene/Co-CZIF 50% variant displayed notable absorption capabilities, attaining a minimum RL of −60.09 dB at 7.36 GHz with 9.3 GHz effective bandwidth (RL <−10 dB). Meanwhile, the MXene/Ni-CZIF 50% counterpart exhibited superior performance metrics, registering a peak RL value of −64.11 dB at 5.12 GHz accompanied by a 4.56 GHz operational bandwidth under equivalent threshold conditions. This methodology provides novel design principles for developing MOF-MXene hybrid absorbers with enhanced functionality.233 Furthermore, Liu et al. successfully synthesized a Ti3C2Tx/CNFs/TiO2/CoNi nanocomposite through electrostatic self-assembly and thermal treatment, demonstrating superior EMW absorption characteristics. In this architecture, Ni-catalyzed CNFs interconnect layered MXene substrates (originally employed for MOF anchoring) to establish an interconnected framework. This structural configuration enhances dielectric loss through the creation of extensive heterogeneous interfaces, while the incorporation of CoNi NPs introduces additional magnetic loss mechanisms, enabling balanced impedance matching and synergistic attenuation. With 40 wt% filler loading, the Ti3C2Tx/CNFs/TiO2/CoNi-70 composite achieves a RL of −54.6 dB and 4.0 GHz EAB at 1.76 mm thickness (Fig. 20a and b). This methodology advances microwave absorption material design through synergistic structural engineering.235 Despite its high electrical conductivity, 2D laminated Ti3C2Tx MXene suffers from an excessively high dielectric constant and a lack of magnetic loss, leading to significant impedance mismatch that restricts its practical application in EMW absorption. To tackle this challenge, Wang et al. successfully synthesized a multilayered Ni/TiO2/C nanocomposite. Remarkable enhancement in EMW absorption properties was observed through precision modulation of calcination parameters. As depicted in Fig. 20c, the optimized specimen treated at 700 °C exhibited exceptional absorption characteristics: a minimum RL of −45.6 dB, an EAB of 3.40 GHz (14.6–18.0 GHz), and ultra-thin 1.5 mm. This study establishes a novel pathway for developing high-performance MXene-derived electromagnetic absorbers through simplified synthesis protocols.287 Like other 2D materials, MXene are prone to severe aggregation and stacking issues that impede their applications. To overcome this issue, Wu et al. assembled a 1D heterostructure without the need for templates or rigid frameworks. The results indicated that the absorber attained a minimum RL of −51.6 dB at a matched thickness of 1.6 mm, with an EAB (RL <−10 dB) extending across 4.5 GHz within the 13.2–17.7 GHz frequency range. These findings advance the methodology for engineering efficient microwave absorption materials.184


image file: d5ra04436g-f20.tif
Fig. 20 (a) Dependence of the minimum RL values on the thickness and (b) EAB of Ti3C2Tx/CNFs/TiO2/CoNi. (c) Schematic illustration of possible EMW absorption mechanisms for the Ni/TiO2/C nanocomposites. (a and b) Reproduced with permission.235 Copyright 2022, Elsevier. (c) Reproduced with permission.287 Copyright 2022, Elsevier.

3.6. Others

MOF/MXene hybrid systems exhibit dual functionalities that extend beyond the previously described chemical interactions, serving as highly effective flame-retardant additives with significant protective capabilities. For instance, Wan et al. employed solvothermal synthesis to fabricate 3D flower-shaped Ni-MOFs while achieving functionalization of multi-terminal layered Ti3C2Tx. The engineered Ni-MOF@Ti3C2Tx integration substantially improved the flame-retardant performance of TPU. Quantitative analysis revealed distinct combustion parameter reductions: TPU composites containing 2% Ni-MOF, 1% Ni-MOF@Ti3C2Tx, and 2% Ni-MOF@Ti3C2Tx exhibited smoke production rate (SPR) reductions of 41.2%, 44.1%, and 35.3% respectively, along with corresponding total smoke production decreases measuring 22.5%, 26.4%, and 23.0% relative to pure TPU. Mechanistic investigations identified three synergistic protective mechanisms: MXene's physical barrier formation, catalytic carbonization capability, and gaseous-phase radical quenching, complemented by MOF-mediated enhancement of MXene dispersion efficiency, thermal resilience, and flame suppression effectiveness.128 To develop an eco-friendly flame-retardant system with multi-component synergy for polymer composites, Gong et al. synthesized a biomimetic MXene@Bi-MOF composite through interfacial engineering and integrated it into epoxy resin (EP). Experimental characterization revealed that the 2 wt% MXene@Bi-MOF modified EP composite exhibited superior fire safety performance compared to pristine EP, demonstrating significant reductions in key combustion parameters: 28.8% peak heat release rate, 45.3% peak smoke production rate, 36.5% total heat release, 30.7% peak CO emission, and 55.3% smoke factor. Additionally, the composite simultaneously achieved decreased char residue carbonization efficiency and reduced fire growth index, indicating enhanced flame inhibition mechanisms. This performance enhancement originates from the complementary interactions between MXene's catalytic carbonization capability and Bi-MOF's gas-phase radical trapping function during thermal degradation processes.199 In another effort to enhance the flame resistance of TPU, Wan and colleagues successfully synthesized ZIF-67@MXene and integrated it into TPU materials. Remarkably, even at a low loading of 0.5 wt%, the composite demonstrated substantial fire safety enhancements: a 26% reduction in peak heat release, 9% decrease in total heat emission, 50% suppression of maximum smoke generation, and 22% mitigation in both total smoke output and carbon monoxide formation relative to unmodified TPU.141 These improvements primarily stem from the hybrid material's barrier mechanism that effectively impedes combustion propagation within the polymer matrix.

Furthermore, MOF/MXene exhibits adsorption and sterilization properties.339 In the design of antibacterial platforms, Guo et al. developed an oxygen-enriched Schottky interface by partially substituting Zn coordination sites in Zn-MOF with Ti atoms derived from Ti3C2Tx nanosheets. This structural innovation facilitated superior separation efficiency and accelerated electron transfer of photogenerated charge carriers under 808 nm near-infrared illumination. The resulting Zn-MOF@Ti3C2Tx composite exhibited remarkable biocompatibility and multimodal antibacterial performance, showing promising therapeutic applications for Staphylococcus aureus-infected wound regeneration.172 Triazole fungicides, widely utilized in agriculture due to their high efficacy, broad-spectrum activity, and chemical stability, are the primary agents for controlling rice blast. To mitigate environmental concerns associated with triazole fungicide residues, Lu et al. fabricated a novel magnetic composite, Fe3O4@MXene@MOF-74, via a facile synthesis approach. Various analytical techniques were employed to characterize the morphology and properties of the fabricated composite, with its adsorption capabilities being methodically evaluated. In practical applications, satisfactory removal effects were achieved for paddy water samples. Importantly, Fe3O4@MXene@MOF-74 exhibited notable antibacterial properties, achieving a 50% effective inhibitory concentration against Magnaporthe oryzae proliferation. These results underscore the composite's dual-functional capability in simultaneously removing pollutants and suppressing phytopathogens, positioning Fe3O4@MXene@MOF-74 as a promising candidate for agricultural water remediation.213 More recently, Zhu et al. synthesized MXene/Cu-MOF composites with varying MXene loadings (3.3%, 6.7%, and 13.3%) through an in situ precipitation method. Compared to pristine MXene and Cu-MOF, the optimized 3.3% MXene/Cu-MOF composite showed a significant improvement in the adsorption performance of methylbenzyl sulfide in model fuels, with an equilibrium adsorption capacity reaching 87.4 mg g−1.340

The continuous progress in science and technology has enabled the development of ion-based flexible actuators capable of sustained shape adaptation, offering a promising alternative to conventional rigid mechanical components. This innovation has generated considerable attention regarding their prospective utilization in emerging metaverse platforms and flexible robotic systems. In this regard, Garai et al. introduced a novel Ti3C2Tx electrode-anchored Mn-based MOF (MnBTC), which can be utilized for super-stable electro-ionic artificial muscles. The Ti3C2Tx-MnBTC composite nanostructure forms coordination bonds, hydrogen bonds, and hydrophilic interactions with the conductive polymer poly(3,4-ethylenedioxythiophene) polystyrene sulfonate, resulting in an electrode with both mechanical flexibility and ionic activity. The Ti3C2Tx-MnBTC composite electrode ion actuator demonstrated a substantial bending displacement of 12.5 mm under 0.5 V at 0.1 Hz. It also exhibited an ultra-fast drive response time of 0.77 s at a DC input signal of 0.5 V, without any backlash phenomenon. Notably, the actuator maintained 98% long-term durability for over 43[thin space (1/6-em)]200 seconds (≈12 h) at a frequency of 1 Hz, with no distortion of the electrodes. These findings highlight the viability of MXene-MOF hybrid constructs as foundational materials for next-generation biomimetic systems requiring both dynamic performance and operational longevity.309 More recently, Memon et al. developed MOF-808@MXene composites for use as triboactive materials in TENG, which exhibited a high triboelectric output of 296.22 V, 30.7 μA, and 67.4 mW m−2. Furthermore, the device was stable in repetitive cycles of 5000 and showed only 34% degradation after a duration of 10 weeks, thus making it suitable as a durable and long-lasting power source based on applications in microcircuit devices.341

In an effort to enhance the flux, antifouling, and dye separation properties of polyether-based membranes, Yao et al. developed mixed matrix membranes using polyether blends combined with ZIF-8@Ti3C2Tx. With a 3% microemulsion content, the membrane achieved an optimal flux of 280 L m−2 h−1 and demonstrated an antifouling performance with a 91% bovine serum albumin rejection rate. Notably, the membrane maintained a 140 L m−2 h−1 high flux in high-salinity wastewater, achieving a dye removal efficiency of up to 98% while only removing 5% of salt ions.173 To address the challenge of antifouling coating degradation and efficacy loss, Wang and his collaborators successfully synthesized a novel MOF, MZ-8, based on mercaptobenzothiazole (MBT). By integrating MZ-8 with MXene nanosheets, they formulated an organic–inorganic nanocomposite, MZ-8/MXene, which was employed as a reinforcing filler in PU resin. The resulting MZ-8/MXene/PU coating exhibited reduced microbial adhesion due to photothermal effects, zinc ions release, and MBT presence, leading to over 99% reduction in surface-adherent bacteria and 93% in microalgae. This advanced coating system, combining robust antifouling performance with intrinsic self-healing capabilities, presents significant potential for applications in marine antifouling and biomedical devices.342

Li and colleagues successfully fabricated a 3D amidoxime-functionalized Ti3C2Tx/ZIF-67 structure that demonstrated a uranium extraction capacity of 2224.54 mg g−1 at 1 V. The material showed exceptional selectivity towards uranium, with a reusability rate of 63.51% after 10 cycles. Furthermore, Ti3C2Tx/ZIF-67 displayed nearly 100% bactericidal efficiency against Escherichia coli and Staphylococcus aureus, underscoring its superior antibacterial properties. Under an electric field, the uranium adsorption rate of Ti3C2Tx/ZIF-67 reached 11.40 mg g−1 within 24 h, which is 5.3 times higher than that of physicochemical adsorption. This study presents a promising method for the efficient capture of uranium resources from seawater through electro adsorption.142 Subas Kashif et al.178 fabricated the UiO-66/MXene composite material via a solvothermal method. A series of studies have shown that the UiO-66/MXene composite exhibits removal rates of 98% for MB, 97% for rhodamine B, and 99% for tetracycline. The maximum adsorption capacities of UiO-66/MXene for MB, rhodamine B, and tetracycline are 312, 285, and 476 mg g−1, respectively. These findings offer a promising outlook for the development of water-refining adsorbents. In another study, Deng et al. developed MXene nanosheets and ZIF-8 functionalized PU sponges (PU-Ti3C2Tx-MOF), resulting in excellent hydrophobicity and a water contact angle of 135°. The sponges demonstrated high adsorption capacities for liquid hazardous chemicals, absorbing up to 48–91 times their weight. Additionally, the materials exhibited excellent recyclability, retaining significant adsorption capabilities toward oils and organic solvents through five consecutive usage cycles. This study proposes a novel engineering approach for fabricating specialized polymeric sponges with significant promise for rapid response to accidental chemical spills.174

Navid Rabiee et al.171 conducted a study on an inorganic MXene/MOF-5 nanostructure for the co-delivery of drugs and genes. The results indicated a complete dose-dependency, with a significant decrease in relative cell viability observed as the treatment time extended from 24 h to 72 h, characterized by a pronounced slope. The chitosan-alginate bilayer coating approach significantly enhanced the mitigation efficiency of nanocarriers. To evaluate their dual delivery capability, the engineered nanosystems were functionalized with pCRISPR for simultaneous tracking and therapeutic assessment. Confocal microscopy analysis demonstrated effective pCRISPR internalization in both HEK-293 and HeLa cellular models, particularly evidenced by 26% elevation in green fluorescent protein signal intensity observed in HeLa populations. These findings suggest that this strategy, along with the optimized MXene/MOF-5 nanosystem, holds promise for further development in drug/gene co-delivery applications in animal models. To address the challenge of efficiently treating bacterially infected wounds, Guo et al. developed a new oxygen-rich vacancy Schottky junction. This junction notably enhances the separation of photo-induced carriers and electron transfer kinetics under 808 nm near-infrared irradiation. The resulting Zn-MOF@Ti3C2Tx composite demonstrates remarkable biocompatibility along with multimodal antimicrobial efficacy against both Staphylococcus aureus and Escherichia coli, while promoting rapid wound healing post-Staphylococcus aureus infection, achieving a wound closure rate of 99%.172 Acknowledging the limited availability of biosensors capable of detecting both chemical and biological signals, Lin et al. designed a multifunctional biosensor. The device integrates three functionalities: sweat-based electrochemical biosensing for UA/Glu monitoring, electrophysiological signal recording, and electrostimulation therapy delivery. Additionally, they fabricated a MOF/MXene electrode suitable for electrical muscle stimulation therapy. This research underscores its considerable potential in muscle therapy and daily noninvasive monitoring, while also offering timely nutritional guidance to individuals with hyperuricemia or hyperglycemia.154 More recently, Alqahtani et al.343 based on MXene coupled with Zn-BTC MOF and coated by chitosan for smart drug delivery, forming a MXene-Zn-BTC@chitosan nanosystem. The nanosystem could enhance cancer cell therapy by combining controlled release with effective cancer treatment and improving the cells sensitivity to therapeutic intervention, thus improving the therapeutic effect of the cisplatin-MXene-Zn-BTC@chitosan nanocomposites as a chemo-photothermal therapy agent against HepG2 cancer cells.

3.7 Summary

The superior performance of MOF/MXene-based materials across a wide range of application fields (from energy storage and sensing to catalysis and electromagnetic interference shielding) highlights their great potential. However, the scalability of these materials from laboratory-scale synthesis to industrial production and real-world deployment remains a critical, application-specific challenge that must be addressed for their widespread adoption.

A critical analysis reveals that the primary bottlenecks are inherent to the synthesis of the constituent materials and the manufacturing of final devices. The production of high-quality MXene typically involves hazardous etchants (e.g., HF) and multi-step processes, which are energy-intensive and difficult to control. This poses a significant obstacle to safe, cost-effective, and environmentally sustainable large-scale production. Similarly, the synthesis of many MOF materials, especially those that require solvothermal conditions, is usually time-consuming and relies on large volumes of organic solvents, increasing the overall cost and environmental pollution. The challenges of scalability manifest differently in various applications:

(1) In energy storage (e.g., batteries, SCs), the performance of electrodes like MXene@ hydroxide and ZIF-67@MXene is highly dependent on precise control of morphology and composition. Scaling up while maintaining this level of control, and ensuring the long-term stability of the MXene component, is a significant engineering hurdle.

(2) In sensing applications, the high selectivity and sensitivity of devices usually rely on complex nanostructures and surface functionalization, which may be difficult to reproduce consistently and affordably on a large scale.

(3) In catalysis, the hydrogen evolution performance of materials depends on the interfacial connectivity and exposure of active sites. Expanding the synthesis scale while maintaining these nanoscale features is a challenge.

(4) In electromagnetic interference shielding, achieving high performance typically requires a high filler loading and precise control of microstructure for the composites to balance impedance matching and attenuation. The dispersion of 2D MXene sheets and prevention of restacking are critical for maintaining performance at scale.

Furthermore, the susceptibility of MXene to oxidation under environmental conditions raises concerns about long-term reliability and shelf life, which are critical for commercial products. It is crucial to develop stable, low-cost, and scalable device fabrication processes (e.g., printing, roll-to-roll manufacturing) to maintain high performance. In conclusion, while MOF/MXene-based materials offer a compelling combination of properties, their path to widespread application depends on overcoming these multifaceted scalability barriers. Future research should prioritize the development of greener, more efficient, and scalable synthesis routes for both MXene and MOF, alongside the design of stable and simplified device architectures. As highlighted in recent perspectives on zinc–air batteries,107 a concerted effort from materials scientists, chemical engineers, and industry partners is required to bridge the gap between the laboratory and the market.

4. Summary and outlook

In summary, this review has systematically presented the synthesis and applications of MOF/MXene-based materials, categorized into pure MOF/MXene, MOF composite/MXene, and MOF derivative/MXene. The synergistic integration of the high surface area and tunable functionality of MOF materials with the exceptional electrical conductivity and rich surface chemistry composition of MXene has led to remarkable advancements across diverse fields, including energy storage, sensing, catalysis, and electromagnetic interference shielding. This powerful combination has established MOF/MXene-based materials as a development direction for next-generation materials. However, to fully exploit the potential of these materials and transform laboratory innovation into influential industrial technologies, a concerted effort is needed to address key challenges and explore new fields.

4.1. Specific research directions

Future research should focus on advancing a deeper understanding and precise control of MOF/MXene-based materials. Building upon the synthesis strategies discussed, a primary direction is advanced interfacial engineering to achieve strong chemical bonds (e.g., Ti–O–C) at the interface, which is crucial for efficient and stable charge transfer. This requires a deeper mechanistic understanding of the nucleation and growth processes, potentially through the application of in situ characterization techniques. Furthermore, it is crucial to develop more environmentally friendly and scalable synthetic strategies. This includes innovating in situ and ex situ methods and ensuring the environmental sustainability of production processes. The field of MOF derivatives offers a rich platform for innovation, with opportunities to design novel bimetallic or multi-functional hybrids with MXene for enhanced catalytic and electrochemical performance. Finally, reinforcing the integration of theoretical and experimental research will be essential to elucidate the structure–property relationships and guide the rational design of these materials.

4.2. Challenges to commercialization

The path to commercialization is fraught with significant hurdles. The large-scale, cost-effective, and environmentally sustainable production of both high-quality MXene and MOF remains the most critical bottleneck. The synthesis of MXene usually involves hazardous etchants and complex, energy-intensive processes, while MOF synthesis can rely on large volumes of organic solvents. The long-term stability of these composites, particularly the susceptibility of MXene to oxidation, is a major concern for device reliability. Additionally, the integration of these nanomaterials into practical devices presents engineering challenges. Combining bottom-up self-assembly approaches with scalable manufacturing technologies, such as 3D printing and roll-to-roll deposition, may facilitate large-scale integration. Fostering collaboration between academia, industry, and government is crucial to translate lab innovations into commercial technologies.

4.3. Interdisciplinary opportunities

The future of MOF/MXene-based materials lies in interdisciplinary collaboration. By merging expertise from materials science, chemistry, physics, and engineering, these composites can be tailored for next-generation technologies. A promising frontier is flexible and wearable electronics, where the mechanical flexibility of MXene and the functional versatility of MOF can be harnessed for integrated health-monitoring devices. Another significant opportunity lies in smart systems and the internet of things, where MOF/MXene composites can serve as multifunctional units combining sensing, energy storage, and signal processing. Exploring their use in solid-state batteries, fuel cells, SCs, and water-splitting devices for hydrogen production represents a crucial interdisciplinary venture. This collaborative effort is indispensable for transforming the exceptional laboratory performance of MOF/MXene-based materials into transformative practical applications.

Author contributions

Shasha Zheng: writing review & editing, data curation, formal analysis, project administration, investigation, supervision. Keke Zhou: writing original draft, software. Xiaoxue Zhang: writing original draft, investigation. Ning Ren: methodology, investigation.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Acknowledgements

This work was supported by the Key Scientific Research Projects Plan of Colleges and Universities in Henan Province (No. 24A150021) and Youth Science and Technology Innovation Special Project of Zhumadian (No. QNZX202308).

References

  1. Z. Pei, H. Lei and L. Cheng, Chem. Soc. Rev., 2023, 52, 2031–2081 RSC.
  2. X. Luo, Y. Zhai, P. Wang, B. Tian, S. Liu, J. Li, C. Yang, V. Strehmel, S. Li, K. Matyjaszewski, G. Yilmaz, B. Strehmel and Z. Chen, Angew. Chem., Int. Ed., 2024, 63, e202316431 CrossRef CAS PubMed.
  3. L. Liu, Y. Pan, L. Ye, C. Liang, X. Mou, X. Dong and Y. Cai, Coord. Chem. Rev., 2024, 517, 216006 CrossRef CAS.
  4. S. Liang, J. Yao, D. Liu, L. Rao, X. Chen and Z. Wang, Adv. Mater., 2023, 35, 2211130 CrossRef CAS PubMed.
  5. I. E. Khalil, J. Fonseca, M. R. Reithofer, T. Eder and J. M. Chin, Coord. Chem. Rev., 2023, 481, 215043 CrossRef CAS.
  6. T. Yao, X. Zeng, X. Tao and H. Xu, Chem. Eng. J., 2024, 487, 150641 CrossRef CAS.
  7. S. Guo, M. Gao, W. Zhang, F. Liu, X. Guo and K. Zhou, Adv. Mater., 2023, 35, 2303065 CrossRef CAS PubMed.
  8. Y. Wang, Y. Yang, J. Liu, X. Zi, H. Zhu, X. Sun, Y. Miao and Y. Fu, Chem. Eng. J., 2024, 500, 156850 Search PubMed.
  9. S. Zhou, F. Jia, Y. Wei, J. Du, J. Liu, W. Dong, J. Sui, W. Xue, Y. Yang, L. Chen, X. Meng and S. Yu, Adv. Funct. Mater., 2024, 34, 2314012 Search PubMed.
  10. J. M. Huang, X. D. Zhang, J. Y. Huang, D. S. Zheng, M. Xu and Z. Y. Gu, Coord. Chem. Rev., 2023, 494, 215333 CrossRef CAS.
  11. S. Li, W. Han, Q. An, K. Yong and M. Yin, Adv. Funct. Mater., 2023, 33, 2303447 CrossRef CAS.
  12. J. Li, C. Hou, C. Chen, W. Ma, Q. Li, L. Hu, X. Lv and J. Dang, ACS Nano, 2023, 17, 10947–10957 CrossRef CAS PubMed.
  13. Y. Song, C. Yu, D. Ma and K. Liu, Coord. Chem. Rev., 2024, 499, 215492 CrossRef CAS.
  14. A. H. Mamaghani, J. Liu, Z. Zhang, R. Gao, Y. Wu, H. Li, M. Feng and Z. Chen, Adv. Energy Mater., 2024, 14, 2402278 CrossRef CAS.
  15. J. Park, M. Lee, D. Feng, Z. Huang, A. C. Hinckley, A. Yakovenko, X. Zou, Y. Cui and Z. Bao, J. Am. Chem. Soc., 2018, 140, 10315–10323 CrossRef CAS PubMed.
  16. A. M. Patil, S. Moon, Y. Seo, S. B. Roy, A. A. Jadhav, D. P. Dubal, K. Kang and S. C. Jun, Small, 2023, 19, 2205491 CrossRef CAS PubMed.
  17. H. Zhu, S. Li, L. Peng, W. Zhong, Q. Wu, S. Cheng and J. Xie, Nano Energy, 2024, 125, 109571 CrossRef CAS.
  18. W. Zhou, Y. Tang, X. Zhang, S. Zhang, H. Xue and H. Pang, Coord. Chem. Rev., 2023, 477, 214949 CrossRef CAS.
  19. P. Wiśniewska, J. Haponiuk, M. R. Saeb, N. Rabiee and S. A. Bencherif, Chem. Eng. J., 2023, 471, 144400 CrossRef PubMed.
  20. H. Abed, R. Sabouni and M. Ghommem, RSC Adv., 2024, 14, 39472–39497 RSC.
  21. T. Zhao, H. Wu, X. Wen, J. Zhang, H. Tang, Y. Deng, S. Liao and X. Tian, Coord. Chem. Rev., 2022, 468, 214642 CrossRef CAS.
  22. B. Xu, Z. Huang, Y. Liu, S. Li and H. Liu, Nano Today, 2023, 48, 101690 CrossRef CAS.
  23. H. Yang, L. Wang and X. Huang, Coord. Chem. Rev., 2023, 495, 215372 Search PubMed.
  24. G. Tatrari, R. An and F. U. Shah, Coord. Chem. Rev., 2024, 512, 215876 CrossRef CAS.
  25. R. Abazari, S. Sanati, W. K. Fan, M. Tahir, S. Nayak, K. Parida, M. El-Shahat, R. M. Abdelhameed, D. S. Nesterov, A. M. Kirillov and J. Qian, Coord. Chem. Rev., 2025, 523, 216256 CrossRef CAS.
  26. W. Zhang, X. Li, X. Ding, K. Hua, A. Sun, X. Hu, Z. Nie, Y. Zhang, J. Wang, R. Li and S. Liu, RSC Adv., 2023, 13, 10800–10817 Search PubMed.
  27. R. Ji, Z. Cui, T. Wu, Y. Xu, M. Yang, R. Zhang, F. Kang, Q. Huang, Q. Zhang and D. Tian, Coord. Chem. Rev., 2025, 540, 216778 Search PubMed.
  28. S. Zheng, X. Li, B. Yan, Q. Hu, Y. Xu, X. Xiao, H. Xue and H. Pang, Adv. Energy Mater., 2017, 7, 1602733 Search PubMed.
  29. Y. Xue, S. Zheng, H. Xue and H. Pang, J. Mater. Chem. A, 2019, 7, 7301–7327 RSC.
  30. Y. Du, X. Jia, L. Zhong, Y. Jiao, Z. Zhang, Z. Wang, Y. Feng, M. Bilal, J. Cui and S. Jia, Coord. Chem. Rev., 2022, 454, 214327 CrossRef CAS.
  31. C. Zhang, Y. Wu, D. Li and H. L. Jiang, Chem. Sci., 2025, 16, 13149–13172 RSC.
  32. L. Jia, S. Zhou, A. Ahmed, Z. Yang, S. Liu, H. Wang, F. Li, M. Zhang, Y. Zhang and L. Sun, Chem. Eng. J., 2023, 475, 146361 CrossRef CAS.
  33. Z. Wang, C. Wei, H. Jiang, Y. Zhang, K. Tian, Y. Li, X. Zhang, S. Xiong, C. Zhang and J. Feng, Adv. Mater., 2024, 36, 2306015 Search PubMed.
  34. W. Zeng, X. Ye, Y. Dong, Y. Zhang, C. Sun, T. Zhang, X. Guan and L. Guo, Coord. Chem. Rev., 2024, 508, 215753 Search PubMed.
  35. Z. Wu, S. Liu, Z. Hao and X. Liu, Adv. Sci., 2023, 10, 2207174 Search PubMed.
  36. W. Cao, J. Nie, Y. Cao, C. Gao, M. Wang, W. Wang, X. Lu, X. Ma and P. Zhong, Chem. Eng. J., 2024, 496, 154097 CrossRef CAS.
  37. H. Wu, M. Almalki, X. Xu, Y. Lei, F. Ming, A. Mallick, V. Roddatis, S. Lopatin, O. Shekhah, M. Eddaoudi and H. N. Alshareef, J. Am. Chem. Soc., 2019, 141, 20037–20042 CrossRef CAS PubMed.
  38. C. Zheng, Y. Yao, X. Rui, Y. Feng, D. Yang, H. Pan and Y. Yu, Adv. Mater., 2022, 34, 2204988 Search PubMed.
  39. D. Lei, N. Liu, T. Su, Q. Zhang, L. Wang, Z. Ren and Y. Gao, Adv. Mater., 2022, 34, 2110608 CrossRef CAS PubMed.
  40. C. Wei, Q. Zhang, Z. Wang, W. Yang, H. Lu, Z. Huang, W. Yang and J. Zhu, Adv. Funct. Mater., 2023, 33, 2211889 CrossRef CAS.
  41. S. Kazim, C. Huang, N. H. Hemasiri, A. Kulkarni, S. Mathur and S. Ahmad, Adv. Funct. Mater., 2024, 34, 2315694 CrossRef CAS.
  42. J. Li, L. Han, Y. Li, J. Li, G. Zhu, X. Zhang, T. Lu and L. Pan, Chem. Eng. J., 2020, 380, 122590 CrossRef CAS.
  43. L. Sun, J. Sun, S. Zhai, T. Dong, H. Yang, Y. Tan, X. Fang, C. Liu, W. Deng and H. Wu, Adv. Energy Mater., 2022, 12, 2200113 Search PubMed.
  44. Y. Chen, H. Yang, Z. Han, Z. Bo, J. Yan, K. Cen and K. K. Ostrikov, Energy Fuels, 2022, 36, 2390–2406 Search PubMed.
  45. B. Zhang, A. Boretti and S. Castelletto, Chem. Eng. J., 2022, 435, 134959 CrossRef CAS.
  46. X. Bai and J. Guan, Small Struct., 2023, 4, 2200354 CrossRef CAS.
  47. M. Bilal, A. K. Singh, H. M. N. Iqbal and G. Boczkaj, Chem. Eng. J., 2023, 474, 145020 CrossRef CAS.
  48. X. Fan, Y. Yang, X. Shi, Y. Liu, H. Li, J. Liang and Y. Chen, Adv. Funct. Mater., 2020, 30, 2007110 CrossRef CAS.
  49. L. Wang, H. Song, L. Yuan, Z. Li, P. Zhang, J. K. Gibson, L. Zheng, H. Wang, Z. Chai and W. Shi, Environ. Sci. Technol., 2019, 53, 3739–3747 CrossRef CAS PubMed.
  50. M. Mozafari, A. Arabi Shamsabadi, A. Rahimpour and M. Soroush, Adv. Mater. Technol., 2021, 6, 2001189 CrossRef CAS.
  51. I. Persson, J. Halim, H. Lind, T. W. Hansen, J. B. Wagner, L. Å. Näslund, V. Darakchieva, J. Palisaitis, J. Rosen and P. O. Å. Persson, Adv. Mater., 2019, 31, 1805472 CrossRef PubMed.
  52. Y. Zhou, Y. Zhang, K. Ruan, H. Guo, M. He, H. Qiu and J. Gu, Sci. Bull., 2024, 69, 2776–2792 CrossRef PubMed.
  53. J. Liu, M. Wang, C. Guo, Z. Tao, M. Wang, L. He, B. Liu and Z. Zhang, Food Chem., 2023, 416, 135839 CrossRef CAS PubMed.
  54. Y. Jo, Y. K. Jo, J. Lee, H. W. Jang, I. Hwang and D. J. Yoo, Adv. Mater., 2023, 35, 2206842 CrossRef CAS PubMed.
  55. R. Qin, J. Nong, K. Wang, Y. Liu, S. Zhou, M. Hu, H. Zhao and G. Shan, Adv. Mater., 2024, 36, 2312761 Search PubMed.
  56. S. Elbeltagi, M. Al-zharani, F. A. Nasr, A. M. Ismail, H. M. El-Tohamy, K. M. Abdelbased and Z. E. Eldin, Int. J. Pharm., 2025, 674, 125492 Search PubMed.
  57. B. Deng, Z. Liu, F. Pan, Z. Xiang, X. Zhang and W. Lu, J. Mater. Chem. A, 2021, 9, 3500–3510 Search PubMed.
  58. M. Liu, Z. Zhang, M. Yang, P. Li, Y. Wang, Y. He and J. Yuan, Composites, Part A, 2022, 161, 107122 CrossRef CAS.
  59. L. Wang, H. Liu, X. Lv, G. Cui and G. Gu, J. Alloys Compd., 2020, 828, 154251 CrossRef CAS.
  60. Y. Ji, Y. Zhang, J. Lu, F. Gao, X. Lv, X. Qu, G. Zhu, T. Tian, H. Pang, Y. Tian and X. Dong, Adv. Healthc. Mater., 2025, 14, 2404723 Search PubMed.
  61. Y. He, Y. Zou, L. Sun, F. Xu, Y. Xia, Y. Chen, S. Wang, Z. Hu, S. Hao and C. Xiang, J. Energy Storage, 2025, 119, 116385 CrossRef.
  62. X. Zhang, G. Jin, Y. Liu, Y. Liu, M. Zhang, C. Li, X. Zhang and D. Cao, Adv. Compos. Hybrid Mater., 2025, 8, 206 CrossRef CAS.
  63. X. Zhuang, S. Zhang, Y. Tang, F. Yu, Z. Li and H. Pang, Coord. Chem. Rev., 2023, 490, 215208 CrossRef CAS.
  64. H. Saini, N. Srinivasan, V. Šedajová, M. Majumder, D. P. Dubal, M. Otyepka, R. Zbořil, N. Kurra, R. A. Fischer and K. Jayaramulu, ACS Nano, 2021, 15, 18742–18776 CrossRef CAS PubMed.
  65. T. Xu, Y. Wang, Y. Xue, J. Li and Y. Wang, Chem. Eng. J., 2023, 470, 144247 CrossRef CAS.
  66. F. Ren, L. Ma, C. Li, T. Wu, J. Zhang, L. Pei, Y. Jin, Z. Sun, Z. Guo, P. Song and P. Ren, Carbohydr. Polym., 2025, 356, 123369 CrossRef CAS PubMed.
  67. A. Hanan, M. N. Lakhan, F. Bibi, A. Khan, I. A. Soomro, A. Hussain and U. Aftab, Chem. Eng. J., 2024, 482, 148776 CrossRef CAS.
  68. I. Ijaz, A. Bukhari, E. Gilani, A. Nazir and H. Zain, RSC Adv., 2023, 13, 5643–5655 RSC.
  69. H. Kefayati, Y. Yamini and M. Ghaemmaghami, Microchim. Acta, 2025, 192, 145 CrossRef CAS PubMed.
  70. S. Zheng, H. Zhou, H. Xue, P. Braunstein and H. Pang, J. Colloid Interface Sci., 2022, 614, 130–137 CrossRef CAS PubMed.
  71. S. Huang, H. Wang, C. Xiang, D. Luo, H. Zeng, A. Cai, X. Ma, C. Fang and J. Deng, J. Membr. Sci., 2025, 719, 123738 CrossRef CAS.
  72. A. Chaouiki, M. Chafiq, R. Salghi, N. Elboughdiri, J. H. Kang, Y. G. Ko and M. Abboud, J. Energy Chem., 2025, 104, 687–715 CrossRef CAS.
  73. S. Bibi, S. S. A. Shah, M. A. Nazir, M. H. Helal, S. M. El-Bahy, Z. M. El-Bahy, S. Ullah, M. A. Wattoo and A. ur Rehman, Adv. Sustain. Syst., 2024, 8, 2400011 CrossRef CAS.
  74. Y. Zhao, J. Zhang, X. Guo, X. Cao, S. Wang, H. Liu and G. Wang, Chem. Soc. Rev., 2023, 52, 3215–3264 RSC.
  75. N. Kitchamsetti and J. S. Cho, J. Energy Storage, 2024, 80, 110293 Search PubMed.
  76. H. Wang, X. Li, Y. Deng, J. Jiang, H. Ma and J. Zou, Coord. Chem. Rev., 2025, 529, 216462 CrossRef CAS.
  77. H. Ji, Y. Liu, G. Du, T. Huang, Y. Zhu, Y. Sun and H. Pang, Chem. Res. Chin. Univ., 2024, 40, 943–963 Search PubMed.
  78. G. T. Tran, T. T. T. Nguyen and T. Van Tran, Coord. Chem. Rev., 2025, 544, 217002 CrossRef CAS.
  79. R. Ghanbari and E. Nazarzadeh Zare, Coord. Chem. Rev., 2024, 518, 216089 CrossRef CAS.
  80. Y. Zheng, S. Zheng, H. Xue and H. Pang, Adv. Funct. Mater., 2018, 28, 1804950 CrossRef.
  81. N. A. I. M. Mokhtar, R. M. Zawawi, W. M. Khairul and N. A. Yusof, Environ. Chem. Lett., 2022, 20, 3099–3131 CrossRef CAS.
  82. S. Ullah, A. ur Rehman, T. Najam, I. Hossain, S. Anjum, R. Ali, M. U. Shahid, S. S. A. Shah and M. A. Nazir, J. Ind. Eng. Chem., 2024, 137, 87–105 CrossRef CAS.
  83. S. Gautam, S. Rialach, S. Paul and N. Goyal, RSC Adv., 2024, 14, 14311–14339 RSC.
  84. Y. Absalan, M. Gholizadeh, E. B. Kim, S. Ameen, Y. Wang, Y. Wang and H. He, Coord. Chem. Rev., 2024, 515, 215972 CrossRef CAS.
  85. A. K. Bindra, D. Wang and Y. Zhao, Adv. Mater., 2023, 35, 2300700 CrossRef CAS PubMed.
  86. Y. Zhang, X. Feng, S. Yuan, J. Zhou and B. Wang, Inorg. Chem. Front., 2016, 3, 896–909 RSC.
  87. F. Parsapour, M. Moradi, V. Safarifard and S. Sojdeh, J. Energy Storage, 2024, 82, 110487 CrossRef.
  88. C. Brahmi, M. Benltifa, C. Vaulot, L. Michelin, F. Dumur, F. Millange, M. Frigoli, A. Airoudj, F. Morlet-Savary, L. Bousselmi and J. Lalevée, Eur. Polym. J., 2021, 154, 110560 CrossRef CAS.
  89. M. Kalaj, K. C. Bentz, S. Ayala, J. M. Palomba, K. S. Barcus, Y. Katayama and S. M. Cohen, Chem. Rev., 2020, 120, 8267–8302 CrossRef CAS PubMed.
  90. I. Salahshoori, M. Golriz, M. Namayandeh Jorabchi, S. Moghari and H. A. Khonakdar, Coord. Chem. Rev., 2025, 543, 216919 CrossRef CAS.
  91. W. Lv, Y. Song, H. Pei and Z. Mo, J. Ind. Eng. Chem., 2023, 128, 17–54 CrossRef CAS.
  92. T. Wu, X. Liu, Y. Liu, M. Cheng, Z. Liu, G. Zeng, B. Shao, Q. Liang, W. Zhang, Q. He and W. Zhang, Coord. Chem. Rev., 2020, 403, 213097 Search PubMed.
  93. J. Aguilera-Sigalat and D. Bradshaw, Coord. Chem. Rev., 2016, 307, 267–291 Search PubMed.
  94. M. Zahran, S. Tian, J. Li, A. H. Marei, Y. Xie, Q. Liu, J. Huang, D. Wang, X. Ning, J. Wang, H. Chi and X. Li, RSC Adv., 2025, 15, 10801–10815 RSC.
  95. Y. Liu, Z. Liu, D. Huang, M. Cheng, G. Zeng, C. Lai, C. Zhang, C. Zhou, W. Wang, D. Jiang, H. Wang and B. Shao, Coord. Chem. Rev., 2019, 388, 63–78 Search PubMed.
  96. M. Mukoyoshi and H. Kitagawa, Chem. Commun., 2022, 58, 10757–10767 Search PubMed.
  97. S. Tang, Y. Wang, P. He, Y. Wang and G. Wei, Materials, 2024, 17, 2660 Search PubMed.
  98. J. Yu, C. Mu, B. Yan, X. Qin, C. Shen, H. Xue and H. Pang, Mater. Horiz., 2017, 4, 557–569 Search PubMed.
  99. Y. Liu, D. Huang, M. Cheng, Z. Liu, C. Lai, C. Zhang, C. Zhou, W. Xiong, L. Qin, B. Shao and Q. Liang, Coord. Chem. Rev., 2020, 409, 213220 Search PubMed.
  100. X. Lian, Y. Fang, E. Joseph, Q. Wang, J. Li, S. Banerjee, C. Lollar, X. Wang and H. C. Zhou, Chem. Soc. Rev., 2017, 46, 3386–3401 Search PubMed.
  101. J. Jiang, Y. Shi, M. Wu, M. Rezakazemi, T. M. Aminabhavi, R. Huang, C. Jia and S. Ge, Chem. Eng. J., 2024, 494, 152932 Search PubMed.
  102. Y. Chen, Y. Shi, G. Song, B. Yang and H. Pang, Next Mater., 2025, 6, 100323 Search PubMed.
  103. Q. Wu, J. Liang, D. Wang, R. Wang and C. Janiak, Chem. Soc. Rev., 2025, 54, 601–622 Search PubMed.
  104. W. K. Fan, A. Sherryna and M. Tahir, ACS Omega, 2022, 7, 38158–38192 Search PubMed.
  105. Y. Ji, Y. You, G. Xu, X. Yang and Y. Liu, Chem. Eng. J., 2024, 483, 149365 Search PubMed.
  106. H. Yang, G. X. Zhang, H. J. Zhou, Y. Y. Sun and H. Pang, Energy Mater. Adv., 2023, 4, 0033 Search PubMed.
  107. T. Ramachandran, R. K. Raji and M. Rezeq, J. Mater. Chem. A, 2025, 13, 12855–12890 RSC.
  108. S. Venkateswarlu, S. Vallem, M. Umer, N. V. V. Jyothi, A. Giridhar Babu, S. Govindaraju, Y. Son, M. Jong Kim and M. Yoon, J. Energy Chem., 2023, 86, 409–436 Search PubMed.
  109. N. Khosroshahi, M. Bakhtian, A. Asadi and V. Safarifard, Nano Express, 2023, 4, 042002 Search PubMed.
  110. B. Tahir, A. Alraeesi and M. Tahir, Front. Chem., 2024, 12, 1448700 CrossRef CAS PubMed.
  111. S. Afzal, A. ur Rehman, T. Najam, I. Hossain, M. A. I. Abdelmotaleb, S. Riaz, M. R. Karim, S. S. A. Shah and M. A. Nazir, Chemosphere, 2024, 364, 143194 Search PubMed.
  112. S. Bagyalakshmi, K. S. Balamurugan, M. Mathankumar, A. Sivakami and N. R. Devi, J. Mol. Struct., 2025, 1334, 141838 CrossRef CAS.
  113. P. Cao, Y. L. Liu, Q. Zhang, C. Guo, J. Li and Q. Zeng, Coord. Chem. Rev., 2025, 545, 217018 Search PubMed.
  114. R. Verma, P. Thakur, A. Chauhan, R. Jasrotia and A. Thakur, Carbon, 2023, 208, 170–190 CrossRef CAS.
  115. S. Feng, S. Wen, R. Wang, X. Yang, X. Yuan, Y. Liu, J. Ma and Z. Li, Nanomaterials, 2025, 15, 841 Search PubMed.
  116. Z. Bai, L. Guo, J. Huang, H. Li, G. An, H. Zheng, N. Wang, Z. Li and Y. Zhu, Chem. Eng. J., 2024, 479, 147932 CrossRef CAS.
  117. Y. Liu, D. Luo, Y. Deng, J. Zheng, F. Wu, X. Dai and C. Deng, J. Energy Storage, 2023, 67, 107641 Search PubMed.
  118. K. Sharma, V. Hasija, S. Patial, P. Singh, V. H. Nguyen, A. K. Nadda, S. Thakur, P. Nguyen-Tri, C. C. Nguyen, S. Y. Kim, Q. Van Le and P. Raizada, Int. J. Hydrogen Energy, 2023, 48, 6560–6574 CrossRef CAS.
  119. Y. Sun, C. Song, X. Guo, S. Hong, J. Choi and Y. Liu, J. Membr. Sci., 2020, 616, 118615 Search PubMed.
  120. C. Liu, Y. Bai, W. Li, F. Yang, G. Zhang and H. Pang, Angew. Chem., Int. Ed., 2022, 61, e202116282 Search PubMed.
  121. M. Mahato, J. Kim, M. Lee, S. Jo, G. Kim, S. Nam, J. Kim, V. H. Nguyen, M. Garai, H. Yoo, D. Saatchi, Z. Ullah, C. W. Ahn, Y. Gogotsi and I. Oh, Adv. Mater., 2025, 2500479 Search PubMed.
  122. X. Zhang, Z. Wang and X. Guo, Sci. Rep., 2025, 15, 3727 CrossRef CAS PubMed.
  123. G. Zeng, Y. Liu, Q. Lin, S. Pu, S. Zheng, M. B. M. Y. Ang and Y. H. Chiao, Sep. Purif. Technol., 2022, 293, 121052 Search PubMed.
  124. J. Zhang, Z. Li, Q. Zhang, L. Zhang, T. Ma, X. Ma, K. Liang, Y. Ying and Y. Fu, ACS Appl. Mater. Interfaces, 2023, 15, 17222–17232 Search PubMed.
  125. J. Zhang, X. Cheng, J. Jiang, A. Pattanateeradetch, S. Shi, C. Chokejaroenrat, Q. Xu and J. Lu, Inorg. Chem., 2025, 64, 4920–4933 CrossRef CAS PubMed.
  126. N. Liu, Y. Zheng, C. Jing, B. Gao, W. Huang, Z. Li, J. Lei, X. Zhang, L. Cui and L. Tang, J. Mol. Liq., 2020, 311, 113201 CrossRef CAS.
  127. L. Yang, Y. Chen, Q. Wen, H. Xu, X. Pan and X. Li, Electrochim. Acta, 2022, 428, 140959 CrossRef CAS.
  128. M. Wan, C. Shi, X. Qian, Y. Qin, J. Jing, H. Che, F. Ren, J. Li and B. Yu, Composites, Part A, 2022, 163, 107187 CrossRef CAS.
  129. D. Cheng, P. Li, X. Zhu, M. Liu, Y. Zhang and Y. Liu, Chin. J. Chem., 2021, 39, 2181–2187 CrossRef CAS.
  130. J. Paul and J. Kim, Appl. Surf. Sci., 2023, 613, 156103 CrossRef CAS.
  131. R. Li, X. Fu, G. Liu, J. Li, G. Zhou, G. Liu and W. Jin, J. Membr. Sci., 2022, 664, 121097 CrossRef CAS.
  132. Y. Wen, Z. Wei, C. Ma, X. Xing, Z. Li and D. Luo, Nanomaterials, 2019, 9, 775 CrossRef CAS PubMed.
  133. P. Xiao, G. Zhu, X. Shang, B. Hu, B. Zhang, Z. Tang, J. Yang and J. Liu, J. Electroanal. Chem., 2022, 916, 116382 CrossRef CAS.
  134. B. Jia, H. Yang, L. Wang, Z. Zhao and X. Wu, Appl. Surf. Sci., 2022, 602, 154386 CrossRef CAS.
  135. H. Li, H. Lin, S. Raza, C. Chen, W. Yu, Q. Zeng, Z. Zhao, X. Huang and L. Shen, J. Membr. Sci., 2025, 718, 123685 CrossRef CAS.
  136. J. Wang, Y. Shen, Y. Zhuang, J. Wang and Y. Zhang, Anal. Chem., 2024, 96, 12102–12111 CrossRef CAS PubMed.
  137. W. Han, Y. Yan, H. Wang, J. Li, P. Zhao, Z. Liu, F. Yu, J. Cui and G. Zhang, Langmuir, 2024, 40, 21767–21779 CrossRef CAS PubMed.
  138. Y. Liu, X. Meng, Z. Ma, H. Gu, X. Luo, X. Yin, H. Yi and Y. Chen, Food Chem., 2025, 463, 141496 CrossRef CAS PubMed.
  139. R. Ramachandran, K. Rajavel, W. Xuan, D. Lin and F. Wang, Ceram. Int., 2018, 44, 14425–14431 CrossRef CAS.
  140. X. Liang, X. Ren, Q. Yang, L. Gao, M. Gao, Y. Yang, H. Zhu, G. Li, T. Ma and A. Liu, Nanoscale, 2021, 13, 2843–2848 RSC.
  141. M. Wan, C. Shi, X. Qian, Y. Qin, J. Jing and H. Che, Nanomaterials, 2022, 12, 1142 CrossRef CAS PubMed.
  142. N. Li, L. Yang, R. Su, N. Shi, J. Wu, J. Zhao, L. Wen and Z. Wang, Desalination, 2023, 566, 116940 CrossRef CAS.
  143. K. S. Ranjith, S. Sonwal, A. Mohammadi, G. Seeta Rama Raju, M.-H. Oh, Y. S. Huh and Y.-K. Han, J. Mater. Chem. A, 2024, 12, 26132–26146 RSC.
  144. D. Roy, R. Singh, S. Mandal and N. Chanda, Anal. Methods, 2024, 16, 6183–6192 Search PubMed.
  145. J. Ding, Q. Wang, X. Liu, S. Li and H. Li, J. Hazard. Mater., 2024, 480, 136261 Search PubMed.
  146. J. Qin, B. Liu, K. H. Lam, S. Song, X. Li and X. Hu, ACS Sustain. Chem. Eng., 2020, 8, 17791–17799 Search PubMed.
  147. L. Cheng, Y. Tang, M. Xie, Y. Sun and H. Liu, J. Alloys Compd., 2021, 864, 158913 Search PubMed.
  148. X. Zhang, S. Yang, W. Lu, D. Lei, Y. Tian, M. Guo, P. Mi, N. Qu and Y. Zhao, J. Colloid Interface Sci., 2021, 592, 95–102 Search PubMed.
  149. H. Zhang, Z. Li, Z. Hou, H. Mei, Y. Feng, B. Xu and D. Sun, Chem. Eng. J., 2021, 425, 130602 CrossRef CAS.
  150. A. Ganiyat Olatoye, W. Li, E. Oluwaseyi Fagbohun, X. Zeng, Y. Zheng and Y. Cui, J. Electroanal. Chem., 2023, 928, 117036 Search PubMed.
  151. X. Yang, Y. Tian, S. Li, Y. P. Wu, Q. Zhang, D. S. Li and S. Zhang, J. Mater. Chem. A, 2022, 10, 12225–12234 Search PubMed.
  152. S. Li, Y. Wang, Y. Li, J. Xu, T. Li and T. Zhang, Nanomaterials, 2023, 13, 610 Search PubMed.
  153. O. P. Nanda, M. Ravipati, L. Durai and S. Badhulika, Mater. Res. Bull., 2023, 168, 112488 Search PubMed.
  154. X. Lin, D. Song, T. Shao, T. Xue, W. Hu, W. Jiang, X. Zou and N. Liu, Adv. Funct. Mater., 2023, 2311637 Search PubMed.
  155. W. Xu, Y. Lu, Z. Zhang, Q. Zhou, W. Zhao, R. Yang, L. Zhang, L. Yang, X. Gao and G. Pan, IEEE Sens. J., 2024, 24, 9437–9448 CAS.
  156. R. Zhu, Y. Gu, L. Liu, J. Lu and H. Pang, New J. Chem., 2024, 48, 10593–10598 RSC.
  157. Y. Ji, W. Li, Y. You and G. Xu, Chem. Eng. J., 2024, 496, 154009 Search PubMed.
  158. D. S. Shivade, A. N. Kurade, R. K. Bhosale, S. S. Kundale, A. R. Shelake, A. D. Patil, P. P. Waifalkar, R. K. Kamat, A. M. Teli and T. D. Dongale, J. Energy Storage, 2024, 100, 113754 Search PubMed.
  159. C. Gu, C. Lu, Y. X. Gao, P. Tan, S. S. Peng, X. Q. Liu and L. B. Sun, Inorg. Chem., 2021, 60, 1380–1387 CrossRef CAS PubMed.
  160. H. Hassan, E. Umar, M. W. Iqbal, A. K. Alqorashi, B. S. Almutair, H. Alrobei, A. M. Afzal and N. Ahmad, New J. Chem., 2024, 48, 6277–6295 RSC.
  161. Z. Liu, H. Lv, Y. Zhang, J. W. He, L. Han, S. Li, L. Yang and Y. Xu, ACS Sens., 2024, 9, 3641–3651 CrossRef CAS PubMed.
  162. A. Sharma, H.-S. Lee, C.-M. Yeom, H. K. Surendran, C. Narayana, S. B. Eadi, R. Zboril, H.-D. Lee and K. Jayaramulu, Chem. Mater., 2024, 36, 8466–8476 CrossRef CAS.
  163. H. Zhu, Y. Chen, T. Zhang, L. Qin, S. Z. Kang and X. Li, J. Solid State Chem., 2023, 327, 124298 Search PubMed.
  164. W. H. Lin, T. Y. Huang, C. H. Bai, C. H. Hung, C. A. Lung, W. H. Hung, K. Gopinadhan and L. H. Yeh, Nano Energy, 2024, 128, 109924 Search PubMed.
  165. Y. Wang, W. Yang, J. Cao, Y. Liu, M. Shao, X. Xu and H. Chen, Sens. Actuators, B, 2024, 417, 136066 Search PubMed.
  166. Q. Zhou, Q. Ding, Z. Geng, C. Hu, L. Yang, Z. Kan, B. Dong, M. Won, H. Song, L. Xu and J. S. Kim, Nano-Micro Lett., 2025, 17, 50 CrossRef CAS PubMed.
  167. Y. Ge, D. Lei, C. Zhang, Q. Zhang, J. Mu and J. Li, Adv. Sci., 2025, 12, 2500027 CrossRef CAS PubMed.
  168. Y. Li, P. Kamdem and X. J. Jin, J. Electrochem. Soc., 2020, 167, 110562 CrossRef CAS.
  169. Y. Yao, X. Han, X. Yang, J. Zhao and C. Chai, Chin. J. Chem., 2021, 39, 330–336 CrossRef CAS.
  170. Y. Wang, W. Sun, Y. Li, X. Zhuang, C. Tian, F. Luan and X. Fu, Microchem. J., 2021, 167, 106332 CrossRef CAS.
  171. N. Rabiee, M. Bagherzadeh, M. Jouyandeh, P. Zarrintaj, M. R. Saeb, M. Mozafari, M. Shokouhimehr and R. S. Varma, ACS Appl. Bio Mater., 2021, 4, 5106–5121 CrossRef CAS PubMed.
  172. C. Guo, F. Cheng, G. Liang, S. Zhang, S. Duan, Y. Fu, F. Marchetti, Z. Zhang and M. Du, Adv. NanoBiomed Res., 2022, 2, 2200064 CrossRef CAS.
  173. Y. Yao, T. Wang, L. Wu and H. Chen, Desalination, 2022, 543, 116116 CrossRef CAS.
  174. G. Deng, M. Sun, Y. Shi, Y. Feng, Y. Lv, L. Tang, J. Gao and P. Song, J. Clean. Prod., 2023, 403, 136887 CrossRef CAS.
  175. L. P. Hao, A. Hanan, R. Walvekar, M. Khalid, F. Bibi, W. Y. Wong and C. Prakash, Catalysts, 2023, 13, 802 CrossRef CAS.
  176. H. Shahriyari Far, M. Najafi, M. Hasanzadeh and R. Rahimi, Inorg. Chem. Commun., 2023, 152, 110680 CrossRef CAS.
  177. C. Wen, D. Guo, X. Zheng, H. Li and G. Sun, ACS Appl. Energy Mater., 2021, 4, 8231–8241 CrossRef CAS.
  178. S. Kashif, S. Akram, M. Murtaza, A. Amjad, S. S. A. Shah and A. Waseem, Diam. Relat. Mater., 2023, 136, 110023 CrossRef CAS.
  179. B. Xiang, J. Gong, Y. Sun, W. Yan, R. Jin and J. Li, J. Membr. Sci., 2024, 691, 122247 CrossRef CAS.
  180. A. Maleki and A. Bozorg, Chemosphere, 2024, 364, 143273 CrossRef CAS PubMed.
  181. Y. Meng, F. Yue, S. Zhang, L. Zhang, C. Li, M. Shi, Y. Ma, M. Berrettoni, X. Zhang and H. Zhang, Carbon Capture Sci. Technol., 2024, 13, 100274 CAS.
  182. Y. Liu, Y. He, E. Vargun, T. Plachy, P. Saha and Q. Cheng, Nanomaterials, 2020, 10, 695 CrossRef CAS PubMed.
  183. C. Du, Q. Song, Q. Liang, X. Zhao, J. Wang, R. Zhi, Y. Wang and H. Yu, ChemNanoMat, 2021, 7, 539–544 CrossRef CAS.
  184. F. Wu, Z. Liu, J. Wang, T. Shah, P. Liu, Q. Zhang and B. Zhang, Chem. Eng. J., 2021, 422, 130591 CrossRef CAS.
  185. T. Xiao, J. Jin, Y. Zhang, W. Xi, R. Wang, Y. Gong, B. He and H. Wang, Electrochim. Acta, 2022, 427, 140851 CrossRef CAS.
  186. L. Yue, L. Chen, X. Wang, D. Lu, W. Zhou, D. Shen, Q. Yang, S. Xiao and Y. Li, Chem. Eng. J., 2023, 451, 138687 CrossRef CAS.
  187. P. Tan, R. Gao, Y. Zhang, N. Han, Y. Jiang, M. Xu, S. J. Bao and X. Zhang, J. Colloid Interface Sci., 2023, 630, 363–371 CrossRef CAS PubMed.
  188. D. Xu, Z. Zhang, K. Tao and L. Han, Dalton Trans., 2023, 52, 2455–2462 RSC.
  189. J. Liu, Q. Xia, L. Wang, Q. Hu, N. M. Shinde and A. Zhou, ACS Appl. Mater. Interfaces, 2024, 16, 49380–49391 CrossRef CAS PubMed.
  190. M. Murtaza, K. Farooq, W. A. Shah, I. Ahmad and A. Waseem, Nanoscale Adv., 2024, 6, 6317–6327 RSC.
  191. K. Yu, J. Zhang, Y. Hu, L. Wang, X. Zhang and B. Zhao, Catalysts, 2024, 14, 184 CrossRef CAS.
  192. Z. Li, S. Ning, F. Hu, H. Zhu, L. Zeng, L. Chen, X. Wang, T. Fujita and Y. Wei, J. Colloid Interface Sci., 2023, 629, 97–110 CrossRef CAS PubMed.
  193. M. Adil, A. G. Olabi, M. A. Abdelkareem, H. Alawadhi, A. Bahaa, K. ElSaid and C. Rodriguez, J. Energy Storage, 2023, 68, 107817 CrossRef.
  194. M. Murtaza, K. Farooq, A. A. Amjad, S. S. A. Shah and A. Waseem, Diam. Relat. Mater., 2024, 147, 111379 CrossRef CAS.
  195. V. A. Thai, T. B. Nguyen, C. W. Chen, X. T. Bui, R. Doong and C. D. Dong, Environ. Sci. Nano, 2024, 11, 3871–3886 RSC.
  196. S. Li, L. Zhang, P. Ye, M. Zhu, Y. Nie, Y. Dai and F. Yang, Cryst. Growth Des., 2024, 24, 7445–7454 CrossRef CAS.
  197. M. Ali, A. M. Afzal, S. Mumtaz, K. Yusuf and A. Deifalla, J. Energy Storage, 2024, 92, 111972 CrossRef.
  198. S. Siva Shalini, Y. P. Fu and A. Chandra Bose, Chem. Eng. J., 2024, 500, 156751 CrossRef CAS.
  199. K. Gong, L. Cai, C. Shi, F. Gao, L. Yin, X. Qian and K. Zhou, Composites, Part A, 2022, 161, 107109 CrossRef CAS.
  200. Y. Chen, S. Li, L. Zhang, T. Jing, J. Wang, L. Zhao, F. Li, C. Li and J. Sun, J. Solid State Chem., 2022, 308, 122919 CrossRef CAS.
  201. S. Li, H. Chai, L. Zhang, Y. Xu, Y. Jiao and J. Chen, J. Colloid Interface Sci., 2023, 642, 235–245 CrossRef CAS PubMed.
  202. Y. Zhang, H. Yuan, X. Chen, Z. Jiang, J. Lu and F. Xin, ACS Appl. Polym. Mater., 2024, 6, 8277–8290 CrossRef CAS.
  203. S. Li, Q. Cheng, P. Ye, Y. Zhang, L. Zhang, F. Liu, H. Qiu, X. Qu and Y. Nie, Appl. Organomet. Chem., 2024, 38, e7547 CrossRef CAS.
  204. K. Zhao, B. Zhang, X. Cui, X. Chao, F. Song, H. Chen and B. He, Food Chem., 2024, 461, 140828 CrossRef CAS PubMed.
  205. F. Yu, X. Yu, P. Jia, X. Zhou, H. Wang, Y. Hou, B. Wang, L. Song and Y. Hu, Chem. Eng. J., 2024, 499, 156145 CrossRef CAS.
  206. Y. Lin, W. Dong, S. Li, S. Zhang, X. Chen, B. Wang and L. Du, Polym. Degrad. Stab., 2025, 234, 111251 CrossRef CAS.
  207. G. Kaur, S. Sharma, N. Bhardwaj, M. K. Nayak and A. Deep, Food Control, 2024, 165, 110694 CrossRef CAS.
  208. W. Wang, Y. Bai, P. Yang, S. Yuan, F. Li, W. Zhao, B. Jin, X. Zhang, S. Liu, D. Yuan and Q. Zhao, Sci. Bull., 2023, 68, 2180–2189 CrossRef CAS PubMed.
  209. G. Dai, Y. Li, Z. Li, J. Zhang, X. Geng, F. Zhang, Q. Wang and P. He, ACS Appl. Nano Mater., 2022, 5, 9201–9208 CrossRef CAS.
  210. H. Huang, J. Qin, C. Liu, L. Luo, Y. Lan, L. Yang, J. Zhang and H. He, Carbon, 2024, 226, 119171 CrossRef CAS.
  211. W. Wen, Y. Dang, J. Tang, C. Su, S. Yu, J. Ma and Y. Zhou, J. Electrochem. Soc., 2024, 171, 057514 CrossRef CAS.
  212. X. Zhang, H. Wang, R. Dai, P. Zhao and Y. Wang, J. Colloid Interface Sci., 2025, 690, 137278 CrossRef CAS PubMed.
  213. Z. H. Lu, Q. Tian, D. D. Zhou, M. Chen, Y. W. Cao, L. Y. Zhuang, X. Liu, Z. H. Yang and I. A. Senosy, J. Environ. Chem. Eng., 2022, 10, 108037 CrossRef CAS.
  214. J. Chi, P. Ju, F. Bi, T. Jiang, S. Wen, Y. Cai, L. Wang and M. Qiu, Adv. Funct. Mater., 2025, 35, 2415174 CrossRef CAS.
  215. X. Tu, F. Gao, X. Ma, J. Zou, Y. Yu, M. Li, F. Qu, X. Huang and L. Lu, J. Hazard. Mater., 2020, 396, 122776 CrossRef CAS PubMed.
  216. S. M. S. Rana, M. T. Rahman, M. A. Zahed, S. H. Lee, Y. Do Shin, S. Seonu, D. Kim, M. Salauddin, T. Bhatta, K. Sharstha and J. Y. Park, Nano Energy, 2022, 104, 107931 CrossRef CAS.
  217. A. Rajan, M. Daniel, J. Rafi, A. R. S. C. Lazuli and B. Neppolian, ChemSusChem, 2024, 18, e202400207 CrossRef PubMed.
  218. Y. Qi, Y. Chen, Q. Li, X. Dang and H. Chen, Anal. Chim. Acta, 2024, 1316, 342876 CrossRef CAS PubMed.
  219. H. Hassan, K. Khan, S. Mumtaz, A. A. Rafi, A. M. Idris, A. Althobaiti and A. Mohammad, J. Power Sources, 2024, 623, 235486 CrossRef CAS.
  220. Z. F. Lin, T. H. Wang, P. Venkatesan and R. A. Doong, Chem. Eng. J., 2025, 505, 159411 CrossRef CAS.
  221. X. Zhao, M. Zhu, C. Tang, K. Quan, Q. Tong, H. Cao, J. Jiang, H. Yang and J. Zhang, J. Colloid Interface Sci., 2022, 620, 478–485 CrossRef CAS PubMed.
  222. L. Guo, L. Tao and T. Wang, Langmuir, 2024, 40, 8862–8871 CrossRef CAS PubMed.
  223. X. Liang, Y. Yan, G. Liu, S. Hou and S. Hou, Chem. Eng. J., 2024, 483, 149301 CrossRef CAS.
  224. S. Hou, M. Wen, J. Wei, L. Li and G. Liu, J. Mol. Struct., 2025, 1321, 140076 CrossRef CAS.
  225. X. Zhu, F. Liu, L. Meng, Q. Gao, X. Wang, M. Lou, X. Xu, W. Zhang, F. Li and B. Van der Bruggen, Nano Lett., 2025, 25, 2810–2819 CrossRef CAS PubMed.
  226. H. Yang, P. Zhang, Q. Zheng, A. Hayat, H. S. M. Abd-Rabboh, S. Raza, D. Li and Y. Sui, J. Mater. Chem. C, 2024, 12, 17620–17634 RSC.
  227. C. Sun, D. Lan, Z. Jia, Z. Gao and G. Wu, Small, 2024, 20, 2405874 CrossRef CAS PubMed.
  228. H. Yang, S. Yin, P. Zhang, Q. Zheng, S. Raza, D. Li and Y. Sui, Sep. Purif. Technol., 2025, 365, 132600 CrossRef CAS.
  229. J. Bian, M. Liu, X. Liu, X. Bian, C. Gu, J. Ma and T. Jiang, Microchem. J., 2025, 210, 113021 CrossRef CAS.
  230. H. Gao, J. You, H. Wu and M. Tian, Food Chem., 2025, 473, 143092 CrossRef CAS PubMed.
  231. J. Wang, Y. Zhang, Z. Chen, Z. Zhou, J. Wang, C. Ma, W. Qiao, Z. Xu and L. Ling, J. Colloid Interface Sci., 2025, 678, 79–88 CrossRef CAS PubMed.
  232. H. B. Parse, I. Patil, B. Kakade and A. Swami, Energy Fuels, 2021, 35, 17909–17918 CrossRef CAS.
  233. X. Han, Y. Huang, L. Ding, Y. Song, T. Li and P. Liu, ACS Appl. Nano Mater., 2021, 4, 691–701 CrossRef CAS.
  234. Z. Xiang, Y. Shi, X. Zhu, L. Cai and W. Lu, Nano-Micro Lett., 2021, 13, 150 CrossRef CAS PubMed.
  235. Z. Liu, J. Chen, M. Que, H. Zheng, L. Yang, H. Yuan, Y. Ma, Y. Li and X. Yang, Chem. Eng. J., 2022, 450, 138442 CrossRef CAS.
  236. W. Liu, X. Liu, X. Tan, M. Zhao, Z. Liu, M. Yi, Q. Ding, Y. Ren and H. Li, Sens. Actuators, B, 2025, 433, 137577 CrossRef CAS.
  237. C. Ji, C. Yan, Z. Wei, T. Yu, J. Fan and Y. Li, Chem. Eng. J., 2024, 483, 149431 CrossRef CAS.
  238. W. Xie, Y. Wang, J. Zhou, M. Zhang, J. Yu, C. Zhu and J. Xu, Appl. Surf. Sci., 2020, 534, 147584 CrossRef CAS.
  239. B. Lei, X. Huang, P. Wu, L. Wang, J. Huang and Z. Wang, Ceram. Int., 2022, 48, 25111–25119 CrossRef CAS.
  240. C. Shi, Z. Long, C. Wu, H. Dai, Z. Li, H. Qiao, K. Liu, Q. H. Fan and K. Wang, Small, 2023, 19, 2303802 CrossRef CAS PubMed.
  241. S. R. Shingte, V. D. Chavan, R. P. Dhavale, D.-K. Kim, H.-H. Park, T. D. Dongale and P. B. Patil, J. Energy Storage, 2024, 92, 112169 CrossRef.
  242. M. Liu, J. Wang, P. Song, J. Ji and Q. Wang, Sens. Actuators, B, 2022, 361, 131755 CrossRef CAS.
  243. S. Zhang, Y. Ding, Q. Wang and P. Song, Sens. Actuators, B, 2023, 393, 134122 CrossRef CAS.
  244. A. Wei, L. Wang and Z. Li, J. Alloys Compd., 2022, 899, 163369 CrossRef CAS.
  245. T. Kshetri, D. D. Khumujam, T. I. Singh, Y. S. Lee, N. H. Kim and J. H. Lee, Chem. Eng. J., 2022, 437, 135338 CrossRef CAS.
  246. W. Wu, C. Zhao, H. Liu, T. Liu, L. Wang and J. Zhu, J. Colloid Interface Sci., 2022, 623, 216–225 CrossRef CAS PubMed.
  247. X. Bu, F. Ma, Q. Wu, H. Wu, Y. Yuan, L. Hu, C. Han, X. Wang, W. Liu and X. Li, Sens. Actuators, B, 2022, 369, 132232 CrossRef CAS.
  248. H. Geun Oh and S. K. Park, Appl. Surf. Sci., 2021, 564, 150415 CrossRef CAS.
  249. Y. Zhou, Y. Wu, D. Guo, J. Li, D. Chu, S. Na, M. Yu, D. Li, G. Sui and D. F. Chai, J. Alloys Compd., 2023, 963, 171133 CrossRef CAS.
  250. M. T. Rahman, S. M. S. Rana, M. Salauddin, M. A. Zahed, S. Lee, E. S. Yoon and J. Y. Park, Nano Energy, 2022, 100, 107454 CrossRef CAS.
  251. J. Wang, Y. Huang, S. Zhang, X. Du, Z. Duan and X. Sun, J. Mater. Sci. Technol., 2023, 147, 112–123 CrossRef CAS.
  252. D. Wang, D. Zhang, Y. Yang, Q. Mi, J. Zhang and L. Yu, ACS Nano, 2021, 15, 2911–2919 CrossRef CAS PubMed.
  253. D. Song, X. Jiang, Y. Li, X. Lu, S. Luan, Y. Wang, Y. Li and F. Gao, J. Hazard. Mater., 2019, 373, 367–376 CrossRef CAS PubMed.
  254. H. Zong, L. Hu, Z. Wang, R. Qi, K. Yu and Z. Zhu, Chem. Eng. J., 2021, 416, 129102 CrossRef CAS.
  255. X. Liu, F. Liu, X. Zhao and L. Z. Fan, J. Mater., 2022, 8, 30–37 Search PubMed.
  256. H. Zong, R. Qi, K. Yu and Z. Zhu, Electrochim. Acta, 2021, 393, 139068 CrossRef CAS.
  257. J. Li, C. Chen, Z. Lv, W. Ma, M. Wang, Q. Li and J. Dang, J. Mater. Sci. Technol., 2023, 145, 74–82 CrossRef CAS.
  258. H. Wang, J. Zhao, J. Yu and Z. Wang, Synth. Met., 2023, 292, 117215 CrossRef CAS.
  259. Y. Ma, H. Zhang, H. Cong, Y. Wang, S. Xin, P. Xu, Y. Qin, G. Xu, X. Kang and F. Wang, J. Alloys Compd., 2023, 968, 171985 CrossRef CAS.
  260. Y. Cui, L. Zhao, D. He, J. Sun, J. Yang, W. Tang, H. Yu, C. Lou, W. Wang, X. Zhang and H. Zhao, ACS Appl. Mater. Interfaces, 2024, 16, 47751–47762 CrossRef CAS PubMed.
  261. M. Zhang, S. Yun, T. Yang, G. Yang, M. Sun, J. Dang, Z. Wang, H. Yang, S. Yuan, A. Arshad and K. Wang, J. Mater. Chem. A, 2024, 12, 29479–29492 RSC.
  262. T. Hou, Z. Jia, B. Wang, H. Li, X. Liu, Q. Chi and G. Wu, Chem. Eng. J., 2021, 422, 130079 CrossRef CAS.
  263. Y. Yang, X. Huang, C. Sheng, Y. Pan, Y. Huang and X. Wang, J. Alloys Compd., 2022, 920, 165908 CrossRef CAS.
  264. H. Li, J. Li, L. Ma, X. Zhang, J. Li, J. Li, T. Lu and L. Pan, J. Mater. Chem. A, 2023, 11, 2836–2847 RSC.
  265. X. Shi, W. Liang, G. Liu, B. Chen, L. Shao, Y. Wu, Z. Sun and F. García, Chem. Eng. J., 2023, 462, 142271 CrossRef CAS.
  266. L. Wang, Z. Liu, Y. Ma, Z. Li, M. Xiao, B. Tu and H. Song, Dalton Trans., 2023, 572–581 Search PubMed.
  267. W. Liang, M. Liu, X. Shi, B. Chen, L. Shao, J. Xu and Z. Sun, Batter. Supercaps, 2023, 7, e202300462 CrossRef.
  268. R. Hu, D. Sha, X. Cao, C. Lu, Y. Wei, L. Pan and Z. M. Sun, Adv. Energy Mater., 2022, 12, 2203118 CrossRef CAS.
  269. J. H. Na, H. G. Oh, S. Lee and S. K. Park, J. Mater. Chem. A, 2023, 12, 2848–2855 RSC.
  270. S. Vallem, S. Song, Y. Oh, J. Kim, M. Li, Y. Li, X. Cheng and J. Bae, J. Colloid Interface Sci., 2024, 665, 1017–1028 CrossRef CAS PubMed.
  271. Y. Xue, W. Li, R. Liu, T. Zhao, Y. Lv, J. Li, Y. Guo, R. Hao, X. Liu and H. He, Batter. Supercaps, 2025, e202400770 CrossRef CAS.
  272. H. Zou, B. He, P. Kuang, J. Yu and K. Fan, ACS Appl. Mater. Interfaces, 2018, 10, 22311–22319 CrossRef CAS PubMed.
  273. J. Wei, F. Hu, C. Lv, L. Bian, X. Quan and Q. Ouyang, Small, 2025, 21, 2411146 CrossRef CAS.
  274. Q. Li, Q. Jiao, Y. Yan, H. Li, W. Zhou, T. Gu, X. Shen, C. Lu, Y. Zhao, Y. Zhang, H. Li and C. Feng, Chem. Eng. J., 2022, 450, 137922 CrossRef CAS.
  275. M. Adil, A. G. Olabi, M. A. Abdelkareem, H. Alawadhi, A. Bahaa, K. ElSaid and C. Rodriguez, J. Energy Storage, 2023, 60, 106537 CrossRef.
  276. K. Farooq, M. Murtaza, Z. Yang, A. Waseem, Y. Zhu and Y. Xia, Nanoscale Adv., 2024, 6, 3169–3180 RSC.
  277. L. Hu, R. Xiao, X. Wang, X. Wang, C. Wang, J. Wen, W. Gu and C. Zhu, Appl. Catal., B, 2021, 298, 120599 CrossRef CAS.
  278. K. Kasinathan, Y. K. Park, B. Ravindran, K. Marimuthu, G. Munuswamy-Ramanujam, S. Woong Chang and J. H. Yim, Chem. Eng. J., 2024, 482, 149088 CrossRef CAS.
  279. A. Alshraim, S. Rahman, M. Al-Gawati, A. N. Grace and A. N. Alodhayb, Int. J. Electrochem. Sci., 2025, 20, 100980 CrossRef CAS.
  280. Y. Zhang, Z. Li, B. Qu, X. Shen, L. Tong, J. Wang, J. Cui, X. Li, Q. Xie and J. Wang, Inorg. Chem., 2025, 64, 6822–6831 CrossRef CAS PubMed.
  281. X. Liu, Y. Zhang and Y. Q. Liu, Appl. Surf. Sci., 2025, 694, 162859 CrossRef CAS.
  282. J. Zhou, G. Zhang, J. Luo, Y. Hu, G. Hao, H. Guo, F. Guo, S. Wang and W. Jiang, Chem. Eng. J., 2021, 426, 130725 CrossRef CAS.
  283. W. Gu, M. Wu, J. Xu and T. Zhao, Int. J. Hydrogen Energy, 2022, 47, 17224–17232 CrossRef CAS.
  284. T. Wu, F. Ren, Z. Guo, J. Zhang, X. Hou, Z. Chen, Y. Jin and P. Ren, J. Alloys Compd., 2024, 976, 172984 CrossRef CAS.
  285. Q. Liao, M. He, Y. Zhou, S. Nie, Y. Wang, B. Wang, X. Yang, X. Bu and R. Wang, Langmuir, 2018, 34, 15854–15863 CrossRef CAS PubMed.
  286. B. Deng, Z. Xiang, J. Xiong, Z. Liu, L. Yu and W. Lu, Nano-Micro Lett., 2020, 12, 55 CrossRef CAS PubMed.
  287. M. Wang, P. Zhou, T. Feng, Z. Song, Q. Ren, H. Gu, X. Shi, Q. Zhang and L. Wang, Ceram. Int., 2022, 48, 22681–22690 CrossRef CAS.
  288. C. Cui, L. Geng, S. Jiang, W. Bai, L. Dai, E. Ren, L. Liu and R. Guo, Ind. Eng. Chem. Res., 2023, 62, 8297–8311 CrossRef CAS.
  289. J. Zheng, C. Xu, Z. Li, C. Gu, X. Li, Z. Li, Y. Li, G. Lou and Y. Chen, J. Mater. Chem. A, 2024, 12, 29103–29112 RSC.
  290. Z. Guo, Y. Zhao, P. Luo, J. Wang, L. Pei, H. Yu, P. Song, F. Ren and P. Ren, Chem. Eng. J., 2024, 497, 155707 CrossRef CAS.
  291. Y. Yao, Y. Ma, C. Chen, K. Zhu, G. Wang, D. Cao and J. Yan, J. Energy Storage, 2025, 114, 115877 CrossRef.
  292. Q. Liu, H. Zhang, J. Xie, F. Yang, Z. Yang, X. Liu, H. Wu, Q. Liu and X. Lu, Nano Energy, 2023, 113, 108567 CrossRef CAS.
  293. Y. Wang, R. Niu, L. Chen, Y. Yang, H. Yu and X. Qiu, Chem. Eng. J., 2025, 512, 162305 CrossRef CAS.
  294. M. Daniel, G. Mathew, M. Anpo and B. Neppolian, Coord. Chem. Rev., 2022, 468, 214627 CrossRef CAS.
  295. B. Mohanty, S. Kumari, P. Yadav, P. Kanoo and A. Chakraborty, Coord. Chem. Rev., 2024, 519, 216102 CrossRef CAS.
  296. Y. Lu, J. Sun, W. Xu, X. Di, W. Zhao, L. Zhang and G. Pan, IEEE Sens. J., 2025, 25, 5294–5303 CAS.
  297. S. Vallem, S. Venkateswarlu, Y. Li, S. Song, M. Li and J. Bae, Energy Storage Mater., 2024, 65, 103159 CrossRef.
  298. K. Farooq, Z. Yang, M. Murtaza, M. A. Naseeb, A. Waseem, Y. Zhu and Y. Xia, Adv. Energy Sustain. Res., 2025, 2400400 CrossRef CAS.
  299. J. Liu, L. Chen, H. Cui, J. Zhang, L. Zhang and C.-Y. Su, Chem. Soc. Rev., 2014, 43, 6011–6061 RSC.
  300. Y. Wang, J. Chen, J. Shi, J. Zhang, Y. Ma, X. Yang and Y. Li, Appl. Surf. Sci., 2025, 710, 163965 CrossRef CAS.
  301. B. Shan, Y. Wang, X. Ji and Y. Huang, Nano-Micro Lett., 2024, 16, 212 CrossRef PubMed.
  302. F. Wu, P. Hu, F. Hu, Z. Tian, J. Tang, P. Zhang, L. Pan, M. W. Barsoum, L. Cai and Z. Sun, Nano-Micro Lett., 2023, 15, 194 CrossRef CAS PubMed.
  303. G. Song, Y. Shi, S. Jiang and H. Pang, Adv. Funct. Mater., 2023, 33, 2303121 CrossRef CAS.
  304. J. Liu, T. A. Goetjen, Q. Wang, J. G. Knapp, M. C. Wasson, Y. Yang, Z. H. Syed, M. Delferro, J. M. Notestein, O. K. Farha and J. T. Hupp, Chem. Soc. Rev., 2022, 51, 1045–1097 RSC.
  305. S. Bai, M. Yang, J. Jiang, X. He, J. Zou, Z. Xiong, G. Liao and S. Liu, npj 2D Mater. Appl., 2021, 5, 78 CrossRef CAS.
  306. Z. Qian, R. Zhang, Y. Xiao, H. Huang, Y. Sun, Y. Chen, T. Ma and X. Sun, Adv. Energy Mater., 2023, 13, 2300086 CrossRef CAS.
  307. A. Fattah-alhosseini, Z. Sangarimotlagh, M. Karbasi and M. Kaseem, Nano-Struct. Nano-Objects, 2024, 38, 101192 CrossRef CAS.
  308. Y. Chen, Y. Qiu, T. Chen and H. Wang, ACS Nano, 2025, 19, 6588–6600 CrossRef CAS PubMed.
  309. M. Garai, M. Mahato, S. Nam, E. Kim, D. Seo, Y. Lee, V. H. Nguyen, S. Oh, P. Sambyal, H. Yoo, A. K. Taseer, S. A. Syed, H. Han, C. W. Ahn, J. Kim and I. K. Oh, Adv. Funct. Mater., 2023, 33, 2212252 CrossRef CAS.
  310. F. Wei, F. Liang, Y. Zhao, Z. Ji, T. Yan, R. Li, H. Liu, Y. Kong, H. He, W. Huang, C. Cao, W. Zhang, B. Fei and M. Ge, J. Mater. Chem. A, 2025, 13, 13070–13080 RSC.
  311. L. Sun, H. Wang, S. Zhai, J. Sun, X. Fang, H. Yang, D. Zhai, C. Liu, W. Q. Deng and H. Wu, J. Energy Chem., 2023, 76, 368–376 CrossRef CAS.
  312. H. Zhang, D. Xiong, Y. Xie, K. Wu, Z. Feng, K. Wen, Z. Li and M. He, Colloids Surf., A, 2023, 656, 130332 CrossRef CAS.
  313. W. Wang, J. Si, L. Yue, W. Zhang, Q. Chen, H. Huang, F. Lin, Y. Zhao, S. Meng and L. Liu, Chem. Eng. J., 2025, 507, 160363 CrossRef CAS.
  314. J. Sun, L. Zhang, F. Li, F. Yang, M. Liu, S. Li and D. Zhang, Adv. Funct. Mater., 2025, 2501181 CrossRef CAS.
  315. J. Li, X. Qiang, B. Jia, L. Wang and X. Wu, Appl. Surf. Sci., 2025, 695, 162867 CrossRef CAS.
  316. M. Saeed, S. Sharif, J. H. Shah, T. T. R. Afzal, M. Shahbaz, A. M. Shehzad, A. Shahzad, O. Şahin and S. Shahzad, RSC Adv., 2025, 15, 17130–17141 RSC.
  317. J. Kunwar, D. Acharya, K. Chhetri, B. Karki, B. Deo, R. M. Bhattarai, S. Neupane, M. P. Adhikari and A. P. Yadav, J. Electroanal. Chem., 2023, 950, 117915 CrossRef CAS.
  318. C. Ruan, D. Zhu, J. Qi, Q. Meng, F. Wei, Y. Ren, Y. Sui and H. Zhang, Surf. Interfaces, 2021, 25, 101274 CrossRef CAS.
  319. I. Pathak, D. Acharya, K. Chhetri, P. Chandra Lohani, S. Subedi, A. Muthurasu, T. Kim, T. H. Ko, B. Dahal and H. Y. Kim, J. Mater. Chem. A, 2023, 11, 5001–5014 RSC.
  320. C. Liu, W. Feng, Y. Bai and H. Pang, Inorg. Chem. Front., 2022, 9, 5463–5468 RSC.
  321. N. Hussain and S. M. Mobin, Chem. Sci., 2025, 16, 8460–8469 RSC.
  322. Y. Shi, F. Dong, A. Rodas-Gonzalez, G. Wang, L. Yang, S. Chen, H.-B. Zheng and S. Wang, Food Chem., 2025, 475, 143362 CrossRef CAS PubMed.
  323. Y. Wang, X. Wang, M. Chu, J. Xin, Y. Liu, H. Pang, G. Yang and H. Ma, Anal. Chim. Acta, 2025, 1346, 343770 CrossRef CAS PubMed.
  324. X. Wang, X. Bai, W. Wang, Z. Zhao and J. Shan, J. Electrochem. Soc., 2023, 170, 037519 CrossRef CAS.
  325. M. Ravipati, D. Ramasamy and S. Badhulika, Electrochim. Acta, 2025, 525, 146154 CrossRef CAS.
  326. X. Zheng, M. Yuan, P. Su, J. Xu and G. Sun, Appl. Surf. Sci., 2025, 692, 162742 CrossRef CAS.
  327. M. Murtaza, K. Farooq, W. A. Shah and A. Waseem, Fuel, 2025, 394, 135095 CrossRef CAS.
  328. Z. H. Zhu, X. Y. Wu, J. F. Lu, H. Xu, S. L. Hou, J. Zhao, S. J. Liu and H. R. Wen, Inorg. Chem., 2025, 64, 8261–8269 CrossRef CAS PubMed.
  329. A. P. Tiwari, P. Chandra, M. S. Rahman, K. A. Mirica and W. J. Scheideler, Nanoscale, 2025, 17, 11028–11036 RSC.
  330. J. Jiang, C. Xi, Y. Zhao, Y. Zhu, K. Hu, Y. Wang, S. Wang, Z. Zhang and S. Han, ACS Catal., 2025, 15, 2561–2575 CrossRef CAS.
  331. I. Pathak, D. Acharya, K. Chhetri, Y. R. Rosyara, A. Muthurasu, T. Kim, T. H. Ko and H. Y. Kim, Composites, Part B, 2025, 296, 112238 CrossRef CAS.
  332. T. Fang, J. Ma, Y. Dai, J. Gou, X. Li, C. Li and J. Tong, ACS Appl. Nano Mater., 2025, 8, 3519–3526 CrossRef CAS.
  333. S. Raza, A. U. Rehman, C. Chen, T. Zhao, A. Hayat, T. Bashir, L. Shen, Y. Orooji and H. Lin, J. Mater. Chem. A, 2025, 13, 4390–4403 RSC.
  334. X. Guo, H. Zhang, Y. Yao, C. Xiao, X. Yan, K. Chen, J. Qi, Y. Zhou, Z. Zhu, X. Sun and J. Li, Appl. Catal., B, 2023, 323, 122136 CrossRef CAS.
  335. W. Zhan, L. Ma and M. Gan, Mater. Today Chem., 2022, 26, 101041 CrossRef CAS.
  336. Y. Zhou, Y. Duan, Y. Liu, Y. Geng, Z. Jin, L. Yu and S. Li, Appl. Surf. Sci., 2025, 691, 162621 CrossRef CAS.
  337. Z. Lin, H. Xiao, Y. Chen and H. Cai, J. Alloys Compd., 2025, 1020, 179368 CrossRef CAS.
  338. Q. Peng, W. Yu, C. Gao, L. Geng, P. Fatehi, S. Wang and F. Kong, Adv. Compos. Hybrid Mater., 2025, 8, 232 CrossRef CAS.
  339. G. Zhang, G. Zhang, H. Zhang, Y. Lu, Q. Shi, R. Li and J. Song, Appl. Surf. Sci., 2025, 688, 162327 CrossRef CAS.
  340. Y. Zhu, Y. Yang, S. Li, C. Zhu, Z. Qiu and Q. Zhu, New J. Chem., 2025, 49, 2889–2897 RSC.
  341. M. H. Memon, A. Mir, F. Rehman, U. Amjad and M. Mustafa, Energy Technol., 2025, 2500041 CrossRef.
  342. P. Wang, B. He, B. Wang, S. Liu, Q. Ye, F. Zhou and W. Liu, Chem. Eng. J., 2023, 474, 145835 CrossRef CAS.
  343. A. S. Alqahtani, A. S. Alahmari, Z. E. Eldin and S. Elbeltagi, Inorg. Chem. Commun., 2025, 176, 114128 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.