Next Article in Journal
Effect of Superabsorbent Polymer Size on Strength and Shrinkage in Concrete Mixtures
Previous Article in Journal
The Influence of the Dispersion and Covalent Functionalization of CNTs on Electrical Conduction Under an Electric Field in LDPE/CNT Composites
Previous Article in Special Issue
Plasticized Starch/Gelatin Blends with Humidity-Activated Shape-Memory Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal–Organic Frameworks as Fillers in Porous Organic Polymer-Based Hybrid Materials: Innovations in Composition, Processing, and Applications

by
Victor Durán-Egido
1,†,
Daniel García-Giménez
2,†,
Juan Carlos Martínez-López
2,†,
Laura Pérez-Vidal
2,† and
Javier Carretero-González
2,*
1
Faculty of Chemistry, Universidad Complutense de Madrid, 28040 Madrid, Spain
2
Institute of Polymer Science and Technology, ICTP, CSIC, 28006 Madrid, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2025, 17(14), 1941; https://doi.org/10.3390/polym17141941
Submission received: 13 June 2025 / Revised: 8 July 2025 / Accepted: 13 July 2025 / Published: 15 July 2025
(This article belongs to the Special Issue Organic-Inorganic Hybrid Materials, 4th Edition)

Abstract

Hybrid materials based on porous organic polymers (POPs) and metal–organic frameworks (MOFs) are increasing attention for advanced separation processes due to the possibility to combine their properties. POPs provide high surface areas, chemical stability, and tunable porosity, while MOFs contribute a high variety of defined crystalline structures and enhanced separation characteristics. The combination (or hybridization) with PIMs gives rise to mixed-matrix membranes (MMMs) with improved permeability, selectivity, and long-term stability. However, interfacial compatibility remains a key limitation, often addressed through polymer functionalization or controlled dispersion of the MOF phase. MOF/COF hybrids are more used as biochemical sensors with elevated sensitivity, catalytic applications, and wastewater remediation. They are also very well known in the gas sorption and separation field, due to their tunable porosity and high electrical conductivity, which also makes them feasible for energy storage applications. Last but not less important, hybrids with other POPs, such as hyper-crosslinked polymers (HCPs), covalent triazine frameworks (CTFs), or conjugated microporous polymers (CMPs), offer enhanced functionality. MOF/HCP hybrids combine ease of synthesis and chemical robustness with tunable porosity. MOF/CTF hybrids provide superior thermal and chemical stability under harsh conditions, while MOF/CMP hybrids introduce π-conjugation for enhanced conductivity and photocatalytic activity. These and other findings confirm the potential of MOF-POP hybrids as next-generation materials for gas separation and carbon capture applications.

1. Introduction

Metal–organic frameworks (MOFs) have emerged as one of the most versatile classes of porous materials, characterized by their tunable crystalline structures and high surface areas (500–7000 m2/g) [1,2,3]. Constructed through the coordination of metal nodes or metal–oxo clusters with organic linkers, MOFs exhibit remarkable potential in areas such as gas storage [4], separation [5], catalysis [6], energy [7], and sensing applications [8,9]. However, despite their structural and functional diversity, MOFs often face limitations such as moderate stability under harsh chemical conditions, poor processability, and challenges in integrating into practical devices [10]. These shortcomings have motivated researchers to combine them with other materials to form hybrids such as different polymer matrixes. For instance, the creation of hybrids made of MOFs and polymers leads to a broader spectrum of favorable properties than do the individual MOF or polymer components separately [11]. Combining polymers and MOFs with well-established functional properties could yield easily processable and ready-to-use formulations [12]. Diverse approaches have been explored by research groups where MOF−polymer composites comprise the leading MOF-based mixed-matrix membranes (MMMs) [13,14], matrix-free hybrids with polymers grafted from the surface of MOFs or through MOF particles [15], MOF-templated polymers where polymerization takes place in the free space of MOF channels [16], polymers templating MOF growth [17,18], and the synthesis of MOFs using polymer ligands (polyMOFs) [17,19].
Porous organic polymers (POPs) represent an ideal complementary material class for hybridization with MOFs. The combination of MOFs with POPs enhances the performance of the hybrid system compared to the individual counterparts and extends their applicability. POPs offer outstanding chemical and thermal stability due to their entirely covalent framework [20]. Their synthesis involves robust polymerization strategies such as condensation reactions, Suzuki coupling, Sonogashira coupling, and Friedel–Crafts alkylation, allowing for precise control over pore structure and functionality [21]. Porous organic polymers encompass a diverse range of materials, each with distinct structural and functional characteristics that make them suitable for hybridization with MOFs.
Covalent organic frameworks (COFs) stand out among POPs due to their crystalline structure, which provides well-defined, ordered porosity [22]. Their tunable pore sizes, high surface areas, and chemical versatility make them particularly suited for integration with MOFs in applications such as catalysis, gas separation, energy storage, and biosensors [23,24]. Another important group is conjugated microporous polymers (CMPs), which combine permanent microporosity with extended π-conjugation [25]. This structure imparts electronic conductivity, making CMPs highly valuable in energy storage [26], photocatalysis [27], and optoelectronics [28]. Their conjugated nature also promotes strong π–π interactions with MOFs, enhancing interfacial adhesion in hybrid materials [29]. In contrast to the ordered networks of COFs and CMPs, hyper-crosslinked polymers (HCPs) are amorphous materials characterized by ultra-microporosity [30]. Typically synthesized via Friedel–Crafts reactions, they offer exceptional chemical stability and hydrophobicity, making them ideal for adsorption-based applications [31]. Their rigid frameworks also reinforce MOFs structurally, enhancing mechanical stability in hybrid systems [32]. Covalent triazine frameworks (CTFs) are another class of POPs, notable for their nitrogen-rich composition and robust structures [33,34]. Synthesized via the trimerization of aromatic nitriles, they exhibit strong interactions with metal sites, facilitating their integration with MOFs. The nitrogen functionalities of CTFs improve interfacial adhesion by coordinating with MOF metal nodes, enhancing hybrid stability and performance [35]. MOF–CTF hybrids are commonly synthesized by growing MOFs on CTF surfaces or embedding MOFs within CTF networks, resulting in hierarchical porosity that benefits applications in catalysis [36], separation [37,38], and energy storage [39]. Polymers of intrinsic microporosity (PIMs) differ from other POPs due to their rigid and contorted backbones, which prevent efficient packing and create permanent free volume [21]. This structure grants them high permeability and excellent gas transport properties. Their combination of porosity and flexibility makes PIMs valuable in MOF-based membranes for separation technologies, where they enhance processability and introduce complementary pore structures [40,41,42].
Therefore, the combination of MOFs and POPs has given rise to a rapidly expanding field of materials research that leverages the strengths of both classes. By integrating these two porous systems, researchers can achieve advanced materials with improved stability, enhanced processability, and synergistic functionality [43].
Beyond their advantages, however, MOF–POP hybrids also present challenges in terms of compromising MOF structure or accessible porosity [15]. One of the primary difficulties is achieving optimum interfacial compatibility between the two components [44]. Because MOFs are constructed by coordination bonds and POPs are covalent networks, ensuring sufficient adhesion and interaction at their interface is crucial [45,46]. Differences in chemical functionality, hydrophobicity, and thermal expansion behavior can lead to phase separation or weak interfacial contact, which may limit hybrid performance [47]. Additionally, while the structural diversity of MOFs allows for significant functionalization, maintaining their crystallinity during polymer integration can be challenging [48]. The synthesis of MOF–POP hybrids often requires precise control over reaction conditions to ensure homogeneous mixing, avoid framework collapse, and optimize interactions between the materials. Furthermore, while hybridization opens new avenues for enhanced performance, it may also increase synthetic complexity [12].
A crucial step in addressing these challenges is the strategic functionalization of MOFs, which enhances their compatibility with polymeric frameworks [49]. Modifications at the metal clusters or organic linkers can introduce functional groups that improve adhesion, promote covalent or non-covalent interactions, and optimize interfacial contact between MOFs and POPs. Likewise, tailoring the chemistry of POPs (e.g., incorporating polar groups or π-conjugated systems) further enhances hybrid formation [50], ensuring efficient charge or mass transfer across the interface. In the following sections, we will delve into the structural diversity of MOFs, highlighting their functionalization strategies and their role in facilitating MOF–POP hybridization. We will then discuss the synthetic methodologies employed in fabricating these hybrids, examine their physicochemical properties, and assess their potential applications in gas separation, catalysis, and energy storage and conversion.

2. Diversity of MOF Structures and Their Functionalization Potential

Metal–organic frameworks (MOFs) exhibit an extraordinary diversity of structures due to the nearly limitless combinations of metal nodes and organic linkers [51]. This tunability enables precise control over pore size, shape, and chemical functionality, making MOFs highly versatile for various applications [52]. The fundamental structural diversity of MOFs stems from the coordination geometry of metal centers and the connectivity of organic ligands, leading to distinct topologies. Among the vast array of MOF architectures, several common structural types have been extensively studied and employed for hybridization with porous organic polymers (POPs), some of which are summarized in Table 1 and further described below.

2.1. Zeolitic Imidazolate Frameworks (ZIFs)

Zeolitic imidazolate frameworks (ZIFs) represent a remarkable subclass of MOFs, blending the structural features of zeolites with the tunability of metal–organic frameworks. Their defining characteristic lies in the tetrahedral coordination of metal centers, usually Zn2+ or Co2+, with imidazolate linkers, resulting in a robust structure that withstands high temperatures and extreme pH conditions (pH 2–12). This extraordinary stability [53] makes ZIFs invaluable for gas separation, catalysis, and chemical sensing, particularly in environments where conventional MOFs may degrade [75]. Among the numerous ZIFs, ZIF-8 and ZIF-67 stand out as the most studied. ZIF-8, featuring zinc centers coordinated with 2-methylimidazolate, is highly hydrophobic and chemically resilient, making it ideal for selective gas adsorption and membrane technologies [53,76]. In contrast, ZIF-67, with cobalt in place of zinc, finds widespread application in electrocatalysis and energy storage [77,78].
One of the most compelling aspects of ZIFs is their ability to be functionalized, allowing fine-tuned properties for specific applications. The imidazolate linkers can be modified with functional groups like methyl (-CH3), nitro (-NO2), or amine (-NH2), which particularly enhances the pervaporation performance of ethanol/water separation processes [79]. This functionalization has a great influence on hydrophilicity, electronic properties, and interactions with other materials [80,81]. Such modifications are particularly beneficial when hybridizing ZIFs with porous organic polymers (POPs) [82]. Functionalized linkers can enhance adhesion through hydrogen bonding, π–π stacking, or acid–base interactions, ensuring a stable and uniform hybrid structure. These composites, which combine the chemical resilience of ZIFs with the processability of polymers, are gaining traction in membrane-based gas separation, hybrid catalysts, and next-generation sensors [83].

2.2. UiO-Series MOFs (UiO-66, UiO-67, UiO-68)

The UiO family of MOFs, named after the University of Oslo, is among the most chemically robust and thermally stable metal–organic frameworks [59]. Their exceptional durability arises from the strong coordination between zirconium clusters and carboxylate linkers, making them highly resistant to water, heat, and harsh chemical conditions. Unlike many MOFs, which suffer from structural collapse in humid environments [84], UiO materials maintain their integrity, making them ideal for water-stable catalysis [85], adsorption [86], and energy storage [87]. Among the widely studied UiO MOFs, UiO-66, UiO-67, and UiO-68 exhibit subtle yet important differences. UiO-66, constructed with terephthalic acid (BDC) linkers, is the most extensively investigated, owing to its high surface area and tunable porosity [88]. Expanding on this design, UiO-67 and UiO-68 feature longer linkers—biphenyl dicarboxylate (BPDC) and terphenyl dicarboxylate (TPDC), respectively—resulting in increased pore sizes and enhanced molecular diffusion [60].
A major advantage of UiO MOFs lies in their versatility in functionalization. The carboxylate linkers can be modified with amines (-NH2) whereupon an increase in ionic conductivity is observed, benefiting from the new positively charged sites [89], as well as an increase in basicity of the Zr6O4(OH)4 clusters, which allows for higher acid adsorption properties [90]. Other modifications, such as functionalizing with hydroxyls (-OH) or sulfonic acids (-SO3H), lead to significant alterations in their surface properties and compatibility with other materials [35]. These functional groups enhance hydrogen bonding, acid–base interactions, and even covalent bonding with polymers, making UiO MOFs excellent candidates for hybridization with POPs. Strategies such as post-synthetic modification (PSM) and linker exchange further refine their properties, enabling the formation of composite materials that combine high porosity with enhanced mechanical flexibility [14,41].

2.3. MIL-Series MOFs (MIL-101, MIL-53, MIL-100)

The MIL (Matériaux de l’Institut Lavoisier) family of MOFs, predominantly constructed from iron (Fe), chromium (Cr), aluminum (Al), or vanadium (V) clusters and polycarboxylates [91,92,93], is distinguished by its large pore volume, high surface area, and remarkable chemical adaptability. These properties make MIL MOFs exceptionally versatile for catalysis, gas adsorption, drug delivery [94], and energy applications [95]. Among the most renowned MIL MOFs, MIL-101 stands out due to its massive mesoporous cages (2.9 nm and 3.4 nm in diameter), which enable exceptional molecular diffusion and guest encapsulation [96]. Another important member, MIL-53, is known for its “breathing effect,” where its pores dynamically expand and contract in response to external stimuli, enhancing its performance in CO2 capture and adsorption-based applications [67].
To enhance their interaction with polymers, MIL MOFs can be post-synthetically modified with sulfonic acid (-SO3H), amine (-NH2), or hydroxyl (-OH) groups. These modifications improve hydrophilicity, chemical reactivity, and interfacial adhesion with POPs, ensuring better mechanical stability in composite materials. Hybridization strategies often exploit MIL-101’s large pores, which allow polymer chains to infiltrate the framework, resulting in highly stable interpenetrated hybrid structures with enhanced gas storage and catalytic properties [97].

2.4. HKUST-1 (Cu3(BTC)2, Copper–Benzene Tricarboxylate MOF)

HKUST-1 (where HKUST stands for Hong Kong University of Science and Technology) is one of the most recognizable MOFs, featuring copper-based Cu2 paddlewheel units bridged by benzene-1,3,5-tricarboxylate (BTC) linkers [98]. With its high porosity (~1800 m2/g) and excellent gas adsorption properties, it has been extensively utilized in gas storage [99,100], catalysis [101], and separation technologies [99]. However, a major limitation of HKUST-1 is its moderate stability in humid environments [84], as water molecules can coordinate with Cu2+ sites, leading to gradual framework degradation. To mitigate this, functionalization strategies involving hydrophilic or hydrophobic modifications are employed [43]. By modifying BTC linkers with hydroxyl (-OH) or carboxyl (-COOH) groups, HKUST-1 can also achieve better compatibility with polymers, improving interfacial adhesion and long-term stability.
Hybridization strategies often leverage ligand modification and direct coordination with polymer chains, ensuring that HKUST-1 retains its high adsorption capacity while improving mechanical strength [102].

2.5. MOFs with Pillared-Layer Structures (MOF-5, DUT-Series)

MOF-5, one of the earliest and most studied MOFs, consists of zinc oxide clusters coordinated with benzene dicarboxylate (BDC) linkers, forming a highly porous cubic framework [70]. This structure has inspired a range of derivatives with enhanced chemical stability and tunable porosity, including the DUT (Dresden University of Technology) series, which incorporates longer and more flexible linkers to optimize gas storage and separation properties [103].
Functionalization strategies (Figure 1), as with all other carboxylate-based MOFs, focus on modifying the BDC linkers with amine (-NH2), hydroxyl (-OH), or sulfonic acid (-SO3H) groups, so that compatibility with polymeric matrices can be improved [104], or combining different ligands (azolates and carboxylates) in the same structure [105], yielding high CO2 selectivity.
Therefore, functionalization plays a pivotal role in improving the compatibility between MOFs and porous organic polymers. The abovementioned covalent functionalization, where reactive groups such as amines, thiols, or sulfonic acids are introduced onto the MOF linkers, provides functional groups that can form strong covalent bonds with polymer chains, creating a more integrated and durable hybrid structure [106]. In addition to covalent bonding, non-covalent interactions can also play a crucial role in enhancing interfacial adhesion between MOFs and POPs. For instance, hydrogen bonding, π–π stacking, and electrostatic interactions help reinforce the hybrid interface, improving the overall mechanical integrity of the material while also maintaining high porosity [48].
Some widely used techniques for this functionalization are post-synthetic modification (PSM) and ligand exchange. Whereas the first allows for the introduction of functional groups onto the MOF framework after its initial synthesis and provides greater flexibility in tailoring the surface chemistry of MOFs to make them more compatible with polymer networks (thereby facilitating stronger interactions and better integration), in the latter some of the original MOF linkers are replaced with new, polymer-compatible moieties [107]. This modification enhances the hybridization potential by introducing linkers that are more chemically suited to interacting with POPs, all while maintaining the structural integrity and porosity of the MOF. In the following sections, we will focus on the latest advances in synthetic approaches and applications in MOF-POP hybrids based on the principal families of these porous organic polymers.

3. MOF-PIM Hybrids (MMMs)

PIMs are a specific type of POP with some specific properties that make them very suitable for making hybrid materials [42,108]. These polymers comprise fused rings and a contorted structure that leads to an inefficient packing of the chain, giving rise to sub-nanometer micropores and high fractional free volume, leading to Brunauer–Emmett–Teller (BET) surface areas of 500–800 m2·g−1. Moreover, thanks to the not-crosslinked structure of these polymers, they are able to generate microporous membranes without any activation or modification and using simple techniques [109], which makes them very attractive for gas separations or solvent purification purposes and for industrial application, since they can be processed into films by simple techniques like film casting [110,111,112].
While MOFs and PIMs have some disadvantages when are used separately, like aging and the low gas selectivity of most PIMs, and the non-processability of MOFs, their combination can lead to enhanced properties, increasing the lower separation efficiency of the polymers and allowing inorganic fillers to be processed into scaled-up materials, leading to advances in fields like gas separation, CO2 adsorption, catalysis, or energy applications (Figure 2) [113,114].
Hybrid materials combining MOFs and PIMs are usually used as membranes, called mixed-matrix membranes (MMMs). These membranes were first developed by Paul and Kemp in 1973 using zeolite and silicon as components [115]. They are usually formed by a filler and a polymer matrix. Due to their properties, MOFs are very versatile fillers for MMMs, since their high porosity and gas permeability allow them to be used in separation processes for industry application [115]. Moreover, these fillers need a polymer matrix with good performance in gas separation and easy processability. PIMs can be processed in solution by using a wide range of solvents and have a high gas permeability, which makes them a good candidate to combine with MOFs.
However, there are some challenges in the fabrication of MMMs. For instance, PIMs are chemically and mechanically stable, but they are prone to aging, which can reduce permeability efficiency and increase selectivity over time due to chain relaxation over time [113,116]. Hybridizing PIMs with MOFs mitigates these drawbacks while enhancing performance [117,118]. Moreover, MOFs restrict polymer chain mobility, reducing swelling and plasticization while preventing long-term aging [119,120].
In gas separation application, the incorporation of highly selective fillers can enhance permeability by increasing the free volume in the membrane and improving the selectivity, breaking with the upper bound trend [121], where normally one of them increases and the other one decreases [122]. However, two of the main problems of MMMs using PIMs are the interface compatibility with the inorganic filler, leading to gaps, poor dispersion, and compromised selectivity and permeability, which can limit the performance of the membrane [123,124], and the particle aggregation, which can lead to inhomogeneity in the hybrid material due to the MOF particle agglomeration in the polymer matrix and poor mechanical properties.
There are some strategies to follow to solve these problems. On the one hand, introducing functional groups like a hydroxyl group (OH) between filler and polymer can improve adhesion, but does not always work as desired [50]; MOFs can be incorporated into polymer matrices using noncovalent interactions such as hydrogen bonds or π–π stacking, which is achieved by dispersing MOF ink into a polymer solution and casting films; or more flexible polymers or even a combination can be chosen, like some PIM–polyimides, which are more flexible but can also create hydrogen bonds with MOF fillers like ZIF-8 [125,126]. On the other hand, to mitigate particle aggregation, optimized dispersion techniques and covalent bonding can improve uniformity and homogeneity.
Taking into account the strategies to improve the interaction between filler and polymer, different fabrication methods have been developed, but two of them are the most used ones:
  • Solution mixing: The polymer is first dissolved in a suitable solvent, and the MOF filler is dispersed in the same solvent, often using ultrasonication to achieve a homogeneous mixture. This approach is simple and widely applied; however, it may involve risks such as partial loss of MOF crystallinity, especially under prolonged ultrasonication or in the presence of aggressive solvents, which can compromise the structural integrity of the framework and reduce its adsorption performance. Similarly, depending on the polymer chemistry, some POPs may undergo chemical degradation or swelling during extended exposure to polar solvents or elevated temperatures used in casting [127,128].
  • In situ method: This involves mixing the filler (typically pre-formed MOFs) with monomers prior to polymerization, allowing the filler to become embedded within the growing polymer network. A variation of this approach is in situ MOF growth, where MOF precursors are added to a pre-formed polymer solution, and crystallization is induced within the polymer matrix. This technique often enhances MOF dispersion and interfacial compatibility but requires precise control of synthesis conditions to avoid defective MOF growth or unintended interactions with functional groups in POPs, which might limit polymer performance or porosity development. In some cases, acidic or basic by-products generated during MOF crystallization can disrupt the polymer structure, especially in POPs containing hydrolytically sensitive linkages such as imines or boronate esters. Moreover, the thermal and chemical conditions required for successful MOF nucleation (such as solvothermal treatment or the presence of modulators) may exceed the stability window of certain polymers, resulting in chain scission, crosslinking, or partial collapse of the porous architecture. Therefore, careful optimization of the precursor ratios, reaction time, and solvent environment is essential to preserve both the structural integrity of the MOF and the intrinsic microporosity of the polymer phase [124,127,128].

Applications of MOF-PIM MMMs

MMMs are versatile composite materials that can be used in many fields, due to the advanced properties obtained from the combination of the filler and the polymer matrix. In the case of MMMs made from MOF–filler and PIM polymers, the main applications are in the gas separation and CO2 capture fields, but others, like hydrogen separation, natural gas purification, organic solvent filtration, and propylene separation, are also described, as can be seen in Figure 3 [105,126,129,130,131,132,133,134,135,136,137,138].
The MMM family formed from CAU-12-ODB and PIM-1 is an example of a hybrid material used in hydrogen permeability applications. Chi Zhang et al. reported a very selective MMM with an H2/N2 selectivity of 127 and a PH2 of 7199 Barrer (Barrer is the typical unit used to quantify gas permeability in polymer membranes. It is defined as the amount of gas (in cm3, at standard temperature and pressure) that passes through a membrane of 1 cm thickness and 1 cm2 area per second, under a pressure gradient of 1 cmHg. Mathematically, 1 Barrer = 10−10 cm3 (STP)·cm/(cm2·s·cmHg). This unit integrates both the diffusivity and solubility of the gas in the membrane, making it a fundamental parameter in characterizing gas transport properties). CAU-12 filler has a pore diameter of 3.3 Å, just between the kinetic diameters of H2 and N2 (2.9 and 3.64, respectively), which, together with the high permeability of the PIM-1 and the good interfacial adhesion between them, makes them very suitable for hydrogen separation [129].
Shaohui Xiong et al.’s research is another example of using MMM hybrid materials for a hydrogen-related application—in this case, oriented toward H2 purification. They constructed an MMM in situ with ZIF-8 and amidoxime-functionalized PIM-1 (AO-PIM-1), achieving remarkable performance in H2/CO2 separations, with a selectivity of 11.97 and an H2 permeability of up to 5688 Barrer. In addition, they demonstrated that defect-free MOF/PIM MMMs are possible when they are made on a SiO2 support by using an in situ growth method [139].
ZIF-8 MOF has also been used to improve the O2/N2 separation performance of PIM-1. Composite membranes were made by growing ZIF particles on the PIM-1 membrane, reaching O2 permeabilities of 1287 Barrer and an O2/N2 selectivity of 3.7, exceeding the Robeson upper bound line [121,128,140,141] (The Robeson Upper Bound was first defined by Robeson in 1991 as an empirical trade-off curve between gas permeability and selectivity for polymeric membranes [140]. It was updated in 2008 to reflect advances such as PIM-1 and PIM-7 [121], with further refinements in 2015 (for O2/N2, H2/N2, H2/CH4) and again in 2019 for CO2/CH4 and CO2/N2 separations [141]. Materials that surpass these boundaries (such as thermally rearranged polymers and ultra-rigid PIMs like benzotriptycene-based variants) demonstrate exceptional separation performance, often via enhanced diffusivity or dual-mode transport. Their ability to redefine the boundary sets new targets, enabling more energy-efficient and cost-effective gas separation technologies). Ma X et al. used ZIF-8/PIM-6FDA-OH MMMs to enhance propylene/propane separation, a very important industrial process. They used an ultra-microporous ZIF-8 and a dihydroxyl-functionalized polyimide (PIM-6FDA-OH), demonstrating strong interactions between the OH-functionalized polyimide and the imidazole framework between H bonding. They reported propylene permeability data significantly above the upper bound, using up to 65% ZIF content and maintaining a C3H3/C3H8 selectivity of 31 [126].
Another example is the one reported by Ma X et al., who fabricated highly propylene/propane-selective mixed-matrix membranes composed of a hydroxyl-functionalized microporous polyimide (PIM-6FDA-OH) and the well-known zeolitic imidazole framework (ZIF-8). They report an excellent compatible MMM loaded up to 65% with filler, which is a high, uncommon load for an MMM, thanks to the hydrogen bonds formed from the nitrogen and the hydroxyl groups. They reach a 3.5 Barrer permeability for the propylene and a propylene/propane selectivity of 30 [126].
However, gas separation properties and CO2 capture are the main applications of PIM/MOF-based MMMs since the high permeability of the MOFs and the adequate gas properties and processability of the PIMs allows for the creation of membranes that combine both high permeability rates with high selectivity. Moreover, the dispersibility and interface compatibility in these materials are improved by using the techniques mentioned above [48,142,143,144].
Zhang et al. reported in 2024 that CO2 permeability is one of the most relevant industrial applications of MMMs. They used PIM-1 combined with the CALF-20 MOF to improve the permeability, which in turn enhanced the overall gas separation performance. In the study, they used ultrasonication to reduce the particle size of the CALF-20. Only a 5% load in a 5%-CALF-20/PIM-1 membrane increased the CO2 permeability by 38.7% (8003 Barrer) compared with the pure PIM-1 membrane. CO2/N2 selectivity also increased up to 25%, surpassing the Robeson upper bound. However, when they increased the load to 10 or 20%, the mixed-matrix membrane did not yield improved gas separation properties, probably due to the agglomeration of the filler into the polymer matrix. They concluded that a 5% load was the limit to obtain an ultrafast CO2 selective membrane, thanks to the stronger attraction of the CALF-20 filler and the rigid PIM-1 molecular chains [105,145].
In another communication, Nguyen Tien-Binh et al. used a novel filler called Mg-MOF-74 (also called CPO-27) to improve the interface compatibility with PIM-1. They used specific temperature and time conditions to mix these two components (65 °C and 24 h) in chloroform, and thanks to the crosslink between the hydroxyl groups of the filler with the fluoride chain-ends of the PIM-1, the interfacial defects were minimized. The enhanced interaction allowed for the generation of interconnected 1-D hexagonal channels of 1.1 nm in diameter, forming an inter-connected micropore network that increased CO2 permeability by 3.2 times, from 6500 Barrer for the pure PIM-1 to 21,000 Barrer for the 20%-CPO-27/PIM-1. Meanwhile CO2/CH4 selectivity improved from 12.3 to 19.1. However, these MMMs showed a lower elasticity (a decrease of 30–70% compared with the pure PIM-1), probably due to the crosslink. They also tested the membrane after removing the MOF with acetic acid, seeing a decrease in CO2 permeability from 21,200 to 7500 Barrer, almost the same as that of the pure PIM-1. This experiment shows clearly the MOF contribution and the high PIM-1 stability. The also saw a decay in the CO2/CH4 and CO2/N2 MMM selectivity from 19.1 and 28.7 to 10.6 and 15.4, respectively, which did not happen in the O2/N2 and CH4/N2 ones, which were still low [146].
Another study published by Guangcan Huang et al. used the MOF filler MIL-101-HNO3 (also known as MIL-101(Cr), where Cr stands for chromium) due to its improved dispersibility properties in solvents and good relationship with the PIM-1 matrix due to its optimal particle size and exposure of crystallographic planes. The 15%-MIL-101-HNO3/PIM-1 MMMs showed an exceptional CO2 permeability and CO2/N2 selectivity, of 14,879 Barrer and 24.3, respectively. They used acid modulators as nitric acid to change the morphology characteristics of the MOF, resulting in reduced particle size with a better filler dispersibility and more [220] crystallographic planes, allowing the polymer to penetrate the filler crystal pores, enhancing even more the interfacial compatibility and facilitating the CO2 transfer inside the mixed-matrix membrane. They also performed some N2 adsorption experiments, with that MIL-101-HNO3/PIM-1 showing the minimum adsorption compared with the other MOFs under study, leading to a better selectivity versus CO2. In addition, they compared different permeability and selectivity results varying the filler load, and they saw that loads of 5–15% led to increased CO2 permeability and CO2/N2 selectivity, higher than those of PIM-1; meanwhile when increasing the filler load from 20 to 40%, the MMMs reached a CO2 permeability of almost 30,000 Barrer, but the selectivity decreased almost 50%. They also evidenced that the fillers did not stop polymer aging of the PIM [147].
MOF-801 was also used to improve PIM-1 CO2 permeability. Wenbo Chen et al. reported an MMM with enhanced CO2 permeability and selectivity, and this time, they improved the aging properties, with 70% of the CO2 permeability remaining after the 90-day aging test. Since the particle size in MOF-801 is 500 nm, the dispersion in the polymer matrix was good, which directly affected the gas separation properties. CO2 permeability increased from 4200 Barrer for the pure PIM-1 membrane to 9686 Barrer for the 5%-MOF-801/PIM-1 membrane. CO2/N2 selectivity also increased from 20 to 27 since PIM-1 membranes have low selectivity. The MOF cavity diameter (7.4 A) and the increased free volume in the polymer matrix led to fast and improved CO2 selective transport, but when the filler load surpassed 5%, selectivity started to decay and permeability increased, which indicates that the best performance was obtained with a 5% load [148]. Lower loads were used by Xuebi Du et al. and his team. They synthesized a 2%-MOF-808/PIM-1 mixed-matrix membrane that increased the CO2 permeability from 4022 Barrer for the pure PIM-1 to 6854 Barrer for the MMM. They also saw an increase in the CO2/N2 and CO2/CH4 selectivity of 27.4 and 1.7%, reaching a 23.2 selectivity value for CO2/N2, which is not that high compared with other MOFs [105,146,147,148] but still increased the pure PIM-1. They also tested a long-term experiment to see the gas property retention, and after 6 days they saw sustained efficacy of the PIM/MOF-808-2% membranes [42].
Another filler used not only to increase the gas permeability and selectivity but also to prevent the aging of the polymer is the MUF-15 MOF (Massey University Framework-15). Hang Yin et al. reported the first attempt to make an MMM with this MOF and PIM-1. Due to its high BET surface area (1260 m2/g) and pore volume (0.46 cm3/g), it has a very high CO2 adsorption capacity. It also possesses remarkable properties as a filler due to its aperture size (0.36 nm) and its crystalline structure, which allows crystalline two-dimensional nanosheets that can cover larger area of the final membrane compared with other morphologies. They saw that more than a 5% filler load led to defects at the polymer interface and agglomeration problems. Moreover, the CO2/N2 and CO2/CH4 selectivity decreased. The 15%-MUF-15/PIM-1 membranes showed a remarkable 98% increase in CO2 permeability compared with the pure PIM-1 but lower selectivity; thus, the 5% filler load was the most equilibrated mixture, since CO2 permeability increased 38% and CO2/N2 selectivity increased to 17.72. However, this result highlights lower values than for previous MMMs with other PIM-1/MOF combinations. On the other hand, they conducted 35-day aging studies and saw that 2% and 15 loads prevented aging, but 2% did so better since the 15% load had interfacial voids and filler agglomeration, which did not prevent the polymer from compacting. Furthermore, the relative CO2 permeability was the best preserved [149].
Yanan Wang et al. observed that combining a 20% NH2-ZIF-7 with PIM-1 provided MMMs with enhanced CO2/CH4 diffusion and selectivity and a strong interface, using experimental and predicted data. They used MOF nanoparticles of both NH2-ZIF-7 and ZIF-7 and PIM-1 to cast the MMMs by the solution mixing method. Thanks to the nano-particle size of the filler, they did not find any aggregation or sedimentation under a 30% filler load, but when they increased beyond this load, big aggregations and interfacial voids could be observed, leading to a deterioration of the separation and selectivity properties. NH2-ZIF-7 loads of 10 and 20% seemed to be good since transparent and homogeneous films were obtained. In addition, they performed gas separations tests using pure PIM-1 and MMMs. When they increased the filler load up to 20%, CO2 permeability decreased from 4533 Barrer for pure PIM-1 to 2953 Barrer for NH2-ZIF-7/PIM-1, but CO2/CH4 selectivity increased from 12.5 to 20.6. When they increased the load to 30%, the permeability increased dramatically, with the corresponding selectivity drop. In conclusion, the study reported a filler able to improve selectivity but not permeability, which makes it less interesting than other MOFs commented on before for gas separation applications. However, aging of the polymer and mechanical properties were improved. They tested ZIF-7/PIM-1 and NH2-ZIF-7/PIM-1 membranes and compared them with pure PIM-1. After 120 days, pure PIM lost 50% of its CO2 permeability, while the MMM with ZIF-7 decreased by 39% and the amino one by just 26%, concluding that aging and plasticization effects were reduced by the introduction of this NH2-ZIF-7 filler into PIM-1 and proving that amino groups enhance interfacial interactions, making the MMMs stronger [150].
In another study, Yang Feng et al. achieved higher values of CO2 permeability and even higher ones for CO2/CH4 selectivity without losing any of them, using the glass phase of ZIF-62 (agZIF-62). They used a thermal treatment that eliminates the interfacial voids between the filler and the polymer while increasing at the same time the PIM-1 free volume [137]. By heating for 30 min at 373 K and then raising the temperature to 693 K and maintaining it for 5 more minutes, they obtained agZIF-62. Then, they used the solution casting method to make two different MMMs, agZIF-62/PIM-1 and ZIF-62/PIM-1. Yang’s group used different filler loads (10%, 20%, 30%, 40%, 50%) to test the differences. When they used the ZIF-62/PIM-1 membranes, they found an increase in the CO2/CH4 selectivity with respect to that of the PIM-1 pure membrane to 37.2, 41.5, 49.8, 39.6, and 30.8, respectively. Moreover, when they used a 30% filler load, they found a maximum increase in selectivity. In addition, the maximum CO2 permeability was found for the 50% filler load. However, when the experiment was repeated with the homologous membrane with agZIF-62/PIM-1 in the same ratios, they found a similar increase for the CO2 permeability, a decrease in the CH4 permeability in proportion to the filler load, and a marked increase in the CO2/CH4 selectivity to up to 67 using a 30% load. For this MMM ratio they also found a CO2 permeability of 5914 Barrer, much higher than the values corresponding to PIM-1 pure membrane (4654 and 18.1). In both cases, they surpassed the Robeson upper bound line of 2019 and most of the state-of-art MOF-based MMM CO2/CH4 selectivity [151].
On the other hand, Chumei Ye and his team incorporated a nano-sized ZIF-67 (ZIF-S) and a submicro-sized ZIF-67 (ZIF-L) with larger surface areas and better compatibility to PIMs to enhance the selectivity. Different loads of filler, from 5 to 20%, were compared. In all cases, they saw by XRD that when increasing the filler load, membrane density increased, giving rise to a reduced fractional free volume, which is in accordance with the variation in the d-spacing value observed, which verifies that polymer chain rigidization occurred that was induced by the well interfacial interactions with the filler. When they performed separation tests for the ZIF-S/PIM-1 membrane, they found a decrease in CO2 permeability with an increase in filler load from 4521 for the pure PIM to 2805 for the 15% load MMMs, with the permeability decrease being slightly lower for the 20% filler content. The maximum CO2/CH4 selectivity was found for a 15% ZIF-S load, increasing from 12.45 to 21.09, which is 69.4% higher. All the other selectivity also increased (CO2/N2: 36.2%, H2/CH4: 99.7%, and H2/N2: 60.7%). Surprisingly, when they repeated the same experiment using ZIF-L as the load, the permeability increased with the filler content, reaching a maximum for 20%, but both selectivity values decreased compared with the ZIF-S load. This is well known since ZIF-8 performs very similarly [128]. Good interfacial compatibility of ZIF-S particles with the polymer matrix could facilitate the formation of partial pore blockage regions, generating improvements in CO2/CH4 selectivity but resulting in lower CO2 permeability due to the narrowed effective aperture size of ZIF-S particles. When the filler content was increased to 20% in the ZIF-S/PIM membrane, the CO2 permeability increased, which can be attributed to the interfacial defects leading to non-selective gas diffusion, which was also proven by the aggregation observed by SEM.
Finally, they performed a CO2 permeability retention test measured after 120 days and observed a 70% permeability maintenance, compared with 50% for the pure PIM-1 membrane [82].
The data collected in Table 2 demonstrates that the performance of MOF/PIM-based MMMs in gas separation applications is highly dependent on both the choice of MOF filler and the targeted gas pair. For hydrogen separations, the CAU-12-ODB/PIM-1 membrane stands out for its outstanding permeability and exceptional H2/N2 selectivity. In the case of O2/N2 separation, although only one example (ZIF-8/PIM-1) is reported, its performance appears promising for further development. For propylene/propane separations, the ZIF-8/PIM-6FDA-OH MMM provides favorable selectivity, even if its permeability is modest. Regarding CO2 separation, a trade-off between permeability and selectivity is observed. On one hand, the Mg-MOF-74/PIM-1 membrane exhibits the highest CO2 permeability, but with only moderate selectivity. On the other hand, MIL-101-HNO3/PIM-1 offers an excellent balance for CO2/N2 separations, combining high permeability with respectable selectivity. Additionally, for CO2/CH4 separation, the agZIF-62/PIM-1 membrane emerges as the best option with the highest selectivity, even though it may not reach the peak permeability values observed in other cases.
Overall, all these data indicate that by carefully selecting the type of MOF filler and optimizing its loading within the PIM matrix, it is possible to tailor the membrane’s properties to meet specific gas separation requirements.

4. MOF-COF Hybrid Materials

The development of hybrid materials based on MOFs and COFs has emerged as a rapidly growing field of materials research. First described in the late 1990s, metal–organic frameworks (MOFs) are crystalline materials composed of metal ions or clusters coordinated to organic linkers, resulting in extremely porous structures [153,154]. MOFs have received a lot of interest for their large surface areas, variable porosity, and functional diversity in applications including gas storage, catalysis, and sensing [155,156,157,158,159]. However, their resistance to strong chemical or heat conditions remains a significant constraint in several circumstances [160,161].
Covalent organic frameworks (COFs), introduced in the mid-2000s, are composed entirely of light elements (e.g., carbon, hydrogen, nitrogen, oxygen) connected through covalent bonds. Their highly ordered, crystalline nature and permanent porosity make them suitable for applications requiring chemical robustness and long-term stability [162,163,164,165]. COFs also exhibit chemical versatility and an ease of functionalization comparable to MOFs. Despite these advantages, COFs often lack the tunable electronic and catalytic properties provided by metal centers like those found in MOFs [166].
The combination of MOFs and COFs benefits from the complementary strengths of both materials, with the high structural diversity and tunable functionalities of MOFs being combined with the chemical stability and well-defined pore architectures of COFs [167]. This allows the limitations inherent to each class of material to be overcome, leading to the design of multifunctional materials with improved performance in various fields [24,168]. As a case in point, the enhanced thermal and chemical stability, greater structural diversity, and optimized porosity offered by MOF–COF hybrids make them a highly promising material for advanced applications such as catalysis, energy storage, and gas separation. The following table (Table 3) presents a comparative overview of the characteristics of MOFs, COFs, and their hybrid materials.

4.1. Type of MOF/COF Hybrids

There are several types of MOF/COF hybrids based on how the MOF and COF components interact with each other.

4.1.1. Strategies for the Synthesis of MOF-COF Hybrid Materials

The combination of metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) into hybrid materials has garnered significant attention due to their complementary properties. Different synthesis strategies have been developed to fabricate MOF-COF hybrids, including the called MOF-first strategy, the COF-first strategy, and post-synthetic linking of pre-synthesized MOFs and COFs (Figure 4) [169]. Each approach offers distinct advantages in terms of structural stability, functional integration, and synthetic control.

4.1.2. MOF-First Strategy

In this approach, the MOF is synthesized first and subsequently serves as a template for COF growth, resulting in MOF@COF hybrid materials. This method enhances the stability and functionality of the final composite by leveraging the inherent porosity and structural integrity of the MOF (Figure 4). Several variations of this strategy have been explored [170,171,172].
One commonly employed method is in situ COF growth, in which the COF is directly formed on the MOF surface. This process can involve either covalent bonds, which provide a strong and stable interaction, or non-covalent interactions, which offer a more flexible but weaker linkage. A typical example of this strategy is the development of core–shell structures, where the MOF forms the core and the COF forms the shell, enhancing the material’s surface area and catalytic activity [106,173].
Another variation involves the transformation of an amorphous layer into a crystalline COF layer. Initially, a non-crystalline layer of COF precursors is deposited on the MOF surface. Through thermal or chemical treatments, this layer undergoes structural rearrangement, leading to the formation of a well-defined crystalline COF. This method ensures better integration between the two frameworks and results in a highly ordered hybrid material [174,175,176].
A third approach under the MOF-first strategy is one-step COF synthesis, where the MOF and COF are synthesized simultaneously within a single reaction process. This simplifies the fabrication method and promotes direct growth of the COF onto the MOF. However, achieving uniform structures requires precise control over reaction parameters to balance the nucleation and growth kinetics of both components [177,178,179].

4.1.3. COF-First Strategy

In contrast to the MOF-first strategy, the COF-first strategy involves the initial synthesis of the COF, which subsequently acts as a scaffold for MOF growth (Figure 4). This method is particularly useful when the COF’s intrinsic porosity and functional groups need to be pre-established before the MOF is introduced. Two main techniques have been developed within this framework.
The layer-by-layer growth approach involves the sequential deposition of MOF layers onto the COF surface. This method enables fine control over the thickness and composition of the MOF layer, leading to tunable hybrid structures with tailored properties. The gradual growth process ensures homogeneous coverage of the COF and facilitates the formation of well-defined MOF domains.
Another technique is sequential MOF-on-COF growth, in which the MOF is grown on the COF surface through controlled chemical reactions. This approach ensures strong adhesion between the two frameworks, leading to a robust hybrid material with enhanced mechanical stability and chemical resilience [180,181,182].

4.1.4. Post-Synthetic Linking of Pre-Synthesized COFs and MOFs

A third strategy for synthesizing MOF-COF hybrid materials involves the post-synthetic linking of separately prepared MOFs and COFs (Figure 4). In this approach, both components are synthesized independently and then integrated through either covalent bonding or physical interactions. Covalent bonding offers a strong and stable linkage between the MOF and the COF, facilitating the formation of durable hybrid materials with well-defined interfaces. Alternatively, physical interactions such as hydrogen bonding or π–π stacking provide a more flexible means of integration, allowing for dynamic hybrid structures that can be tuned for specific applications. This post-synthetic method provides high versatility, as different MOF-COF combinations can be designed and assembled based on desired properties and functionalities [170,171,172].
The selection of the appropriate synthetic strategy for MOF-COF hybrids depends on the desired outcome for the material’s stability, structural integrity, and functionality. The MOF-first approach is best for enhancing the stability and performance of MOFs. The COF-first strategy is ideal for structural control and design flexibility, and post-synthetic linking allows for modular, customizable assembly of hybrid materials (Table 4).

4.2. Applications of MOF/COF Hybrids

Hybrid materials derived from the integration of metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) have emerged as promising candidates for a wide range of advanced applications. Owing to their complementary structural and chemical features, these hybrid systems exhibit enhanced performance in various fields. The following sections provide an overview of their potential in key application areas, including pollutant adsorption, catalysis, energy storage and conversion, sensing technologies, biomedical applications, and gas adsorption and separation (Figure 5) [175,177,182,188,189,190].

4.2.1. Sensors

The combination of metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) has led to the development of advanced sensor materials with enhanced sensitivity and selectivity. For instance, a hybrid material composed of Ce-MOF@MCA was synthesized by combining a melamine–cyanuric acid COF with a cerium-based MOF. This hybrid demonstrated high electrochemical activity and strong bio-affinity, leading to the development of an aptasensor for oxytetracycline (OTC) detection with an extremely low detection limit of 35.0 fM [182]. Another example is the Co-MOF@TPN-COF-based aptasensor, which was used for detecting ampicillin, showing a remarkable limit of detection (LOD) of 0.217 fg/mL [181]. Furthermore, a UiO-66-NH2@COF hybrid was employed in electrochemical analytical methods for adenosine triphosphate (ATP) and chloramphenicol (CAP) detection, achieving a detection limit of 5.04 fg/mL [191]. Another hybrid with the same MOF, UiO-66-NH2@TpTt-COF (UNT), a core–shell MOF-COF composite, is used for tetracycline detection due to its dual emissions and its ability to suppress aggregation-caused quenching (ACQ), resulting in high sensitivity and specificity. UNT is used for the detection of tetracycline in soil and river water, with recoveries of 92.96–98.44% and 88.58–104.31%, respectively [192].

4.2.2. Catalysis

MOF–COF hybrids have demonstrated remarkable potential in catalytic applications, particularly in heterogeneous and photocatalysis applications. They are perfect candidates for catalytic reactions because of their structural characteristics, which include stability, well-defined active sites, and large surface areas.
The catalytic and photocatalytic applications of MOF-COF hybrids have been extensively explored, showcasing their exceptional potential in various reactions. A prime example is the PCN-222-Co@TpPa-1 hybrid, where a covalent organic framework (COF), TpPa-1, is grown on a cobalt-based metal–organic framework (MOF), PCN-222-Co. This hybrid demonstrated remarkable recyclability and high catalytic activity in the deacetalization–Knoevenagel reaction [193]. Similarly, the sandwiched Pd/UiO-66-NH2@COF composite exhibited selective catalytic activity in hydrogenation reactions, driven by the molecular size, offering a tailored approach to catalytic processes [175].
In the field of photocatalysis, MOF-COF hybrids have also garnered significant attention. The NH2-UiO-66@TFPT-DETH hybrid achieved a hydrogen evolution rate of 7178 μmol/g/h, while another hybrid, NH2-UiO-66/TpPa-1-COF, demonstrated an even superior photocatalytic hydrogen evolution rate of 23.41 mmol/g/h [168]. These materials highlight the impressive photocatalytic efficiency of MOF-COF hybrids. In addition to these materials, the IR-MOF3@COF-LZU1 hybrid exhibited high photocatalytic activity for the degradation of p-nitrophenol [194].
Another study demonstrated the efficiency of MOF/COF hybrids in wastewater remediation, where methylene blue pollutants were effectively degraded using visible light and hydrogen peroxide as co-reactants. NH2-MOF-5/MCOF hybrid showed high visible light absorption (up to 596 nm) and a reduced band gap (2.20 eV), facilitating the photodegradation of dyes such as methylene blue (69% degradation in 5 h), which significantly improved with the addition of H2O2 (degradation rate of 0.728 h−1) [195]. These results indicate their utility for environmental applications. Moreover, the integration of metal nodes in MOFs with nitrogen-rich COF frameworks has been employed to develop materials capable of catalyzing CO2 reduction reactions, showcasing their potential in addressing environmental challenges related to greenhouse gas emissions. The PCN-222-Cu@TpPa-1 (1:2) hybrid showed a CO yield of 90.57 μmol g−1h−1 and a CH4 yield of 21.27 μmol g−1h−1 in the photocatalytic reduction of CO2, outperforming other materials. CH4 selectivity could be improved by adjusting the ratio of the MOF and COF. This material had a specific surface area of 577.8 m2g−1, demonstrating high catalytic efficiency. In addition, the stability of the material was observed after eight cycles, with minimal variations in its structure [196].
These findings underscore the versatility and high performance of MOF-COF hybrids in both catalytic and photocatalytic applications.

4.2.3. Gas Sorption and Separation

Gas sorption and separation are among the most well-researched applications of MOF–COF hybrids due to their tunable porosity and high adsorption capacities. Recent advancements have demonstrated their use in CO2 capture, hydrogen storage, and natural gas purification.
The use of MOF-COF hybrids for gas storage and separation has been investigated due to their high surface areas and tunable porosities. A UiO-66-NH2@COF-TAPB-BTCA composite displayed enhanced N2 and H2O uptake, making it suitable for water sorption applications [197]. Similarly, COF-300 was grown on UiO-66 membranes to produce a [COF-300]-[UiO-66] composite membrane with high gas selectivity [198]. Another hybrid membrane, [COF-300]-[Zn2(bdc)2(dabco)], showed improved separation selectivity for H2/CO2 mixtures [180].
Pyridine-based COFs grown on NH2-UiO-66 cores showed large specific surface areas and favorable chemical stability, exhibiting significantly enhanced CO2 adsorption, with MOF@COF-0.05 reaching an uptake of 4.53 mmol⋅g−1. Functionalization with bromoethanol created MOF@COF-0.15-Br, which facilitated efficient catalytic conversion of CO2 to cyclic carbonates with high yields. This functionalized catalyst demonstrated stability, with negligible loss of activity after six cycles. Compared to other materials, MOF@COF-0.15-Br did not require co-catalysts, metals, or solvents for CO2 conversion [199]. Regarding H2 adsorption, the hybrid material MIL-101(Cr)@UiO-66(Zr) showed a higher H2 adsorption capacity than the base materials (MIL-101 and UiO-66). This material showed an increase in adsorption of up to 26% compared to MIL-101 and 60% more than UiO-66 [200].

4.2.4. Adsorption and Removal of Pollutants

MOF-COF hybrids have demonstrated excellent adsorption capabilities for environmental remediation. Mn-1,4-BDC/COF (Mn-1,4-benzenedicarboxylate MOF combined with an imine-linked melamine and terephthalaldehyde COF) composite is a robust adsorbent for the rapid and simultaneous removal of auramine O (AO) and rhodamine B (RB) dyes from aqueous media. It exhibited superior adsorption capacity (25.13 mg/g for AO and 21.53 mg/g for RB) compared to individual MOF Mn-1,4-BDC and COF. This improvement was due to synergistic effects that allow the simultaneous use of multiple adsorption mechanisms [177]. NH2-MIL-125(Ti)@TpPa-1 hybrid exhibited high adsorption capacity for uranium (536.73 mg/g) and europium (593.97 mg/g) [201]. Additionally, a Zn-MOF-5 and melamine–terephthaldehyde COF hybrid showed outstanding performance in removing organic pollutants from wastewater [168].

4.2.5. Energy Storage and Conversion

The high surface area, tunable pore sizes, and electrical conductivity of MOF–COF hybrids make them excellent candidates for energy storage and conversion applications, including batteries, supercapacitors, and electrocatalysis.
The incorporation of MOFs into COFs enhances the electrochemical performance of supercapacitors. For instance, a novel MOF@COF-TCNQ material demonstrated a high area capacitance of 78.36 mF cm−2 and a volumetric energy density of 4.46 F cm−3, thanks to TCNQ infiltration increasing the quantum capacitance, with MOF being UiO-66-NH2 and COF being TpPa [202]. For electrodes in supercapacitors, the UiO-66-NH2@TFP-TABQ hybrid combines the structural advantages of both MOFs and COFs. Although the MOF possesses the largest surface area (478 m2 g−1), the hybrid exhibits superior porosity, with an area of 236 m2 g−1, six times larger than that of the COF (39 m2 g−1). In addition, it exhibits outstanding pseudocapacitive behavior, with a b-value of 0.85 (compared to 0.59 for MOF and 0.83 for COF), indicating a higher contribution of redox reactions in its charge storage capacity [203]. Another aza-MOF@COF composite, synthesized via an aza-Diels–Alder cycloaddition, combines the porosity of MOF (UiO-66-NH2) with the stability of COF (LZU-1) and is further enhanced by quinoline modification. This hybrid material demonstrated a high specific capacitance of 20.35 mF/cm2 and a volumetric energy density of 1.16 F/cm3, making it highly effective for supercapacitor applications. It also exhibited excellent cycling stability, retaining 89.3% of its capacitance after 2000 cycles. The quinoline modification plays a key role in improving energy storage performance and structural robustness [204].
An interlinked hybrid of imine-based COFs and Mn-based MOFs (COF/Mn-MOF) demonstrated superior lithium storage performance, with a specific capacity of 1015 mAh/g after 650 cycles [205].

4.2.6. Biomedical Applications

A series of MOF@COF hybrid materials have been developed for biomedical and electrochemical applications, combining the properties of metal–organic frameworks (MOFs) and covalent organic frameworks (COFs). The NMCTP-TTA nanozyme integrates a peroxidase-like MOF (NH2-MIL-88B(Fe)) with a COF (COFTP-TTA) to enhance bacterial inhibition. The MOF serves as the catalytic center, while the COF’s hierarchical nanocavities activate substrates, mimicking natural enzymes. Pseudopodia-like structures on the COF facilitate bacterial capture. NMCTP-TTA generates reactive oxygen species (ROS), exhibiting a 7.9-fold increase in peroxidase-like activity over NM-88. Functional groups such as triazine and phenol boost proton and electron transfer, significantly improving bacterial inhibition and wound healing outcomes [190]. A Cu-MOF@TpBD hybrid material has been designed for electrochemical detection of PDGF-BB in human serum. The Cu-MOF core generates electrochemical signals without mediators, while the COF shell stabilizes the sensor and enables aptamer immobilization via π–π stacking and hydrogen bonding. This combination enhances bioanalyte sensitivity and signal stability. The aptasensor demonstrated excellent recovery rates (91.0–107.9%) and low RSD values (2.9–4.5%) in human serum samples. The hybrid material provides a highly selective and reproducible platform for PDGF-BB detection [206]. Targeting GL-3 antigen, a Ti-MOF@COF composite has been developed for a sandwich-type immunosensor. The hybrid material enhances biomolecule compatibility and provides a stable antibody immobilization framework. Its large surface area (1403.3 m2 g−1) amplifies signal detection, resulting in an ultra-sensitive response. The immunosensor achieves a linear range of 0.0001–20.0 ng mL−1 with a sensitivity of 0.025 pg mL−1. Ti-MOF-modified glassy carbon electrodes (GCE) present the highest amperometric response, confirming the material’s superior electrochemical performance [207]. A Cu-MOF@CuPc-TA-COF hybrid material has been designed for dual photoelectrochemical (PEC) and electrochemical (EC) detection of HIV-1 DNA. The Cu-MOF core enhances electrochemical and photoactivity, while the COF framework improves stability and probe immobilization. The hybrid structure provides a high BET surface area (514.7 m2·g−1) for efficient DNA detection. Detection limits reach 0.07 fM (PEC) and 0.18 fM (EC) within a 1 fM–1 nM linear range. The biosensor demonstrates high accuracy, with recovery rates of 90.2–114.8% and low RSD values (0.2–2.3%) in human serum samples [208].

5. MOF/HCP, MOF/CTF, and MOF/CMP Hybrid Materials

As already discussed in the first part of the review, due to the results in different research fields [209,210,211,212,213], MOFs have been hybridized with other materials [24,48,214] in order to improve their properties. Among these materials are porous materials—more specifically, POPs, which have recently been postulated as a promising alternative to improve the properties of MOFs [215]. PIMs [110] and COFs [216] are the most widely used compounds to synthesize hybrid compounds with MOFs and offer great results in different applications, such as catalysis, gas storage, and molecular separation [215]. However, relatively recently, crosslinker polymers [25,28], covalent triazine frameworks [33,34,217], and hyper-crosslinker polymers [218]—another set of POPs with very good properties for different applications that have received considerably less attention because they have not been studied as thoroughly as the other POPs, but have great potential for hybridization with MOFs—have begun to be used [215]. The hybridization of the MOFs CMP, CTF, and HCP allows the structural and functional advantages of these porous materials to be combined with the inherent properties of MOFs.
  • MOF/HCP hybrids:
The hybridization of MOFs with HCPs brings together the tunable porosity and chemical characteristics of MOFs with the thermal stability and versatility of HCPs [31]. These hybrids are particularly promising in gas capture and catalysis applications [219,220,221]. In addition, the ease of preparation of HCPs and their compatibility with various chemical conditions broaden the range of possible applications for these materials [219,222].
2.
MOF/CTF hybrids:
Hybrids of MOFs with CTFs take advantage of their structural rigidity and ability to handle extreme conditions, making them interesting materials for catalytic and gas separation applications [33,217]. The thermal and chemical stability of CTFs, coupled with the tunable functionality and high porosity of MOFs, make them particularly attractive for catalytic and gas separation applications [36,38]. However, their synthesis and the control of the MOF-CFT interfaces are areas of current active research.
3.
MOF/CMP hybrids:
MOFs can also be combined with CMPs to create hybrid materials that take advantage of their π-conjugated systems to allow for good electronic conductivity, making them ideal for applications related to photocatalytic catalysis and gas storage [27,223,224]. Such hybrids have shown potential in energy-related processes such as hydrogen production by photocatalysis, as well as in electronics applications [26,29].
As discussed above, HPC/MOF, CMP/MOF, and CTF/MOF systems have been less studied, despite offering complementary properties such as greater structural diversity, electronic tunability, and chemical stability. Below, we highlight the most relevant research on the hybrids between these compounds with a diversity of applications, with the aim of improving the individual properties of each one and overcoming the limitations of each of the separate components.
One of the first studies on these POPs hybridized with MOFs to improve their properties was carried out in 2013 when they used a hyper-crosslinked polymer to encapsulate MOFs in the micropores of the polymer to store hydrogen. They used hyper-crosslinked styrene–maleic acid beads decorated with a MOF to synthesize SMA-MOF-5, SMA-Zn-BPDC, and SMA-Zn-NDC. The last hybrid composite performed the highest, with a storage capacity of 0.61 wt% at 300 K and 0.69 bar for 24 h [219]. One of the applications where both POPs and MOFs are most used is in the adsorption of compounds, thanks to their structure, which is why researchers have recently been trying to find synergies between the two to improve their properties, as mentioned above. Li S et al. succeeded in introducing MOF MIL-101(Fe)-NH2 into a hyper-crosslinked IRA-900 resin to improve its phosphonate adsorption properties by enhancing the properties of each component separately. Separately, MIL-101(Fe)-NH2’s adsorption capacity is 7.6 mg P/g in 240 min, IRA-900’s adsorption capacity is 5.6 mg in 360 min, and the MOF/HPC hybrid has a capacity of 12.9 mg in 120 min, in addition to a high selectivity to other anions [220]. On the other hand, MIL-101(Cr) together with hyper-crosslinked aniline polymer was also used for the separation of Cr(VI), forming a series of compounds called HPAN@M-x. Specifically, HPAN@M-b (b = 5 mmol aniline) achieved a Cr(VI) adsorption of 290.0 mg/g in 7 min and could be transformed into BaCrO4 for use, and very good stability was observed, maintaining the Cr(VI) removal during five cycles [222].
Another approach for improving the properties of these hybrid composites is the introduction of carbon or the preparation of carbon from microporous solids as precursors, such as HPCs. For example, the use of graphene oxide together with MOFs has been extensive, improving selectivity, resistance to acids and bases, and basically modifying the pore structure, among other characteristics, thus allowing for its use in a wide variety of applications [225,226]. For example, for CO2 separation, Ning H et al. succeeded in synthesizing a new hyper-crosslinked polyimide with UiO together with graphene oxide (PI-UiO/GO-1), which increased the amount of CO2 adsorbed by three times and four times compared to the CO2 selectivity against the nitrogen in UiO-NH2 MOF. The CO2 adsorption of the hybrid described in the study was 8.24 mmol/g at 298 K and 30 bar versus 2.8 mmol/g for UiO66-NH2. On the other hand, the PI-UiO/GO-1 hybrid showed a selectivity of 64.71 vs. 15.43 CO2/N2 for the MOF UiO-66-NH2 [227]. From a different approach, Bauza et al. synthesized a magnetic hybrid carbon MOF from a benzene hyper-crosslinked polymer together with MIL-100(Fe), with diclofenac adsorption properties of 210 mg/g for diclofenac adsorption due to improved surface area and pore interactions [32].
These compounds can also be used as catalysts for reactions that, as in other cases, obtain the advantages of MOFs on the one hand and porous polymers on the other hand. For example, Karimi et al. synthesized PC4RA@PrSi-CA/Cu-MOF. This compound substantially reduces side reactions by directing the use of organic solvents, and it can be reused up to 10 times, improving the yield of organic reactions substantially [221]. Another very current topic is the self-healing of polymers, especially in the field of elastomers, as the increasing demand for this type of compound has led to an increase in the maximum lifetime of these compounds. In the work carried out by R. Wang et al., they synthesized hybrid membranes using a hyper-crosslinked hyperbranched polymer such as HPMA together with the MOF UIO-66-NH2, diethylenetriamine, and terephthaldehyde, forming a so-called hyper-crosslinked hyperbranched polymer @ metal–organic polyhedral (HHMOP). These membranes achieved 91.35% strength recovered after 24 h, extending the use of MOFs and polymeric membranes [228].
CTFs are relatively recently synthesized compounds, but have great potential in different applications. This is why the synthesis of CTF/MOF hybrids is being pursued for the improvement of the separate materials. The evolution of the catalysis of hydrogen is crucial for sustainable hydrogen production, energy storage, and other reasons, which is why much research focuses on it. For example, Zr-UiO-66-NH2 (UN) is a MOF with good photocatalytic capabilities for H2 evolution. In 2023, Dong, S. et al. used this compound together with benzoic acid-modified covalent triazine-based frameworks. The photocatalytic H2 ratio was 378 µ/h-g for 30NUBC, a much higher result than with the MOF and the CTF alone [35]. Shao et al. were looking for a new compound to improve the catalytic properties for this reaction, and they synthesized a series of compounds using Ni-CAT, a conductive MOF, together with CTF-1, a photocatalyst widely used individually. At a 1:19 ratio, Ni-CAT-1/CTF-1 increased the photocatalytic activity, with a 8.03 mmol/h-g H2 evolution rate versus 1.96 mmol/h-g for CTF-1 [36]. This demonstrates once again the potential of hybrid materials, especially in the field of catalysis. With greenhouse gases such as SO2 and NO2 increasing on the planet, solutions are being sought in different ways, and one with great potential is the use of MOFs and POPs, which, thanks to their structure, offer promising results. For this purpose, Sumin Li et al. used CTF to synthesize Zr-MOF-NH2 in situ by adding CU2+ to obtain more active centers, obtaining the porous hybrid Zr-MOF-NH2/CTF-CU2+. The SO2 capture was 39.3 mg/g, which was 2.6 times more than the SO2 capture with the MOF, while the NO2 capture was 27.3 mg/g, which was 3.1 times more than with the pristine MOF [38].
Crosslinker polymers are a group of POPs that are booming because of their incredible surface area diversity, building blocks permanently, which has led to the use of these composites for a wide variety of applications [25,28,229]. Hybrids of CMP and MOFs have great potential for the synergistic properties that they can provide. There is very little research that focuses on these hybrids, as hybrids with PIMs and COFs are of much greater importance. Catalysis, as has been observed in other hybrids, is a very important topic nowadays. For example, Bo Chen et al. managed to synthesize the hybrid NU/NUF/FNBZ. NU is a MOF with a good CO2 adsorption capacity, and FNNZ is a CMP with visible light absorption capacity whose synergistic properties allow for an increase in its catalytic activity. The 40NU/FNBZ composite showed a CO formation rate of 18.78 µmol/h-g, four times higher than that of FNBZ and 48 times higher than that of NU [230].

6. Perspectives on MOF/PIM, MOF/COF, MOF/HCP, MOF/CTF, and MOF/CMP Hybrids

While MOF/PIM and MOF/COF hybrids have demonstrated superior development and a wider variety of applications, hybrids based on HCPs, CTFs, and CMPs are distinguished by their unique properties, which position them as alternatives with significant potential. These materials possess the capacity to overcome specific challenges, such as demonstrating superior thermal stability in extreme situations or exhibiting enhanced electric conductivity. Even though research on MOF/HCP, MOF/CTF, and MOF/CMP hybrids is in its early stages, their complementary characteristics indicate significant potential for future technological applications. As synthesis and characterization methods are developed and improved, these systems are expected to become increasingly influential in fields such as catalysis, energy storage, and gas separation.
Looking to the future, the field of MOF-based hybrids shows tremendous promise. MOF/PIM and MOF/COF hybrids, for example, benefit from relatively mature synthetic methodologies and have already demonstrated remarkable performance in applications such as selective gas separation, membrane technologies, and advanced catalysis. Their processability and tunable porosity also make them ideal candidates for integration into scalable technologies.
However, hybrids with HCPs, CTFs, and CMPs offer complementary benefits, including exceptional thermal and chemical stability, higher electric conductivity, and the potential for tailored electric or catalytic functions. These features are of relevance in the context of challenging applications, including environments characterized by aggressive catalysis, electrochemistry, and next-generation energy storage systems.
Despite these promising perspectives, several challenges must be addressed to fully unlock the potential of MOF-based hybrid materials. Achieving precise control over hybrid architectures, improving compatibility between MOF and polymeric phases, and ensuring reproducibility and scalability are critical. Furthermore, a more comprehensive understanding of the molecular-level interactions that give rise to synergistic properties will be essential to guide rational design. The real-world applicability of these materials will be determined by several factors, including environmental considerations. Such considerations may take the form of sustainable synthesis routes and recyclability.
In conclusion, the continuous evolution of MOF-based hybrids incorporating PIMs, COFs, HCPs, CTFs, and CMPs presents a significant opportunity to extend the scope of porous materials beyond their current limitations. The combination of the structural tunability and high porosity of metal–organic frameworks (MOFs) with the diverse functionalities and robustness of organic porous networks has the potential to drive innovation across multiple sectors. Advances in synthetic techniques, characterization methods, and fundamental understanding have led researchers to believe that these materials have the potential to become key enablers of future technologies in the fields of catalysis, energy storage, and separation science.

Author Contributions

Conceptualization, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; methodology, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; validation, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; formal analysis, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; investigation, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; resources, J.C.-G.; data curation, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; writing—original draft preparation, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; writing—review and editing, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; visualization, V.D.-E., D.G.-G., J.C.M.-L. and L.P.-V.; supervision, J.C.-G.; project administration, J.C.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
MOFMetal–organic framework
POPPorous organic polymer
COFCovalent organic Framework
MMMsMixed-matrix membranes
ZIFZeolitic imidazole framework
HCPHyper-crosslinked polymers
CMPConjugated microporous polymers
SBETSpecific surface area (Brunauer–Emmett–Teller)
FFVFree fraction volume
UiOUniversitetet i Oslo
MILMaterials Institute Lavoisier
HKUSTHong Kong University of Science and Technology
DUTDresden University of Technology
PSMPost-synthetic modification
BDCBenzene dicarboxylate
BPDCBiphenyl dicarboxylate
BTCBenzene tricarboxylate
TPDCTerphenyl dicarboxylate
MTNMobil Thirty-Nine, related to the structure of the zeolite ZSM-39
VOCVolatile organic compound
NGANegative gas adsorption

References

  1. Kitagawa, S.; Uemura, K. Dynamic Porous Properties of Coordination Polymers Inspired by Hydrogen Bonds. Chem. Soc. Rev. 2005, 34, 109–119. [Google Scholar] [CrossRef]
  2. Kitagawa, S.; Kawata, S. Coordination Compounds of 1,4-Dihydroxybenzoquinone and Its Homologues. Structures and Properties. Coord. Chem. Rev. 2002, 224, 11–34. [Google Scholar] [CrossRef]
  3. Zhou, H.C.; Long, J.R.; Yaghi, O.M. Introduction to Metal-Organic Frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef] [PubMed]
  4. Henke, S. Metal-Organic Frameworks with Additional Flexible Substituents-Modulating Responsiveness, Gas Sorption Selectivity & Network Topologies. Ph.D. Thesis, Technical University Dortmund, Dortmund, Germany, 2011. [Google Scholar]
  5. Peralta, D.; Chaplais, G.; Simon-Masseron, A.; Barthelet, K.; Chizallet, C.; Quoineaud, A.A.; Pirngruber, G.D. Comparison of the Behavior of Metal-Organic Frameworks and Zeolites for Hydrocarbon Separations. J. Am. Chem. Soc. 2012, 134, 8115–8126. [Google Scholar] [CrossRef] [PubMed]
  6. Malekshah, R.E.; Moharramnejad, M.; Gharanli, S.; Shahi, M.; Ehsani, A.; Haribabu, J.; Ouachtak, H.; Mirtamizdoust, B.; Kamwilaisak, K.; Sillanpää, M.; et al. MOFs as Versatile Catalysts: Synthesis Strategies and Applications in Value-Added Compound Production. ACS Omega 2023, 8, 31600–31619. [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. Zuliani, A.; Khiar, N.; Carrillo-Carrión, C. Recent Progress of Metal–Organic Frameworks as Sensors in (Bio)Analytical Fields: Towards Real-World Applications. Anal. Bioanal. Chem. 2023, 415, 2005–2023. [Google Scholar] [CrossRef]
  9. Avci, C.; Imaz, I.; Carné-Sánchez, A.; Pariente, J.A.; Tasios, N.; Pérez-Carvajal, J.; Alonso, M.I.; Blanco, A.; Dijkstra, M.; López, C.; et al. Self-Assembly of Polyhedral Metal-Organic Framework Particles into Three-Dimensional Ordered Superstructures. Nat. Chem. 2018, 10, 78–84. [Google Scholar] [CrossRef]
  10. 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]
  11. Duan, S.; Qian, L.; Zheng, Y.; Zhu, Y.; Liu, X.; Dong, L.; Yan, W.; Zhang, J. Mechanisms of the Accelerated Li+ Conduction in MOF-Based Solid-State Polymer Electrolytes for All-Solid-State Lithium Metal Batteries. Adv. Mater. 2024, 36, 2314120. [Google Scholar] [CrossRef]
  12. Pander, M.; Gil-San-Millan, R.; Delgado, P.; Perona-Bermejo, C.; Kostrzewa, U.; Kaczkowski, K.; Kubicki, D.J.; Navarro, J.A.R.; Bury, W. MOF/Polymer Hybrids through in Situ Free Radical Polymerization in Metal-Organic Frameworks. Mater. Horiz. 2023, 10, 1301–1308. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Feng, X.; Yuan, S.; Zhou, J.; Wang, B. Challenges and Recent Advances in MOF-Polymer Composite Membranes for Gas Separation. Inorg. Chem. Front. 2016, 3, 896–909. [Google Scholar] [CrossRef]
  14. Husna, A.; Hossain, I.; Jeong, I.; Kim, T.H. Mixed Matrix Membranes for Efficient CO2 Separation Using an Engineered UiO-66 MOF in a Pebax Polymer. Polymers 2022, 14, 655. [Google Scholar] [CrossRef]
  15. Barcus, K.; Lin, P.A.; Zhou, Y.; Arya, G.; Cohen, S.M. Influence of Polymer Characteristics on the Self-Assembly of Polymer-Grafted Metal-Organic Framework Particles. ACS Nano 2022, 16, 18168–18177. [Google Scholar] [CrossRef]
  16. Uemura, T.; Uchida, N.; Higuchi, M.; Kitagawa, S. Effects of Unsaturated Metal Sites on Radical Vinyl Polymerization in Coordination Nanochannels. Macromolecules 2011, 44, 2693–2697. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Nguyen, H.T.H.; Miller, S.A.; Cohen, S.M. PolyMOFs: A Class of Interconvertible Polymer-Metal-Organic-Framework Hybrid Materials. Angew. Chem. 2015, 127, 6250–6255. [Google Scholar] [CrossRef]
  18. Gu, Y.; Huang, M.; Zhang, W.; Pearson, M.A.; Johnson, J.A. PolyMOF Nanoparticles: Dual Roles of a Multivalent PolyMOF Ligand in Size Control and Surface Functionalization. Angew. Chem. 2019, 131, 16829–16834. [Google Scholar] [CrossRef]
  19. Yazaki, K.; Takahashi, M.; Miyajima, N.; Obata, M. Construction of a PolyMOF Using a Polymer Ligand Bearing the Benzenedicarboxylic Acid Moiety in the Side Chain. New J. Chem. 2020, 44, 5182–5185. [Google Scholar] [CrossRef]
  20. Yu, S.B.; Lin, F.; Tian, J.; Yu, J.; Zhang, D.W.; Li, Z.T. Water-Soluble and Dispersible Porous Organic Polymers: Preparation, Functions and Applications. Chem. Soc. Rev. 2022, 51, 434–449. [Google Scholar] [CrossRef]
  21. Song, W.; Zhang, Y.; Tran, C.H.; Choi, H.K.; Yu, D.G.; Kim, I. Porous Organic Polymers with Defined Morphologies: Synthesis, Assembly, and Emerging Applications. Prog. Polym. Sci. 2023, 142, 101691. [Google Scholar] [CrossRef]
  22. Wu, M.; Zhou, Z. Covalent Organic Frameworks as Electrode Materials for Rechargeable Metal-ion Batteries. Interdiscip. Mater. 2023, 2, 231–259. [Google Scholar] [CrossRef]
  23. Li, Y.; Karimi, M.; Gong, Y.N.; Dai, N.; Safarifard, V.; Jiang, H.L. Integration of Metal-Organic Frameworks and Covalent Organic Frameworks: Design, Synthesis, and Applications. Matter 2021, 4, 2230–2265. [Google Scholar] [CrossRef]
  24. Altintas, C.; Erucar, I.; Keskin, S. MOF/COF Hybrids as next Generation Materials for Energy and Biomedical Applications. CrystEngComm 2022, 24, 7360–7371. [Google Scholar] [CrossRef]
  25. Xu, Y.; Jin, S.; Xu, H.; Nagai, A.; Jiang, D. Conjugated Microporous Polymers: Design, Synthesis and Application. Chem. Soc. Rev. 2013, 42, 8012–8031. [Google Scholar] [CrossRef] [PubMed]
  26. Liao, Y.; Cheng, Z.; Trunk, M.; Thomas, A. Targeted Control over the Porosities and Functionalities of Conjugated Microporous Polycarbazole Networks for CO2-Selective Capture and H2 Storage. Polym. Chem. 2017, 8, 7240–7247. [Google Scholar] [CrossRef]
  27. Xie, Y.; Li, Z.; Liu, C.; Song, Q.; Liu, Y.; Zhang, Z.; Han, B. Bipyridine-Based Conjugated Microporous Polymers for Boosted Photocatalytic U(VI) Separation. Chem. Commun. 2024, 60, 13348–13351. [Google Scholar] [CrossRef] [PubMed]
  28. Luo, S.; Zeng, Z.; Wang, H.; Xiong, W.; Song, B.; Zhou, C.; Duan, A.; Tan, X.; He, Q.; Zeng, G.; et al. Recent Progress in Conjugated Microporous Polymers for Clean Energy: Synthesis, Modification, Computer Simulations, and Applications. Prog. Polym. Sci. 2021, 115, 101374. [Google Scholar] [CrossRef]
  29. Guo, W.; Zhang, Z.; Chen, P.; Luan, X.; Dou, J.; Bai, J.; Fang, H.; Shi, W.; Zhou, B. CMP-on-MOF Bimetallic Hybrids Derived Sheet-on-Rod Heterostructure as Bifunctional Oxygen Electrocatalyst for Rechargeable Zn-Air Batteries. Microporous Mesoporous Mater. 2022, 331, 111639. [Google Scholar] [CrossRef]
  30. Moradi, M.R.; Torkashvand, A.; Ramezanipour Penchah, H.; Ghaemi, A. Amine Functionalized Benzene Based Hypercrosslinked Polymer as an Adsorbent for CO2/N2 Adsorption. Sci. Rep. 2023, 13, 9214. [Google Scholar] [CrossRef]
  31. Song, W.; Tang, Y.; Moon, B.Y.; Liao, Q.; Xu, H.; Hou, Q.; Zhang, H.; Yu, D.G.; Liao, Y.; Kim, I. Green Synthesis of Hypercrosslinked Polymers for CO2 Capture and Conversion: Recent Advances, Opportunities, and Challenges. Green. Chem. 2024, 26, 2476–2504. [Google Scholar] [CrossRef]
  32. Bauzá, M.; Palomino, G.T.; Cabello, C.P. Hypercrosslinked Polymer Derived Carbon@MIL-100 Magnetic Material for the Enhanced Extraction of Diclofenac. Separ. Purif. Technol. 2022, 303, 122211. [Google Scholar] [CrossRef]
  33. Qian, Z.; Wang, Z.J.; Zhang, K.A.I. Covalent Triazine Frameworks as Emerging Heterogeneous Photocatalysts. Chem. Mater. 2021, 33, 1909–1926. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Jin, S. Recent Advancements in the Synthesis of Covalent Triazine Frameworks for Energy and Environmental Applications. Polymers 2018, 11, 31. [Google Scholar] [CrossRef] [PubMed]
  35. Dong, S.; Liu, X.; Kong, X.; Dong, F.; Yu, Y.; Wang, L.; Wang, D.; He, Z.; Song, S. Boosting Photocatalytic H2 Evolution UIO-66-NH2/Covalent Triazine-Based Frameworks Composites by Constructing a Covalent Heterojunction. Environ. Sci. Pollut. Res. 2023, 30, 111039–111050. [Google Scholar] [CrossRef] [PubMed]
  36. Shao, L.H.; Huang, A.X.; Yan, X.C.; Liu, Y.H.; Wang, Y.; Jin, X.; Zhang, F.M. Constructing Tightly Integrated Conductive Metal-Organic Framework/Covalent Triazine Framework Heterostructure by Coordination Bonds for Photocatalytic Hydrogen Evolution. J. Colloid. Interface Sci. 2023, 633, 233–242. [Google Scholar] [CrossRef] [PubMed]
  37. Yuan, N.; Zhang, C.; Zhang, X.; Zhang, R. Covalent Integration of Fe-Based Metal–Organic Framework and Triazine-Containing Covalent Organic Framework for Enhanced Adsorptive Removal of Antibiotics. J. Clean. Prod. 2024, 434, 140259. [Google Scholar] [CrossRef]
  38. Li, S.; Dai, Y.; Ye, P.; Liu, F.; Li, L.; Zhang, D.; Zhu, M. Hierarchical Porous MOF/CTF Hybrid Frameworks Used as Protection against Acidic Harmful Gases. J. Chem. Eng. 2024, 491, 152035. [Google Scholar] [CrossRef]
  39. Vadiyar, M.M.; Kim, J.Y.; Bae, J.H.; Nam, K.W. Imidazole Linker-Induced Covalent Triazine Framework–ZIF Hybrids for Confined Hollow Carbon Super-Heterostructures toward a Long-Life Supercapacitor. Carbon. Energy 2023, 5, e344. [Google Scholar] [CrossRef]
  40. Tahir, Z.; Alomar, M.; Sarfraz, M.; Waheed, A.; Ayub, H.M.U. Carbon Capturing Composite Membranes Comprising Cu-MOF and PIM-1. J. Appl. Polym. Sci. 2024, 141, e55709. [Google Scholar] [CrossRef]
  41. Mondal, S.K.; Aina, P.O.; Rownaghi, A.A.; Rezaei, F. Design and Development of UiO-67-Coated PIM-1-Based Composites and Demonstration of Their Detoxification Performance. J. Chem. Eng. 2024, 493, 152269. [Google Scholar] [CrossRef]
  42. Du, X.; Shao, D.; Wan, Y. PIM-1 Based Mixed Matrix Membranes with MOF-808(Ce) for CO2 Separations. Gas. Sci. Eng. 2024, 131, 205476. [Google Scholar] [CrossRef]
  43. Rong, S.; Su, P.; Chen, S.; Jia, M.; Li, W. Sub-5 Nm Porous Polymer Decoration toward Superhydrophobic MOFs with Enhanced Stability and Processability. Chin. Chem. Lett. 2022, 33, 2134–2138. [Google Scholar] [CrossRef]
  44. Yu, S.; Li, C.; Zhao, S.; Chai, M.; Hou, J.; Lin, R. Recent Advances in the Interfacial Engineering of MOF-Based Mixed Matrix Membranes for Gas Separation. Nanoscale 2024, 16, 7716–7733. [Google Scholar] [CrossRef] [PubMed]
  45. Doǧan, E.B.; Maurin, G.; Ahunbay, M.G. Atomistic Insight into the Interfacial Structuring of ZIF-67 MOF/Polymer Composites and Their Propylene-Propane Adsorption Properties. J. Phys. Chem. C 2023, 127, 20491–20502. [Google Scholar] [CrossRef]
  46. Fan, D.; Ozcan, A.; Shekhah, O.; Semino, R.; Eddaoudi, M.; Maurin, G. Engineering MOF Surface Defects in Mixed Matrix Membranes: An Effective Strategy to Enhance MOF/Polymer Adhesion and Control Interfacial Gas Transport. J. Membr. Sci. Lett. 2022, 2, 100029. [Google Scholar] [CrossRef]
  47. Jaid, G.M.; AbdulRazak, A.A.; Meskher, H.; Al-Saadi, S.; Alsalhy, Q.F. Metal-Organic Frameworks (MOFs), Covalent Organic Frameworks (COFs), and Hydrogen-Bonded Organic Frameworks (HOFs) in Mixed Matrix Membranes. Mater. Today Sustain. 2024, 25, 100672. [Google Scholar] [CrossRef]
  48. 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]
  49. Pu, Y.; Yang, Z.; Wee, V.; Wu, Z.; Jiang, Z.; Zhao, D. Amino-Functionalized NUS-8 Nanosheets as Fillers in PIM-1 Mixed Matrix Membranes for CO2 Separations. J. Membr. Sci. 2022, 641, 119912. [Google Scholar] [CrossRef]
  50. Carja, I.D.; Tavares, S.R.; Shekhah, O.; Ozcan, A.; Semino, R.; Kale, V.S.; Eddaoudi, M.; Maurin, G. Insights into the Enhancement of MOF/Polymer Adhesion in Mixed-Matrix Membranes via Polymer Functionalization. ACS Appl. Mater. Interfaces 2021, 13, 29041–29047. [Google Scholar] [CrossRef]
  51. Sadakiyo, M.; Kitagawa, H. Ion-Conductive Metal-Organic Frameworks. Dalton Trans. 2021, 50, 5385–5397. [Google Scholar] [CrossRef]
  52. Mínguez Espallargas, G.; Coronado, E. Magnetic Functionalities in MOFs: From the Framework to the Pore. Chem. Soc. Rev. 2018, 47, 533–557. [Google Scholar] [CrossRef]
  53. Morris, W.; Stevens, C.J.; Taylor, R.E.; Dybowski, C.; Yaghi, O.M.; Garcia-Garibay, M.A. NMR and X-Ray Study Revealing the Rigidity of Zeolitic Imidazolate Frameworks. J. Phys. Chem. C 2012, 116, 13307–13312. [Google Scholar] [CrossRef]
  54. Bux, H.; Liang, F.; Li, Y.; Cravillon, J.; Wiebcke, M.; Caro, J. Zeolitic Imidazolate Framework Membrane with Molecular Sieving Properties by Microwave-Assisted Solvothermal Synthesis. J. Am. Chem. Soc. 2009, 131, 16000–16001. [Google Scholar] [CrossRef]
  55. Hu, L.; Yan, Z.; Mo, X.; Peng, X.; Chen, L. Morphology Control Synthesis of ZIF-8 as Highly Efficient Catalyst for the Cycloaddition of CO2 to Cyclic Carbonate. ChemCatChem 2019, 11, 3212–3219. [Google Scholar] [CrossRef]
  56. Park, K.S.; Ni, Z.; Cô, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’keeffe, M.; Yaghi, O.M. Exceptional Chemical and Thermal Stability of Zeolitic Imidazolate Frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef]
  57. Qian, J.; Sun, F.; Qin, L. Hydrothermal Synthesis of Zeolitic Imidazolate Framework-67 (ZIF-67) Nanocrystals. Mater. Lett. 2012, 82, 220–223. [Google Scholar] [CrossRef]
  58. Zhao, P.; Fu, S.; Cheng, L.; Jiao, Z.; Wu, M. Modulating Zeolitic Imidazolate Framework-67 and Its Derivatives as Advanced Oxygen Evolution Reaction Electrocatalysts. Coord. Chem. Rev. 2024, 498, 215452. [Google Scholar] [CrossRef]
  59. Winarta, J.; Shan, B.; McIntyre, S.M.; Ye, L.; Wang, C.; Liu, J.; Mu, B. A Decade of UiO-66 Research: A Historic Review of Dynamic Structure, Synthesis Mechanisms, and Characterization Techniques of an Archetypal Metal-Organic Framework. Cryst. Growth Des. 2020, 20, 1347–1362. [Google Scholar] [CrossRef]
  60. Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. [Google Scholar] [CrossRef]
  61. Øien, S.; Wragg, D.; Reinsch, H.; Svelle, S.; Bordiga, S.; Lamberti, C.; Lillerud, K.P. Detailed Structure Analysis of Atomic Positions and Defects in Zirconium Metal-Organic Frameworks. Cryst. Growth Des. 2014, 14, 5370–5372. [Google Scholar] [CrossRef]
  62. Liu, Y.; Zhao, Y.; Zhou, Z.; Shen, Y.; Jiang, L. Pore Size Effect of Zirconium-Based Metal-Organic Frameworks for Encapsulation and Release of Volatile Monoterpenes. Colloids Surf. A Physicochem. Eng. Asp. 2023, 674, 131907. [Google Scholar] [CrossRef]
  63. Ye, X.; Liu, D. Metal-Organic Framework UiO-68 and Its Derivatives with Sufficiently Good Properties and Performance Show Promising Prospects in Potential Industrial Applications. Cryst. Growth Des. 2021, 21, 4780–4804. [Google Scholar] [CrossRef]
  64. Dong, J.; Wee, V.; Zhao, D. Stimuli-Responsive Metal–Organic Frameworks Enabled by Intrinsic Molecular Motion. Nat. Mater. 2022, 21, 1334–1340. [Google Scholar] [CrossRef]
  65. Férey, C.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. Chemistry: A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef]
  66. Llewellyn, P.L.; Bourrelly, S.; Serre, C.; Vimont, A.; Daturi, M.; Hamon, L.; De Weireld, G.; Chang, J.S.; Hong, D.Y.; Hwang, Y.K.; et al. High Uptakes of CO2 and CH4 in Mesoporous Metal-Organic Frameworks MIL-100 and MIL-101. Langmuir 2008, 24, 7245–7250. [Google Scholar] [CrossRef] [PubMed]
  67. Boutin, A.; Coudert, F.X.; Springuel-Huet, M.A.; Neimark, A.V.; Férey, G.; Fuchs, A.H. The Behavior of Flexible MIL-53(Al) upon CH4 and CO2 Adsorption. J. Phys. Chem. C 2010, 114, 22237–22244. [Google Scholar] [CrossRef]
  68. Horcajada, P.; Surblé, S.; Serre, C.; Hong, D.Y.; Seo, Y.K.; Chang, J.S.; Grenèche, J.M.; Margiolaki, I.; Férey, G. Synthesis and Catalytic Properties of MIL-100(Fe), an Iron(III) Carboxylate with Large Pores. ChemComm. 2007, 27, 2820–2822. [Google Scholar] [CrossRef] [PubMed]
  69. Tran, D.T.; To, T.D.; Dao, Q.T.; Nghiem, L.D. Facile Synthesis of MIL-100(Fe) Adsorbents for Boosting Organic Dye Removal. Environ. Sci. Pollut. Res. 2025, 32, 4975–4988. [Google Scholar] [CrossRef]
  70. 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]
  71. Rowsell, J.L.C.; Yaghi, O.M. Metal-Organic Frameworks: A New Class of Porous Materials. Microporous Mesoporous Mater. 2004, 73, 3–14. [Google Scholar] [CrossRef]
  72. Millward, A.R.; Yaghi, O.M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef]
  73. S-Y Chui, S.; M-F Lo, S.; H Charmant, J.P.; Guy Orpen, A.; Williams, I.D. A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n. Science 1999, 283, 1148–1150. [Google Scholar] [CrossRef]
  74. Krause, S.; Bon, V.; Senkovska, I.; Stoeck, U.; Wallacher, D.; Többens, D.M.; Zander, S.; Pillai, R.S.; Maurin, G.; Coudert, F.X.; et al. A Pressure-Amplifying Framework Material with Negative Gas Adsorption Transitions. Nature 2016, 532, 348–352. [Google Scholar] [CrossRef]
  75. Kouser, S.; Hezam, A.; Khadri, M.J.N.; Khanum, S.A. A Review on Zeolite Imidazole Frameworks: Synthesis, Properties, and Applications. J. Porous Mater. 2022, 29, 663–681. [Google Scholar] [CrossRef]
  76. Troyano, J.; Carné-Sánchez, A.; Avci, C.; Imaz, I.; Maspoch, D. Colloidal Metal-Organic Framework Particles: The Pioneering Case of ZIF-8. Chem. Soc. Rev. 2019, 48, 5534–5546. [Google Scholar] [CrossRef] [PubMed]
  77. Nazir, M.A.; Ullah, S.; Shahid, M.U.; Hossain, I.; Najam, T.; Ismail, M.A.; ur Rehman, A.; Karim, M.R.; Shah, S.S.A. Zeolitic Imidazolate Frameworks (ZIF-8 & ZIF-67): Synthesis and Application for Wastewater Treatment. Sep. Purif. Technol. 2025, 356, 129828. [Google Scholar]
  78. Duan, C.; Yu, Y.; Hu, H. Recent Progress on Synthesis of ZIF-67-Based Materials and Their Application to Heterogeneous Catalysis. Green Energy Environ. 2022, 7, 3–15. [Google Scholar] [CrossRef]
  79. Xiong, Y.; Deng, N.; Wu, X.; Zhang, Q.; Liu, S.; Sun, G. De Novo Synthesis of Amino-Functionalized ZIF-8 Nanoparticles: Enhanced Interfacial Compatibility and Pervaporation Performance in Mixed Matrix Membranes Applying for Ethanol Dehydration. Sep. Purif. Technol. 2022, 285, 120321. [Google Scholar] [CrossRef]
  80. Neubertová, V.; Švorčík, V.; Kolská, Z. Amino-Modified ZIF-8 for Enhanced CO2 Capture: Synthesis, Characterization and Performance Evaluation. Microporous Mesoporous Mater. 2024, 366, 112956. [Google Scholar] [CrossRef]
  81. Behera, D.; Priyadarshini, P.; Parida, K. ZIF-8 Metal-Organic Frameworks and Their Hybrid Materials: Emerging Photocatalysts for Energy and Environmental Applications. Dalton Trans. 2025, 54, 2681–2708. [Google Scholar] [CrossRef]
  82. Ye, C.; Wu, X.; Wu, H.; Yang, L.; Ren, Y.; Wu, Y.; Liu, Y.; Guo, Z.; Zhao, R.; Jiang, Z. Incorporating Nano-Sized ZIF-67 to Enhance Selectivity of Polymers of Intrinsic Microporosity Membranes for Biogas Upgrading. Chem. Eng. Sci. 2020, 216, 115497. [Google Scholar] [CrossRef]
  83. Wu, R.; Li, Y.; Huang, A. Synthesis of High-Performance Co-Based ZIF-67 Membrane for H2 Separation by Using Cobalt Ions Chelated PIM-1 as Interface Layer. J. Membr. Sci. 2021, 620, 118841. [Google Scholar] [CrossRef]
  84. Li, C.; Chandresh, A.; Zhang, Z.; Moulai, S.; Heinke, L. Stability and Degradation of Metal–Organic-Framework Films under Ambient Air Explored by Uptake and Diffusion Experiments. Adv. Mater. Interfaces 2022, 9, 2101947. [Google Scholar] [CrossRef]
  85. Kadhom, M.; Al-Furaiji, M.; Salih, S.; Al-Obaidi, M.A.; Abdullah, G.H.; Albayati, N. A Review on UiO-66 Applications in Membrane-Based Water Treatment Processes. J. Water Process. Eng. 2023, 51, 103402. [Google Scholar] [CrossRef]
  86. Sun, Y. UiO-66 Metal-Organic Framework Membranes: Structural Engineering for Separation Applications. Membranes 2025, 15, 8. [Google Scholar] [CrossRef]
  87. Okba, D.; Hassan, S.; Abdel Aleem, A.A.H.; Shehab El-Din, M.T.; El Tantawy El Sayed, I.; Abou-Elyazed, A.S. Enhanced Performance of UiO-66 for Supercapacitor Applications through Oxidation via the Hummers’ Method. RSC Adv. 2025, 15, 795–805. [Google Scholar] [CrossRef]
  88. Tatay, S.; Martínez-Giménez, S.; Rubio-Gaspar, A.; Gómez-Oliveira, E.; Castells-Gil, J.; Dong, Z.; Mayoral, Á.; Almora-Barrios, N.; Padial, M.N.; Martí-Gastaldo, C. Synthetic Control of Correlated Disorder in UiO-66 Frameworks. Nat. Commun. 2023, 14, 6962. [Google Scholar] [CrossRef]
  89. Miao, C.L.; Wang, X.X.; Guan, D.H.; Li, J.X.; Wang, H.F.; Xu, J.J. Directional Modification-Functionalized Metal–Organic Framework Solid-State Electrolytes for Highly Stable Li–O2 Batteries. Adv. Funct. Mater. 2024, 34, 2307150. [Google Scholar] [CrossRef]
  90. Timofeev, K.L.; Kulinich, S.A.; Kharlamova, T.S. NH2-Modified UiO-66: Structural Characteristics and Functional Properties. Molecules 2023, 28, 3916. [Google Scholar] [CrossRef]
  91. Surblé, S.; Serre, C.; Mellot-Draznieks, C.; Millange, F.; Férey, G. A New Isoreticular Class of Metal-Organic-Frameworks with the MIL-88 Topology. ChemComm. 2006, 3, 284–286. [Google Scholar] [CrossRef]
  92. Hendon, C.H.; Tiana, D.; Fontecave, M.; Sanchez, C.; D’Arras, L.; Sassoye, C.; Rozes, L.; Mellot-Draznieks, C.; Walsh, A. Engineering the Optical Response of the Titanium-MIL-125 Metal-Organic Framework through Ligand Functionalization. J. Am. Chem. Soc. 2013, 135, 10942–10945. [Google Scholar] [CrossRef]
  93. Palomino Cabello, C.; Gómez-Pozuelo, G.; Opanasenko, M.; Nachtigall, P.; Čejka, J. Metal–Organic Frameworks M-MOF-74 and M-MIL-100: Comparison of Textural, Acidic, and Catalytic Properties. Chempluschem 2016, 81, 828–835. [Google Scholar] [CrossRef]
  94. Le Vuong, M.D.; Christodoulou, I.; Porcino, M.; Dong, S.T.; Lassalle-Kaiser, B.; Haouas, M.; Gref, R.; Martineau-Corcos, C. Degradation Mechanism of Metal-Organic Framework Drug Nanocarriers Studied by Solid-State Nuclear Magnetic Resonance and X-Ray Absorption Near-Edge Structure Spectroscopy. Chem. Mater. 2022, 34, 8178–8189. [Google Scholar] [CrossRef]
  95. Férey, G.; Millange, F.; Morcrette, M.; Serre, C.; Doublet, M.L.; Grenèche, J.M.; Tarascon, J.M. Mixed-Valence Li/Fe-Based Metal-Organic Frameworks with Both Reversible Redox and Sorption Properties. Angew. Chem. Int. Ed. 2007, 46, 3259–3263. [Google Scholar] [CrossRef] [PubMed]
  96. Celeste, A.; Paolone, A.; Itié, J.P.; Borondics, F.; Joseph, B.; Grad, O.; Blanita, G.; Zlotea, C.; Capitani, F. Mesoporous Metal-Organic Framework MIL-101 at High Pressure. J. Am. Chem. Soc. 2020, 142, 15012–15019. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, Y.Y.; Zhou, M.L.; Cui, Y.H.; Yang, M.; Bao, Y.S.; Ye, Y.; Tian, D.M.; Liu, L.Y.; Han, Z.B. Polymelamine Formaldehyde-Coated MIL-101 as an Efficient Dual-Functional Core-Shell Composite to Catalyze the Deacetalization-Knoevenagel Tandem Reaction. Inorg. Chem. 2022, 61, 13678–13684. [Google Scholar] [CrossRef] [PubMed]
  98. Sharma, D.; Rasaily, S.; Pradhan, S.; Baruah, K.; Tamang, S.; Pariyar, A. HKUST-1 Metal Organic Framework as an Efficient Dual-Function Catalyst: Aziridination and One-Pot Ring-Opening Transformation for Formation of β-Aryl Sulfonamides with C-C, C-N, C-S, and C-O Bonds. Inorg. Chem. 2021, 60, 7794–7802. [Google Scholar] [CrossRef]
  99. Grande, C.A.; Kaiser, A.; Andreassen, K.A. Methane Storage in Metal-Organic Framework HKUST-1 with Enhanced Heat Management Using 3D Printed Metal Lattices. ChERD 2023, 192, 362–370. [Google Scholar] [CrossRef]
  100. Lin, K.S.; Adhikari, A.K.; Ku, C.N.; Chiang, C.L.; Kuo, H. Synthesis and Characterization of Porous HKUST-1 Metal Organic Frameworks for Hydrogen Storage. Int. J. Hydrogen Energy 2012, 37, 13865–13871. [Google Scholar] [CrossRef]
  101. Shen, T.; Liu, T.; Mo, H.; Yuan, Z.; Cui, F.; Jin, Y.; Chen, X. Cu-Based Metal-Organic Framework HKUST-1 as Effective Catalyst for Highly Sensitive Determination of Ascorbic Acid. RSC Adv. 2020, 10, 22881–22890. [Google Scholar] [CrossRef]
  102. Fuoco, A.; Khdhayyer, M.R.; Attfield, M.P.; Esposito, E.; Jansen, J.C.; Budd, P.M. Synthesis and Transport Properties of Novel MOF/PIM-1/MOF Sandwich Membranes for Gas Separation. Membranes 2017, 7, 7. [Google Scholar] [CrossRef]
  103. Bönisch, N.; Maliuta, M.; Senkovska, I.; Bon, V.; Petkov, P.; Plätzer, C.; Müller, P.; Kaskel, S. Linker Expansion and Its Impact on Switchability in Pillared-Layer MOFs. Inorg. Chem. 2021, 60, 1726–1737. [Google Scholar] [CrossRef] [PubMed]
  104. Yang, G.; Wang, Y.; Sun, M.; Xu, P.; Wang, C.; Huang, K.; Jiang, H.; Mintova, S.; Guo, H. Is Amino-Modification of HKUST-1 in PEI Mixed-Matrix Membranes Always Favorable to CO2 Separation? Microporous Mesoporous Mater. 2023, 359, 112649. [Google Scholar] [CrossRef]
  105. Drwęska, J.; Roztocki, K.; Janiak, A.M. Advances in Chemistry of CALF-20, a Metal-Organic Framework for Industrial Gas Applications. Chem. Commun. 2024, 61, 1032–1047. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, H.Y.; Yang, Y.; Li, C.C.; Tang, H.L.; Zhang, F.M.; Zhang, G.L.; Yan, H. A New Strategy for Constructing Covalently Connected MOF@COF Core-Shell Heterostructures for Enhanced Photocatalytic Hydrogen Evolution. J. Mater. Chem. A Mater. 2021, 9, 16743–16750. [Google Scholar] [CrossRef]
  107. Chen, T.; Zhao, D. Post-Synthetic Modification of Metal-Organic Framework-Based Membranes for Enhanced Molecular Separations. Coord. Chem. Rev. 2023, 491, 215259. [Google Scholar] [CrossRef]
  108. Jue, M.L.; Lively, R.P. PIM Hybrids and Derivatives: How to Make a Good Thing Better. Curr. Opin. Chem. Eng. 2022, 35, 100750. [Google Scholar] [CrossRef]
  109. Ma, C.; Urban, J.J. Polymers of Intrinsic Microporosity (PIMs) Gas Separation Membranes: A Mini Review. Proc. Nat. Res. Soc. 2018, 2, 02002. [Google Scholar] [CrossRef]
  110. McKeown, N.B. Polymers of Intrinsic Microporosity. ISRN Mater. Sci. 2012, 2012, 1–16. [Google Scholar] [CrossRef]
  111. Mc Keown, N.B.; Budd, P.M. Polymers of Intrinsic Microporosity (PIMs): Organic Materials for Membrane Separations, Heterogeneous Catalysis and Hydrogen Storage. Chem. Soc. Rev. 2006, 35, 675–683. [Google Scholar] [CrossRef]
  112. Wang, Y.; Ma, X.; Ghanem, B.S.; Alghunaimi, F.; Pinnau, I.; Han, Y. Polymers of Intrinsic Microporosity for Energy-Intensive Membrane-Based Gas Separations. Mater. Today Nano 2018, 3, 69–95. [Google Scholar] [CrossRef]
  113. Low, Z.X.; Budd, P.M.; McKeown, N.B.; Patterson, D.A. Gas Permeation Properties, Physical Aging, and Its Mitigation in High Free Volume Glassy Polymers. Chem. Rev. 2018, 118, 5871–5911. [Google Scholar] [CrossRef]
  114. Gao, G.K.; Wang, Y.R.; Zhu, H.J.; Chen, Y.; Yang, R.X.; Jiang, C.; Ma, H.; Lan, Y.Q. Rapid Production of Metal–Organic Frameworks Based Separators in Industrial-Level Efficiency. Adv. Sci. 2020, 7, 2002190. [Google Scholar] [CrossRef]
  115. Paul, D.R.; Kemp, D.R. Diffusion Time Lag in Polymer Membranes Containing Adsorptive Fillers. J. Polym. Sci. Part. C Polym. Symp. 1973, 41, 79–93. [Google Scholar] [CrossRef]
  116. Yong, W.F.; Kwek, K.H.A.; Liao, K.S.; Chung, T.S. Suppression of Aging and Plasticization in Highly Permeable Polymers. Polymer 2015, 77, 377–386. [Google Scholar] [CrossRef]
  117. Salahshoori, I.; Seyfaee, A.; Babapoor, A. Recent Advances in Synthesis and Applications of Mixed Matrix Membranes. Synth. Sinter. 2021, 1, 1–27. [Google Scholar] [CrossRef]
  118. Castro-Muñoz, R.; Fíla, V.; Dung, C.T. Mixed Matrix Membranes Based on PIMs for Gas Permeation: Principles, Synthesis, and Current Status. Chem. Eng. Commun. 2017, 204, 295–309. [Google Scholar] [CrossRef]
  119. Ahmad, M.Z.; Castro-Munõz, R.; Budd, P.M. Boosting Gas Separation Performance and Suppressing the Physical Aging of Polymers of Intrinsic Microporosity (PIM-1) by Nanomaterial Blending. Nanoscale 2020, 12, 23333–23370. [Google Scholar] [CrossRef] [PubMed]
  120. Hossain, I.; Kim, K.I.; Husna, A.; Kang, J.H.; Kim, T.H.; Park, H.B. Metal-Coordinated, Dual-Crosslinked PIM Polymer Membranes for Upgraded CO2 Separation: Aging and Plasticization Resistance. Small 2024, 21, 2407973. [Google Scholar] [CrossRef]
  121. Robeson, L.M. The Upper Bound Revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  122. Budd, P.M.; Msayib, K.J.; Tattershall, C.E.; Ghanem, B.S.; Reynolds, K.J.; Mckeown, N.B.; Fritsch, D. Gas Separation Membranes from Polymers of Intrinsic Microporosity. J. Membr. Sci. 2005, 251, 263–269. [Google Scholar] [CrossRef]
  123. Monsalve-Bravo, G.M.; Bhatia, S.K. Modeling Permeation through Mixed-Matrix Membranes: A Review. Processes 2018, 6, 172. [Google Scholar] [CrossRef]
  124. Goh, S.H.; Lau, H.S.; Yong, W.F. Metal–Organic Frameworks (MOFs)-Based Mixed Matrix Membranes (MMMs) for Gas Separation: A Review on Advanced Materials in Harsh Environmental Applications. Small 2022, 18, 2107536. [Google Scholar] [CrossRef] [PubMed]
  125. Knebel, A.; Bavykina, A.; Datta, S.J.; Sundermann, L.; Garzon-Tovar, L.; Lebedev, Y.; Durini, S.; Ahmad, R.; Kozlov, S.M.; Shterk, G.; et al. Solution Processable Metal–Organic Frameworks for Mixed Matrix Membranes Using Porous Liquids. Nat. Mater. 2020, 19, 1346–1353. [Google Scholar] [CrossRef] [PubMed]
  126. Ma, X.; Swaidan, R.J.; Wang, Y.; Hsiung, C.E.; Han, Y.; Pinnau, I. Highly Compatible Hydroxyl-Functionalized Microporous Polyimide-Zif-8 Mixed Matrix Membranes for Energy Efficient Propylene/Propane Separation. ACS Appl. Nano Mater. 2018, 1, 3541–3547. [Google Scholar] [CrossRef]
  127. Zhuang, G.L.; Tseng, H.H.; Wey, M.Y. Preparation of PPO-Silica Mixed Matrix Membranes by in-Situ Sol-Gel Method for H2/CO2 Separation. Int. J. Hydrogen Energy 2014, 39, 17178–17190. [Google Scholar] [CrossRef]
  128. Liu, Y.; Zhang, J.; Tan, X. High Performance of PIM-1/ZIF-8 Composite Membranes for O2/N2 Separation. ACS Omega 2019, 4, 16572–16577. [Google Scholar] [CrossRef]
  129. Zhang, C.; Liu, B.; Wang, G.; Yu, G.; Zou, X.; Zhu, G. Small-Pore CAU-21 and Porous PIM-1 in Mixed-Matrix Membranes for Improving Selectivity and Permeability in Hydrogen Separation. Chem. Commun. 2019, 55, 7101–7104. [Google Scholar] [CrossRef]
  130. Ghalei, B.; Sakurai, K.; Kinoshita, Y.; Wakimoto, K.; Isfahani, A.P.; Song, Q.; Doitomi, K.; Furukawa, S.; Hirao, H.; Kusuda, H.; et al. Enhanced Selectivity in Mixed Matrix Membranes for CO2 Capture through Efficient Dispersion of Amine-Functionalized MOF Nanoparticles. Nat. Energy 2017, 2, 17086. [Google Scholar] [CrossRef]
  131. Zhang, C.; Lively, R.P.; Zhang, K.; Johnson, J.R.; Karvan, O.; Koros, W.J. Unexpected Molecular Sieving Properties of Zeolitic Imidazolate Framework-8. J. Phys. Chem. Lett. 2012, 3, 2130–2134. [Google Scholar] [CrossRef]
  132. Fairen-Jimenez, D.; Moggach, S.A.; Wharmby, M.T.; Wright, P.A.; Parsons, S.; Düren, T. Opening the Gate: Framework Flexibility in ZIF-8 Explored by Experiments and Simulations. J. Am. Chem. Soc. 2011, 133, 8900–8902. [Google Scholar] [CrossRef]
  133. Liu, G.; Chernikova, V.; Liu, Y.; Zhang, K.; Belmabkhout, Y.; Shekhah, O.; Zhang, C.; Yi, S.; Eddaoudi, M.; Koros, W.J. Mixed Matrix Formulations with MOF Molecular Sieving for Key Energy-Intensive Separations. Nat. Mater. 2018, 17, 283–289. [Google Scholar] [CrossRef] [PubMed]
  134. Aliyev, E.; Warfsmann, J.; Tokay, B.; Shishatskiy, S.; Lee, Y.J.; Lillepaerg, J.; Champness, N.R.; Filiz, V. Gas Transport Properties of the Metal-Organic Framework (MOF)-Assisted Polymer of Intrinsic Microporosity (PIM-1) Thin-Film Composite Membranes. ACS Sustain. Chem. Eng. 2021, 9, 684–694. [Google Scholar] [CrossRef]
  135. Qian, Q.; Asinger, P.A.; Lee, M.J.; Han, G.; Mizrahi Rodriguez, K.; Lin, S.; Benedetti, F.M.; Wu, A.X.; Chi, W.S.; Smith, Z.P. MOF-Based Membranes for Gas Separations. Chem. Rev. 2020, 120, 8161–8266. [Google Scholar] [CrossRef] [PubMed]
  136. Cook, M.; Gaffney, P.R.J.; Peeva, L.G.; Livingston, A.G. Roll-to-Roll Dip Coating of Three Different PIMs for Organic Solvent Nanofiltration. J. Membr. Sci. 2018, 558, 52–63. [Google Scholar] [CrossRef]
  137. Li, N.; Ma, C.; Li, D.; Li, P.; Ye, M.; Wang, Z.; Qiao, Z.; Zhong, C. In-Situ Glass Transition of ZIF-62 Based Mixed Matrix Membranes for Enhancing H2 Fast Separation. Sep. Purif. Technol. 2025, 353, 128500. [Google Scholar] [CrossRef]
  138. Baloochiyan, A.; Öztürk, H.; Erucar, I. Atomistic Investigation of Porous Amorphous Materials for CH4/H2 Separation. Chem. Eng. Sci. 2025, 301, 120741. [Google Scholar] [CrossRef]
  139. Xiong, S.; Pan, C.; Dai, G.; Liu, C.; Tan, Z.; Chen, C.; Yang, S.; Ruan, X.; Tang, J.; Yu, G. Interfacial Co-Weaving of AO-PIM-1 and ZIF-8 in Composite Membranes for Enhanced H2 Purification. J. Membr. Sci. 2022, 645, 120217. [Google Scholar] [CrossRef]
  140. Robeson, L.M. Correlation of Separation Factor versus Permeability for Polymeric Membranes. J. Membr. Sci. 1991, 62, 165–185. [Google Scholar] [CrossRef]
  141. Comesaña-Gándara, B.; Chen, J.; Bezzu, C.G.; Carta, M.; Rose, I.; Ferrari, M.C.; Esposito, E.; Fuoco, A.; Jansen, J.C.; McKeown, N.B. Redefining the Robeson Upper Bounds for CO2/CH4 and CO2/N2 Separations Using a Series of Ultrapermeable Benzotriptycene-Based Polymers of Intrinsic Microporosity. Energy Environ. Sci. 2019, 12, 2733–2740. [Google Scholar] [CrossRef]
  142. Tanh Jeazet, H.B.; Staudt, C.; Janiak, C. Metal-Organic Frameworks in Mixed-Matrix Membranes for Gas Separation. Dalton Trans. 2012, 41, 14003–14027. [Google Scholar] [CrossRef]
  143. Guo, X.; Qiao, Z.; Liu, D.; Zhong, C. Mixed-Matrix Membranes for CO2 Separation: Role of the Third Component. J. Mater. Chem. A Mater. 2019, 7, 24738–24759. [Google Scholar] [CrossRef]
  144. Wu, W.-N. Microporous Polymer-Metal Organic Framework (MOF) Hybrid Materials for Separations. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 2024. [Google Scholar]
  145. Zhang, H.; Sheng, L.; Chen, J.; Wang, X.; Tao, P.; Ren, D.; Cui, H.; Yang, K.; Tang, Z.; Zhang, Z.; et al. Scalable MOF-Based Mixed Matrix Membranes with Enhanced Permeation Processes Facilitate the Scale Application of Membrane-Based Carbon Capture Technologies. Carbon Capture Sci. Technol. 2024, 13, 100276. [Google Scholar] [CrossRef]
  146. Tien-Binh, N.; Vinh-Thang, H.; Chen, X.Y.; Rodrigue, D.; Kaliaguine, S. Crosslinked MOF-Polymer to Enhance Gas Separation of Mixed Matrix Membranes. J. Membr. Sci. 2016, 520, 941–950. [Google Scholar] [CrossRef]
  147. Huang, G.; Liu, T.; Peng, H.; Xie, Y.; Chen, Q.; Li, J.R. Enhancement of CO2/N2 Separation in MMMs by the Regulation of Nanofiller Microenvironment. Sep. Purif. Technol. 2025, 354, 128737. [Google Scholar] [CrossRef]
  148. Chen, W.; Zhang, Z.; Hou, L.; Yang, C.; Shen, H.; Yang, K.; Wang, Z. Metal-Organic Framework MOF-801/PIM-1 Mixed-Matrix Membranes for Enhanced CO2/N2 Separation Performance. Sep. Purif. Technol. 2020, 250, 117198. [Google Scholar] [CrossRef]
  149. Yin, H.; Alkaş, A.; Zhang, Y.; Zhang, Y.; Telfer, S.G. Mixed Matrix Membranes (MMMs) Using an Emerging Metal-Organic Framework (MUF-15) for CO2 Separation. J. Membr. Sci. 2020, 609, 118245. [Google Scholar] [CrossRef]
  150. Wang, Y.; Ren, Y.; Wu, H.; Wu, X.; Yang, H.; Yang, L.; Wang, X.; Wu, Y.; Liu, Y.; Jiang, Z. Amino-Functionalized ZIF-7 Embedded Polymers of Intrinsic Microporosity Membrane with Enhanced Selectivity for Biogas Upgrading. J. Membr. Sci. 2020, 602, 117970. [Google Scholar] [CrossRef]
  151. Feng, Y.; Yan, W.; Kang, Z.; Zou, X.; Fan, W.; Jiang, Y.; Fan, L.; Wang, R.; Sun, D. Thermal Treatment Optimization of Porous MOF Glass and Polymer for Improving Gas Permeability and Selectivity of Mixed Matrix Membranes. J. Chem. Eng. 2023, 465, 142873. [Google Scholar] [CrossRef]
  152. Budd, P.M.; Foster, A.B. Seeking Synergy in Membranes: Blends and Mixtures with Polymers of Intrinsic Microporosity. Curr Opin. Chem. Eng. 2022, 36, 100792. [Google Scholar] [CrossRef]
  153. Zhou, H.C.J.; Kitagawa, S. Metal–Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef] [PubMed]
  154. Li, B.; Wen, H.M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Emerging Multifunctional Metal-Organic Framework Materials. Adv. Mater. 2016, 28, 8819–8860. [Google Scholar] [CrossRef] [PubMed]
  155. Jiao, L.; Wang, Y.; Jiang, H.L.; Xu, Q. Metal-Organic Frameworks as Platforms for Catalytic Applications. Adv. Mater. 2018, 30, 1703663. [Google Scholar] [CrossRef] [PubMed]
  156. Shanmugam, M.; Agamendran, N.; Sekar, K.; Natarajan, T.S. Metal–Organic Frameworks (MOFs) for Energy Production and Gaseous Fuel and Electrochemical Energy Storage Applications. Phys. Chem. Chem. Phys. 2023, 25, 30116–30144. [Google Scholar] [CrossRef]
  157. Virender, V.; Pandey, V.; Singh, G.; Sharma, P.K.; Bhatia, P.; Solovev, A.A.; Mohan, B. Hybrid Metal-Organic Frameworks (MOFs) for Various Catalysis Applications. Top. Curr. Chem. 2024, 383, 1–36. [Google Scholar] [CrossRef]
  158. Wei, Y.S.; Zhang, M.; Zou, R.; Xu, Q. Metal-Organic Framework-Based Catalysts with Single Metal Sites. Chem. Rev. 2020, 120, 12089–12174. [Google Scholar] [CrossRef]
  159. Liu, Z.; Lv, H.; Xu, Y. Etching Strategies Induced Multihierarchical Structures of MOFs and Their Derivatives for Gas Sensing Applications: A Review. J. Mater. Chem. C Mater. 2025, 13, 3653–3668. [Google Scholar] [CrossRef]
  160. Healy, C.; Patil, K.M.; Wilson, B.H.; Hermanspahn, L.; Harvey-Reid, N.C.; Howard, B.I.; Kleinjan, C.; Kolien, J.; Payet, F.; Telfer, S.G.; et al. The Thermal Stability of Metal-Organic Frameworks. Coord. Chem. Rev. 2020, 419, 213388. [Google Scholar] [CrossRef]
  161. Ding, M.; Cai, X.; Jiang, H.L. Improving MOF Stability: Approaches and Applications. Chem. Sci. 2019, 10, 10209–10230. [Google Scholar] [CrossRef]
  162. Ding, S.Y.; Wang, W. Covalent Organic Frameworks (COFs): From Design to Applications. Chem. Soc. Rev. 2012, 42, 548–568. [Google Scholar] [CrossRef]
  163. Beagle, L.K.; Fang, Q.; Tran, L.D.; Baldwin, L.A.; Muratore, C.; Lou, J.; Glavin, N.R. Synthesis and Tailored Properties Towards Designer Covalent Organic Framework Thin Films and Heterostructures. arXiv 2021, arXiv:2105.12904. [Google Scholar]
  164. Geng, K.; He, T.; Liu, R.; Dalapati, S.; Tan, K.T.; Li, Z.; Tao, S.; Gong, Y.; Jiang, Q.; Jiang, D. Covalent Organic Frameworks: Design, Synthesis, and Functions. Chem. Rev. 2020, 120, 8814–8933. [Google Scholar] [CrossRef]
  165. Diercks, C.S.; Yaghi, O.M. The atom, the molecule, and the covalent organic framework. Science 2017, 355, eaal1585. [Google Scholar] [CrossRef] [PubMed]
  166. Banerjee, T.; Bennett, T.; Butler, K.; Easun, T.L.; Eddaoudi, M.; Forgan, R.; Gagliardi, L.; Hendon, C.; Jorge, M.; Lamberti, C.; et al. Electronic, Magnetic and Photophysical Properties of MOFs and COFs: General Discussion. Faraday Discuss. 2017, 201, 87–99. [Google Scholar] [CrossRef] [PubMed]
  167. Guo, C.; Duan, F.; Zhang, S.; He, L.; Wang, M.; Chen, J.; Zhang, J.; Jia, Q.; Zhang, Z.; Du, M. Heterostructured Hybrids of Metal–Organic Frameworks (MOFs) and Covalent–Organic Frameworks (COFs). J. Mater. Chem. A Mater. 2022, 10, 475–507. [Google Scholar] [CrossRef]
  168. Xue, R.; Guo, H.; Yang, W.; Huang, S.L.; Yang, G.Y. Cooperation between Covalent Organic Frameworks (COFs) and Metal Organic Frameworks (MOFs): Application of COFs-MOFs Hybrids. Adv. Compos. Hybrid Mater. 2022, 5, 1595–1611. [Google Scholar] [CrossRef]
  169. Ahmadijokani, F.; Ghaffarkhah, A.; Molavi, H.; Dutta, S.; Lu, Y.; Wuttke, S.; Kamkar, M.; Rojas, O.J.; Arjmand, M. COF and MOF Hybrids: Advanced Materials for Wastewater Treatment. Adv. Funct. Mater. 2024, 34, 2305527. [Google Scholar] [CrossRef]
  170. Ma, M.; Lu, X.; Guo, Y.; Wang, L.; Liang, X. Combination of Metal-Organic Frameworks (MOFs) and Covalent Organic Frameworks (COFs): Recent Advances in Synthesis and Analytical Applications of MOF/COF Composites. TrAC 2022, 157, 116741. [Google Scholar] [CrossRef]
  171. Chen, Z.; Li, X.; Yang, C.; Cheng, K.; Tan, T.; Lv, Y.; Liu, Y. Hybrid Porous Crystalline Materials from Metal Organic Frameworks and Covalent Organic Frameworks. Adv. Sci. 2021, 8, 2101883. [Google Scholar] [CrossRef]
  172. Wu, M.X.; Wang, Y.; Zhou, G.; Liu, X. Sparks from Different Worlds: Collaboration of MOFs and COFs. Coord. Chem. Rev. 2021, 430, 213735. [Google Scholar] [CrossRef]
  173. Peng, Y.; Zhao, M.; Chen, B.; Zhang, Z.; Huang, Y.; Dai, F.; Lai, Z.; Cui, X.; Tan, C.; Zhang, H. Hybridization of MOFs and COFs: A New Strategy for Construction of MOF@COF Core-Shell Hybrid Materials. Adv. Mater. 2018, 30, 1705454. [Google Scholar] [CrossRef]
  174. Cheng, Y.; Ying, Y.; Zhai, L.; Liu, G.; Dong, J.; Wang, Y.; Christopher, M.P.; Long, S.; Wang, Y.; Zhao, D. Mixed Matrix Membranes Containing MOF@COF Hybrid Fillers for Efficient CO2/CH4 Separation. J. Membr. Sci. 2019, 573, 97–106. [Google Scholar] [CrossRef]
  175. Zhou, W.; Liu, Y.; Teo, W.L.; Chen, B.; Jin, F.; Zhang, L.; Zeng, Y.; Zhao, Y. Construction of a Sandwiched MOF@COF Composite as a Size-Selective Catalyst. Cell Rep. Phys. Sci. 2020, 1, 100272. [Google Scholar] [CrossRef]
  176. Sun, D.; Kim, D.P. Hydrophobic MOFs@Metal Nanoparticles@COFs for Interfacially Confined Photocatalysis with High Efficiency. ACS Appl. Mater. Interfaces 2020, 12, 20589–20595. [Google Scholar] [CrossRef]
  177. Firoozi, M.; Rafiee, Z.; Dashtian, K. New MOF/COF Hybrid as a Robust Adsorbent for Simultaneous Removal of Auramine O and Rhodamine B Dyes. ACS Omega 2020, 5, 9420–9428. [Google Scholar] [CrossRef]
  178. Zhang, F.M.; Sheng, J.L.; Di Yang, Z.; Sun, X.J.; Tang, H.L.; Lu, M.; Dong, H.; Shen, F.C.; Liu, J.; Lan, Y.Q. Rational Design of MOF/COF Hybrid Materials for Photocatalytic H2 Evolution in the Presence of Sacrificial Electron Donors. Angew. Chem. Int. Ed. Engl. 2018, 57, 12106–12110. [Google Scholar] [CrossRef]
  179. Troyano, J.; Çamur, C.; Garzón-Tovar, L.; Carné-Sánchez, A.; Imaz, I.; Maspoch, D.; Maspoch, D. Spray-Drying Synthesis of MOFs, COFs, and Related Composites. Acc. Chem. Res. 2020, 53, 1206–1217. [Google Scholar] [CrossRef] [PubMed]
  180. Fu, J.; Das, S.; Xing, G.; Ben, T.; Valtchev, V.; Qiu, S. Fabrication of COF-MOF Composite Membranes and Their Highly Selective Separation of H2/CO2. J. Am. Chem. Soc. 2016, 138, 7673–7680. [Google Scholar] [CrossRef] [PubMed]
  181. Liu, X.; Hu, M.; Wang, M.; Song, Y.; Zhou, N.; He, L.; Zhang, Z. Novel Nanoarchitecture of Co-MOF-on-TPN-COF Hybrid: Ultralowly Sensitive Bioplatform of Electrochemical Aptasensor toward Ampicillin. Biosens. Bioelectron. 2019, 123, 59–68. [Google Scholar] [CrossRef] [PubMed]
  182. Zhou, N.; Ma, Y.; Hu, B.; He, L.; Wang, S.; Zhang, Z.; Lu, S. Construction of Ce-MOF@COF Hybrid Nanostructure: Label-Free Aptasensor for the Ultrasensitive Detection of Oxytetracycline Residues in Aqueous Solution Environments. Biosens. Bioelectron. 2019, 127, 92–100. [Google Scholar] [CrossRef]
  183. Chen, Y.; Yang, D.; Shi, B.; Dai, W.; Ren, H.; An, K.; Zhou, Z.; Zhao, Z.; Wang, W.; Jiang, Z. In Situ Construction of Hydrazone-Linked COF-Based Core–Shell Hetero-Frameworks for Enhanced Photocatalytic Hydrogen Evolution. J. Mater. Chem. A Mater. 2020, 8, 7724–7732. [Google Scholar] [CrossRef]
  184. Sun, D.; Jang, S.; Yim, S.J.; Ye, L.; Kim, D.P. Metal Doped Core–Shell Metal-Organic Frameworks@Covalent Organic Frameworks (MOFs@COFs) Hybrids as a Novel Photocatalytic Platform. Adv. Funct. Mater. 2018, 28, 1707110. [Google Scholar] [CrossRef]
  185. Cai, M.; Li, Y.; Liu, Q.; Xue, Z.; Wang, H.; Fan, Y.; Zhu, K.; Ke, Z.; Su, C.-Y.; Li, G.; et al. One-Step Construction of Hydrophobic MOFs@COFs Core–Shell Composites for Heterogeneous Selective Catalysis. Adv. Sci. 2019, 6, 1802365. [Google Scholar] [CrossRef]
  186. Yang, R.G.; Fu, Y.M.; Wang, H.N.; Zhang, D.P.; Zhou, Z.; Cheng, Y.Z.; Meng, X.; He, Y.O.; Su, Z.M. ZIF-8/Covalent Organic Framework for Enhanced CO2 Photocatalytic Reduction in Gas-Solid System. J. Chem. Eng. 2022, 450, 138040. [Google Scholar] [CrossRef]
  187. Yue, B.; Liu, J.; Li, G.; Wu, Y. Synthesis of Magnetic Metal Organic Framework/Covalent Organic Framework Hybrid Materials as Adsorbents for Magnetic Solid-Phase Extraction of Four Endocrine-Disrupting Chemicals from Milk Samples. Rapid Commun. Mass Spectrom. 2020, 34, e8909. [Google Scholar] [CrossRef] [PubMed]
  188. Fan, H.; Peng, M.; Strauss, I.; Mundstock, A.; Meng, H.; Caro, J. MOF-in-COF Molecular Sieving Membrane for Selective Hydrogen Separation. Nat. Commun. 2021, 12, 38. [Google Scholar] [CrossRef] [PubMed]
  189. Lin, T.; Wang, H.; Du, X.; Zhang, D.; Zhang, Z.; Liu, G. A COF-Coated MOF Framework Polysulfide Barrier Design for Enhanced Performance in Lithium-Sulfur Batteries. Electrochim. Acta 2022, 412, 140156. [Google Scholar] [CrossRef]
  190. Zhang, L.; Liu, Z.; Deng, Q.; Sang, Y.; Dong, K.; Ren, J.; Qu, X. Nature-Inspired Construction of MOF@COF Nanozyme with Active Sites in Tailored Microenvironment and Pseudopodia-Like Surface for Enhanced Bacterial Inhibition. Angew. Chem. 2020, 60, 3469–3474. [Google Scholar] [CrossRef]
  191. Zhang, H.W.; Zhu, Q.Q.; Yuan, R.; He, H. Crystal Engineering of MOF@COF Core-Shell Composites for Ultra-Sensitively Electrochemical Detection. Sens. Actuators B Chem. 2021, 329, 129144. [Google Scholar] [CrossRef]
  192. Li, C.; Xu, X.; Xing, J.; Wang, F.; Shi, Y.; Zhao, X.; Liu, J.; Yang, Y.; Zhao, Z. A Fluorescent Dual-Emissions UiO-66-NH2@TpTt-COF Core-Shell Composite for Sensitive and Optosmart Sensing of Tetracycline. Appl. Surf. Sci. 2023, 616, 156455. [Google Scholar] [CrossRef]
  193. Gao, M.L.; Qi, M.H.; Liu, L.; Han, Z.B. An Exceptionally Stable Core–Shell MOF/COF Bifunctional Catalyst for a Highly Efficient Cascade Deacetalization–Knoevenagel Condensation Reaction. Chem. Commun. 2019, 55, 6377–6380. [Google Scholar] [CrossRef]
  194. Zhao, J.; Jin, B.; Peng, R. New Core-Shell Hybrid Material IR-MOF3@COF-LZU1 for Highly Efficient Visible-Light Photocatalyst Degrading Nitroaromatic Explosives. Langmuir 2020, 36, 5665–5670. [Google Scholar] [CrossRef]
  195. Abeysinghe, A.K.; Peng, Y.-P.; Huang, P.-J.; Chen, K.-F.; Chen, C.-H.; Chen, W.-X.; Liang, F.Y.; Chien, P.-Y. Enhancing Visible-Light-Driven Photocatalysis: Unveiling the Remarkable Potential of H2O2-Assisted MOF/COF Hybrid Material for Organic Pollutant Degradation. Environ. Sci. Pollut. Res. 2024, 31, 50983–50999. [Google Scholar] [CrossRef]
  196. Wu, Z.; Li, W.; Hou, L.; Wei, Q.; Yang, H.; Jiang, Y.; Tang, D. A Novel Sunflower-like MOF@COF for Improved Photocatalytic CO2 Reduction. Sep. Purif. Technol. 2023, 311, 123322. [Google Scholar] [CrossRef]
  197. Garzón-Tovar, L.; Pérez-Carvajal, J.; Yazdi, A.; Hernández-Muñoz, J.; Tarazona, P.; Imaz, I.; Zamora, F.; Maspoch, D. A MOF@COF Composite with Enhanced Uptake through Interfacial Pore Generation. Angew. Chem. Int. Ed. Engl. 2019, 58, 9512–9516. [Google Scholar] [CrossRef] [PubMed]
  198. Das, S.; Ben, T. A [COF-300]-[UiO-66] Composite Membrane with Remarkably High Permeability and H2/CO2 Separation Selectivity. Dalton Trans. 2018, 47, 7206–7212. [Google Scholar] [CrossRef] [PubMed]
  199. Liu, P.; Cai, K.; Tao, D.J.; Zhao, T. The Mega-Merger Strategy: M@COF Core-Shell Hybrid Materials for Facilitating CO2 Capture and Conversion to Monocyclic and Polycyclic Carbonates. Appl. Catal. B 2024, 341, 123317. [Google Scholar] [CrossRef]
  200. Fakhraie, S.; Rajabi, H.R.; Rashidi, A. Fabrication and Application of Novel Core–Shell MIL-101(Cr)@UiO-66(Zr) Nanocrystals for Highly Selective Separation of H2S and CO2. J. Chem. Eng. 2023, 452, 139001. [Google Scholar] [CrossRef]
  201. Zhong, X.; Liu, Y.; Liang, W.; Zhu, Y.; Hu, B. Construction of Core-Shell MOFs@COF Hybrids as a Platform for the Removal of UO22+and Eu3+Ions from Solution. ACS Appl. Mater. Interfaces 2021, 13, 13883–13895. [Google Scholar] [CrossRef]
  202. Peng, H.; Huang, S.; Tranca, D.; Richard, F.; Baaziz, W.; Zhuang, X.; Samorì, P.; Ciesielski, A. Quantum Capacitance through Molecular Infiltration of 7,7,8,8-Tetracyanoquinodimethane in Metal-Organic Framework/Covalent Organic Framework Hybrids. ACS Nano 2021, 15, 18580–18589. [Google Scholar] [CrossRef]
  203. Pakulski, D.; Veró Nica Montes-Garcí, A.; Czepa, W.; Marcinkowski, D.; Peng, H.; Chudziak, T.; Gorczyński, A.; Gorczyński, G.; Kukułka, W.; Valentini, C.; et al. MOF (UiO-66-NH2)@COF (TFP-TABQ) Hybrids via on-Surface Condensation Reactions for Sustainable Energy Storage. Chem. Commun. 2024, 60, 412. [Google Scholar] [CrossRef]
  204. Peng, H.; Raya, J.; Richard, F.; Baaziz, W.; Ersen, O.; Ciesielski, A.; Samorì, P. Synthesis of Robust MOFs@COFs Porous Hybrid Materials via an Aza-Diels-Alder Reaction: Towards High-Performance Supercapacitor Materials. Angew. Chem. Int. Ed. Engl. 2020, 59, 19602–19609. [Google Scholar] [CrossRef]
  205. Sun, W.; Tang, X.; Yang, Q.; Xu, Y.; Wu, F.; Guo, S.; Zhang, Y.; Wu, M.; Wang, Y.; Sun, W.; et al. Coordination-Induced Interlinked Covalent- and Metal–Organic-Framework Hybrids for Enhanced Lithium Storage. Adv. Mater. 2019, 31, 1903176. [Google Scholar] [CrossRef]
  206. Li, Y.; Liu, Z.; Lu, W.; Zhao, M.; Xiao, H.; Hu, T.; Ma, J.; Zheng, Z.; Jia, J.; Wu, H. A Label-Free Electrochemical Aptasensor Based on the Core-Shell Cu-MOF@TpBD Hybrid Nanoarchitecture for the Sensitive Detection of PDGF-BB. Analyst 2021, 146, 979–988. [Google Scholar] [CrossRef] [PubMed]
  207. Yola, M.L.; Atar, N. Amperometric Galectin-3 Immunosensor-Based Gold Nanoparticle-Functionalized Graphitic Carbon Nitride Nanosheets and Core-Shell Ti-MOF@COFs Composites. Nanoscale 2020, 12, 19824–19832. [Google Scholar] [CrossRef] [PubMed]
  208. Xu, M.; Chen, K.; Zhu, L.; Zhang, S.; Wang, M.; He, L.; Zhang, Z.; Du, M. MOF@COF Heterostructure Hybrid for Dual-Mode Photoelectrochemical-Electrochemical HIV-1 DNA Sensing. Langmuir 2021, 37, 13479–13492. [Google Scholar] [CrossRef] [PubMed]
  209. Li, B.; Wen, H.M.; Wang, H.; Wu, H.; Tyagi, M.; Yildirim, T.; Zhou, W.; Chen, B. A Porous Metal-Organic Framework with Dynamic Pyrimidine Groups Exhibiting Record High Methane Storage Working Capacity. J. Am. Chem. Soc. 2014, 136, 6207–6210. [Google Scholar] [CrossRef]
  210. Xie, L.S.; Skorupskii, G.; Dincǎ, M. Electrically Conductive Metal-Organic Frameworks. Chem. Rev. 2020, 120, 8536–8580. [Google Scholar] [CrossRef]
  211. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Metal-Organic Frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef]
  212. Bavykina, A.; Kolobov, N.; Khan, I.S.; Bau, J.A.; Ramirez, A.; Gascon, J. Metal-Organic Frameworks in Heterogeneous Catalysis: Recent Progress, New Trends, and Future Perspectives. Chem. Rev. 2020, 120, 8468–8535. [Google Scholar] [CrossRef]
  213. Lawson, H.D.; Walton, S.P.; Chan, C. Metal-Organic Frameworks for Drug Delivery: A Design Perspective. ACS Appl. Mater. Interfaces 2021, 13, 7004–7020. [Google Scholar] [CrossRef] [PubMed]
  214. Liu, C.; Wang, J.; Wan, J.; Yu, C. MOF-on-MOF Hybrids: Synthesis and Applications. Coord. Chem. Rev. 2021, 432, 213743. [Google Scholar] [CrossRef]
  215. Daliran, S.; Oveisi, A.R.; Peng, Y.; López-Magano, A.; Khajeh, M.; Mas-Ballesté, R.; Alemán, J.; Luque, R.; Garcia, H. Metal–Organic Framework (MOF)-, Covalent-Organic Framework (COF)-, and Porous-Organic Polymers (POP)-Catalyzed Selective C–H Bond Activation and Functionalization Reactions. Chem. Soc. Rev. 2022, 51, 7810–7882. [Google Scholar] [CrossRef] [PubMed]
  216. Feng, X.; Ding, X.; Jiang, D. Covalent Organic Frameworks. Chem. Soc. Rev. 2012, 41, 6010–6022. [Google Scholar] [CrossRef]
  217. Liu, M.; Guo, L.; Jin, S.; Tan, B. Covalent Triazine Frameworks: Synthesis and Applications. J. Mater. Chem. A Mater. 2019, 7, 5153–5172. [Google Scholar] [CrossRef]
  218. Raza, S.; Nazeer, S.; Abid, A.; Kanwal, A. Recent Research Progress in the Synthesis, Characterization and Applications of Hyper Cross-Linked Polymer. J. Polym. Res. 2023, 30, 1–26. [Google Scholar] [CrossRef]
  219. Gonte, R.R.; Deb, P.C.; Balasubramanian, K. Hydrogen Sorption onto Hypercrosslinked Polymer Decorated with Metal-Organic Framework. J. Polym. 2013, 2013, 684584. [Google Scholar] [CrossRef]
  220. Li, S.; Zhang, Y.; Zhao, S.; Zhang, L.; Qiao, S.; Zhou, J. New Design to Enhance Phosphonate Selective Removal from Water by MOF Confined in Hyper-Cross-Linked Resin. Sci. Total Environ. 2024, 914, 169760. [Google Scholar] [CrossRef]
  221. Karimi, Z.; Karami, B.; Farahi, M.; Mahmoudi Asl, A. Reinforced MOF/Polymer Composite Based on a Copper Metal-Organic Framework Supported on Nanoporous Hyper Cross-Linked Polycalix[4]Resorcinarene for the Preparation of Phthalazine and 4H-Pyran Heterocycles. Mater. Today Chem. 2024, 38, 102085. [Google Scholar] [CrossRef]
  222. Ci, E.; Chen, Q.; Wang, Y.; Liu, T.; Xu, Z.; Liu, F.; Chen, P.; Zhao, T. The Hyper-Crosslinked Aniline Polymer@MOFs Hybrid Materials Reinforced with Active Ionic Sites for Efficient and Fast Cr(VI) Removal. Sep. Purif. Technol. 2025, 352, 128144. [Google Scholar] [CrossRef]
  223. Bariki, R.; Joseph, R.G.; El-Kadri, O.M.; Al-Sayah, M.H. The Development of Metal-Free Porous Organic Polymers for Sustainable Carbon Dioxide Photoreduction. Nanomaterials 2024, 14, 1432. [Google Scholar] [CrossRef]
  224. Zhang, Y.; Zhang, C.; Shi, W.; Zhang, Z.; Zhao, Y.; Luo, X.; Liu, X. Pyridine-Based Conjugated Microporous Polymers as Adsorbents for CO2 Uptake via Weak Supramolecular Interaction. New J. Chem. 2022, 46, 6394–6397. [Google Scholar] [CrossRef]
  225. Petit, C.; Bandosz, T.J. MOF–Graphite Oxide Composites: Combining the Uniqueness of Graphene Layers and Metal–Organic Frameworks. Adv. Mater. 2009, 21, 4753–4757. [Google Scholar] [CrossRef]
  226. 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]
  227. Ning, H.; Yang, Z.; Yin, Z.; Wang, D.; Meng, Z.; Wang, C.; Zhang, Y.; Chen, Z. A Novel Strategy to Enhance the Performance of CO2Adsorption Separation: Grafting Hyper-Cross-Linked Polyimide onto Composites of UiO-66-NH2 and GO. ACS Appl. Mater. Interfaces 2021, 13, 17781–17790. [Google Scholar] [CrossRef] [PubMed]
  228. Wang, R.; Lei, Y.; Ke, L.; Bai, L.; Zheng, Y.; Xu, Z.; Li, T.; Chen, S.; Zhang, D. Construction of Self-Healing Multifunctional Hyper-Crosslinked Hyperbranched Polymer @ Metal–Organic Polyhedra (HHMOP) Membranes. Eur. Polym. J. 2024, 211, 112966. [Google Scholar] [CrossRef]
  229. Naz, N.; Manzoor, M.H.; Naqvi, S.M.G.; Ehsan, U.; Aslam, M.; Verpoort, F. Porous Organic Polymers; an Emerging Material Applied in Energy, Environmental and Biomedical Applications. Appl. Mater. Today 2024, 38, 102198. [Google Scholar] [CrossRef]
  230. Chen, B.; Chen, W.; Wang, M.; Peng, C.; Chen, B.; Chen, W.; Wang, M.; Peng, C. Unravelling the Multiple Synergies in MOF/CMP Supramolecular Heterojunction for Enhanced Artificial Photosynthesis. Adv. Mater. Interfaces 2023, 10, 2201971. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the two most common types of metal–organic frameworks based on the nature of the ligand, whether it is azolate-based (ZIFs) or carboxylate-based (UiO, MIL, MOF, HKUST and DUT), and modification by combining ligands or by functionalization of their ligands to later enhance MOF–polymer interaction. Adapted with permission from Ref. [53] copyright American Chemical Society 2025, [79] copyright Elsevier 2021, [89] copyright Wiley-VCH 2023, [90] copyright MDPI 2023, [94] copyright American Chemical Society 2022, and [105] copyright Royal Society of Chemistry 2025.
Figure 1. Schematic representation of the two most common types of metal–organic frameworks based on the nature of the ligand, whether it is azolate-based (ZIFs) or carboxylate-based (UiO, MIL, MOF, HKUST and DUT), and modification by combining ligands or by functionalization of their ligands to later enhance MOF–polymer interaction. Adapted with permission from Ref. [53] copyright American Chemical Society 2025, [79] copyright Elsevier 2021, [89] copyright Wiley-VCH 2023, [90] copyright MDPI 2023, [94] copyright American Chemical Society 2022, and [105] copyright Royal Society of Chemistry 2025.
Polymers 17 01941 g001
Figure 2. Schematic representation of a hybrid membrane composed of a PIM polymer matrix (shown in orange) and dispersed MOF particles (shown in green). This integration combines the high intrinsic microporosity of PIMs with the crystalline structure of MOFs, mitigating the individual limitations while enhancing performance in applications like gas separation, CO2 adsorption, and catalysis.
Figure 2. Schematic representation of a hybrid membrane composed of a PIM polymer matrix (shown in orange) and dispersed MOF particles (shown in green). This integration combines the high intrinsic microporosity of PIMs with the crystalline structure of MOFs, mitigating the individual limitations while enhancing performance in applications like gas separation, CO2 adsorption, and catalysis.
Polymers 17 01941 g002
Figure 3. Main applications of MOF/PIM MMMs in different fields.
Figure 3. Main applications of MOF/PIM MMMs in different fields.
Polymers 17 01941 g003
Figure 4. Schematic illustrating the synthetic strategies of COF–MOF hybrids. Reproduced under the terms of the Creative Commons CC-BY 4.0 license (https://creativecommons.org/licenses/by/4.0/) [169]. Copyright 2023 the authors. Advanced Functional Materials published by Wiley-VCH GmbH.
Figure 4. Schematic illustrating the synthetic strategies of COF–MOF hybrids. Reproduced under the terms of the Creative Commons CC-BY 4.0 license (https://creativecommons.org/licenses/by/4.0/) [169]. Copyright 2023 the authors. Advanced Functional Materials published by Wiley-VCH GmbH.
Polymers 17 01941 g004
Figure 5. Applications of MOF-COF hybrids. (a) ZIF-67-in-TpPa-1 membrane for hydrogen separation. Reproduced under the terms of the Creative Commons CC-BY 4.0 license (https://creativecommons.org/licenses/by/4.0/) [188]. Copyright 2021 the authors. Nature Communications published by Springer Nature. (b) MOF-5/COF (M5C) for removal of auramine O and rhodamine B [177]. (c) Pd/UiO-66-NH2@COF composite for olefin hydrogenation catalysis. Reproduced under the terms of the Creative Commons BY-NC-ND 4.0 license (https://creativecommons.org/licenses/by-nc-nd/4.0/) [175]. Copyright 2020 the authors. Cell Reports Physical Science published by Elsevier. (d) S/Co/Zr-NC@TpPa cathode for Li-S battery [189]. (e) Ce-MOF@MCA sensor for OTC [182]. (f) MOF@COF nanoenzime with bacterial inhibition abilities [190].
Figure 5. Applications of MOF-COF hybrids. (a) ZIF-67-in-TpPa-1 membrane for hydrogen separation. Reproduced under the terms of the Creative Commons CC-BY 4.0 license (https://creativecommons.org/licenses/by/4.0/) [188]. Copyright 2021 the authors. Nature Communications published by Springer Nature. (b) MOF-5/COF (M5C) for removal of auramine O and rhodamine B [177]. (c) Pd/UiO-66-NH2@COF composite for olefin hydrogenation catalysis. Reproduced under the terms of the Creative Commons BY-NC-ND 4.0 license (https://creativecommons.org/licenses/by-nc-nd/4.0/) [175]. Copyright 2020 the authors. Cell Reports Physical Science published by Elsevier. (d) S/Co/Zr-NC@TpPa cathode for Li-S battery [189]. (e) Ce-MOF@MCA sensor for OTC [182]. (f) MOF@COF nanoenzime with bacterial inhibition abilities [190].
Polymers 17 01941 g005
Table 1. Most used pristine MOF types and their surface area, main characteristics, advantages and disadvantages, and functional applications with their corresponding references.
Table 1. Most used pristine MOF types and their surface area, main characteristics, advantages and disadvantages, and functional applications with their corresponding references.
MOFSurface Area
(m2/g)
CharacteristicsAdvantagesDisadvantagesApplicationsRef.
ZIF-8~1600Zn2+ + 2-methylimidazolate; sodalite-typeHigh thermal/chemical stability; simple synthesisHydrophobicity limits polar gas captureCO2/CH4 separation, membranes, catalysis[53,54,55,56]
ZIF-67~1500Co2+ analog of ZIF-8; same topologyRedox-active; magnetic; electrocatalyticProne to oxidationElectrocatalysis, batteries, sensors[57,58]
UiO-66~1100Zr6O4(OH)4 + BDC; fcu topologyHigh defect tolerance; stable in aqueous/acidic mediaSmaller pore sizeDrug delivery, catalysis, CO2 capture[59,60]
UiO-67~2500Zr6 + BPDC linker; larger poresAccommodates larger guestsCostlier than UiO-66Guest encapsulation, pollutant adsorption[61,62]
UiO-68~3300Zr6 + TPDC; very large poresSuperior pore volume; tunable linkerIncreased structural defectsPhotocatalysis, dye adsorption, enzyme carriers[63,64]
MIL-101(Cr)~4100Cr3+ + BDC; mesoporous cagesVery high porosity; chemical stabilityChromium toxicity; expensiveMethane/H2 storage, catalysis[65,66]
MIL-53(Al)~1100Al3+ + BDC; breathing frameworkFlexible; scalablePore instability on activationCO2 separation, VOC adsorption[67]
MIL-100(Fe)~2200–2900Fe3+ + BTC; MTN topology; mesoporousBiocompatible; water-stableModerate crystallinityDrug delivery, wastewater treatment[68,69]
MOF-5 (IRMOF-1)~3800Zn4O + BDC; cubic poresHistoric benchmark; easy to modifyVery moisture-sensitiveH2 storage, small-molecule adsorption[70,71]
HKUST-1~1800Cu2+ paddlewheels + BTCOpen metal sites; strong π-interactionsMoisture instabilityCO2/CH4 storage, catalysis, sensing[72,73]
DUT-49~5470Cu+ framework; flexible with NGA behaviorVery high surface area; responsive poresMechanically unstableGas adsorption, negative gas adsorption (NGA)[74]
Table 2. Summary of gas separation performance in various MOF/PIM-based mixed-matrix membranes (MMMs), highlighting the permeability (in Barrer) and selectivity for different gas pairs, along with the corresponding reference number.
Table 2. Summary of gas separation performance in various MOF/PIM-based mixed-matrix membranes (MMMs), highlighting the permeability (in Barrer) and selectivity for different gas pairs, along with the corresponding reference number.
Hybrid MaterialType of Gas SeparationPermeability (Barrer)SelectivityRef. nº
CAU-12-ODB/PIM-1H2/N27199 (H2)127[129]
ZIF-8/AO-PIM-1H2/CO25688 (H2)11.97[139]
ZIF-8/PIM-1O2/N21287 (O2)3.7[128]
ZIF-8/PIM-6FDA-OH MMMsC3H6/C3H8 (propylene/propane)3.5 (propylene)30[126]
CALF-20/PIM-1 MMMCO2/N28003 (CO2)25[105]
Mg-MOF-74 (CPO-27)/PIM-1 MMMCO2/CH421,000 (CO2)19.1[124,146]
Mg-MOF-74 (CPO-27)/PIM-1 MMMCO2/N221,000 (CO2)15.4[124,146]
MIL-101-HNO3/PIM-1 MMMCO2/N214,879 (CO2)24.3[147]
MOF-801/PIM-1 MMM (5% load)CO2/N29686 (CO2)27[148]
MOF-808/PIM-1 MMM (2% load)CO2/N26854 (CO2)23.2[42]
MUF-15/PIM-1 MMM (optimal 5% load)CO2/N2(38% increase over pure PIM-1) *17.72[149]
NH2-ZIF-7/PIM-1 MMM (20% load)CO2/CH42953 (CO2)20.6[150,152]
agZIF-62/PIM-1 MMM (30% load)CO2/CH45914 (CO2)67[151]
ZIF-S/PIM-1 MMM (15% load)CO2/CH42805 (CO2)21.09[82]
* In this case, permeability is expressed as the percentage increase relative to the pure PIM-1 membrane.
Table 3. Comparison of MOFs, COFs, and MOF–COF hybrids.
Table 3. Comparison of MOFs, COFs, and MOF–COF hybrids.
Property/AspectMOFsCOFsMOF–COF Hybrids
CompositionMetal ions or clusters linked by organic ligandsCovalent bonds between light elements (C, H, N, O)Combination of metal ions (MOFs) and covalent organic frameworks (COFs)
PorosityHigh, tunablePermanent, highly orderedSynergistic, with hierarchical porosity
Thermal/chemical stabilityModerate to high, but limited in harsh environmentsHigh, with chemical robustnessEnhanced stability due to combined frameworks
Electronic propertiesStrongly tunable via metal centersLimited electronic propertiesTunable properties combining MOF’s metal nodes and COF’s organic framework
FunctionalizationBroad, via choice of metal ions and linkersHigh, via diverse organic chemistryEnhanced functionalization through integration of metal centers and organic scaffolds
Synthesis challengesComplex, often requiring specific solvothermal conditionsModerate, with solvent-based or solid-state approachesComplex, requiring careful hybridization strategies
ApplicationsGas storage, catalysis, sensingEnergy storage, membranes, photocatalysisAdvanced applications: catalysis, gas separation, energy storage, and biomedicine
Advantages of hybrids--Synergy of MOF and COF properties: high surface area, tunable functionality, stability, and hierarchical pores
Table 4. Examples for each type of synthesis of MOF/COF hybrids.
Table 4. Examples for each type of synthesis of MOF/COF hybrids.
StrategyExample MaterialApplicationReference
MOF-first (in situ COF growth)NH2-UiO-66@TFPT-DETHPhotocatalyst for hydrogen evolution[183]
MOF-first (amorphous → crystalline COF)NH2-MIL-125(Ti)@COF-LZU1Photocatalyst for hydrogenation of styrene[184]
MOF-first (one-step COF synthesis)NH2-MIL-101(Fe)@NTU-COFStyrene oxidation catalyst[185]
COF-first (layer-by-layer growth)COF-300@ZIF-8Gas separation[180]
COF-first (sequential MOF-on-COF growth)Tp-Pa-1 COF@ZIF-8Photocatalyst for CO2 reduction[186]
Post-synthetic linkingFe3O4@A-TpBD@NH2-MIL-125(Ti) EDC absorption and selectivity[187]
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

Durán-Egido, V.; García-Giménez, D.; Martínez-López, J.C.; Pérez-Vidal, L.; Carretero-González, J. Metal–Organic Frameworks as Fillers in Porous Organic Polymer-Based Hybrid Materials: Innovations in Composition, Processing, and Applications. Polymers 2025, 17, 1941. https://doi.org/10.3390/polym17141941

AMA Style

Durán-Egido V, García-Giménez D, Martínez-López JC, Pérez-Vidal L, Carretero-González J. Metal–Organic Frameworks as Fillers in Porous Organic Polymer-Based Hybrid Materials: Innovations in Composition, Processing, and Applications. Polymers. 2025; 17(14):1941. https://doi.org/10.3390/polym17141941

Chicago/Turabian Style

Durán-Egido, Victor, Daniel García-Giménez, Juan Carlos Martínez-López, Laura Pérez-Vidal, and Javier Carretero-González. 2025. "Metal–Organic Frameworks as Fillers in Porous Organic Polymer-Based Hybrid Materials: Innovations in Composition, Processing, and Applications" Polymers 17, no. 14: 1941. https://doi.org/10.3390/polym17141941

APA Style

Durán-Egido, V., García-Giménez, D., Martínez-López, J. C., Pérez-Vidal, L., & Carretero-González, J. (2025). Metal–Organic Frameworks as Fillers in Porous Organic Polymer-Based Hybrid Materials: Innovations in Composition, Processing, and Applications. Polymers, 17(14), 1941. https://doi.org/10.3390/polym17141941

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