Next Article in Journal
Unsupported NiO Nanoflowers for Enhanced Methane Activation and Direct Conversion to C2–C6 Hydrocarbons
Previous Article in Journal
Photocatalytic Degradation of Ant Trail Pheromones by P25 TiO2
Previous Article in Special Issue
Thiazolo[5,4-d]thiazole-Based Covalent Organic Frameworks for the Rapid Removal of RhB
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Preparations and Multifunctional Applications Towards MOF/GDY Composites and Their Derivative Materials

1
School of Science, China University of Geosciences (Beijing), Beijing 100083, China
2
Hubei Key Laboratory of Processing and Application of Catalytic Materials, Huanggang Normal University, Huanggang 438000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1041; https://doi.org/10.3390/catal15111041
Submission received: 28 September 2025 / Revised: 21 October 2025 / Accepted: 30 October 2025 / Published: 2 November 2025
(This article belongs to the Special Issue Multifunctional Metal–Organic Framework Materials as Catalysts)

Abstract

Metal–organic frameworks (MOFs) are novel porous crystalline materials formed through the self-assembly of metal ions and organic ligands. They have various advantages, including tunable chemical and electronic structures, high porosity, and large specific surface areas. Owing to their unique structural and physicochemical properties, MOFs have been widely applied in the fields of catalysis, supercapacitors, sensors, and drug recognition/delivery. However, the intrinsic poor stability and low electrical conductivity of conventional MOFs severely hinder their practical implementation. Graphdiyne (GDY), a unique carbon allotrope, features a new structure composed of both sp2- and sp-hybridized carbon atoms. Its distinct chemical and electronic configuration endow it with exceptional properties such as natural bandgap, uniform in-plane cavities, and excellent electronic conductivity. Integrating MOFs with GDY can effectively overcome the intrinsic limitations of MOFs and expand their potential applications. As emerging hybrid materials, MOF/GDY composites and their derivatives have attracted increasing attention in recent years. This article reviews recent advances in the synthesis strategies of MOF/GDY composites and their derivatives, along with their performance and applications in catalysis, energy storage, and biological sensors. It also discusses the future opportunities and challenges faced in the development of these promising composite materials, aiming to inspire interest and provide scientific guidance.

1. Introduction

Metal–organic frameworks (MOFs) are a class of crystalline porous materials constructed from transition metal ions or clusters coordinated with organic ligands, forming unique inorganic/organic hybrid structures [1,2,3,4]. Owing to their well-defined pore architectures, high specific surface areas, and abundant active sites, MOFs have shown great promise in a range of applications, including batteries [5,6,7], catalysis [8,9,10], gas storage and separation [11,12], and chemical sensing [13,14,15]. Nevertheless, their practical implementation is often hindered by poor electrical conductivity and limited structural stability [16,17,18]. For instance, as the most typical representative of MOFs, the zeolitic imidazolate frameworks (ZIFs) series [19,20,21,22] exhibit outstanding thermal stability and porosity, while the materials of the institute Lavoisier (MIL) series [23,24,25,26,27,28] are characterized by large surface areas and ordered pores. However, they are also facing significant challenges [29]. To overcome these limitations, MOFs have been integrated with other advanced materials to leverage the strengths of each component and generate synergistic effects. In recent years, researchers have achieved notable progress by compositing MOFs with diverse functional materials. This strategy leverages relatively straightforward synthetic approaches to combine material benefits and overcome their inherent limitations. Common examples include MOF/metal nanoparticle composites [30,31,32,33,34], MOF/organic polymer composites [35,36,37,38,39], and MOF/carbon-based composites [40,41,42,43,44]. Among these, MOF/carbon composites, involving activated carbon [45,46,47], graphene [48,49,50,51,52,53], carbon nanotubes [54,55,56], and fullerenes [57,58], have attracted considerable research interest. The combination of MOFs with carbon materials effectively integrates the high porosity, tunable structure, and abundant active sites of MOFs with the excellent electrical conductivity, chemical stability, and mechanical strength of carbon materials. Such hybridization not only enhances the intrinsic conductivity and structural stability of MOFs but also promotes overall durability and electrochemical performance. As a result, MOF/carbon composites exhibit superior synergistic effects and have shown great potential in various applications [59].
Graphdiyne (GDY) [60,61,62,63], a novel carbon allotrope, features a unique structure composed of both sp2- and sp-hybridized carbon atoms, distinguishing it from conventional carbon materials. This unique configuration imparts distinct electronic and chemical properties, offering advantages in terms of controlled synthesis and performance. Over the past decade, GDY has emerged as a promising material across diverse domains including catalysis [64,65,66,67,68], energy storage [69,70,71,72,73], and electronic devices [74,75,76]. GDY-based composites have emerged as a vibrant advancement in materials chemistry. Owing to diverse synthesis strategies and improved properties, research on GDY composites has entered a phase of rapid advancement [77,78,79,80,81], leading to significant progress being made in terms of both fundamental understanding and practical applications.
Recent studies have confirmed that integrating MOFs with GDY is an effective strategy for significantly enhancing conductivity and structural stability while also unlocking novel synergistic properties. These improvements broaden the potential of MOF-based materials in energy storage/conversion and biorecognition. While many reviews have summarized MOF/traditional carbon composites, few have focused on MOF/GDY composites and their derivatives. This review systematically summarizes the recent research advancements in MOF/GDY composite materials and their derivatives (Scheme 1), especially regarding the preparation strategies and their applications in catalysis, energy storage, and biological sensors. By summarizing the composite methods and applications for MOF/GDY composites and derivative materials, we can conclude that this should further promote the maturation of composite materials based on inorganic/organic advanced materials and unique GDY-based carbon composites. Finally, the article concludes with a discussion of the development prospects and challenges. Despite promising early findings, research in this area is still in its infancy. Further work is needed to refine composite strategies, deepen performance evaluation, and expand application domains. We hope that this review will inspire broader interest and more in-depth investigations into the emerging composite materials.

2. Preparations of MOF/GDY Composite Materials

Various synthesis strategies have been developed to prepare MOF/GDY-based composites with diverse structures, morphologies, and applications. Generally, these methods can be categorized into a physical mixing strategy and in situ growth strategy. Based on the growth sequence, the in situ growth can be further classified into two types: MOFs in situ anchored on GDY and GDY in situ grown on MOFs.

2.1. Physical Mixing Strategy

The physical mixing strategy involves combining pre-synthesized MOFs with GDY to form composite materials, typically through stirring, grinding, or ultrasonic treatment. This method is straightforward and remains the most widely used and reported approach in current scientific research. In 2023, Wang et al. [82] reported the preparation of a novel 2D/3D GDY/ZnCo-ZIF composite via a simple low-temperature physical mixing method (Figure 1a). In this work, a specific amount of prepared GDY and ZnCo-ZIF was dispersed in a solution and thoroughly mixed using ultrasonic treatment to produce GDY/ZnCo-ZIF composite material. The composite can be easily adjusted by varying the ratio of GDY to ZnCo-ZIF, offering great convenience and flexibility, demonstrating the advantages of the composite strategy. The study demonstrated that integrating GDY with ZnCo-ZIF successfully formed an S-scheme heterojunction, which effectively suppressed the recombination of photoexcited electron–hole pairs and enhanced their separation efficiency. Furthermore, the GDY/ZnCo-ZIF composite exhibited increased charge carrier availability at active sites for proton reduction, thereby improving its hydrogen evolution performance.
Due to the stability of the materials, the physical mixing strategy was widely applicable to the composite of various MOFs and GDY, resulting in reported GDY/CoMo-MOF (Figure 1b) [83], Fe-MOF@GDY (Figure 1c) [84], and GDY/CuS/ZnCoS2 [85] composite materials and derivatives. The primary advantage of the physical mixing strategy lies in its simplicity and ease of operation. In the composite process, achieving a fully integrated material relies on the uniform dispersion of MOFs and GDY materials in solution. However, it is difficult to use this simple strategy to achieve perfect regulation and compositing, which significantly limits the performance and potential applications of the resulting composite materials.

2.2. MOFs In Situ Anchored on GDY

In situ growth is one of the most widely used strategies for preparing MOF nanocomposites, particularly MOF/graphene-based composites. The presence of various functional groups (such as carboxyl and hydroxyl groups) on the surface of modified graphene enables uniform in situ MOF growth, leading to improved adhesion between MOF and graphene substrate. This strategy is also applicable to MOF/GDY composites, where MOFs can nucleate and grow uniformly on GDY substrates.
Recently, a three-dimensional NiCo-MOF was grown and anchored at the edges of a two-dimensional GDY substrate via an in situ growth method (Figure 2a) [86]. In this approach, GDY was introduced during the synthesis of NiCo-MOF and applied as a growth substrate. Owing to the zeta potential of GDY (−15.65 mV), Ni2+ and Co2+ ions can adsorb onto its surface through electrostatic interactions. Following the addition of 1,3,5-trimesic acid, NiCo-MOF can gradually form on the GDY substrate surface and be anchored at the edges, establishing a strongly coupled electron transport interface with stronger interfacial bonding and enhanced structural stability. By adjusting the amount of GDY, composite materials with varying component ratios can be obtained, showing excellent controllability.

2.3. GDY In Situ Grown on MOFs

In addition to the in situ growth of the MOF organic precursor on pre-synthesized GDY, the precursor of GDY or derivatives can also be grown in situ on the MOF surface to construct the composite material. In 2024, Zahir Abbas et al. [87] reported a crumpled, fibrous hydrogen-substituted GDY (HsGDY) in situ wrapped on the surface of Ni-MOFs (Figure 2b). During synthesis, Ni-MOF was introduced into the reaction mixture along with 1,3,5-triethynylbenzene as the reaction precursor, forming a MOF@HsGDY composite material after the Glaser coupling reaction at a relatively low temperature. In particular, the nanofibrous HsGDY was successfully constructed around the Ni-MOF, which was thoroughly characterized using spectroscopic and microscopic techniques. This novel configuration established an efficient charge transfer interface between the two materials along with enhanced electrical conductivity, thereby improving the electrons/ions accessibility to active sites. Similarly, using this in situ growth strategy, a thin layer of HsGDY was conformally seamlessly coated on the surface of NiBDC, resulting in a tandem catalyst named Ni-MOFs@HsGDY@Cu (Figure 2c) [88]. In this composite, GDY derivative ultrathin HsGDY could serve as a bridge with improved electron/ion conductivity, which connected the coordinated unsaturated Ni2+ with Cu single atoms/clusters on the surface of the Ni-MOFs. The uniform preparation strategy can greatly improve the mechanical properties of the material, which is also a factor that must be considered for practical applications.

2.4. MOF/GDY Derivative Composites

Owing to their diverse coordination geometries, MOFs present a promising platform for synthesizing materials with excellent electrochemical properties. In particular, metal compounds derived from MOFs retain their original framework with a high specific surface area and porous structure. Similarly, the combination concept has been further explored by combining MOF-derived metal compounds with GDY carbon materials. In 2023, nitrogen- and fluorine-co-doped GDY (named N-F-GDY) was integrated with NiCo2O4-Co3O4 hollow multi-shelled nanocages (Figure 3a) [89], derived from ZIF-67/Ni-Co layered double hydroxides. The resulting composite, NiCo2O4-Co3O4/N-F-GDY/GCE, was developed into an ultrasensitive electrochemical biosensor for detecting pesticide residues. Additionally, Lu et al. [90] synthesized blade-shaped Co3S4 on nickel foam (NF) via sulfurization of ZIF-67/NF, with GDY coated onto the surface through an in situ coupling reaction, forming GDY/Co3S4/NF composite material (Figure 3b), demonstrating significantly enhanced electrocatalytic properties for the oxygen evolution reaction (OER).
In summary, the in situ growth strategy allows for the more uniform integration of MOFs in situ anchored on GDY or GDY in situ grown on MOFs, resulting in stronger interfacial interactions between the components. However, this method involves a relatively complex preparation process. In contrast, the physical mixing approach offers the advantage of simplicity but often leads to the uneven distribution of MOFs and GDY derivatives, as well as weak and unstable interfacial bonding. Meanwhile, large-scale and structurally controlled synthesis of GDY materials remains a challenge. The limitation hinders the combination of GDY with other advanced materials. Due to the unique nature of novel GDY carbon materials and the current limitations of effective composite strategies, the development of GDY with MOFs and other inorganic/organic composites remains in its initial stages. In the future, more in-depth research is needed (both in materials selection and in the design of composite strategies) to achieve targeted properties and enhanced performance.

3. Applications of MOF/GDY Composites and Derivative Materials

MOF-based composites combine the strengths of MOFs with those of diverse functional materials, effectively mitigating the inherent limitations of each component. This integration creates synergistic effects and unlocks new functionalities that cannot be achieved by a single component. In particular, combining MOFs with carbon materials significantly enhances their electrical conductivity and mechanical strength, thereby broadening their applications in electrochemistry. Recently, MOFs and derivatives were combined with the new carbon allotrope-GDY. Composite MOF/GDY materials have demonstrated great promising potential in catalysis, energy storage, and biological sensors (Table 1), which will be introduced in the following section.

3.1. Application in Catalysis

Due to their tunable pore sizes, high surface areas, and remarkable catalytic properties, MOFs have garnered significant attention in the scientific field. Notably, MOF-based photocatalysts have demonstrated advantages such as excellent adsorption/desorption capabilities, particularly in photocatalytic hydrogen evolution reactions (HERs). Despite these promising attributes, the practical application of MOFs in HER is limited by inherent drawbacks, including low electrical conductivity and inadequate long-term cycling stability. To overcome these challenges, the integration of electrically conductive materials with MOFs has emerged as a promising strategy. GDY is a new carbon material that incorporates both sp– and sp2–carbon atoms into a highly conjugated framework with uniformly distributed pores and an extensive π conjugation system. These distinctive features grant GDY exceptional chemical reactivity, high electron mobility, and anisotropic electron transport, rendering it an excellent electron-conducting interface for improving HER performance. As a result, its integration into MOF-based photocatalysts [82] enhances charge transport between the composite components, thereby enhancing the overall photocatalytic efficiency. After being combined with ZnCo-ZIF, the heterojunction GDY/ZnCo-ZIF can facilitate the photocatalytic reduction reaction, with the hydrogen production of 171.79 μmol and improved photocatalytic efficiency (Figure 4a–c).
Interface engineering is an efficient method for enhancing the performance of MOFs. With its inherent ability to facilitate rapid interfacial electron transfer, GDY has been incorporated into the ZIF-9(Co)/Cu3BTC2 interface to regulate carrier migration at the interface (Figure 4d–f) [91]. The introduction of GDY significantly improved the transport efficiency of photogenerated carriers while retaining the original active sites and high specific surface area and also demonstrated the important role of material composites in improving performance. The Cu3BTC2/ZIF-9(Co)/GDY composite catalyst achieved a hydrogen production rate of 1126 μmol g−1 h−1, approximately 30 and 16 times higher than that of Cu3BTC2 and ZIF-9(Co), respectively. This outstanding performance can be attributed to the formation of an interfacial S-scheme heterojunction between ZIF-9(Co) and Cu3BTC2, which enabled strong reduction capability through band bending and the generation of an internal electric field, thereby facilitating electron transfer. Moreover, the combined GDY can act as an efficient electron transport layer, which accelerates charge carrier migration across the heterojunction. NiCo–MOF, which possesses abundant redox-active sites arising from the coexistence of bimetals and the synergistic interaction between Ni and Co, offers advantages such as facile synthesis, low cost, and excellent electrical conductivity. However, its catalytic performance remains limited, mainly due to structural instability and the intrinsic constraints of the MOF. To address these challenges, GDY was employed as a support material, leveraging its high electronegativity to in situ anchor spherical NiCo-MOF [86]. This integration formed a strongly coupled electron transport interface, significantly enhancing the spatial separation of photogenerated electron–hole pairs. Consequently, the resulting composite photocatalyst exhibited not only improved structural stability but also enhanced photocatalytic hydrogen evolution activity (Figure 4g–i). Overall, the synergistic interaction between MOFs and GDY is demonstrated to play a crucial role in enabling highly efficient hydrogen evolution.
MOF/GDY composites and derivative materials also present outstanding electrocatalytic properties [90,92]. A MOF-derived/GDY composite material, GDY/Co3S4/NF [90], can be employed as a self-supported electrode, exhibiting significantly improved electrocatalytic property for the OER. Notably, it achieved a low overpotential (223 mV at a current density of 10 mA cm−2) and a small Tafel slope (46.5 mV dec−1). The superior performance was also attributed to the synergistic interaction between GDY- and MOF-derived Co3S4, which facilitated the adsorption of key intermediate species such as OOH*, thereby enhancing the overall electrocatalytic activity for OER (Figure 5a–c).
Despite HER and OER, MOF/GDY composites and derivative materials have also shown great potential in the catalysis of nitrate-to-ammonia conversion. In the Ni-MOFs@HsGDY@Cu composite [88], HsGDY served as an efficient bridge facilitating synergistic catalysis. Hydrogen radicals generated at Ni2+ sites can spill over to adjacent Cu sites to participate in the reduction of *HNO3 to NH3, while NO2- intermediates released from Cu sites can diffuse to Ni2+ sites and further reduce to NH3 through coupling with H. As a result, the Ni-MOFs@HsGDY@Cu composite exhibited significantly improved catalytic activity and reduced overpotential for nitrate-to-ammonia conversion (Figure 5d–f). This compositing strategy, utilizing GDY-based carbon material as a conductive/reactive bridge, offered an effective platform for designing tandem catalysts with cooperative active sites, addressing the complex requirements of nitrate reduction and other multi-electron, multi-intermediate transformations.
Previous reports have demonstrated that MOFs can be applied as catalysts in many organic reactions and transformations [93], such as phenol oxidation. However, since the industrial oxidation of phenol typically occurs in aqueous media, the water stability of the MOF becomes a critical factor influencing both catalytic performance and practical applicability. Hydrophobic surface modification is a promising strategy to enhance the stability of MOFs without substantially altering their pore structures, thereby preserving their high catalytic activity in aqueous-phase reactions. To address this challenge, a novel MOF/GDY composite membrane, named HKUST-1/GDY, was fabricated via a coupling reaction using conductive copper foam (CF) as the substrate [94]. The resulting polymeric HKUST-1/GDY/CF composite membrane exhibited efficient catalytic performance for the oxidation reaction of phenol and benzyl alcohol in a fixed-bed reactor, which can be attributed to the presence of GDY facilitating electron transfer and significantly accelerating reactions such as wet peroxide oxidation of phenol, N-bromosuccinimide (NBS) oxidation of benzyl alcohol, and ring-opening of epoxides (Figure 5g–i). The result expanded the potential applications of MOF/GDY composite materials in heterogeneous catalysis, particularly in aqueous-phase systems.
These findings reveal that MOF/GDY composites effectively integrate the complementary strengths of their constituent components, offering remarkable potential across a wide range of catalytic applications. Beyond photocatalysis and electrocatalysis, their applicability can also extend to ammonia conversion and organic transformations. Continued in-depth and systematic studies in this field are expected to further advance their development.

3.2. Application in Energy Storage

Both MOF materials and GDY-based carbon materials have achieved significant and innovative progress in energy storage applications, including their utilization as electrode materials, separators, and electrolytes. Although research on MOF/GDY composites in energy storage remains relatively limited, these materials still exhibit tremendous potential advantages for future development in this field. After the crumpled fibrous HsGDY wrapped around the surface of Ni-MOFs, the MOF@HsGDY composites [87] can be applied for supercapacitor applications. The introduction of HsGDY carbon coating confers a range of advantages. First, HsGDY serves as a conductive support, facilitating efficient interfacial charge transfer. Second, the uniform porous structure of HsGDY enables the formation of abundant charge-transfer nanochannels, significantly enhancing the specific capacitance when combined with MOFs. Moreover, HsGDY offers a promising solution to the long-standing issue of MOF instability in electrochemical environments owing to its uniform porosity and excellent chemical stability. As a result, when tested using three-electrode systems in 2 M KOH as an electrolyte, MOF@HsGDY achieved remarkable electrochemical performance. The MOF@HsGDY-based electrode had a specific capacitance of 982 F g−1 and an energy density of 43.3 Wh kg−1, along with exceptional cycling stability, outperforming previously hybrid electrodes based on traditional carbons (Figure 6). The finding demonstrated that unique GDY carbon or derivatives provide a viable strategy for improving the electrochemical stability and performance of MOFs.
Recently, Shang et al. reported a novel MOF/GDY composite material in which GDY was uniformly and continuously in situ grown on the surface of ZnCo-ZIF, forming a ZnCo-ZIF@GDY composite structure [95]. This composite was employed as a modified membrane separator in lithium–sulfur (Li-S) batteries. Compared to the pristine ZnCo-ZIF, the GDY-coated composite exhibited an increased surface area and reduced pore size, attributed to the unique sp/sp2-hybridized carbon triangular pores of GDY, which can effectively suppress the polysulfide shuttle effect. Additionally, after the in situ coating of GDY, the composite structure facilitated rapid lithium-ion diffusion and enhanced ionic conductivity. As a result of these advantages, the Li-S battery equipped with Ketjen Black/sulfur composite cathode and the ZnCo-ZIF@GDY/polypropylene (PP) separator demonstrated significantly improved electrochemical performance. It delivered a high initial specific capacity of 1126.1 mA h g−1 at 0.2 C and exhibited excellent long-term cycling stability, maintaining a reversible capacity of 552.1 mA h g−1 after 300 cycles at a high rate of 2 C. This work presents a promising strategy for designing carbon/GDY-coated ZIF-based composites, with strong potential for integration into next-generation energy storage systems alongside other advanced materials. Looking forward, GDY-based materials can be further functionalized with MOFs or covalent organic frameworks (COFs), opening up new avenues for research in advanced energy storage and conversion systems.

3.3. Application in Biological Sensors

MOFs have emerged as promising porous materials for biological sensing due to their high surface area, tunable pore structures, and versatile chemical functionality. In recent years, extensive research has focused on utilizing MOFs as sensing platforms or functional components for detection. Despite advances, challenges remain in achieving high biocompatibility, reproducibility, and real-time detection performance. Continued efforts in material composition and interfacial modification are expected to further expand the applicability of MOFs in advanced biological sensing technologies.
In the synthesis of protein, chloramphenicol (CAP) is a broad-spectrum antibiotic that has an inhibitory effect. However, its excessive use poses serious risks to both environmental and human health, and achieving ultrasensitive detection of CAP remains a significant challenge. A novel composite material, formed by chelating Fe-MOFs with GDY, denoted as Fe-MOF@GDY [84], exhibited strong interfacial interactions and enabled high electrochemical sensing of CAP. In the Fe-MOF@GDY structure, the unique electronic distribution of GDY improved the charge transfer capabilities, thereby significantly improving the overall conductivity of the composite. Meanwhile, Fe-MOF nanoparticles were uniformly distributed and chelated on the GDY surface, offering abundant active sites for the efficient electrochemical reduction of CAP. The Fe-MOF@GDY-based sensor demonstrated outstanding CAP detection performance, with a broad linear range from 1 pM to 24 mM and an ultralow detection limit of just 0.54 pM (Figure 7a–c). Additionally, it exhibited excellent reproducibility and long-term storage stability, highlighting its strong potential for practical applications in trace antibiotic detection. A similar strategy was also applied to construct a novel MOF/GDY@polyaniline structure, which could be used as a photoactive material for sensitive prostate-specific antigen (PSA) detection [96].
Hollow-structured materials are widely used as sensing materials due to their high specific surface area and abundance of electroactive sites, which significantly enhance the sensitivity of biosensors. However, most reported hollow materials used in sensing platforms feature single-shell architectures with simple compositions, limiting their functionality and application scope. To overcome these limitations, the development of hollow multi-shelled structures with complex compositions is essential for assembling ultrasensitive biosensors. MOFs have emerged as ideal templates for synthesizing transition metal oxide-based hollow multi-shelled structures. Song et al. [89] developed a novel electrochemical biosensor by combining N-F-GDY with MOF-derived hollow multi-shelled NiCo2O4-Co3O4 nanocages through coordination interactions. Owing to its highly conjugated framework, large specific surface area, porous architecture, and excellent electronic conductivity, N-F-GDY greatly enhanced the interfacial stability of the sensor platform and facilitated effective enzyme immobilization. Utilizing acetylcholinesterase (AChE) as a biorecognition element for the detection of organophosphorus pesticides, the resulting electrochemical biosensor exhibited outstanding performance, with a wide linear detection range of 0.448 pM to 44.8 nM and an ultralow detection limit of 0.0166 fM for monocrotophos (Figure 7d–f). Notably, this platform also demonstrated strong versatility, as it can be adapted for the detection of various harmful substances simply by substituting the specific biorecognition molecule, underscoring its potential in advanced electrochemical analysis.
MOFs have garnered significant attention for electrochemiluminescence (ECL) signal amplification due to porous architectures and numerous active sites. Recently, NH2-Zr-MOF/GDY/GCE was designed and employed as an ECL sensor, in which the layered GDY served as the luminophore, while through electrostatic adsorption, NH2-Zr-MOF was assembled onto the surface of a glassy carbon electrode (Figure 7g–i) [97]. Via amide bond coordination, a carbonylated aptamer (COOH-Apt) was then immobilized onto the NH2-Zr-MOF, constructing an enhanced ECL aptamer sensor for the ultrasensitive detection of di(2-ethylhexyl) phthalate (DEHP). The developed aptamer sensor demonstrated a wide linear detection range, high selectivity, and a low detection limit. It effectively detected DEHP in both river water and municipal drinking water samples, achieving ultrasensitive quantification. For the first time, nitrogen micro-nanobubbles, featuring superior mass transfer capabilities and high stability, were applied in the ECL sensor, further boosting performance. Using the aptamer as a selective recognition unit, this system achieved ultrasensitive detection of DEHP with a low detection limit and demonstrated the strong potential of the composite MOF/GDY-based ECL aptamer sensor for the environmental monitoring of endocrine-disrupting compounds.
Despite MOFs, metal oxides derived from MOFs are also widely utilized in ECL sensors. Among them, zinc oxide (ZnO) nanomaterials—particularly nanoflower structures—have attracted considerable attention owing to their wide bandgap, good biocompatibility, and abundant catalytic sites. However, the relatively large particle size of flower-like ZnO can hinder electron transfer on the surface of GCEs, thereby limiting their overall ECL performance. To overcome this limitation, a novel composite material based on MOF-derived ZnO (M-ZnO) and GDY oxide quantum dots (GDYO QDs) was developed [98]. M-ZnO, derived from ZIF-8, enabled the effective adsorption of GDYO QDs with intrinsic ECL properties. This GDYO QDs@M-ZnO-modified electrode not only enhanced catalytic activity but also significantly amplified the ECL signal, exhibiting an ECL intensity that was five times greater than that of an electrode modified with GDYO QDs alone (Figure 7j–l). These findings highlight the promising application potential of the composite GDY/MOF-based ECL sensor in the detection of free radical scavenging drugs. These studies demonstrate that the synergistic integration of MOF and GDY significantly enhances the electronic conductivity, structural stability, and catalytic activity of the resulting nanocomposites, making them highly effective for sensing applications.
Table 1. Detailed performance of MOF/GDY composites and related materials.
Table 1. Detailed performance of MOF/GDY composites and related materials.
Materials PerformanceRole of GDYRefs.
GDY/ZnCo-ZIF-0.9CatalysisH2 evolution: 171.79 µmol
(6.67× vs. GDY)
(13.5× vs. ZnCo-ZIF)
S-scheme heterojunction[82]
GDY/CoMo-MOF-60CatalysisH2 evolution: 300 µmol
(19.61× vs. GDY)
(9.03× vs. CoMo-MOF)
S-scheme heterojunction[83]
NiCo-MOFCatalysisH2 evolution: 6.89 µmol [86]
NiCo-MOF/Cu-GDY-15H2 evolution: 112.22 µmol
(47.0× vs. Cu-GDY)
(16.3× vs. NiCo-MOF)
Strongly coupled electronic interface and ohmic contact
NiBDCCatalysisOnset potential: −142 mV vs. RHE [88]
HsGDY@CuOnset potential: 65 mV vs. RHE
Ni-MOFs@HsGDY@CuOnset potential: 90.5 mV vs. RHEConductive and reactive bridge
Co3S4/NFCatalysisOverpotential: 230 mV @ 10 mA cm−2
Tafel Slope: 78.9 mV dec−1
[90]
GDY/Co3S4/NFOverpotential: 223 mV @ 10 mA cm−2
Tafel Slope: 46.5 mV dec−1
Modulating the electronic configuration and enhancing durability
Cu3BTC2/ZIF-9(Co)/GDYCatalysisH2 Evolution Rate: 1126 μmol g−1 h−1
(~30× vs. Cu3BTC2) (~16× vs. ZIF-9(Co))
Electron transport layer[91]
Polymer-HKUST-1/CFCatalysisH2O2 conversions: ~68% [93]
HKUST-1/GDY/CFH2O2 conversions: ~94%Facilitate electron transfer
HsGDYEnergy
storage
Specific Capacitance: 154 F g−1 @ 1 A g−1 [87]
Ni-MOFSpecific Capacitance: 426 F g−1 @ 1 A g−1
Ni-MOF@HsGDYSpecific Capacitance: 982 F g−1 @ 1 A g−1
(~6.4× vs. HsGDY)
(~2.3× vs. Ni-MOF)
Conductive substrate
ZnCo-ZIF/PPEnergy
storage
Initial Specific Capacity: 862.2 mAh g−1 @ 0.2C [95]
ZnCo-ZIF@GDYInitial Specific Capacity: 1126.1 mAh g−1 @ 0.2CEnhance electrical conductivity;
Suppress polysulfide shuttle effect
Fe-MOFBiological sensorsLimit of detection: 0.011μM [84]
Fe-MOF@GDYLimit of detection: 0.54 pM
Detection Range: 1 pM–24 mM
Charge transfer enhancer and anchoring substrate
IL1-MWCNTs/AChE/GCEBiological sensorsLimit of detection: 33 pM
Detection Range: 0.1–5000 nM
[89]
AChE-NF/NiCo2O4–Co3O4/N-F-GDY/GCELimit of detection: 0.0166 fM
Detection Range: 0.448 pM–44.8 nM
Synergistic signal amplification
NH2-Zr-MOF/GDY/GCEBiological sensorsECL intensity: ~1.5 × vs. GDYECL Luminophore[97]
GDYO QDs @M-ZnOBiological sensorsECL intensity: ~5 × vs. GDYO QDsECL Luminophore[98]

4. Conclusions

MOFs and MOF-derived materials have been extensively studied in recent decades due to their large specific surface areas, high porosity, and tunable ordered structures. However, their practical application has been hindered by poor electrical conductivity and limited structural stability. To address these challenges and enhance performance, MOF-based composites incorporating functional materials have been developed and proposed. GDY, a novel carbon allotrope, has attracted substantial scientific interest following its successful synthesis, owing to its remarkable performance in diverse applications, particularly in electrocatalysis, semiconductor technology, and energy storage. Furthermore, its relatively facile synthetic route, combined with the precise tunability of precursor molecular structures, has enabled the rational design and fabrication of functional derivatives through hybridization and combination. The integration of MOFs/MOF derivatives with GDY/GDY derivatives into composite materials effectively combines the complementary advantages and overcomes the intrinsic limitations, resulting in enhanced functionality and broader applicability for diverse applications. In this review, we have provided a comprehensive summary of recent advances in MOF/GDY composites and derivative materials, mainly focusing on composite strategies and applications in catalysis, energy storage, and biosensing, which can guide future composite structure design and performance improvement.
However, there are still some issues that need to be considered. (1) The preparation of highly crystalline MOFs and GDY is crucial for effectively achieving MOF/GDY composite materials. Currently, the high-quality synthesis of MOFs has been achieved. However, large-scale, structurally controlled GDY materials remain significant challenges, which hinder the combination of GDY with other advanced materials. Additionally, innovative methods and strategies for the controlled combination of MOF/GDY composites and derivative materials are also needed. (2) Although notable progress has been made in addressing fundamental challenges, further research is needed to construct a wider variety of MOF/GDY and related composite materials, as well as to understand their combined structural properties. This includes the selection of suitable MOFs and GDY precursors, comprehensive analysis of their structures across different dimensions, exploration of diverse combination techniques, advancements in characterization technologies, and evaluation of application potential. (3) Theoretical modeling plays a vital role in guiding experimental work and interpreting the results. It is essential to develop more sophisticated and reliable theoretical models to elucidate the complex mechanisms underlying device performance based on MOF/GDY composites and derivative materials.
We hope that this review provides valuable guidance for the future development and exploration of MOF/GDY composites and derivative materials. By engineering specific MOF architectures, optimizing integration methods, applying advanced characterization techniques, and exploring broader application scenarios, we can further advance MOF and GDY composite materials and broaden their applications in the field of catalysis, energy storage, biological sensors, and beyond. Furthermore, this should further promote the maturation of composites based on inorganic/organic advanced materials and unique carbon composites, paving the way for their precision designations, practical applications, and a brighter future.

Author Contributions

Conceptualization, H.S. and J.W.; formal analysis, J.P.; investigation, J.P., T.Z., and Z.T.; writing—original draft preparation, J.P. and H.S.; writing—review and editing, H.S. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Open Project Funding of Hubei Key Laboratory of Processing and Application of Catalytic materials, Huanggang Normal University (202442304).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O.M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402, 276–279. [Google Scholar] [CrossRef]
  2. Cai, G.; Yan, P.; Zhang, L.; Zhou, H.C.; Jiang, H.L. Metal-organic framework-based hierarchically porous materials: Synthesis and applications. Chem. Rev. 2021, 121, 12278–12326. [Google Scholar] [CrossRef]
  3. Jiao, L.; Seow, J.Y.R.; Skinner, W.S.; Wang, Z.U.; Jiang, H.L. Metal-organic frameworks: Structures and functional applications. Mater. Today 2019, 27, 43–68. [Google Scholar] [CrossRef]
  4. Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; Sun, Y.; Qin, J.; Yang, X.; Zhang, P.; et al. Stable metal-organic frameworks: Design, synthesis, and applications. Adv. Mater. 2018, 30, 1704303. [Google Scholar] [CrossRef]
  5. Zheng, H.; Yan, W.; Zhang, J. Porous organic framework-based materials (MOFs, COFs and HOFs) for lithium-/sodium-/potassium-/zinc-/aluminum-/calcium-ion batteries: A review. Electrochem. Energy Rev. 2025, 8, 3. [Google Scholar] [CrossRef]
  6. Gao, X.; Dong, Y.; Li, S.; Zhou, J.; Wang, L.; Wang, B. MOFs and COFs for batteries and supercapacitors. Electrochem. Energy Rev. 2019, 3, 81–126. [Google Scholar] [CrossRef]
  7. Ye, Z.; Jiang, Y.; Li, L.; Wu, F.; Chen, R. Rational design of MOF-based materials for next-generation rechargeable batteries. Nano-Micro Lett. 2021, 13, 203. [Google Scholar] [CrossRef]
  8. Jiao, L.; Wang, Y.; Jiang, H.L.; Xu, Q. Metal-organic frameworks as platforms for catalytic applications. Adv. Mater. 2017, 30, 1703663. [Google Scholar] [CrossRef]
  9. Zhu, L.; Liu, X.Q.; Jiang, H.L.; Sun, L.B. Metal-organic frameworks for heterogeneous basic catalysis. Chem. Rev. 2017, 117, 8129–8176. [Google Scholar] [CrossRef]
  10. Dhakshinamoorthy, A.; Li, Z.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134–8172. [Google Scholar] [CrossRef]
  11. Zhang, H.; Zhou, Z.; Yin, Y.; Xu, H.; Wang, Y.; Yang, K.; Zhang, Z.; Wang, J.; He, X. Efficient capture and separation of CO2-Boosted carbon neutralization enabled by tailorable metal-organic frameworks: A review. EcoEnergy 2023, 1, 217–247. [Google Scholar] [CrossRef]
  12. Fan, W.; Zhang, X.; Kang, Z.; Liu, X.; Sun, D. Isoreticular chemistry within metal-organic frameworks for gas storage and separation. Coord. Chem. Rev. 2021, 443, 213968. [Google Scholar] [CrossRef]
  13. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2011, 112, 1105–1125. [Google Scholar] [CrossRef]
  14. Olorunyomi, J.F.; Geh, S.T.; Caruso, R.A.; Doherty, C.M. Metal-organic frameworks for chemical sensing devices. Mater. Horiz. 2021, 8, 2387–2419. [Google Scholar] [CrossRef]
  15. Hu, Z.; Deibert, B.J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815–5840. [Google Scholar] [CrossRef]
  16. Xu, G.; Zhu, C.; Gao, G. Recent progress of advanced conductive metal-organic frameworks: Precise synthesis, electrochemical energy storage applications, and future challenges. Small 2022, 18, 2203140. [Google Scholar] [CrossRef]
  17. Xie, L.S.; Skorupskii, G.; Dincă, M. Electrically conductive metal-organic frameworks. Chem. Rev. 2020, 120, 8536–8580. [Google Scholar] [CrossRef]
  18. Sun, L.; Campbell, M.G.; Dincă, M. Electrically conductive porous metal-organic frameworks. Angew. Chem. Int. Ed. 2016, 55, 3566–3579. [Google Scholar] [CrossRef]
  19. Kaneti, Y.V.; Dutta, S.; Hossain, M.S.A.; Shiddiky, M.J.A.; Tung, K.L.; Shieh, F.K.; Tsung, C.K.; Wu, K.C.W.; Yamauchi, Y. Strategies for improving the functionality of zeolitic imidazolate frameworks: Tailoring nanoarchitectures for functional applications. Adv. Mater. 2017, 29, 1700213. [Google Scholar] [CrossRef]
  20. Bennett, T.D.; Keen, D.A.; Tan, J.C.; Barney, E.R.; Goodwin, A.L.; Cheetham, A.K. Thermal amorphization of zeolitic imidazolate frameworks. Angew. Chem. Int. Ed. 2011, 50, 3067–3071. [Google Scholar] [CrossRef]
  21. Hughes, J.T.; Bennett, T.D.; Cheetham, A.K.; Navrotsky, A. Thermochemistry of zeolitic imidazolate frameworks of varying porosity. J. Am. Chem. Soc. 2012, 135, 598–601. [Google Scholar] [CrossRef]
  22. Zhou, K.; Mousavi, B.; Luo, Z.; Phatanasri, S.; Chaemchuen, S.; Verpoort, F. Characterization and properties of Zn/Co zeolitic imidazolate frameworks vs. ZIF-8 and ZIF-67. J. Mater. Chem. A 2017, 5, 952–957. [Google Scholar] [CrossRef]
  23. Yang, S.; Li, X.; Zeng, G.; Cheng, M.; Huang, D.; Liu, Y.; Zhou, C.; Xiong, W.; Yang, Y.; Wang, W.; et al. Materials institute lavoisier (MIL) based materials for photocatalytic applications. Coord. Chem. Rev. 2021, 438, 213874. [Google Scholar] [CrossRef]
  24. Zhang, H.; Hu, X.; Li, T.; Zhang, Y.; Xu, H.; Sun, Y.; Gu, X.; Gu, C.; Luo, J.; Gao, B. MIL series of metal organic frameworks (MOFs) as novel adsorbents for heavy metals in water: A review. J. Hazard. Mater. 2022, 429, 128271. [Google Scholar] [CrossRef]
  25. Zorainy, M.Y.; Gar Alalm, M.; Kaliaguine, S.; Boffito, D.C. Revisiting the MIL-101 metal-organic framework: Design, synthesis, modifications, advances, and recent applications. J. Mater. Chem. A 2021, 9, 22159–22217. [Google Scholar] [CrossRef]
  26. Huang, L.; Yang, Z.; Alhassan, S.I.; Luo, Z.; Song, B.; Jin, L.; Zhao, Y.; Wang, H. Highly efficient fluoride removal from water using 2D metal-organic frameworks MIL-53(Al) with rich Al and O adsorptive centers. Environ. Sci. Ecotechnol. 2021, 8, 100123. [Google Scholar] [CrossRef]
  27. Adly, M.S.; El-Dafrawy, S.M.; Ibrahim, A.A.; El-Hakam, S.A.; El-Shall, M.S. Efficient removal of heavy metals from polluted water with high selectivity for Hg (I) and Pb (II) by a 2-imino-4-thiobiuret chemically modified MIL-125 metal-organic framework. RSC Adv. 2021, 11, 13940–13950. [Google Scholar] [CrossRef]
  28. Khosravi, M.J.; Hosseini, S.M.; Vatanpour, V. Polyamidoamine dendrimers-Mil-125(Ti) MOF embedded polyethersulfone membrane for enhanced removal of heavy metal, antibiotic and dye from water. J. Environ. Chem. Eng. 2022, 10, 108644. [Google Scholar] [CrossRef]
  29. Tatrari, G.; An, R.; Shah, F.U. Designed metal-organic framework composites for metal-ion batteries and metal-ion capacitors. Coord. Chem. Rev. 2024, 512, 215876. [Google Scholar] [CrossRef]
  30. Li, B.; Ma, J.G.; Cheng, P. Integration of metal nanoparticles into metal-organic frameworks for composite catalysts: Design and synthetic strategy. Small 2019, 15, 1804849. [Google Scholar] [CrossRef]
  31. Yu, J.; Mu, C.; Yan, B.; Qin, X.; Shen, C.; Xue, H.; Pang, H. Nanoparticle/MOF composites: Preparations and applications. Mater. Horiz. 2017, 4, 557–569. [Google Scholar] [CrossRef]
  32. Deng, X.; Yang, L.; Huang, H.; Yang, Y.; Feng, S.; Zeng, M.; Li, Q.; Xu, D. Shape-defined hollow structural Co-MOF-74 and metal nanoparticles@Co-MOF-74 composite through a transformation strategy for enhanced photocatalysis performance. Small 2019, 15, 1902287. [Google Scholar] [CrossRef]
  33. Mukoyoshi, M.; Kitagawa, H. Nanoparticle/metal-organic framework hybrid catalysts: Elucidating the role of the MOF. Chem. Commun. 2022, 58, 10757–10767. [Google Scholar] [CrossRef]
  34. Yang, Q.; Xu, Q.; Jiang, H.L. Metal-organic frameworks meet metal nanoparticles: Synergistic effect for enhanced catalysis. Chem. Soc. Rev. 2017, 46, 4774–4808. [Google Scholar] [CrossRef]
  35. Kalaj, M.; Bentz, K.C.; Ayala, S.; Palomba, J.M.; Barcus, K.S.; Katayama, Y.; Cohen, S.M. MOF-polymer hybrid materials: From simple composites to tailored architectures. Chem. Rev. 2020, 120, 8267–8302. [Google Scholar] [CrossRef]
  36. Brahmi, C.; Benltifa, M.; Vaulot, C.; Michelin, L.; Dumur, F.; Millange, F.; Frigoli, M.; Airoudj, A.; Morlet-Savary, F.; Bousselmi, L.; et al. New hybrid MOF/polymer composites for the photodegradation of organic dyes. Eur. Polym. J. 2021, 154, 110560. [Google Scholar] [CrossRef]
  37. Liang, X.; Zhang, F.; Feng, W.; Zou, X.; Zhao, C.; Na, H.; Liu, C.; Sun, F.; Zhu, G. From metal-organic framework (MOF) to MOF-polymer composite membrane: Enhancement of low-humidity proton conductivity. Chem. Sci. 2013, 4, 983–992. [Google Scholar] [CrossRef]
  38. Wang, D.; Li, T. Toward MOF@polymer core-shell particles: Design principles and potential applications. Acc. Chem. Res. 2023, 56, 462–474. [Google Scholar] [CrossRef]
  39. Parsapour, F.; Moradi, M.; Safarifard, V.; Sojdeh, S. Polymer/MOF composites for metal-ion batteries: A mini review. J. Energy Storage 2024, 82, 110487. [Google Scholar] [CrossRef]
  40. Liu, X.W.; Sun, T.J.; Hu, J.L.; Wang, S.D. Composites of metal-organic frameworks and carbon-based materials: Preparations, functionalities and applications. J. Mater. Chem. A 2016, 4, 3584–3616. [Google Scholar] [CrossRef]
  41. Zheng, Y.; Zheng, S.; Xue, H.; Pang, H. Metal-Organic Frameworks/Graphene-Based Materials: Preparations and Applications. Adv. Funct. Mater. 2018, 28, 1804950. [Google Scholar] [CrossRef]
  42. Petit, C.; Bandosz, T.J. Engineering the surface of a new class of adsorbents: Metal-organic framework/graphite oxide composites. J. Colloid Interface Sci. 2015, 447, 139–151. [Google Scholar] [CrossRef]
  43. Wang, K.; Hui, K.N.; Hui, K.S.; Peng, S.; Xu, Y. Recent progress in metal-organic framework/graphene-derived materials for energy storage and conversion: Design, preparation, and application. Chem. Sci. 2021, 12, 5737–5766. [Google Scholar] [CrossRef]
  44. Wang, Z.; Huang, J.; Mao, J.; Guo, Q.; Chen, Z.; Lai, Y. Metal-organic frameworks and their derivatives with graphene composites: Preparation and applications in electrocatalysis and photocatalysis. J. Mater. Chem. A 2020, 8, 2934–2961. [Google Scholar] [CrossRef]
  45. Kayal, S.; Chakraborty, A. Activated carbon (type Maxsorb-III) and MIL-101(Cr) metal organic framework based composite adsorbent for higher CH4 storage and CO2 capture. Chem. Eng. J. 2018, 334, 780–788. [Google Scholar] [CrossRef]
  46. Govindaraju, S.; Arumugasamy, S.K.; Chellasamy, G.; Yun, K. Zn-MOF decorated bio activated carbon for photocatalytic degradation, oxygen evolution and reduction catalysis. J. Hazard. Mater. 2022, 421, 126720. [Google Scholar] [CrossRef]
  47. Wang, X.; Kukkar, D.; Younis, S.A.; Vikrant, K.; Ahmadi, Y.; Boukhvalov, D.; Kim, K.H. The co-adsorption potential of metal-organic framework/activated carbon composites against both polar and non-polar volatile organic compounds in air. Sep. Purif. Technol. 2023, 306, 122594. [Google Scholar] [CrossRef]
  48. Jayaramulu, K.; Mukherjee, S.; Morales, D.M.; Dubal, D.P.; Nanjundan, A.K.; Schneemann, A.; Masa, J.; Kment, S.; Schuhmann, W.; Otyepka, M.; et al. Graphene-based metal-organic framework hybrids for applications in catalysis, environmental, and energy technologies. Chem. Rev. 2022, 122, 17241–17338. [Google Scholar] [CrossRef]
  49. Zhang, X.; Zhang, S.; Tang, Y.; Huang, X.; Pang, H. Recent advances and challenges of metal-organic framework/graphene-based composites. Compos. Part B Eng. 2022, 230, 109532. [Google Scholar] [CrossRef]
  50. Gao, C.; Wang, P.; Wang, Z.; Kær, S.K.; Zhang, Y.; Yue, Y. The disordering-enhanced performances of the Al-MOF/graphene composite anodes for lithium ion batteries. Nano Energy 2019, 65, 104032. [Google Scholar] [CrossRef]
  51. Jahan, M.; Bao, Q.; Loh, K.P. Electrocatalytically active graphene-porphyrin MOF composite for oxygen reduction reaction. J. Am. Chem. Soc. 2012, 134, 6707–6713. [Google Scholar] [CrossRef]
  52. Jayaramulu, K.; Dubal, D.P.; Schneemann, A.; Ranc, V.; Perez-Reyes, C.; Stráská, J.; Kment, Š.; Otyepka, M.; Fischer, R.A.; Zbořil, R. Shape-assisted 2D MOF/graphene derived hybrids as exceptional lithium-ion battery electrodes. Adv. Funct. Mater. 2019, 29, 1902539. [Google Scholar] [CrossRef]
  53. Yang, C.; You, X.; Cheng, J.; Zheng, H.; Chen, Y. A novel visible-light-driven in-based MOF/graphene oxide composite photocatalyst with enhanced photocatalytic activity toward the degradation of amoxicillin. Appl. Catal. B Environ. 2017, 200, 673–680. [Google Scholar] [CrossRef]
  54. Huang, H.; Chen, Y.; Chen, Z.; Chen, J.; Hu, Y.; Zhu, J.J. Electrochemical sensor based on Ce-MOF/carbon nanotube composite for the simultaneous discrimination of hydroquinone and catechol. J. Hazard. Mater. 2021, 416, 125895. [Google Scholar] [CrossRef]
  55. Jabbari, V.; Veleta, J.M.; Zarei-Chaleshtori, M.; Gardea-Torresdey, J.; Villagrán, D. Green synthesis of magnetic MOF@GO and MOF@CNT hybrid nanocomposites with high adsorption capacity towards organic pollutants. Chem. Eng. J. 2016, 304, 774–783. [Google Scholar] [CrossRef]
  56. Wang, X.; Ma, Z.; Chai, L.; Xu, L.; Zhu, Z.; Hu, Y.; Qian, J.; Huang, S. MOF derived N-doped carbon coated CoP particle/carbon nanotube composite for efficient oxygen evolution reaction. Carbon 2019, 141, 643–651. [Google Scholar] [CrossRef]
  57. Bhadra, B.N.; Shrestha, L.K.; Ma, R.; Hill, J.P.; Yamauchi, Y.; Ariga, K. Metal-organic framework on fullerene (MOFOF) as a hierarchical composite by the integration of coordination chemistry and supramolecular chemistry. ACS Appl. Mater. Interfaces 2024, 16, 41363–41370. [Google Scholar] [CrossRef]
  58. Han, J.; Wu, H.; Fan, H.; Ding, L.; Hai, G.; Caro, J.; Wang, H. Tuning the phase composition of metal-organic framework membranes for helium separation through incorporation of fullerenes. J. Am. Chem. Soc. 2023, 145, 14793–14801. [Google Scholar] [CrossRef]
  59. Zhai, Y.; Dou, Y.; Zhao, D.; Fulvio, P.F.; Mayes, R.T.; Dai, S. Carbon materials for chemical capacitive energy storage. Adv. Mater. 2011, 23, 4828–4850. [Google Scholar] [CrossRef]
  60. Li, G.; Li, Y.; Liu, H.; Guo, Y.; Li, Y.; Zhu, D. Architecture of graphdiyne nanoscale films. Chem. Commun. 2010, 46, 3256–3258. [Google Scholar] [CrossRef]
  61. Fang, Y.; Liu, Y.; Qi, L.; Xue, Y.; Li, Y. 2D graphdiyne: An emerging carbon material. Chem. Soc. Rev. 2022, 51, 2681–2709. [Google Scholar] [CrossRef]
  62. Gao, X.; Liu, H.; Wang, D.; Zhang, J. Graphdiyne: Synthesis, properties, and applications. Chem. Soc. Rev. 2019, 48, 908–936. [Google Scholar] [CrossRef]
  63. Jia, Z.; Li, Y.; Zuo, Z.; Liu, H.; Huang, C.; Li, Y. Synthesis and properties of 2D carbon-graphdiyne. Acc. Chem. Res. 2017, 50, 2470–2478. [Google Scholar] [CrossRef]
  64. Zuo, Z.; Wang, D.; Zhang, J.; Lu, F.; Li, Y. Synthesis and applications of graphdiyne-based metal-free catalysts. Adv. Mater. 2018, 31, 1803762. [Google Scholar] [CrossRef]
  65. Qi, L.; Gao, Y.; Gao, Y.; Zheng, Z.; Luan, X.; Zhao, S.; Chen, Z.; Liu, H.; Xue, Y.; Li, Y. Controlled growth of metal atom arrays on graphdiyne for seawater oxidation. J. Am. Chem. Soc. 2024, 146, 5669–5677. [Google Scholar] [CrossRef]
  66. Zheng, X.; Xue, Y.; Chen, S.; Li, Y. Advances in hydrogen energy conversion of graphdiyne-based materials. EcoEnergy 2023, 1, 45–59. [Google Scholar] [CrossRef]
  67. Zhu, Y.; Yang, L.; Ma, J.; Fang, Y.; Yang, J.; Chen, X.; Zheng, J.; Zhang, S.; Chen, W.; Pan, C.; et al. Rapid ozone decomposition over water-activated monolithic MoO3/graphdiyne nanowalls under high humidity. Angew. Chem. Int. Ed. 2023, 62, e202309158. [Google Scholar] [CrossRef]
  68. Ma, K.; Wu, J.; Wang, X.; Sun, Y.; Xiong, Z.; Dai, F.; Bai, H.; Xie, Y.; Kang, Z.; Zhang, Y. Periodically interrupting bonding behavior to reformat delocalized electronic states of graphdiyne for improved electrocatalytic hydrogen evolution. Angew. Chem. Int. Ed. 2022, 61, e202211094. [Google Scholar] [CrossRef]
  69. Wang, N.; He, J.; Wang, K.; Zhao, Y.; Jiu, T.; Huang, C.; Li, Y. Graphdiyne-based materials: Preparation and application for electrochemical energy storage. Adv. Mater. 2019, 31, 1803202. [Google Scholar] [CrossRef]
  70. Du, Y.; Zhou, W.; Gao, J.; Pan, X.; Li, Y. Fundament and application of graphdiyne in electrochemical energy. Acc. Chem. Res. 2020, 53, 459–469. [Google Scholar] [CrossRef]
  71. An, J.; Zhang, H.; Wang, Y.; Kong, Z.; Li, W.; Gao, X.; Song, J.; Yao, Y. Application and prospects of interface engineering in energy storage and conversion of graphdiyne-based materials. EcoEnergy 2025, 3, 77–104. [Google Scholar] [CrossRef]
  72. Li, H.; Lim, J.H.; Lv, Y.; Li, N.; Kang, B.; Lee, J.Y. Graphynes and graphdiynes for energy storage and catalytic utilization: Theoretical insights into recent advances. Chem. Rev. 2023, 123, 4795–4854. [Google Scholar] [CrossRef]
  73. Chen, Z.; Zhang, D.; Yang, Z.; Xu, Y.; Wang, X.; Huang, H.; Qiu, F.; Huang, C. Doped-graphdiyne: Synthesis, theoretical prediction and application for electrochemical energy storage. Nanoscale Horiz. 2025, 10, 2239–2261. [Google Scholar] [CrossRef]
  74. Zhang, M.; Li, Y.; Li, X.; Wang, N.; Huang, C. Graphdiyne ink for ionic liquid gated printed transistor. Adv. Electron. Mater. 2020, 6, 2000157. [Google Scholar] [CrossRef]
  75. Niu, G.; Wang, Y.; Yang, Z.; Cao, S.; Liu, H.; Wang, J. Graphdiyne and its derivatives as efficient charge reservoirs and transporters in semiconductor devices. Adv. Mater. 2023, 35, 2212159. [Google Scholar] [CrossRef]
  76. Do, D.P.; Hong, C.; Bui, V.Q.; Pham, T.H.; Seo, S.; Do, V.D.; Phan, T.L.; Tran, K.M.; Haldar, S.; Ahn, B.W.; et al. Highly efficient van der waals heterojunction on graphdiyne toward the high-performance photodetector. Adv. Sci. 2023, 10, 2300925. [Google Scholar] [CrossRef]
  77. Rana, S.; Kumar, A.; Dhiman, P.; Sharma, G.; Amirian, J.; Stadler, F.J. Progress in graphdiyne and phosphorene based composites and heterostructures as new age materials for photocatalytic hydrogen evolution. Fuel 2024, 356, 129630. [Google Scholar] [CrossRef]
  78. Wang, R.; Shi, M.; Xu, F.; Qiu, Y.; Zhang, P.; Shen, K.; Zhao, Q.; Yu, J.; Zhang, Y. Graphdiyne-modified TiO2 nanofibers with osteoinductive and enhanced photocatalytic antibacterial activities to prevent implant infection. Nat. Commun. 2020, 11, 4465. [Google Scholar] [CrossRef]
  79. Gao, N.; Ren, G.; Zhang, M.; Mao, L. Electroless deposition of palladium nanoparticles on graphdiyne boosts electrochemiluminescence. J. Am. Chem. Soc. 2024, 146, 3836–3843. [Google Scholar] [CrossRef]
  80. He, T.; Matta, S.K.; Will, G.; Du, A. Transition-metal single atoms anchored on graphdiyne as high-efficiency electrocatalysts for water splitting and oxygen reduction. Small Methods 2019, 3, 1800419. [Google Scholar] [CrossRef]
  81. Pei, Y.; Wu, X.; Ren, Z.; Lv, Y.; Xue, R.; Guo, J.; Jia, D. Reinforcement of C-NFO@GDY membranes via the synergistic effect of the graphdiyne honeycomb nanostructure and electronegativity for high-efficiency oil-in-water emulsion separation. Adv. Fiber Mater. 2025, 7, 1195–1207. [Google Scholar] [CrossRef]
  82. Wang, K.; Kong, X.; Xie, H.; Li, S.; Wang, M.; Jin, Z. High performance of visible-light driven hydrogen production over graphdiyne (g-CnH2n−-2)/MOF S-scheme heterojunction. Dalton Trans. 2023, 52, 8716–8727. [Google Scholar] [CrossRef]
  83. Ding, L.; Lei, M.; Wang, T.; Wang, J.; Jin, Z. Graphdiyne coordinated CoMo-MOF formed S-scheme heterojunction boosting photocatalytic hydrogen production. Carbon Lett. 2024, 34, 2099–2112. [Google Scholar] [CrossRef]
  84. Zhang, K.; Shi, Z.; Li, Y.; Wu, X.; Ma, Z.; Sun, W.; Ming Li, C.; Guo, C. Graphdiyne chelated iron-based metal–organic frameworks for electrochemical sensing of antibiotic chloramphenicol with ultralow detection limit. Microchem. J. 2024, 201, 110526. [Google Scholar] [CrossRef]
  85. Zhang, L.; Wang, X.; Li, Y.; Jin, Z. Graphdiyne (CnH2n−2) coupled with ZnCo-MOF double S-scheme heterojunction forms an efficient electron transport layer and its characterization via in situ XPS. Catal. Sci. Technol. 2023, 13, 3667–3681. [Google Scholar] [CrossRef]
  86. Liu, Z.; Wang, J.; Liu, G.; Jin, Z.; Tsubaki, N. NiCo-MOFs in situ anchored on graphdiyne with metal-like properties form a strongly coupled electron transport interface and construct an ohmic contact to achieve efficient charge-hole spatial separation. Nanoscale 2024, 16, 19322–19334. [Google Scholar] [CrossRef]
  87. Abbas, Z.; Hussain, N.; Mobin, S.M. Crumpled fibrous graphdiyne network decorated metal-organic framework: A promising heterostructure for improved energy storage performance. J. Mater. Chem. A 2024, 12, 20170–20178. [Google Scholar] [CrossRef]
  88. Wang, B.; Ma, J.; Yang, R.; Meng, B.; Yang, X.; Zhang, Q.; Zhang, B.; Zhuo, S. Bridging nickel-MOF and copper single atoms/clusters with H-substituted graphdiyne for the tandem catalysis of nitrate to ammonia. Angew. Chem. Int. Ed. 2024, 63, e202404819. [Google Scholar] [CrossRef]
  89. Song, D.; Xu, X.; Huang, X.; Li, G.; Wang, X.; Zhao, Y.; Gao, F. Nitrogen and fluoride co-doped graphdiyne with metal-organic framework (MOF)-derived NiCo2O4-Co3O4 nanocages as sensing layers for ultra-sensitive pesticide detection. Anal. Chim. Acta 2023, 1252, 341012. [Google Scholar] [CrossRef]
  90. Lu, M.; Zhao, X.; Zhang, S.; Jian, H.; Wang, M.; Lu, T. Stabilization of MOF-derived Co3S4 nanoparticles via graphdiyne coating for efficient oxygen evolution. Sci. China Mater. 2024, 67, 1882–1890. [Google Scholar] [CrossRef]
  91. Jin, F.; Yang, B.; Wang, X.; Li, T.; Tsubaki, N.; Jin, Z. Facilitating efficient photocatalytic hydrogen evolution via enhanced carrier migration at MOF-on-MOF S-scheme heterojunction interfaces through a graphdiyne (CnH2n−–2) electron transport layer. Chin. J. Struct. Chem. 2023, 42, 100198. [Google Scholar] [CrossRef]
  92. Cui, J.; Liu, J.; Wang, C.; Rong, F.; He, L.; Song, Y.; Zhang, Z.; Fang, S. Efficient electrocatalytic water oxidation by using the hierarchical 1D/2D structural nanohybrid of CoCu-based zeolitic imidazolate framework nanosheets and graphdiyne nanowires. Electrochim. Acta 2020, 334, 135577. [Google Scholar] [CrossRef]
  93. Sun, H.P.; Hye, M.K.; Mariana, L.D.; Sunggi, L.; Nak, C.J. Hydroquinone-treated Cu3(BTC)2: A mixed-valence Cu(I/II) MOF catalyst for efficient cycloadditions. Chem. Commun. 2024, 60, 14577. [Google Scholar] [CrossRef]
  94. Wu, Y.; Yang, Y.; Ke, Z. Metal-organic frameworks/graphdiyne/copper foam composite membranes for catalytic applications. ACS Appl. Mater. Interfaces 2023, 15, 40933–40941. [Google Scholar] [CrossRef]
  95. Guo, L.; Shang, H.; Peng, J.; Li, H.; Pan, H. In-situ growth of graphdiyne on ZnCo-ZIF for enhanced lithium-sulfur battery performance. J. Mater. Chem. A 2025, 13, 36436–36443. [Google Scholar] [CrossRef]
  96. Hosseini, S.S.; Mansouri Majd, S.; Salimi, A. Polarity-switchable dual-mode photoelectrochemical cancer marker immunoassay based on a metal-organic framework@nitrogen-doped graphdiyne heterojunction. ACS Appl. Mater. Interfaces 2025, 17, 27759–27771. [Google Scholar] [CrossRef]
  97. Dong, M.; Jiang, D.; Cao, Q.; Wang, W.; Shiigi, H.; Chen, Z. A metal-organic framework regulated graphdiyne-based electrochemiluminescence sensor with a electrocatalytic self-acceleration effect for the detection of di-(2-ethylhexyl) phthalate. Analyst 2023, 148, 4470–4478. [Google Scholar] [CrossRef]
  98. Lin, Y.; Wu, J.; Tan, X.; Huang, K. Electrochemiluminescence sensor based on graphdiyne quantum dots functionalized MOF-derived ZnO with self-enhancing effect for the detection of radical scavenging drugs. Sensors Actuat. B-Chem. 2023, 382, 133541. [Google Scholar] [CrossRef]
Scheme 1. Overview of preparations of MOF/GDY composites and derivative materials and their applications.
Scheme 1. Overview of preparations of MOF/GDY composites and derivative materials and their applications.
Catalysts 15 01041 sch001
Figure 1. The strategies of physical mixing to prepare MOF/GDY composite materials. (a) Preparation of GDY/ZnCo-ZIF [82]; (b) preparation of GDY/CoMo-MOF [83]; (c) preparation of Fe-MOF@GDY [84].
Figure 1. The strategies of physical mixing to prepare MOF/GDY composite materials. (a) Preparation of GDY/ZnCo-ZIF [82]; (b) preparation of GDY/CoMo-MOF [83]; (c) preparation of Fe-MOF@GDY [84].
Catalysts 15 01041 g001
Figure 2. The strategies of in situ growth to prepare MOF/GDY composite materials. (a) MOFs in situ anchored on GDY [86]; (b,c) HsGDY in situ grown on MOFs [87,88].
Figure 2. The strategies of in situ growth to prepare MOF/GDY composite materials. (a) MOFs in situ anchored on GDY [86]; (b,c) HsGDY in situ grown on MOFs [87,88].
Catalysts 15 01041 g002
Figure 3. MOF/GDY derivative composites. (a) The preparation of N-F-GDY and NiCo2O4-Co3O4 hollow double-shelled nanocage [89]; (b) the preparation of GDY/Co3S4/NF composite [90].
Figure 3. MOF/GDY derivative composites. (a) The preparation of N-F-GDY and NiCo2O4-Co3O4 hollow double-shelled nanocage [89]; (b) the preparation of GDY/Co3S4/NF composite [90].
Catalysts 15 01041 g003
Figure 4. Application in photocatalytic hydrogen evolution reactions. (ac) The hydrogen production mechanism of GDY/ZnCo-ZIF and its performance [82]; (df) Fermi-level position diagram and photogenerated charge carrier transfer of Cu3BTC2/ZIF-9(Co)/GDY and its performance [91]; (gi) the potential photocatalytic hydrogen production of NiCo-MOF/GDY and its performance [86].
Figure 4. Application in photocatalytic hydrogen evolution reactions. (ac) The hydrogen production mechanism of GDY/ZnCo-ZIF and its performance [82]; (df) Fermi-level position diagram and photogenerated charge carrier transfer of Cu3BTC2/ZIF-9(Co)/GDY and its performance [91]; (gi) the potential photocatalytic hydrogen production of NiCo-MOF/GDY and its performance [86].
Catalysts 15 01041 g004
Figure 5. (ac) The performance of GDY/Co3S4/NF in catalyze OER [90]; (df) the performance of Ni-MOFs@HsGDY@Cu hybrid in catalyze nitrate-to-ammonia conversion [88]; (gi) the performance of HKUST-1/GDY/CF membrane in catalyze organic reactions [93].
Figure 5. (ac) The performance of GDY/Co3S4/NF in catalyze OER [90]; (df) the performance of Ni-MOFs@HsGDY@Cu hybrid in catalyze nitrate-to-ammonia conversion [88]; (gi) the performance of HKUST-1/GDY/CF membrane in catalyze organic reactions [93].
Catalysts 15 01041 g005
Figure 6. Application in energy storage. (ac) MOF@HsGDY as the positive electrode in asymmetric supercapacitor and its performance [87]; (d,e) ZnCo-ZIF@GDY/PP as separator in Li-S battery and its performance [95].
Figure 6. Application in energy storage. (ac) MOF@HsGDY as the positive electrode in asymmetric supercapacitor and its performance [87]; (d,e) ZnCo-ZIF@GDY/PP as separator in Li-S battery and its performance [95].
Catalysts 15 01041 g006
Figure 7. Application in biological sensors. (ac) Fe-MOF@GDY-based chip device used as CAP detection sensor [84]; (df) NiCo2O4-Co3O4 hollow double-shelled nanocages used as organophosphorus pesticide detection sensor [89]; (gi) NH2-Zr-MOF/GDY composite used as ECL aptamer sensor [97]; (jl) GDYO QDs@M-ZnO-modified electrode and its ECL performance [98].
Figure 7. Application in biological sensors. (ac) Fe-MOF@GDY-based chip device used as CAP detection sensor [84]; (df) NiCo2O4-Co3O4 hollow double-shelled nanocages used as organophosphorus pesticide detection sensor [89]; (gi) NH2-Zr-MOF/GDY composite used as ECL aptamer sensor [97]; (jl) GDYO QDs@M-ZnO-modified electrode and its ECL performance [98].
Catalysts 15 01041 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peng, J.; Tian, Z.; Zhao, T.; Shang, H.; Wu, J. Recent Progress in Preparations and Multifunctional Applications Towards MOF/GDY Composites and Their Derivative Materials. Catalysts 2025, 15, 1041. https://doi.org/10.3390/catal15111041

AMA Style

Peng J, Tian Z, Zhao T, Shang H, Wu J. Recent Progress in Preparations and Multifunctional Applications Towards MOF/GDY Composites and Their Derivative Materials. Catalysts. 2025; 15(11):1041. https://doi.org/10.3390/catal15111041

Chicago/Turabian Style

Peng, Jia, Zhiwei Tian, Tonghe Zhao, Hong Shang, and Jing Wu. 2025. "Recent Progress in Preparations and Multifunctional Applications Towards MOF/GDY Composites and Their Derivative Materials" Catalysts 15, no. 11: 1041. https://doi.org/10.3390/catal15111041

APA Style

Peng, J., Tian, Z., Zhao, T., Shang, H., & Wu, J. (2025). Recent Progress in Preparations and Multifunctional Applications Towards MOF/GDY Composites and Their Derivative Materials. Catalysts, 15(11), 1041. https://doi.org/10.3390/catal15111041

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

Article Metrics

Back to TopTop