Next Article in Journal
A Heterobimetallic Au(I)–Ru(II) Complex Bridged by dppb: Synthesis, Structural and Solution Characterization, BSA Interaction and In Vivo Toxicity Evaluation in Wistar Rats
Previous Article in Journal
Progress and Prospect of Sm-Fe-N Magnets
Previous Article in Special Issue
Preparation and Photocatalytic Hydrogen Production of Pink ZnS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MIL Series in MOFs for the Removal of Emerging Contaminants: Application and Mechanisms

1
School of Energy and Environment, Inner Mongolia University of Science and Technology, Baotou 014010, China
2
School of Energy and Environmental Engineering, University of Science & Technology Beijing, Beijing 100083, China
3
Beijing Key Laboratory of Resource-Oriented Treatment of Industrial Pollutants, University of Science & Technology Beijing, Beijing 100083, China
4
Tangshan Jidong Cement Co., Ltd., Beijing 100020, China
5
Cooperative Innovation Center of Ecological Protection and Comprehensive Utilization in Inner Mongolia Section of the Yellow River Basin, Baotou 014010, China
6
Beijing Shouke Xingye Engineering Technology Co., Ltd., Beijing 102300, China
7
National Engineering Laboratory for Site Remediation Technologies, Beijing 100015, China
*
Authors to whom correspondence should be addressed.
Inorganics 2025, 13(10), 324; https://doi.org/10.3390/inorganics13100324
Submission received: 23 August 2025 / Revised: 24 September 2025 / Accepted: 26 September 2025 / Published: 29 September 2025
(This article belongs to the Special Issue Nanocomposites for Photocatalysis, 2nd Edition)

Abstract

In global economic integration and rapid urbanization, the equilibrium between resource utilization efficiency and ecological preservation is confronted with significant challenges. Emerging contaminants have further exacerbated environmental pressures and posed threats to the ecosystem and human health. Metal–organic frameworks (MOFs) have emerged as a prominent area of research in ecological remediation, owing to their distinctive porous configuration, substantial specific surface area, and exceptional chemical stability. The Materials Institute Lavoisier (MIL) series (e.g., MIL-53, MIL-88, MIL-100, MIL-101, and MIL-125) has been shown to effectively promote the separation and migration of photogenerated carriers and significantly enhance the degradation of organic contaminants. This property renders it highly promising for the photocatalytic degradation of emerging contaminants. This paper provides a concise overview of the classification, synthesis methods, modification strategies, and application effects of MIL series MOFs in the removal of emerging contaminants. The advantages and limitations of MIL series MOFs in environmental remediation are further analyzed. Particularly, we offer insights and support for innovative strategies in the treatment of emerging contaminants, including POPs, PPCPs, VOCs, and microplastics, contributing to technological innovation and development in environmental remediation. Future development of MOFs includes the optimization of the performance of the MILs, reducing the high synthesis costs of MILs, applying MILs in real-environment scenarios, and accurate detection of degradation products of environmental pollutants.

1. Introduction

With the acceleration of global industrialization and continued population growth, environmental governance faces unprecedented challenges [1,2], particularly with the emergence of new contaminants that seriously threaten ecosystems and human health [3]. The term “emerging contaminants” is a broad category that includes Persistent Organic Pollutants (POPs), Pharmaceuticals and Personal Care Products (PPCPs), Volatile Organic Compounds (VOCs), microplastics (MPs) and nanoplastics (NPs), and other novel organic contaminants that have not yet been thoroughly investigated [4,5]. Emerging contaminants are characterized by biotoxicity, environmental persistence, and bioaccumulation [6], and their hazards far exceed those of traditional pollutants. Their complex molecular structure and extremely low ecological concentrations (usually at the level of micrograms/liter or even nanograms/liter) have exceeded the detection range of traditional monitoring technologies [7].
These contaminants have become a significant challenge for global environmental governance due to their environmental persistence, bioaccumulation, and potential ecological and health hazards. For instance, POPs, including perfluorinated compounds (PFCs) and polychlorinated biphenyls (PCBs), are resistant to environmental degradation and can bioaccumulate through the food chain, resulting in immune system disorders and reproductive toxicity [5]. PPCPs, derived from antibiotics and cosmetics, are widespread in aquatic environments, potentially contributing to the dissemination of drug-resistant genes [8]. VOCs are not only involved in atmospheric photochemical reactions but also pose potential health hazards due to bioaccumulation [9]. Microplastics and nanoplastics are ubiquitous in oceans, rivers, and soils, exhibiting strong transport capacity and bioaccumulation, which may trigger complex health risks [10,11]. The complexity and diversity of these contaminants necessitate the development of more efficacious management strategies to address the pressing environmental challenges of the present moment.
Removing emerging contaminants, which vary in nature and characteristics, has historically posed a formidable challenge in environmental science. The predominant removal methodologies currently encompass biological treatment, adsorption, and membrane technology [3,12]. Biological treatment relies on microbial metabolism to degrade contaminants; however, the high chemical stability and poor biodegradability of many emerging contaminants limit its efficiency. Adsorption technology utilizes the surface properties of adsorbents to remove contaminants; nevertheless, traditional adsorbents, such as activated carbon, do not have sufficient adsorption capacity for polar or highly water-soluble pollutants. Although it can effectively retain some emerging contaminants, it faces the challenges of membrane fouling and the high cost of membrane contamination. In this context, advanced oxidation technologies (AOTs) have emerged as a pivotal solution for addressing emerging challenges in pollutant removal. These technologies can to degrade complex organic contaminants into smaller molecules or mineralize them completely into carbon dioxide and water by generating strong oxidizing free radicals, such as hydroxyl and superoxide radicals [13]. Among these technologies, photocatalytic oxidation technology has emerged as a particularly noteworthy solution due to its notable advantages, including high efficiency, environmental sustainability, and reduced environmental impact. It has been demonstrated to efficiently degrade contaminants by utilizing photogenerated carriers produced by photocatalysts when exposed to light. These carriers catalyze the generation of free radicals, exhibiting potent oxidizing properties [14].
In recent years, with the development of advanced oxidation technology, metal–organic frameworks (MOFs) have become a research hotspot in the field of photocatalysis for environmental remediation due to their unique porous structure, high specific surface area, and excellent chemical stability [15]. Among them, the Materials Institute Lavoisier (MIL) series (e.g., MIL-53, MIL-88, MIL-100, MIL-101, and MIL-125), as an important branch of MOFs, has attracted widespread attention due to their flexible pore structure, abundant active sites, and diverse synthesis pathways [14,16]. The application of MOFs in photocatalysis is mainly achieved in three ways: First, they are directly used as photocatalysts to degrade contaminants by using their light absorption and charge separation abilities [17]; second, they are compounded with other semiconductor materials to form heterojunction structures to improve the separation efficiency of photogenerated carriers; and third, they are modified by functionalization to enhance their light absorption properties and catalytic activities [18]. However, despite the excellent performance of MIL series MOFs in laboratory studies, they still face many challenges in practical applications. For example, MIL materials have weak visible-light absorption and mainly absorb UV light, which limits the efficiency of utilizing sunlight. In addition, the lack of stability of MIL materials in complex environmental media limits their application in practical pollution control. Therefore, the development of efficient synthesis methods, the optimization of material properties, and the exploration of the application of MIL series MOFs in emerging contaminants treatment have become key research directions in environmental science.
A substantial corpus of approximately 600 journal articles has been published on the application of MILs to emerging contaminants (as demonstrated in Figure 1). These studies underscore the profound research interest in the MIL series of MOFs within emerging contaminants management and reveal their considerable potential in environmental remediation. This review aims to provide a comprehensive and mechanistically focused overview of the application of MIL series MOFs (e.g., MIL-53, MIL-88, MIL-100, MIL-101, and MIL-125) for the remediation of diverse emerging contaminants, including POPs, PPCPs, VOCs, and microplastics, thus highlighting their broad-spectrum potential. Particular attention is given to correlating the structural features of these MIL frameworks—such as the metal node type, ligand functionality, and pore architecture—with their adsorption, photocatalytic, and Fenton-like removal performance, thereby elucidating the structure–activity relationship. In addition, this work integrates recent findings on the hydrolytic stability, regeneration behavior, and performance under environmentally relevant conditions, offering a more realistic understanding of their applicability. Beyond summarizing the laboratory results, this review discusses green synthesis strategies, cost considerations, and challenges in scaling up, providing a forward-looking perspective for the rational design and practical deployment of MIL-based materials in sustainable environmental remediation.

2. Classification of MILs

To date, MILs (MIL = Materials Institute Lavoisier) have attracted considerable attention due to their flexible pore structures, excellent stability, abundant active sites, and diverse synthetic routes [19]. The numerical suffixes of the MILs denote the presence of different metal ions and specific structures in the oxidation state [20]. MILs include a variety of series (MIL-47, MIL-53, MIL-68, MIL-88, MIL-97, MIL-100, MIL-101, MIL-125, MIL-127). These five materials, MIL-53, MIL-88, MIL-100, MIL-101, and MIL-125, were chosen because they are representative in terms of the structural diversity, photocatalytic potential, research maturity, and experimental feasibility (as shown in Table 1).

2.1. MIL-53

Férey et al. synthesized and characterized the first three-dimensional chromium (III) dicarboxylate, MIL-53(Cr3+) [26]. One of its structures consists of an infinite purely inorganic chain composed of octahedrally coordinated Cr3+ that share opposite corners so that Cr3+ and OH alternate and extend between neighboring chains via benzene-1,4-dicarboxylic acid cross-linking, and the other structure has the dicarboxylates bonded in the same way. Each metal center is linked to the oxygen of four different carboxylic acid groups from four different linkers [27]. The MIL-53 series is a widely known porous flexible framework for MOFs. Researchers have employed various common metal salts (Al, Fe, Cr) and linkers (terephthalic acid) in the synthesis of these frameworks [28].

2.2. MIL-88

The MIL-88 series has grown in popularity in research circles, owing to its demonstrated capacity to catalyze oxidation with remarkable efficiency [29]. Its structural composition consists of a metal ortho-octahedron, comprising diverse organic linear dicarboxylic acid ligands interconnected with an oxygen-centered trimer. The specific designations MIL-88A, B, C, and D correspond to the following ligands: fumarate, terephthalic acid, 2,6-naphthalene dicarboxylate (2,6-NDC), and 4-4′ biphenyldicarboxylic acid (4-4′ BPDC), respectively [30]. MIL-88, prepared using an iron precursor, became a potential Fenton-like catalyst. Among the various forms of MIL-88, those designated as MIL-88A (Fe) and MIL-88B (Fe) have been observed to possess both exceptional chemical stability and hydrolytic stability [29]. These characteristics have demonstrated their superior removal performance in the context of contaminant removal in water.

2.3. MIL-100

The MIL-100 series is a porous trivalent metal carboxylate, and research reports published the synthesis and structural examination of macro-porous chromium (III) carboxylic acids [31]. Subsequent studies successfully synthesized MIL-100(Cr) by combining Cr3+ ions with 1,3,5-benzenetricarboxylic acid (BTC). The metal unit of MIL-100 features an organically connected structure similar to other trivalent metal frameworks, enabling its extension to MIL-100(M) (M = Cr, Fe, Sc, Al, V) [32]. Horcajada et al. initially synthesized iron-based MIL-100 over the course of six days [33]. Following multiple iterations of method refinement, the researchers prepared iron-based MIL-100 under mild conditions, using FeCl3-H2O as the iron source and 1,3,5-benzenetricarboxylic acid as the ligand [32].

2.4. MIL-101

The porous chromium terephthalate metal–organic framework MIL-101 was first synthesized and reported in [34]. MIL-101(Cr) and MIL-101(Fe) in the MIL-101 series have been the focus of systematic and extensive investigation, particularly MIL-101(Cr), which is among the most representative and well-studied MOFs in the MIL series [35]. MIL-101 possesses a number of advantageous physicochemical properties, including elevated chemical and thermal stability and a substantial number of coordinated unsaturated metal sites, thereby rendering it superior to a series of materials [36].

2.5. MIL-125

Dan-Hardi’s pioneering synthesis of MIL-125 utilized Ti8O8(OH)4 clusters and terephthalic acid as precursors [37]. MIL-125 has undergone extensive investigation and application as a highly regarded photocatalyst, exhibiting high stability and photocatalytic activity and a rigid framework conducive to strong metal–ligand bonding [38]. MIL-125, a titanium-based metal–organic framework (MOF), exhibits remarkable photocatalytic activity, with the Ti component transforming from Ti4+ to reduced Ti3+ under light irradiation, positioning it as a leading photocatalyst among its Ti-based MOF counterparts [39]. In recent studies, the photocatalytic activity of MIL-125 was further enhanced by heterojunction, loading metal ions, or making functionalization changes [40].
The five metal–organic framework materials previously mentioned possess distinct structural and performance characteristics. Structurally, MIL-53 exhibits a flexible structure and a “breathing effect.” MIL-88 features multiple interconnected pores, while MIL-100 and MIL-101 adopt an MTN topology. MIL-101 further demonstrates a larger pore size. MIL-125 presents a three-dimensional network composed of titanium centers and terephthalic acid ligands. In terms of performance, MIL-101 is distinguished by its notably high specific surface area and favorable thermal stability. Conversely, MIL-53 and MIL-88 exhibit distinctive advantages in the synergistic integration of adsorption and photocatalytic properties. In conjunction with the imperative for managing emerging contaminants, these materials exhibit considerable promise in the degradation of such compounds, a property attributable to their adaptable porosity, substantial surface area, and plentiful active sites. In practical applications, the rational selection and tailored modification of MOFs according to the contaminant type and treatment objectives are essential for achieving optimal remediation efficiency.

3. Synthesis Method of MILs

To date, a wide range of synthesis strategies for MILs has been developed and progressively refined. Among the many synthesis methods, the solvothermal/hydrothermal method has attracted much attention due to its simple operation and other advantages, and it has become the primary method of choice for the synthesis of MILs. In addition to solvothermal/hydrothermal methods, some commonly used synthesis methods, such as steam-assisted synthesis, microwave-assisted synthesis, electrochemical synthesis, and mechanical grinding synthesis (as shown in Figure 2), are also used for the synthesis of MILs, which have their advantages and disadvantages. The target MILs synthesized by different methods may have other shapes and crystal structures. Fallah prepared mercerized zeolite/MIL-101(Cr) using hydrothermal, solvent-thermal, and reflux methods. The characterizations of the mercerized zeolite/MIL-101(Cr) synthesized by different synthetic methods are different [41]. These variations subsequently affect the contaminant removal efficiency. This review will discuss and summarize the aforementioned methods for the preparation of MILs. After this section, a list of green, economical, and efficient synthesis methods will be provided to provide further reference value for reading. Table 2 lists the preparation methods, synthesis strategies, and properties of the MIL series of materials.

3.1. Solvothermal Synthesis/Hydrothermal Synthesis

Solvothermal and hydrothermal syntheses are two frequently employed synthetic methods for the preparation of MILs. The fundamental principles underpinning these methods are analogous, with both relying on elevated temperatures and pressures within a closed system to facilitate the desired reactions [26,70]. The solvothermal/hydrothermal synthesis method can efficiently synthesize complex materials, adjust reaction conditions, perform large-scale synthesis, and be applicable to a wide range of materials. Valekar et al. used the organic ligand H2BDC (terephthalic acid), the metal source Al(NO3)3⋅9H2O (Aluminum nitrate nonahydrate), and the solvent DIW (deionized water), which were stirred and dissolved. The mixture was transferred to a stainless steel autoclave (lined with Teflon) to provide a hermetically sealed high-pressure environment at 493 K for 3 days [71]. In the synthesis of certain MIL series, auxiliary reagents are added to regulate the reaction environment, such as to control the acidity or alkalinity of the solution or promote the formation of crystals. Examples include NaOH [72] (alkaline modifier, used in the synthesis of MIL-88) and HF [73] (hydrofluoric acid, used in the morphological control of MIL-101). However, it should be noted that the synthesis process may result in the production of byproducts or impurities, which can potentially compromise the purity of the final product [74]. It has been demonstrated that when organic solvents are employed, solvent residues and impurities can deleteriously affect the final photocatalytic performance, stability, and other relevant parameters [75].

3.2. Microwave-Assisted Synthesis

Microwave-assisted synthesis is a technique that facilitates the synthesis of MILs by directly heating solvents and reactants in the reaction system through microwave radiation. In the microwave process, the reaction system is rapidly heated by dielectric heating (the solvent or reactant molecules absorb energy in the microwave electric field, and the polar molecules (e.g., water, DMF) heat up because of the polarization oscillation) and dipole heating [76] (the polar molecules in the alternating electric field due to the change of direction and friction and collision, the release of heat energy), acting directly on the solvent and substances in the reaction system. In comparison with conventional hydrothermal and hot-solvent methodologies, microwave-assisted synthesis has been shown to markedly reduce the reaction times, enhance the energy utilization efficiency, render the synthesis process more environmentally friendly and economical, and yield MILs with high crystallinity [52]. This method has garnered significant attention in MIL research and development in recent years, and it is particularly well-suited for the rapid preparation of high-performance MIL materials [77]. Pouyanfar et al. employed a modified microwave synthesis to prepare MIL-100 (Fe) by dispersing ferric chloride and benzene tricarboxylate in deionized water. They optimized the synthesis conditions to obtain the best crystallinity and particle size by adjusting the microwave maintenance time (5.5–15 min) at 1000 W power [17]. In the synthesis of MILs using microwave, it is imperative to meticulously control the reaction parameters of temperature, time, and power level to ensure the success of the synthesis and the quality of the material.

3.3. Mechanical Grinding Synthesis

Mechanical grinding is a green synthesis technology that induces chemical reactions in solid reactants through mechanical forces (e.g., grinding, ball milling, etc.) [78], and its core principle lies in the use of friction, shear, and compression by mechanical forces to disrupt the crystal structure of the reactants and to promote the coordination reaction between the metal ions and the organic ligands, while the localized high-temperature and high-pressure environment can further accelerate the bond cleavage and reorganization [79]. The method mainly comprises two forms: solvent-free mechanical grinding and liquid-assisted mechanical grinding (LAG) [78]. The former relies entirely on mechanical force to drive the reaction without solvent intervention, while the latter involves the addition of trace amounts of polar solvents (e.g., water, ethanol) as a reaction medium to enhance the efficiency of the reaction and the crystallinity of the products. In one study, Souza et al. synthesized MIL-100(Fe) through two mechanochemical methods: manual grinding (MG) and vortex grinding (VG). The MG method entailed the manual grinding of 1,3,5-benzenetricarboxylic acid (H3BTC), ferric nitrate (Fe(NO3)3), and 5-FU for a duration of 1–2 h using an onyx mortar and pestle. VG involves the processing of the mixture in a 3D-printed, customized, and fixture-fixed polypropylene container. These containers undergo rotation at elevated speeds in a vortex mixer, thereby increasing the frequency of collisions and facilitating the reaction [59]. The mechanical grinding synthesis has been demonstrated to have significant advantages in MIL synthesis. However, it has also exhibited certain shortcomings, including low crystallinity of the product, poor experimental reproducibility due to the wear and tear of the equipment, an unclarified reaction mechanism, and structural inhomogeneity and dust problems that may occur in large-scale production [80].

3.4. Electrochemical Synthesis

The electrochemical synthesis of MILs is a green synthesis technique that generates MILs directly on the surface of electrodes or in solution through electrochemical reactions [81]. The core principle of this technique is to utilize anodic oxidation to dissolve metal electrodes (e.g., Zn, Cu, Fe, etc.) [62,82,83], releasing metal ions that subsequently coordinate with organic ligands (e.g., carboxylate and imidazole ligands) in solution [84,85]. The coordination reaction occurs, and MILs crystals are eventually formed. The method principally consists of two forms: anodic dissolution and cathodic reduction. The former generates metal ions through anodic oxidation, while the latter modulates the reaction environment through cathodic reduction to promote MILs’ growth. The electrochemical synthesis boasts several notable advantages, including mild reaction conditions, controllability, the absence of the need for elevated temperatures and pressures, and its environmental friendliness. This renders it particularly well-suited for the synthesis of heat-sensitive MILs [64]. Larasati et al. synthesized NH2-MIL-101(Fe), 2-aminoterephthalic acid (NH2-H2BDC), and tetrabutylammonium tetrafluoroborate (TBATFB), where the initial step involved the dissolution of these components in N, N-dimethylformamide (DMF). The mixture was then subjected to stirring for a duration of 30 min, which was deemed sufficient for complete dissolution. Subsequently, the mixture was transferred into an electrochemical cell comprising two iron plates serving as electrodes, where it underwent electrolysis for a duration of 30 min at an applied voltage of 15 V. During electrolysis, Fe2+ was oxidized to Fe3+, which coordinated with the NH2-H2BDC ligand to produce NH2-MIL-101 (Fe) [62]. However, its limitations include a slower reaction rate [81], potentially lower crystallinity of the product [86], limited choice of electrode materials, and difficulty in scale-up production. In recent years, electrochemical synthesis has demonstrated promising applications in the fields of functionalized MILs synthesis [64], thin-film MILs preparation, and electrochemical sensors. It is expected that, through the optimization of electrode materials, reaction conditions, and electrolyte composition, this technology will further promote its practical application in the fields of catalysis and separation [87].

3.5. Ultrasound-Assisted Synthesis

Ultrasound-assisted synthesis is a green synthesis technique based on ultrasound energy-driven synthesis. This technique significantly accelerates the dynamic coordination process of metal ions with organic ligands [88] through the cavitation effect triggered by its propagation in liquids, microjets, and localized extreme physical conditions (e.g., transient high temperature of ~5000 K and high pressure of ~1000 atm) [89]. The result of this process is the efficient preparation of MILs. The core mechanism encompasses three aspects: the cavitation effect (release of energy from the violent collapse of bubbles in the liquid, which promotes molecular collisions and chemical bond reorganization of reactants); the microjet effect (enhancement of the mixing homogeneity of the reaction system, avoiding local concentration inhomogeneity) [90]; and mechanical vibrational effect (disruption of the crystal structure of the reactants, increasing the reactivity) [91]. The experimental procedure typically involves the dissolution of metal salts and organic ligands in a solvent, followed by ultrasonic irradiation treatment (frequency 20–40 kHz, power 50–500 W, duration minutes to hours). This is followed by centrifugal separation of the products and subsequent drying and activation steps [92]. The reaction time is brief; for instance, Ge can be synthesized within 25 min. The synthesis of MIL-53 is much faster than the 24 h required by the traditional solvothermal synthesis, with mild conditions [93] (room temperature or low-temperature operation for heat-sensitive ligands or substrates). The product’s high crystallinity is promoted by the cavitation effect, which facilitates homogeneous nucleation. The method is also environmentally friendly, as it requires reduced solvent usage, making it particularly suitable for rapidly preparing various MOFs and their composites and their composites, such as ZIF, MIL, and UiO, among others. [89]. However, its limitations include the complex optimization of ultrasound parameters (e.g., frequency, power, time) (94) and difficulty in scale-up production [94].

3.6. Other Synthetic Methods

In addition to the widely used synthesis methods mentioned above, various innovative strategies exist for synthesizing MIL series. These emerging synthetic methods offer significant efficiency, cost, and environmental protection advantages. Furthermore, they can optimize the properties and structures of MILs to meet the needs of applications in different fields. As research progresses, scientists continue to explore more efficient, controllable, and environmentally friendly synthesis pathways to overcome the limitations of traditional methods. For instance, Baluk and his team employed a thermal injection method for synthesizing NH2-MIL-125(Ti) in octahedral shape and compared it with the conventional solvothermal synthesis. The thermal injection method substantially reduced the synthesis time, from the traditional 24 h to a mere 4 h, while preserving high crystallinity and specific surface area. This method enhanced the synthesis efficiency and was carried out at reduced temperatures and pressures, exhibiting reduced energy consumption and increased environmental sustainability. However, concerning the photocatalytic performance, both exhibited comparable efficiencies (phenol photocatalytic degradation efficiencies of 59.2% and 56.0%, respectively) [95]. Tan et al. successfully prepared MIL-100(Fe) using an aqueous-based non-thermal synthesis method. This synthesis method does not require high temperatures or pressures, nor does it necessitate strong corrosive reagents (e.g., hydrofluoric acid). It is characterized by its environmental sustainability, simplicity, and efficiency. The synthesized MIL-100(Fe) material exhibits a high specific surface area (1893 m2/g), excellent thermal stability, and good aqueous stability [16]. Unlike conventional methods, emerging synthesis methods offer distinct advantages, including reduced processing time, mild reaction conditions, controllable crystallinity and morphology, and streamlined post-processing.

4. Modification of MILs

Conventional MILs have demonstrated removal efficiency, selectivity, and stability limitations. Consequently, there is a growing need to optimize their performance through different modification strategies. Researchers have made substantial advancements in the removal capacity of MILs for organic contaminants and heavy metal ions through various methodologies, including surface modification, post-synthesis modification, doping, and heterojunction. In the domain of environmental remediation, for instance, the specific surface area of MIL-88A(Fe) before modification was 15.9 m2/g, exhibiting weak visible light absorption and low efficiency in the photocatalytic degradation of Rhodamine B (RhB). However, the modification of MIL-88A(Fe) with ultrathin graphite oxide (GO) has substantially enhanced its surface area, reaching 408.9 m2/g. This modification has resulted in enhanced visible-light absorption, accompanied by an 8.4-fold increase in photocatalytic efficiency and the retention of high efficiency over multiple cycles. This enhancement in performance is particularly noteworthy when compared to the performance of the unmodified material [96].

4.1. Elemental Doping

4.1.1. Transition Metals

Transition metals exhibit a range of oxidation states and distinctive electronic configurations. Introducing dopants into MILs can result in the formation of novel active sites. These active sites can adsorb and activate reactant molecules, thereby reducing the activation energy of the reaction and increasing the reaction rate. Liu et al. demonstrated a substantial enhancement in tetracycline (TC) degradation efficiency by applying activated PMS in conjunction with Co-doping MIL-53(Al). The degradation efficiency of TC reached 94% in the 25% Co-MIL-53(Al)/PMS system, which was significantly higher than that of the undoped MIL-53(Al)/PMS system (66%) and PMS alone (26.9%) [97]. In the domain of photocatalysis, the d-orbital electronic properties in MIL-doped transition metals have been observed to trap photogenerated electrons or holes, thereby forming shallow potential wells and hindering the formation of photogenerated carrier recombination.
Furthermore, transition metals, as electron acceptors, have been demonstrated to facilitate carrier separation, thereby enhancing the photogenerated charge migration efficiency through the metal–ligand charge transfer (MLCT) mechanism [98]. Wang et al. synthesized the composite catalyst Ru@NH2-MIL-125/MnO, doped with Ru nanoparticles, via hydrothermal synthesis. The composite catalyst exhibited a 10-fold increase in photocatalytic efficiency compared to the unaltered pristine NH2-MIL-125 [99]. The doped Ru became a reactive site for photocatalysis, increasing the photogenerated electron and hole transfer rates and effectively reducing the electron–hole recombination rate. The transition metal doping enhanced the light absorption, augmented the effective charge transfer, and facilitated separation.

4.1.2. Main Group Elements

Significant progress has been made in doping main-group elements (e.g., sulfur, nitrogen, phosphorus) into MILs in recent years. These doping modifications have enhanced the catalytic performance of the materials and optimized their applications in environmental treatment, gas adsorption, and energy storage, among others. Du et al. have synthesized a novel catalyst of S-MIL-53(Fe) via sulfur sublimation (Figure 3), and the S doping has been shown to increase both the pore size of the catalyst and its degradation capacity for the pollutant oxytetracycline (OTC) [100]. While the doping modifications have become a prevalent strategy to enhance the performance of catalysts, their impact is not universally beneficial. This complexity stems from the interaction between the dopant element and the catalyst matrix and the effect of doping on various aspects of the catalyst’s electronic structure, surface morphology, and pore structure. For instance, George et al. examined the impact of cation doping (Li+, Na+, K+) on the photocatalytic performance of MIL-53(Fe), focusing on the disparities in photocatalytic activity before and after doping [101]. The results obtained from this investigation demonstrate that the photocatalytic performance of MIL-53(Fe) diminishes considerably, despite cation doping reducing the band gap energy in the band gap energy. This phenomenon is attributed to the doping-induced structural distortion and the weakening of electron transport capacity.

4.1.3. Lanthanides

Doping lanthanide elements (e.g., lanthanum, cerium, neodymium, gadolinium) into MILs represents an effective strategy to enhance their functionality by modulating the electronic structure, pore environment, and active sites. Notably, certain lanthanides (e.g., Ce3+/Ce4+, Eu2+/Eu3+) possess variable valence states, which can be utilized as redox-active sites. Their substantial ionic radii (e.g., the ionic radii of La3+ are approximately 1.06 Å) which facilitates the formation of stable high coordination number (typically 8–12) coordination bonds with oxygen- and nitrogen-containing ligands. Jiang et al. have demonstrated that lanthanum (La)-doped NH2-MIL-53 significantly modified the structure and performance of the initial catalyst (Figure 4), exhibiting remarkable efficacy in the removal of arsenic (V) and phosphate from water [102]. Wan et al. optimized the electronic structure of iron-based metal–organic framework material (NH2-MIL-101) by gadolinium (Gd) doping to improve the degradation performance of arsenic (V) and phosphorus from water and to realize electrochemical regeneration [103]. While lanthanide doping has been shown to enhance the performance of materials, it does result in a cost increase. However, given the performance enhancement observed (e.g., higher catalytic activity, enhanced photoluminescence properties, and improved stability), the additional cost can be justified in specific application scenarios where high-performance materials are required.

4.2. Ligand Functionalization

The ligand functionalization modification of the MIL series is a critical strategy for optimizing their properties by introducing specific organic ligands. This modification approach enables the regulation of the pore structure and surface properties of MILs, thereby significantly enhancing their applications in fields such as gas adsorption, separation, catalysis, and environmental remediation. As illustrated by MIL-125(Ti) [95], the post-synthesis modification synthesis involving the grafting of reactive group-containing ligands onto the backbone of MILs can remarkably enhance the photoresponsivity and electron transfer efficiency of the material. The introduction of these ligands, which exhibit exceptional light absorption properties, was facilitated by a solvothermal synthesis, and the successful loading of the ligands was confirmed using characterization techniques, including Fourier transform infrared spectroscopy (FTIR) and X-ray diffraction (XRD).

4.2.1. Amino-Functionalized Ligand

In environmental remediation, amino-functionalized ligand-modified MIL series have substantially enhanced their capacity to degrade organic contaminants by introducing amino (-NH2) functional groups. The fundamental premise of this modification strategy is that the introduction of amino functional groups can modify the materials’ surface electron distribution and chemical properties, thereby enhancing their interactions with target molecules. Mo et al. examined phenol degradation by photocatalytic materials with amino-functionalized groups (NH2-MIL-101(Fe, Ce)) in their investigation. These materials exhibited an accelerated generation of multiple reactive radicals due to the presence of an amino-functionalized group (-NH2) as an electron donor. The subsequent interaction of these radicals resulted in the efficient degradation of phenol [18]. In a separate study, Mo et al. investigated the defect level of NH2-MIL-125(Ti) by systematically controlling the duration of the heat treatment. The 125(Ti) defect level was successfully managed by regulating the duration of the heat treatment process. This resulted in the gradual decarboxylation of ligands in NH2-MIL-125(Ti), occurring from the surface to the interior [104]. Consequently, this process led to the formation of ligand vacancy defects with varying concentrations. This, in turn, enhanced its photocatalytic performance, thereby facilitating the efficient photocatalytic reduction of Cr (VI).

4.2.2. Sulfonated-Functionalized Ligands

The sulfonic acid functional group, which has the chemical formula -SO3H, contains a sulfur atom at its core. This atom is attached to three oxygen atoms, forming a planar structure. One of the oxygen atoms is connected to a hydrogen atom via a single bond, forming the acidic -SO3H group. In a study by Nguyen et al., the epichlorohydrin conversion was 68.9% when using SO3H-MIL-101 as the catalyst alone. This conversion was further increased to 100% after functionalization with amine groups. This observation indicates that the sulfonic acid group possesses intrinsic catalytic activity and the capacity to potentiate the catalytic efficiency when combined with other functional groups, such as amine groups [105]. The functionalization of sulfonic acid groups has been demonstrated to enhance the water stability of the material. In the study by Zhao et al., a sulfonic acid group-functionalized metal–organic framework was utilized to remove iodide ions (I-) from water. The sulfonic acid group (-SO3H) was bonded to the MIL-101(Cr) framework through a strong covalent bond. The stability of this chemical bond ensured the material’s structural integrity in the aqueous environment and reduced its susceptibility to chemical degradation [106]. The functionalization of sulfonate and amino groups has been demonstrated to enhance the performance of photocatalysts through various chemical interaction mechanisms. The functionalization of the sulfonate group has been demonstrated to enhance the charge transfer kinetics and surface active sites, primarily through electron-withdrawal-induced effects. In contrast, amino group functionalization has been shown to improve the electron transfer efficiency and stability by providing electron donors and forming hydrogen bonds [107]. These functionalization approaches offer essential strategies for designing and developing efficient photocatalysts.

4.2.3. Other Group-Functionalized Ligands

In addition to amino (-NH2) and sulfonic (-SO3H) groups, the MIL series can be modified with various functionalized groups to modulate their physicochemical properties and expand their applications. Yu et al. synthesized the hydroxyl (-OH)-functionalized MIL-100 by selecting hydroxyl-rich 2,5-dihydroxyterephthalic acid (H4DODBC) as the ligand functionalized MIL-100. The experimental results demonstrated that NiCo-ZIF@MIL exhibited excellent stability in KOH solution, with nearly unchanged potentials after 1000 cycles, and the catalytic activity could be maintained for at least 24 h [108]. Anindita et al. prepared a catalyst (p-TSA-MIL-100@SBC) functionalized with toluene sulfonic acid (p-TSA). The purpose of introducing P-TSA was twofold: firstly, to enhance the structural stability of MIL-100(Fe) in an aqueous environment, thereby preventing its collapse; secondly, to inhibit the recombination of photogenerated electron–hole pairs through its sulfonic acid moiety, which acts as an electronic hub. This effectively inhibits the recombination of photogenerated electron–hole pairs, prolongs the lifetime of photogenerated carriers, and significantly enhances the photocatalytic activity [109].

4.3. Structural Heterogeneity

A heterojunction is an interfacial structure composed of two or more different materials (typically semiconductors) that differ in energy band structure, lattice constants, or electronic properties. This results in unique physicochemical properties at the interface. The categorization of these junctions can be approached through various criteria, including the type of conductivity (homojunctions and heterojunctions) [110,111], the energy band structure (type I, II, III) [112], the structural form (planar and perpendicular) [113,114], the material composition (semiconductor–semiconductor, metal–semiconductor, 2D materials, and nanomaterials) [115], and exceptional functionality (e.g., Z-scheme and S-scheme heterojunctions) [116,117]. These classifications reflect the diverse properties of heterojunctions in terms of the charge separation, light absorption, and electron transport, which have enabled them to demonstrate a wide range of applications in the fields of optoelectronic devices, semiconductor devices, sensors, and catalyst design [118]. The fundamental properties of these junctions are attributed to the electric potential barriers [119], energy band bending, quantum effects, and spectral properties present at the interfaces. These elements directly influence their performance in photovoltaic conversion, current control, and catalytic activity. In recent years, with the advent of two-dimensional materials and nanotechnology, there has been a remarkable surge in research endeavors concerning heterojunctions, particularly in the domain of novel structure design (e.g., S- and Z-scheme heterojunctions) and high-performance applications.

4.3.1. The Z-Scheme Heterojunctions’ Structure

Z-scheme heterojunctions comprise two or more semiconductor materials that, upon photoexcitation, facilitate the migration of electrons and holes to the valence band (VB) and conduction band (CB) of another semiconductor, thereby establishing a transmission path that resembles the shape of a Z. The design of this structure, which draws inspiration from the electron transport chain in plant photosynthesis, aims to emulate the efficient light energy conversion mechanism observed in nature [120]. This structure retains the high redox capacity of semiconductors, which allows preserving the strong redox potential by spatially separating electrons and holes, thus improving the separation efficiency of photogenerated carriers. Z-scheme heterojunction photocatalysts have been shown to be highly effective in the treatment of contaminants, such as antibiotic residues in water, with both high efficiency and environmental friendliness [116]. Li et al. prepared Z-scheme heterojunction materials (MIL-100(Fe)/TpPa-1 COF, MT2 for short) by in situ solvothermal synthesis (Figure 5). Before light irradiation, charge redistribution at the interface resulted in the formation of an interfacial electric field. In the presence of light, the photogenerated electrons in MIL-100(Fe) were transferred to the valence band of TpPa-1 COF due to the action of interfacial electric field, and the recombination with holes was suppressed, thereby achieving efficient charge separation. Furthermore, MT2 demonstrated its efficacy by removing 91% of TC (20 mg/L) within 120 min, thereby exhibiting its superior photocatalytic degradation performance [121].

4.3.2. The S-Scheme Heterojunctions’ Structure

The S-scheme heterojunction is a recently discovered semiconductor heterostructure that has demonstrated the capacity to enhance the performance of photocatalysis and other photovoltaic applications by optimizing the charge separation and transport pathways. In comparison with conventional type-II and Z-scheme heterojunctions, S-scheme heterojunctions have been demonstrated to enhance the photocatalytic efficiency by recombining electron–hole pairs with low reduction/oxidation capacity and retaining carriers with high reduction/oxidation capacity to convert organic contaminants into nontoxic inorganic substances [122]. Dong et al. employed a hydrothermal synthesis to synthesize MIL-53(Fe)/FeOCl composites. MIL-53(Fe) is an n-type semiconductor with a high degree of light absorption and a large number of active sites for iron oxide clusters. FeOCl is also an n-type semiconductor that exhibits excellent photovoltaic conversion properties and a suitable energy-band structure. The SEM and TEM results demonstrate that FeOCl is distributed uniformly as nano-sheets on the surface of MIL-53(Fe), forming a close contact heterojunction structure. XPS analysis reveals that the introduction of FeOCl results in alterations in the binding energies of Fe 2p and O 1s, confirming the formation of a heterojunction and electron transfer. The composites demonstrated remarkable photocatalytic degradation performance under visible light conditions, with a degradation efficiency of oxytetracycline (OTC) reaching 90%, which was significantly higher than that of MIL-53(Fe) alone (54%) [123].

4.3.3. Double S-Scheme Heterojunctions’ Structure

The double S-scheme heterojunction is a specialized structure composed of two S-scheme heterojunctions, typically involving three or more semiconductor materials [124]. The core feature of this structure is the efficient separation and transfer of photogenerated carriers (electrons and holes) by constructing two stepped energy band structures. In the double S-scheme heterojunction, the photogenerated electrons and holes are transferred between the conduction band (CB) and valence band (VB) of different materials, forming two independent charge transfer paths. This configuration enhances the separation efficiency of the photogenerated carriers and augments the redox capacity of the photocatalyst, thereby significantly improving the photocatalytic performance [125]. Patial et al. successfully synthesized Co3O4/MIL-88A/Mn-STO composites. The material consists of three semiconductors: MIL-88A and Mn-STO were prepared hydrothermally to form a single S-scheme heterojunction; then, Co3O4 was introduced to form a double S-scheme heterojunction. The XRD and FTIR results demonstrated that the components maintained their crystalline structure, and the diffraction peaks corresponded to those of the individual materials, indicating the successful construction of heterojunctions (Figure 6). In the photocatalytic degradation of sulfamethoxazole (SMX), the composite exhibited a degradation efficiency of 95.5% within 90 min, which was significantly higher than that of the material alone and the binary composite. The catalytic degradation rate constant (k) of the composite was determined to be 0.0337 min−1, which is notably higher than those of Co3O4 (8×), Mn-STO (5.2×), MIL-88A (4.6×), and MIL/Mn-STO (3.6×), respectively [126].

4.3.4. Stability and Recyclability

Through the aforementioned strategies of elemental doping, ligand functionalization, and structural heterogeneity, the adsorption capacity, photocatalytic efficiency, and selectivity of MIL series MOFs toward specific pollutants have been significantly enhanced. However, the ultimate goal of these modification strategies is not only to improve the single-cycle reaction efficiency but also to ensure that the materials maintain long-term structural stability and reusability under complex aqueous environments, which is crucial for practical engineering applications. Therefore, it is necessary to systematically evaluate the chemical stability and recyclability of the modified MIL series materials. In an electro-Fenton system, the D-MIL-53(Fe/Ce)/400@GF catalyst exhibited excellent cyclic stability, maintaining 90.52% of the pollutant removal efficiency after eight consecutive degradation cycles, while the leaching of Fe and Ce remained well below the internationally accepted safety thresholds, indicating outstanding chemical robustness (Figure 7a). This stability is primarily attributed to the introduction of framework defects and the optimization of electron transfer pathways, which effectively preserved the catalytic activity during repeated use. Similar improvements in stability have been observed for other modified MOF-based photocatalysts. For instance, the DNMB-5 photocatalyst exhibited only a slight decrease in degradation efficiency after four consecutive cycles, owing to defect engineering and the synergistic effect of oxygen vacancies, which not only suppressed active-site deactivation but also enhanced the separation efficiency of photogenerated charge carriers (Figure 7b). Likewise, the MTO heterojunction catalyst maintained high photocatalytic activity during 36 h of continuous degradation tests, a result ascribed to the construction of a Z-scheme heterostructure and the incorporation of conductive additives, which significantly promoted electron–hole separation and transfer (Figure 7c). Notably, the MIL-100(Fe)/TiO2/CoOx20 composite photocatalyst demonstrated excellent durability in photocatalytic oxygen evolution experiments; when the Co loading corresponded to 20 atomic layer deposition (ALD) cycles, the material achieved the highest oxygen evolution rate, and after 18 h of cycling, it still retained 91.3% of its activity with no significant morphological changes (Figure 7d). These results suggest that the synergistic combination of structural heterogeneity design and precise ALD control greatly enhances the cycling stability and structural integrity of the material.

5. The Utilization of MIL Materials in the Removal of Emerging Contaminants

The MIL series of metal–organic framework (MOF) materials remove emerging contaminants through a multi-level multi-mechanism synergistic process. Their highly porous three-dimensional framework provides an ultrahigh specific surface area and abundant active sites, allowing pharmaceuticals, microplastics, volatile organic compounds, and other organic pollutants to be strongly adsorbed via hydrogen bonding, π–π stacking, electrostatic interactions, hydrophobic interactions, and van der Waals forces within the pores and on the surface. This adsorption not only concentrates pollutants locally at the material surface but also positions them at catalytic reaction sites for subsequent degradation. Under light irradiation, the metal nodes (e.g., Fe3+, Cr3+) and organic ligands of MIL form photoactive centers, where photon energy excites electrons from the valence band to the conduction band, generating photogenerated electron–hole pairs (e/h+). The photogenerated electrons react with dissolved or surface oxygen to produce superoxide radicals (O2·), which can further generate hydrogen peroxide (H2O2) and subsequently hydroxyl radicals (·OH), while photogenerated holes oxidize water or hydroxide ions to generate additional ·OH; part of the energy can also generate singlet oxygen (1O2). These reactive oxygen species (ROS) attack chemical bonds such as C–C, C–N, C–O, and aromatic rings in pollutant molecules, breaking them down into smaller intermediates and ultimately mineralizing them to CO2, H2O, and nontoxic ions. The framework structure and the synergy between metal nodes and organic linkers promote electron–hole separation, suppress recombination, and enhance ROS generation, thereby significantly improving the photocatalytic efficiency. Moreover, the adsorption and photocatalytic processes of MIL can synergize with biodegradation or chemical oxidation, for example by activating H2O2 or ozone, or enhancing microbial degradation, further promoting pollutant removal. Specific pollutant interactions include pharmaceuticals, which are adsorbed through hydrogen bonding and π–π stacking and oxidatively degraded at aromatic rings and functional groups; microplastics, whose surfaces are adsorbed and oxidatively cleaved by ROS; and VOCs or dyes, which undergo photocatalytic oxidation or mineralization. In summary, MIL materials achieve rapid and complete pollutant degradation through a multi-level multi-mechanism process—comprising efficient adsorption, ROS-mediated photocatalysis, pollutant decomposition, and synergistic enhancement—making them highly effective for a broad range of emerging contaminants and providing a theoretical basis for designing high-performance multifunctional MOFs (see Table 3 and Figure 8).

5.1. Persistent Organic Pollutants (POPs)

The Stockholm Convention’s objective is to eliminate or limit the production, use, and release of Persistent Organic Pollutants (POPs) on a global scale [149]. Metal–organic frameworks (MOFs), specifically MIL materials, have emerged as a novel class of materials with the potential to address the treatment of POPs. MILs can function as catalysts, facilitating the conversion of POPs into less harmful or nontoxic substances via catalytic degradation processes. This approach not only mitigates the ecological and human health hazards posed by POPs but also fosters global collaborative efforts to combat POP pollution.

5.1.1. Pesticides

In recent years, pesticides and insecticides have played an important role in ensuring food security and controlling pests and diseases [150]. However, there is growing concern about their negative impact on the environment. The persistence, bioaccumulation, and biomagnification of pesticides in the environment enable their transmission through the food chain, thereby posing risks to human health [151]. In order to address this challenge, researchers are developing efficient catalysts for the removal of pesticide residues from the environment. These materials demonstrate not only the efficient adsorption of POPs, including organochlorine pesticides, but also the capacity to convert them into less harmful or nontoxic substances through the process of photocatalytic degradation [152]. For instance, Reda et al. successfully prepared composites Fe3O4@NH2-MIL-125 and CuO/Cu2O@NH2-MIL-125, which completely degraded 1.23 × 10−4 mol/L solutions of metanaphos and methylpyrimidinphos, respectively, within 150 min [153]. Vigneshwaran et al. synthesized the TiO2/CS/MIL-88(Fe) (TCS@MOF) composites via solvothermal synthesis. The TCS@MOF composites demonstrated remarkable photocatalytic performance, achieving 98.79% degradation efficiency of Monocrotophos (MCP) within a span of 30 min. Furthermore, the material exhibited notable durability (Figure 9), with a mere 4.1% decrease in photocatalytic efficiency after five cycles of experimentation [154]. However, the MIL is confronted with challenges, including high removal costs, limited regeneration capacity, and insufficient stability in complex environmental media. In practical applications, photocatalytic technology can be used synergistically with other water treatment technologies (e.g., adsorption, filtration, biological treatment, etc.) to improve the treatment efficiency and lower costs.

5.1.2. Industrial Chemicals

The harm caused by industrial chemicals to current ecosystems is multidimensional and far-reaching. The presence of industrial chemicals in the environment, resulting from various anthropogenic activities such as industrial effluents, agricultural runoff, and atmospheric deposition, has the potential to disrupt the ecological balance and contribute to biodiversity loss [155]. In light of the deleterious effects of industrial chemicals on ecosystems, the development of effective removal technologies, such as MIL, is crucial for mitigating these impacts. In industry, perfluorooctanoic acid (PFOA), an organic compound with unique chemical properties, is utilized in the synthesis of fluoropolymers, including polytetrafluoroethylene (PTFE), and functions as an emulsifier [156]. However, its environmental persistence has led to its classification as a new type of POP, and researchers are exploring more efficient techniques for PFOA removal. Su et al. achieved a degradation efficiency of 60.5% for PFOA, which was significantly higher than that of pure C3N4, through the use of MIL-100(Fe)@C3N4 via the synergistic effect of adsorption and photocatalysis [129]. This provides an effective technological means to deal with difficult-to-degrade organic contaminants [157]. Kong et al. prepared the material F- through the in situ growth of TiO2 and surface fluorine functionalization. The photocatalytic degradation of perfluorooctanoic acid (PFOA) was studied, and it was found that the reaction rate constant for TiO2@MIL-125 was 1.221 × 10−4/s [158]. This result indicates that the in situ growth of TiO2 and surface fluorine functionalization in MIL-125 is an effective method for achieving efficient photocatalysis. As demonstrated by Wen et al., NH2-MIL-125 exhibited remarkable efficacy in the photocatalytic degradation of PFOA, achieving a PFOA degradation rate of up to 98.9% within a 24 h period. This study also reported an overall defluorination rate of 66.7% [132].
In the primary process of the photocatalytic degradation of PFOA (Figure 10), the photocatalyst generates photogenerated holes (h+) and electrons (e) upon excitation by UV or visible light (EP.1). The photogenerated holes initially oxidize (C7F15COO) to an unstable PFOA radical (C7F15COO·) (Figure 10 (1)). This radical subsequently undergoes decarboxylation, resulting in the formation of a perfluoroalkyl radical (C7F15·) (Figure 10 (2)). C7F15· can undergo oxidation through two distinct pathways: pathway I, representing the hydroxyl radical oxidation route (Figure 10 (3)), produces a perfluorinated primary alcohol (C7F15OH); pathway II, denoting the oxygen-involved oxidation route (Figure 10 (4–7)), generates a similar intermediate. The subsequent decomposition of C7F15OH yields reactive acyl fluoride (C6F13COF) (Figure 10 (8)) and hydrolyses to short chain perfluorocarboxylic acids (e.g., PFHpA) (Figure 10 (9)). By repeating the aforementioned steps, PFOA is progressively degraded to shorter chain perfluorocarboxylic acids and eventually mineralized to CO2 and F [159,160].
M + hve + h+
While MIL materials demonstrate considerable promise in the degradation of POPs, certain photocatalytic materials exhibit deficiencies in terms of their stability in practical applications and are prone to deactivation by light, acidity, alkalinity, and other factors. Furthermore, the photocatalytic degradation process may yield toxic byproducts, which, if inadequately treated, can lead to secondary environmental contamination. In order to address these challenges, researchers have recently implemented improved measures. For instance, Mahmoud Babalar et al. designed a PES@PDA&β-CD@PMACZ-MOF composite membrane that has achieved efficient removal of polychlorinated biphenyls (PCBs) by combining multiple materials with a 100% retention rate and very low release rate (0.3804% to 0.8544%), effectively preventing the secondary release of contaminants [161].

5.2. Pharmaceuticals and Personal Care Products (PPCPs)

Pharmaceuticals and Personal Care Products (PPCPs) have emerged as an emerging class of environmental contaminants of great concern. Their widespread use and relatively stable chemical properties make them difficult to degrade rapidly in the environment, resulting in potential long-term impacts on ecosystems and human health [162]. PPCPs consist mainly of various types of pharmaceuticals, such as antibiotics, anti-inflammatory drugs, and painkillers. They enter water bodies, soils, and sediments through various pathways, such as domestic sewage, hospital wastewater, and agricultural runoff, and continue to accumulate in the environment [5]. Due to their high bioaccumulation and possible ecotoxicity, PPCPs not only persist in the environment but may also be transmitted through the food chain, posing a potential threat to ecosystems and human health [8].
In the context of emerging technologies for environmental governance, the relationship between MILs and PPCPs is evident not only in the degradation mechanism but also in the synergistic effect on environmental governance. The high photocatalytic capacity and catalytic activity inherent in MIL materials enable the precise degradation of specific functional groups in PPCPs, thereby significantly reducing the environmental mobility and bioavailability of these contaminants [163]. This property demonstrates the significant potential of MILs in mitigating the risks posed by PPCPs to ecosystems and human health. An example is sulfadiazine (SDZ), which is a sulfonamide antibiotic that is widely present in the aquatic environment. Nguyen et al. investigated Z-scheme-based Ti-MIL@g-C3N5 composite catalysts for the activation of peroxymonosulfate (PMS) for the removal of SDZ from wastewater under visible light. In the context of optimized conditions (100 mg/L catalyst and 0.4 mM PMS), the catalyst demonstrated a 95.6% degradation efficiency for 0.04 mM SDZ within a span of 15 min. Notably, the removal efficiency persisted beyond five cycles, retaining over 76.25% efficacy [134]. It has been determined that sulfonamides and tetracyclines are among the most prevalent antibiotic contaminants in the environment, exhibiting high levels of residual activity. In their study, Li et al. examined a novel Z-scheme heterojunction photocatalyst, designated as MIL-100(Fe)/TpPa-1 COF (abbreviated as MT2), for the photocatalysis of tetracycline (TC) in water. MT2 was utilized to photocatalyze 20 milligrams per liter of tetracycline (TC) in water over the course of 120 min. MT2 demonstrated a 91% degradation efficiency of TC molecules within 120 min, which is 1.26 times that of MIL-100(Fe) alone. The TC molecule generates intermediates through the pathways of hydroxylation, demethylation, ring-opening, and oxidation and is ultimately mineralized to CO2 and H2O. The genotoxicity of the degradation intermediates decreases with the progress of the reaction, as analyzed by the Toxicity Evaluation Software, version 5.1.2. (T.E.S.T.) [121].
In the primary process of the photocatalytic degradation of tetracycline (TC) (Figure 11), the TC molecule undergoes a series of chemical reactions, including demethylation, ring cleavage (decarbonylation), and deamination, resulting in the conversion of the molecule into two intermediates, TC 1 and TC 2 (Figure 11 (1) (2)). Subsequently, TC 2 is further decomposed by reactions of dehydroxylation and decarbonylation to form TC 3 and TC 4 (Figure 11 (3) (4)). The reaction continues, resulting in the formation of TC 5, TC 6, and TC 7 through cyclic esterification, ring-opening reactions, and demethylation steps (Figure 11 (5) (6) (7)). These intermediates undergo further degradation, ultimately resulting in the formation of the ring-opening structure of TC 8 (Figure 11 (8)). TC 8, a pivotal intermediate in the degradation chain, undergoes oxidation and decomposition, yielding H2O and carbon dioxide (CO2) (Figure 10 (9)), thereby completing the mineralization of tetracycline [164].
The aforementioned studies provide novel concepts for the development of efficient photocatalytic degradation of antibiotics and offer material design paradigms and process optimization strategies for scaled-up water treatment technologies. These strategies will promote the transition of photocatalytic technology from laboratory to engineering applications.

5.3. Volatile Organic Compounds (VOCs)

Volatile organic compounds (VOCs), including hydrocarbons, oxygenated organics, halogenated hydrocarbons, and nitrogen- and sulfur-containing organics, contribute to atmospheric photochemical reactions that generate ozone and PM2.5 and pose health risks through bioaccumulation and toxicity. Developing efficient and environmentally friendly VOC treatment technologies is therefore a key challenge. The MIL series of metal–organic frameworks (MOFs) has emerged as a promising solution due to their porous structure, high specific surface area, and tunable chemical composition, which facilitate both the efficient adsorption and catalytic degradation of VOC molecules. Numerous studies have demonstrated the superior adsorption performance of MIL series MOFs compared with conventional adsorbents. For example, MIL-101 exhibits significantly higher adsorption capacities for VOCs such as acetone, benzene, and toluene than zeolites or activated carbon, with enhanced selectivity toward molecular size and polarity [165], while MIL-101(Cr) achieves markedly greater toluene uptake than silica gel under low-concentration conditions, showing improved breakthrough and saturation behavior [166]. These results collectively highlight the substantial advantages of MIL series MOFs in both adsorption capacity and selectivity, underscoring their potential for efficient removal of gaseous and aqueous VOC pollutants.

5.3.1. Hydrocarbons

Hydrocarbon VOCs, which include alkanes, olefins, alkynes, and aromatic hydrocarbons—are present in industrial emissions, fuel combustion, and daily life activities. These compounds are important precursors of air pollution, photochemical smog, and fine particulate matter (PM2.5), due to their high volatility and chemical stability. Aromatic hydrocarbons, including benzene and toluene, are not only involved in photochemical reactions but also pose a threat to human health due to their toxicity [167]. In the study of the catalytic degradation of benzene, Zhao et al. prepared disc-shaped mesoporous N-TiO2 photocatalysts (MM-500) by high-temperature calcination of a mixture of MIL-125(Ti) and melamine. The photocatalytic degradation efficiency of MM-500 reached 99.1% after 480 min of visible light irradiation. Following the completion of ten cycles of experimentation, the degradation efficiency of benzene was found to be 84.8%, thereby demonstrating its stability and reusability. The findings indicated an impressive mineralization efficiency of 72.0%, suggesting that benzene could be efficiently decomposed into carbon dioxide and water [139]. This observation also reflected the effective separation of photogenerated carriers, further emphasizing the system’s functionality. As a benzene congener, toluene can also be catalytically degraded efficiently with the MIL series catalysts. Sun et al. investigated the degradation performance of TiO2@MIL-101(Cr)-composite photocatalysts for toluene by experimental analysis, which showed a degradation efficiency 23.28% higher than that of MIL-101(Cr) alone and 26.88% higher than that of TiO2 alone. During the degradation process, a significant increase in degradation efficiency was observed during the initial 40 min. Subsequently, a gradual decline in the growth rate was noted, suggesting that TiO2@MIL-101(Cr) predominantly underwent adsorption as the primary mechanism for toluene degradation in the early stages of the reaction. This was followed by a transition to photocatalysis as the dominant degradation pathway [136]. The MILs exhibited notable efficacy in the treatment of aromatic hydrocarbon contaminants, effectively adsorbing and degrading aromatic hydrocarbon VOCs, including benzene, toluene, xylene, and others [168]. However, relatively few studies have been conducted on non-aromatic VOCs, such as alkanes, olefins, and alkynes. This paucity of research is likely attributable to the strong chemical inertness of these compounds and the absence of π-π interactions, which enhance adsorption and degradation efficiency, as observed with aromatic hydrocarbons [169].

5.3.2. Oxygenated Organic Compounds (OVOCs)

Oxygenated Volatile Organic Compounds (OVOCs) are important components of VOCs, mainly including aldehydes (e.g., formaldehyde, acetaldehyde), ketones (e.g., acetone, methyl ethyl ketone), alcohols (e.g., ethanol, isopropanol), ethers (e.g., diethyl ether, ethylene glycol monomethyl ether), and acids (e.g., acetic acid). These compounds have low boiling points and high volatility and are widely present in the atmosphere, posing direct or indirect hazards to human health [170]. Formaldehyde, for example, is a known carcinogen, while substances such as acetone and acetaldehyde are highly irritating and toxic. Jiang et al. constructed Z-type heterojunction photocatalysts MIL-88A(Fe)-CA/BiVO4-OVs (MIBVOS-8) enriched with oxygen vacancies for formaldehyde degradation. In the context of simulated sunlight irradiation, formaldehyde degradation efficiency attained 91.0% within 120 min, and MIBVOS-8 demonstrated 86.3% degradation efficiency after five cycle experiments, indicating notable stability and reusability. DRIFTS analysis revealed that formaldehyde was adsorbed on the catalysts’ surface and converted to dioxymethylene (DOM) and formate intermediates, which were subsequently oxidized to CO2 and H2O [135]. For acetone, which is recognized as a class I carcinogen by the International Agency for Research on Cancer (IARC), Yang et al. designed a phosphomolybdic acid (PMA)@NH2-MIL-125 composite bridged by Ti-O-Mo bonds for acetone degradation. With respect to the degradation performance, the degradation efficiency of PMA@NH2-MIL-125 was 62% for 400 ppm acetone, which was higher than the 45% efficiency of pure NH2-MIL-125. Furthermore, the degradation efficiency for 25 ppm acetone increased from 64% to 78%. PMA@NH2-MIL-125 demonstrated a pseudo-primary reaction rate constant of 0.0426 min−1, which was higher than that of NH2-MIL-125 (0.0278 min−1). These results suggest that PMA loading significantly enhances the photocatalytic performance of the material [140].
Degradation mechanism studies have demonstrated that acetone (CH3COCH3) is initially oxidized to acetaldehyde (CH3CHO) (Figure 12 (1)), which can subsequently be further oxidized to yield acetic acid (CH3COOH) and formic acid (HCOOH) (Figure 12 (2) (3)). Acetic acid and formaldehyde (HCHO) are also oxidized to formic acid (Figure 12 (4) (5)), and eventually formic acid is mineralized to carbon dioxide (CO2) and water (H2O) (Figure 12 (5) (6)). Throughout the process, the photocatalyst generates hydroxyl radicals (·OH) and peroxyl radicals (·O2−) in the presence of light (Figure 12). These reactive species drive the successive oxidation reactions of organic matter [171].
While MILs demonstrate considerable potential in their capacity to degrade OVOCs, it is important to note that they are also subject to certain limitations. In environments characterized by elevated levels of oxygenated OVOCs, particularly in the context of treating substantial concentrations of industrial exhaust gases, the adsorption sites of MILs may be rapidly saturated. This phenomenon has been observed to result in a substantial decline in the degradation efficiency of these materials. To address this challenge, the specific surface area and porosity of MIL materials can be enhanced through the optimization of synthesis or the incorporation of additional materials, such as activated carbon or carbon nanotubes [172,173]. These modifications provide additional reaction sites, thereby promoting the degradation of OVOCs.

5.3.3. Chlorinated Volatile Organic Compounds (CVOCs)

Chlorinated Volatile Organic Compounds (CVOCs) consist primarily of chlorinated alkanes, chlorinated olefins, and aromatic hydrocarbons. Due to their high chemical stability, toxicity, and resistance to natural degradation, these compounds have wide application in industrial contexts, including pharmaceuticals, chemicals, pesticides, and solvent production. However, it is important to note that many chlorinated alkanes (e.g., 1,2-dichloroethane) have been identified as carcinogenic, and prolonged exposure to these substances may result in an elevated risk of developing cancer. The substance has been demonstrated to be acutely toxic to both the central nervous system and the respiratory system. Potential symptoms associated with its use include but are not limited to headaches, fatigue, and respiratory irritation [174].
Furthermore, during catalytic degradation, the incomplete degradation of CVOCs has been observed to result in the generation of more toxic byproducts, such as dioxins, further exacerbating the environmental risks associated with CVOCs [175]. In recent years, the MIL series of MOFs materials has demonstrated noteworthy efficacy in the adsorption and degradation of chlorinated volatile organic compounds (CVOCs). Zhu et al. prepared praseodymium (Pr)-doped Cr2O3 catalysts (M-PrCr) by in situ pyrolysis of MIL-101(Cr) and applied them to the efficient catalytic oxidative degradation of 1,2-dichloroethane (1,2-DCE). The 1,2-DCE was subjected to complete oxidation by M-PrCr, forming environmentally benign CO2 and HCl. The M-PrCr catalyst demonstrated satisfactory overall durability, notwithstanding a decline in activity attributable to water vapor. The PrCr catalyst demonstrated satisfactory overall durability following 100 h of operation at 290 °C, with a conversion of 1,2-DCE that remained above 70%, despite a decline in activity attributable to the impact of water vapor [137]. In Liu et al., carbon nanolayer-encapsulated TiO2 nanocomposites were prepared by the precise regulation of the pyrolysis process of NH2-MIL-125(Ti) composites (CNWT-x). These composites were subsequently utilized in a photocatalytic dichloromethane (DCM) degradation process under UV–Vis irradiation. Among the samples examined, those designated as CNWT-2, following a two-hour exposure to air, exhibited remarkable photocatalytic performance. CNWT-2 exhibited 85% DCM conversion and 90% CO2 selectivity under 5 h of UV–visible irradiation. Furthermore, DRIFTS analysis revealed that no polychlorinated byproducts (e.g., CHCl2 and CCl4) were detected during the degradation process, suggesting that CNWT-2 effectively prevented the generation of hazardous byproducts and ensured that the degradation process was environmentally friendly [141].
MILs are confronted with the challenge of chlorine toxicity in the degradation of CVOCs. This challenge is characterized by the strong adsorption of chlorine species, the formation of chloride compounds, the inhibition of reactive oxygen species, the corrosion of the catalyst, and the accumulation of chlorine species. The combined effect of these factors is the deactivation of the catalyst’s active sites. In order to enhance the resistance of catalysts to chlorine poisoning, researchers have proposed various strategies, including the construction of a protective layer [176], the promotion of chlorine species desorption [174], the enhancement of catalyst stability [177], the introduction of sacrificial sites, and acidification treatment [178]. These strategies have been demonstrated to offer effective protection for the catalyst’s active sites, thereby enhancing its stability and catalytic efficiency.

5.3.4. Sulfur-Containing Volatile Organic Compounds (S-VOCs)

Sulfur-Containing Volatile Organic Compounds (S-VOCs) from petroleum have been identified as a primary contributor to air pollution. These compounds are constantly escaping into the environment during the extraction, processing, storage, and transportation of petroleum [179]. Due to their remarkable volatility and chemical activity, they are capable of rapidly spreading in the atmosphere and becoming involved in a series of complex chemical reactions. These reactions have been demonstrated to directly degrade air quality and potentially generate secondary contaminants, including sulfate aerosols and ozone, under the influence of sunlight. Additionally, these reactions have been observed to exhibit synergistic effects with other air contaminants [9]. These secondary contaminants have been identified as contributing factors to environmental issues such as haze and acid rain, which have been demonstrated to exert deleterious effects on human health, ecological systems, and infrastructure [180]. During the atmospheric distillation and decompression distillation of crude oil, sulfur-containing compounds in crude oil are volatilized with the steam [179]. Methyl mercaptan, a low-boiling point mercaptan, is readily released into the gas phase during distillation. He et al. investigated copper-doped Cu/NH2-MIL-125 photocatalytic materials for the efficient degradation of methyl mercaptan under sunlight. It has been demonstrated that under the condition of visible light (λ > 420 nm), the degradation rate constant of 0.2Cu/80% NH2-MIL-125 (300 °C) was found to be 0.264 min−1, which is 4.65 times higher than that of pure NH2-MIL-125 (Figure 13a–c). NH2-MIL-125 exhibited consistent and efficient methyl mercaptan removal over the course of five cycles (Figure 13d), demonstrating its stability and effective degradation [142]. This result indicates an efficient degradation process. This property underscores its considerable potential for practical applications. Thiophene, a prevalent constituent in petroleum refining, is a notable pollutant that must be addressed and eliminated during petroleum processing. In the study by McNamara et al., TBHP was utilized as an oxidizing agent, and MIL-125 was employed as a catalyst for thiophene degradation. Furthermore, MIL-125 demonstrated notable catalytic activity at 80 °C, exhibiting a rate constant of 0.7 × 10−3 min−1 for the thiophene degradation reaction. Subsequent to the reaction, thiophene underwent a transformation to the more polar thiophene sulfoxide and thiophene sulfone, thereby enabling its separation from the reaction system and facilitating the effective removal of sulfur-containing compounds [181].
The utilization of MILs in the fields of adsorption and catalysis is promising due to their high specific surface area and porous structure. However, concerns regarding their stability in practical applications must be addressed to fully realize their potential. In complex environments, such as those characterized by high humidity and high concentrations of S-VOC, the structures of these enzymes are prone to collapse. Active sites become easily inactivated, and the catalytic efficiency is reduced. Additionally, organic ligands exhibit sensitivity to strong acids, strong bases, or high concentrations of organics, which further exacerbate the stability problems. The structural stability of MILs can be enhanced by post-synthesis treatment or the introduction of more stable ligands to improve their adaptability and service life in complex environments.

5.4. Microplastics

As an emerging global pollutant, microplastics have been found in various environments, including oceans, rivers, soils, and the atmosphere [182]. Their sources are highly diverse, including, but not limited to, the fragmentation of large plastic wastes in the natural environment, the leakage of plastic particles during industrial production, the addition of microbeads to personal care products (e.g., cosmetics, toothpaste, etc.), and the abrasion of synthetic fibers during use [183]. The classification of microplastics is determined by their size, with particles ranging from the micrometer scale (1 µm to 5 mm) to the nanometer scale (less than 1 µm) [184]. In recent years, the MIL series of metal–organic framework materials (MOFs) has demonstrated considerable potential in microplastic pollution control. This is due to their distinctive porous structure and photocatalytic properties. MIL materials exhibit not only the capacity to adsorb microplastic particles with efficiency but also the capability to expedite the oxidative degradation process of microplastics through the effective separation of electron–hole pairs and the generation of free radicals under light conditions. Consequently, MILs facilitate the efficient degradation and removal of microplastics.

5.4.1. Micrometer Scale

The term “microplastics” (MPs) is commonly used to refer primarily to micron-sized particles, with particle sizes ranging from 1 micron to 5 mm [184]. This variety of microplastics is the most prevalent and extensively studied type in the contemporary environment. Due to its comparatively substantial particle size, the migration and diffusion of MPs in the environment is relatively straightforward to observe and investigate. Nevertheless, they represent a potential hazard to the ecosystem and human health [185]. The utilization of MIL adsorption and the degradation of MPs possesses a wide range of possible applications, garnering the interest and investment of numerous researchers. Yang et al. synthesized a titanium-based metal–organic framework material (MX@MIL-125(Ti)) on the surface of Ti3C2 MXene utilizing a hydrothermal synthesis. The study’s findings demonstrated that the composite material exhibited the capacity to attain a removal efficiency that exceeded 75% for polyethylene (PE) within 4 h under conditions of ultrasonic radiation [147]. However, achieving this level of efficiency necessitates elevated ultrasonic power (650 W), an extended action time (4 h), and a heightened catalyst concentration (2.0 g/L). These exacting conditions result in a substantial escalation in energy consumption and treatment costs, thereby constraining its feasibility and economic viability in large-scale water treatment applications. Feng et al. prepared heterostructured photocatalysts via the hydrothermal synthesis, namely NH2-MIL-88B(Fe)/MoS2, which exhibited highly efficient photocatalytic degradation of HDPE under simulated sunlight irradiation. This process also resulted in a significant change in the surface morphology and chemical structure of HDPE (Figure 14) [144].
Gu et al. prepared BiOI/MIL-101 composites for the photocatalytic degradation of PE using hydrothermal and co-precipitation synthesis. The carbonyl index (CI) of the composite exhibited an increase of 0.127 after six hours of exposure to light conditions, which is 5.3 and 3.7 times higher than that of BiOI and MIL-101 alone, respectively. The efficient degradation was primarily attributed to the photo-Fenton reaction, whereby Fe (II) catalyzes the decomposition of H2O2 to generate strongly oxidizing hydroxyl radicals (·OH) within the composite [146]. In the photocatalytic degradation of PE (Figure 15), the hydroxyl radical initially oxidizes the C-C bond of PE, thereby initiating the degradation process (EP.2). In the presence of oxygen, the generated C-C bond is rapidly converted to peroxyl radicals, which further react with other polymer chains to form hydrogen peroxide (EP.3,4). The decomposition of hydrogen peroxide under light exposure results in the formation of alkoxy radicals and hydroxyl radicals, a process facilitated by the weak O-O bonds present in the compound (EP.5). Subsequently, the alkoxy radicals promote the formation of polymers containing ketone and aldehyde groups by reacting with oxygen and undergoing β-cleavage (EP.6). Under UV irradiation, these ketone and aldehyde groups undergo α-cleavage, primarily through Norrish-type reactions, which generate carbon monoxide and result in mass loss (EP.7,8). This process is identified as the rate-determining step in the degradation process. However, the high efficiency of the Fenton reaction is contingent upon acidic conditions and elevated H2O2 concentrations, which can increase the operational complexity and cost in practical applications. Furthermore, the iron ions produced by the Fenton reaction have the potential to induce secondary pollution in water bodies, necessitating additional treatment.
Inorganics 13 00324 i006
Inorganics 13 00324 i007
Inorganics 13 00324 i008
Inorganics 13 00324 i009
Inorganics 13 00324 i010
Inorganics 13 00324 i011
Inorganics 13 00324 i012

5.4.2. Nanoscale

The unique physicochemical properties and environmental behavior of nanoplastics (NPs) result from their minute size, which is less than 1 micron. These particles demonstrate augmented mobility within the environment and possess the capacity to penetrate cell membranes, gain entry into organisms, and accumulate within these hosts, which can manifest more intricate health hazards [186]. The elevated specific surface area of NPs enables their capacity to adsorb a greater quantity of contaminants and heavy metals, thereby amplifying their toxicity [187]. To effectively manage NP pollution, MILs have also demonstrated significant potential in the photocatalytic degradation of NPs. This process involves the activation of peroxymonosulfate (PMS) under light conditions, which results in the generation of highly reactive free radicals (·SO42−) capable of efficiently degrading NPs. Sharmin et al. have developed a novel catalytic membrane that combines a mixed metal oxide (CoFeLMO) with MIL-100 (Fe) and polyethylene glycol (PEG) to promote the PMS-activated advanced oxidation process, to degrade NPs efficiently. The removal efficiency of NPs was 98.5% when the PMS concentration was increased to 0.2 mM [133]. This result indicated that the LMO-MOF-PEG catalytic membrane exhibited high efficiency in removing NPs, even at lower PMS concentrations, demonstrating its remarkable efficacy in addressing NPs contamination. However, the regeneration of these MILs after adsorption saturation is challenging; their reuse is limited, and their reuse may lead to clogging, due to membrane contamination, increasing the operating cost. To enhance their viability for practical water treatment applications, further optimization of their performance and cost reduction are necessary.

6. Summary

This paper provides a comprehensive overview of the current state of MIL series MOFs materials’ application in emerging contaminants’ management. A thorough examination of MIL series MOFs materials offers a specific perspective on future research directions. A comprehensive compendium of MIL series classification, synthesis methods, and modification strategies, along with their application effects in novel pollutant management was presented. This compendium also revealed the advantages and limitations of the MIL series in environmental remediation. The findings of this study offer a multifaceted contribution to the field. On the one hand, they provide a robust theoretical foundation and practical guidance for managing emerging contaminants. On the other hand, they serve as a valuable reference point and a source of inspiration for future research. In the following section, we outline four key areas that we believe will be the focus of future research:
Firstly, the optimization of the performance of the MIL series is pivotal to achieving its wide application. In the contemporary context, the catalytic properties of MILs can be substantially enhanced by implementing elemental doping, ligand functionalization, and heterostructure design. For instance, transition metal doping can introduce new active sites, thereby improving the catalytic performance of the materials. In addition, ligand functionalization can modulate the surface properties of the materials by introducing specific functional groups (e.g., amino and sulfonic acid groups) to improve their adsorption capacity for particular contaminants. Future research should explore novel modification methods, such as multi-element synergistic doping and multi-ligand functionalization, for more efficient pollutant degradation. In addition, the construction of heterojunction structures has been demonstrated to enhance the separation efficiency of photogenerated carriers, thereby improving the photocatalytic performance.
While various synthesis methods have been employed to enhance the MIL series, it is noteworthy that the materials synthesized by different methods exhibit substantial disparities in performance. For instance, the solvothermal synthesis is characterized by its simplicity and low energy consumption; however, it is susceptible to impurity introduction. In contrast, microwave-assisted synthesis significantly reduces the reaction time, yet its equipment costs are substantial. Future research endeavors should prioritize the development of greener, more efficient, and scalable synthesis techniques to reduce production costs and enhance the material reproducibility and consistency. For instance, the crystallinity and purity of MILs can be further enhanced by optimizing the reaction parameters of the microwave-assisted synthesis. Furthermore, integrating synthesis methods, such as mechanical grinding and solvothermal techniques, is anticipated to improve material properties while reducing costs.
Secondly, a cost–benefit comparison of the MIL series with other established technologies (e.g., activated carbon adsorption, bioremediation, etc.) in treating emerging contaminants is warranted. Despite the notable advantages of MILs in adsorption efficiency and selectivity, their economic viability in practical applications is hindered by their relatively high synthesis costs. For instance, a conventional adsorbent, activated carbon, possesses a comparatively modest adsorption capacity; however, its cost-effectiveness and ease of regeneration make it the preferred choice in numerous application scenarios. Conversely, the synthesis of the MIL series necessitates the utilization of costly organic ligands and metal salts, and the energy consumption and solvent usage during the synthesis process contribute to the overall cost. Nevertheless, MIL materials generally exhibit superior overall performance compared to other adsorption materials, including higher stability, adsorption capacity, and recyclability, which can justify their use in targeted applications. A critical factor in enhancing the economic viability of MIL series synthesis is the reduction in the price of raw materials. A significant body of research is currently exploring inexpensive industrial waste as a source of ligands. For instance, a recent study successfully synthesized MIL-101 using waste plastics as a ligand source, substantially reducing the synthesis costs. Additionally, using more economical metal salts, such as iron salts, can contribute to the cost reduction.
Thirdly, the majority of research in the MIL series has centered on laboratory simulation experiments, and its application in real environments continues to face significant challenges. For instance, a limited array of contaminants in actual environments has been demonstrated to impede the efficacy of detection and removal processes. To address these challenges, future research should prioritize the translation of MILs from laboratory settings to real-world applications. This transition necessitates the development of efficient detection techniques and large-scale treatment processes. The development of composite materials and multi-technology coupling processes holds great potential for enhancing the effectiveness of MILs in real-world applications by facilitating a synergistic treatment of multiple contaminants. Achieving enhanced environmental adaptability in MILs is imperative for their practical applications. For instance, in conditions of high humidity, high salinity, and strong acid or alkali, the structure and performance of MILs may be significantly impacted. Future research endeavors should prioritize enhancing material stability in complex environments through optimizing synthesis and modification strategies. For instance, the water and salt resistance of MILs can be substantially improved by incorporating hydrophobic ligands and constructing heterojunction structures. Furthermore, developing efficient regeneration techniques, such as heat treatment, chemical cleaning, and ultrasonic treatment, is crucial for restoring the activity of the materials and enhancing their reuse.
Fourthly, the accurate detection of degradation products is paramount for evaluating the environmental remediation efficacy of MILs. Presently, the detection of degradation products is chiefly accomplished through techniques such as gas chromatography–mass spectrometry (GC-MS) and high-performance liquid chromatography (HPLC). However, these techniques have limitations in detecting low concentrations and complex degradation products. To this end, future research should prioritize the development of more efficient detection means, such as high-resolution mass spectrometry (HRMS) and nanosensors, to achieve accurate detection of the degradation products. Furthermore, enhancement of microscopic characterization tools is imperative for comprehending the degradation mechanism of MIL materials. Currently, techniques such as X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectroscopy (FTIR), and Raman spectroscopy are widely used to analyze the surface and structure of materials.
Nevertheless, these techniques are still inadequate for resolving complex reaction mechanisms. Future research endeavors should prioritize the in-depth elucidation of the degradation mechanism of MILs through the integration of in situ characterization techniques and computational simulations. For instance, in situ XPS and FTIR techniques enable real-time observation of surface alterations and the breaking and formation of chemical bonds within the materials during the reaction process.
Furthermore, developing MILs with multifunctional capabilities and integrating adsorption, catalysis, and photocatalysis into a unified structure holds promise for the synergistic management of multiple contaminants. This, in turn, would enhance the overall performance of MILs in practical applications. By pursuing research in the four directions above, it is anticipated that significant advancements will be made in applying MILs to treat emerging contaminants. These advancements will facilitate the widespread implementation of these materials from the laboratory to the field of environmental remediation. Consequently, this will provide substantial support for the sustainable development of the global environment.

Funding

This research was funded by National Natural Science Foundation of China (42177359), National Key Research and Development Program of China (2023YFC3707800, 2023YFC3706700), the Open Fund of National Engineering Laboratory for Site Remediation Technologies (NEL-SRT201907), and the Research and Demonstration of Key Technologies for Restoration of Water Ecological Service Function in Inner Mongolia Section of Yellow River Basin (2023CXPT005).

Conflicts of Interest

Author Yusheng Jiang was employed by the company Tangshan Jidong Cement Co., Ltd. Authors Shanqing Wang and Zhongshan Qian were employed by the company Tangshan Beijing Shouke Xingye Engineering Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Shah, S.A.R.; Abbas, N.; Serbanescu, L.; Niu, R.; Nassani, A.A. The key challenges and best alternatives to environmental sustainability: A comprehensive study. Sci. Rep. 2025, 15, 7042. [Google Scholar] [CrossRef]
  2. Zhang, X.; Yao, Z.; Yang, W.; Zhang, W.; Liu, Y.; Wang, Z.; Li, W. Distribution, sources, partition behavior and risk assessment of polycyclic aromatic hydrocarbons (PAHs) in the waters and sediments of Lake Ulansuhai, China. Mar. Pollut. Bull. 2024, 200, 116072. [Google Scholar] [CrossRef] [PubMed]
  3. Fan, W.; Shen, X.; Zhu, Z.; Liu, X.; Zhang, H.; Qiu, Y.; Yin, D. Rsearch progress on biochar materials for new pollutants removal in the aquatic environment: A mini-review. Front. Environ. Sci. Eng. 2024, 19, 33. [Google Scholar] [CrossRef]
  4. Wang, H.; Gao, X.; Zuo, Y. Research and Application of Water Treatment Technologies for Emerging Contaminants (ECs): A Pathway to Solving Water Environment Challenges. Water 2024, 16. [Google Scholar] [CrossRef]
  5. Guillotin, S.; Delcourt, N. Studying the Impact of Persistent Organic Pollutants Exposure on Human Health by Proteomic Analysis: A Systematic Review. Int. J. Mol. Sci. 2022, 23, 14271. [Google Scholar] [CrossRef]
  6. Rauseo, J.; Spataro, F.; Patrolecco, L. Persistent and Emerging Organic Contaminants in Natural Environments. Water 2025, 17, 436. [Google Scholar] [CrossRef]
  7. Li, X.; Shen, X.; Jiang, W.; Xi, Y.; Li, S. Comprehensive review of emerging contaminants: Detection technologies, environmental impact, and management strategies. Ecotoxicol. Environ. Saf. 2024, 278, 116420. [Google Scholar] [CrossRef]
  8. Cheng, M.; Li, M.; Zhang, Y.; Gu, X.; Gao, W.; Zhang, S.; Liu, J. Exploring the mechanism of PPCPs on human metabolic diseases based on network toxicology and molecular docking. Environ. Int. 2025, 196, 109324. [Google Scholar] [CrossRef]
  9. David, E.; Niculescu, V.C. Volatile Organic Compounds (VOCs) as Environmental Pollutants: Occurrence and Mitigation Using Nanomaterials. Int. J. Environ. Res. Public Health 2021, 18, 13147. [Google Scholar] [CrossRef]
  10. Liu, J.; Zheng, L. Microplastic migration and transformation pathways and exposure health risks. Environ. Pollut. 2025, 368, 125700. [Google Scholar] [CrossRef]
  11. Lv, M.; Zhang, T.; Ya, H.; Xing, Y.; Wang, X.; Jiang, B. Effects of heavy metals on the adsorption of ciprofloxacin on polyethylene microplastics: Mechanism and toxicity evaluation. Chemosphere 2023, 315, 137745. [Google Scholar] [CrossRef] [PubMed]
  12. Luo, Z.; Yao, B.; Yang, X.; Wang, L.; Xu, Z.; Yan, X.; Tian, L.; Zhou, H.; Zhou, Y. Novel insights into the adsorption of organic contaminants by biochar: A review. Chemosphere 2022, 287, 132113. [Google Scholar] [CrossRef] [PubMed]
  13. Priyadarshini, M.; Das, I.; Ghangrekar, M.M.; Blaney, L. Advanced oxidation processes: Performance, advantages, and scale-up of emerging technologies. J. Environ. Manag. 2022, 316, 115295. [Google Scholar] [CrossRef]
  14. Yang, S.; Li, X.; Zeng, G.; Cheng, M.; Huang, D.; Liu, Y.; Zhou, C.; Xiong, W.; Yang, Y.; Wang, W.; et al. Materials Institute Lavoisier (MIL) based materials for photocatalytic applications. Coord. Chem. Rev. 2021, 438, 213874. [Google Scholar] [CrossRef]
  15. Mourdikoudis, S.; Dutta, S.; Kamal, S.; Gómez-Graña, S.; Pastoriza-Santos, I.; Wuttke, S.; Polavarapu, L. State-of-the-Art, Insights, and Perspectives for MOFs-Nanocomposites and MOF-Derived (Nano) Materials. Adv. Mater. 2025, 2415399. [Google Scholar] [CrossRef]
  16. Tan, K.L.; Foo, K.Y. Preparation of MIL-100 via a novel water-based heatless synthesis technique for the effective remediation of phenoxyacetic acid-based pesticide. J. Environ. Chem. Eng. 2021, 9, 104923. [Google Scholar] [CrossRef]
  17. Pouyanfar, N.; Farnam, G.; Ahmadi, M.; Masoudifar, R.; Banan, K.; Asadian, E.; Shahhosseini, S.; Shahbazi, M.-A.; Shirazi, F.H.; Ghorbani-Bidkorpeh, F. Synthesis, modification, characterization, and in vitro evaluation of chitosan-hyaluronic acid coated MIL-100 (Fe) nanoparticles for methotrexate delivery in rheumatoid arthritis. Int. J. Biol. Macromol. 2024, 283, 137715. [Google Scholar] [CrossRef]
  18. Mo, L.; Chen, G.; Xu, B. Degradation of phenol by peroxymonosulfate catalyzed by cerium-doped amino-functionalized metal-organic frameworks (NH2-MIL-101 (Fe, Ce)). J. Environ. Chem. Eng. 2024, 12, 113256. [Google Scholar] [CrossRef]
  19. Zhang, H.; Hu, X.; Li, T.; Zhang, Y.; Xu, H.; Sun, Y.; Gu, X.; Gu, C.; Luo, J.; Gao, B. MIL series of metal organic frameworks (MOFs) as novel adsorbents for heavy metals in water: A review. J. Hazard. Mater. 2022, 429, 128271. [Google Scholar] [CrossRef]
  20. Khan, S.; Guan, Q.; Liu, Q.; Qin, Z.; Rasheed, B.; Liang, X.; Yang, X. Synthesis, modifications and applications of MILs Metal-organic frameworks for environmental remediation: The cutting-edge review. Sci. Total Environ. 2022, 810, 152279. [Google Scholar] [CrossRef]
  21. Salvador, F.E.; Tegudeer, Z.; Locke, H.; Gao, W.-Y. Facile mechanochemical synthesis of MIL-53 and its isoreticular analogues with a glance at reaction reversibility. Dalton Trans. 2024, 53, 4406–4411. [Google Scholar] [CrossRef] [PubMed]
  22. Smoljan, C.S.; Sha, F.; Campitelli, P.; Xie, H.; Eddaoudi, M.A.; Mian, M.R.; Di Nicola, C.; Kirlikovali, K.O.; Snurr, R.Q.; Farha, O.K. Constraining Flexibility in the MIL-88 Topology through Integration of 3-Dimensional Linkers. Cryst. Growth Des. 2024, 24, 3941–3948. [Google Scholar] [CrossRef]
  23. Quijia, C.R.; Lima, C.; Silva, C.; Alves, R.C.; Frem, R.; Chorilli, M. Application of MIL-100(Fe) in drug delivery and biomedicine. J. Drug Deliv. Sci. Technol. 2021, 61, 102217. [Google Scholar] [CrossRef]
  24. Pederneira, N.; Aina, P.O.; Rownaghi, A.A.; Rezaei, F. Performance of MIL-101(Cr) and MIL-101(Cr)-Pore Expanded as Drug Carriers for Ibuprofen and 5-Fluorouracil Delivery. ACS Appl. Bio Mater. 2024, 7, 1041–1051. [Google Scholar] [CrossRef]
  25. Omrani, A.; Deliballi, Z.; Kaya, K.; Kiskan, B.; Akgun, M. Catalytic Role of Nanoconfinement inside MIL-125 (Ti) on the Ring-Opening Polymerization of Simple Benzoxazines. ACS Appl. Polym. Mater. 2023, 6, 253–264. [Google Scholar] [CrossRef]
  26. Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.; Férey, G. Very Large Breathing Effect in the First Nanoporous Chromium(III)-Based Solids: MIL-53 or CrIII(OH)·{O2C−C6H4−CO2}·{HO2C−C6H4−CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124, 13519–13526. [Google Scholar] [CrossRef]
  27. Millange, F.; Walton, R.I. MIL-53 and its Isoreticular Analogues: A Review of the Chemistry and Structure of a Prototypical Flexible Metal-Organic Framework. Isr. J. Chem. 2018, 58, 1019–1035. [Google Scholar] [CrossRef]
  28. Tomar, S.; Singh, V.K. Review on synthesis and application of MIL-53. Mater. Today Proc. 2021, 43, 3291–3296. [Google Scholar] [CrossRef]
  29. Xiao, K.; Shu, B.; Lv, K.; Huang, P.; Chang, Q.; Wu, L.; Wang, S.; Cao, L. Recent Progress of MIL MOF Materials in Degradation of Organic Pollutants by Fenton Reaction. Catalysts 2023, 13, 734. [Google Scholar] [CrossRef]
  30. Serre, C.; Mellot-Draznieks, C.; Surblé, S.; Audebrand, N.; Filinchuk, Y.; Férey, G. Role of Solvent-Host Interactions That Lead to Very Large Swelling of Hybrid Frameworks. Science 2007, 315, 1828–1831. [Google Scholar] [CrossRef]
  31. Serre, C.; Millange, F.; Surblé, S.; Férey, G. A Route to the Synthesis of Trivalent Transition-Metal Porous Carboxylates with Trimeric Secondary Building Units. Angew. Chem. Int. Ed. 2004, 43, 6285–6289. [Google Scholar] [CrossRef]
  32. Zhong, G.; Liu, D.; Zhang, J. Applications of Porous Metal–Organic Framework MIL-100(M) (M = Cr, Fe, Sc, Al, V). Cryst. Growth Des. 2018, 18, 7730–7744. [Google Scholar] [CrossRef]
  33. Patricia, H.; Suzy, S.; Christian, S.; Do-Young, H.; You-Kyong, S.; Jong-San, C.; Jean-Marc, G.; Irene, M.; Gérard, F. Synthesis and catalytic properties of MIL-100(Fe), an iron(III) carboxylate with large pores. Chem. Commun. 2007, 2820–2822. [Google Scholar] [CrossRef]
  34. Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  35. Zou, M.; Dong, M.; Zhao, T. Advances in Metal-Organic Frameworks MIL-101(Cr). Int. J. Mol. Sci. 2022, 23, 9396. [Google Scholar] [CrossRef] [PubMed]
  36. Bhattacharjee, S.; Chen, C.; Ahn, W.-S. Chromium terephthalate metal–organic framework MIL-101: Synthesis, functionalization, and applications for adsorption and catalysis. RSC Adv. 2014, 4, 52500–52525. [Google Scholar] [CrossRef]
  37. Dan-Hardi, M.; Serre, C.; Frot, T.; Rozes, L.; Maurin, G.; Sanchez, C.; Férey, G. A New Photoactive Crystalline Highly Porous Titanium(IV) Dicarboxylate. J. Am. Chem. Soc. 2009, 131, 10857–10859. [Google Scholar] [CrossRef]
  38. Ockwig, N.W.; Delgado-Friedrichs, O.; O’Keeffe, M.; Yaghi, O.M. Reticular Chemistry:  Occurrence and Taxonomy of Nets and Grammar for the Design of Frameworks. Acc. Chem. Res. 2005, 38, 176–182. [Google Scholar] [CrossRef]
  39. Yan, Y.; Li, C.; Wu, Y.; Gao, J.; Zhang, Q. From isolated Ti-oxo clusters to infinite Ti-oxo chains and sheets: Recent advances in photoactive Ti-based MOFs. J. Mater. Chem. A 2020, 8, 15245–15270. [Google Scholar] [CrossRef]
  40. Yang, S.; Yu, J.; Lu, G.; Song, G.; Shi, G.; Wang, Y.; Xie, X.; Yuan, H.; Ren, X.; Sun, J. Effect of NH2-functionalization of MIL-125 on photocatalytic degradation of o-xylene and acetaldehyde. Chem. Eng. J. 2024, 498, 155251. [Google Scholar] [CrossRef]
  41. Fallah, M.; Sohrabnezhad, S. Study of synthesis of mordenite zeolite/MIL-101 (Cr) metal–organic framework compounds with various methods as bi-functional adsorbent. Adv. Powder Technol. 2019, 30, 336–346. [Google Scholar] [CrossRef]
  42. Luo, T.; Jeppesen, H.S.; Schoekel, A.; Bönisch, N.; Xu, F.; Zhuang, R.; Huang, Q.; Senkovska, I.; Bon, V.; Heine, T.; et al. Photocatalytic Dehalogenation of Aryl Halides Mediated by the Flexible Metal–Organic Framework MIL-53(Cr). Angew. Chem. Int. Ed. 2025, 64, e202422776. [Google Scholar] [CrossRef] [PubMed]
  43. Zhao, H.; Zhu, T.; Sun, Z. Citric acid-induced defective MIL-53(Fe/Ce) for efficient amoxicillin degradation in electro-Fenton system. Sep. Purif. Technol. 2025, 363, 131993. [Google Scholar] [CrossRef]
  44. Qin, Z.; Wu, R.; Gao, S.; Tai, R.; Li, P.; Sui, X.; Song, X.; Wang, Q.; Chen, S. Photocatalytic degradation of pollutants by self-Fenton reaction on defective NH2-MIL-88B with on-site production of hydrogen peroxide. Appl. Surf. Sci. 2025, 695, 162870. [Google Scholar] [CrossRef]
  45. Li, Q.; Zhou, H.; Li, Z.; Liu, A.; Wang, E.; Wu, Y.; Tang, X.; Du, H.; Jin, L.; Zhu, H.; et al. Efficient photocatalytic degradation of antibiotics using Z-scheme MIL-88(Fe)/Ti3C2/MoO3: Mechanistic insights and toxicity assessment. J. Hazard. Mater. 2025, 486, 137051. [Google Scholar] [CrossRef]
  46. Mi, X.; Zhou, X.; Zhan, T.; Chen, Y.; Ma, N.; Dai, W. Interfacial engineering-mediated electronic structure regulation of homologous S-scheme LaFeO3@MIL-100(Fe) heterojunction: Enhancement of photocatalytic activity towards levofloxacin. Sep. Purif. Technol. 2025, 363, 132126. [Google Scholar] [CrossRef]
  47. Li, Y.; Peng, H.; Li, H.; Ma, Q.; Zhang, X.; Chen, Q.; Li, J.R. Elimination of Trace Tetracycline with Alkyl Modified MIL-101 in Water. Small 2024, 20, e2405436. [Google Scholar] [CrossRef]
  48. Dai, Q.; Gao, G.; Tang, J.; Jiang, R.; Sun, S.; Ye, Y.; Li, S.; Xie, R.; Zhang, J. The MIL-125(Ti)/Co3O4 towards efficiently removing tetracycline by synergistic adsorption-photocatalysis roles. Mater. Des. 2025, 250, 113608. [Google Scholar] [CrossRef]
  49. Chakhtouna, H.; Benzeid, H.; Zari, N.; Qaiss, A.e.k.; Bouhfid, R. Microwave-assisted synthesis of MIL–53(Fe)/biochar composite from date palm for ciprofloxacin and ofloxacin antibiotics removal. Sep. Purif. Technol. 2023, 308, 122850. [Google Scholar] [CrossRef]
  50. Zorainy, M.Y.; Kaliaguine, S.; Gobara, M.; Elbasuney, S.; Boffito, D.C. Microwave-Assisted Synthesis of the Flexible Iron-based MIL-88B Metal–Organic Framework for Advanced Energetic Systems. J. Inorg. Organomet. Polym. Mater. 2022, 32, 2538–2556. [Google Scholar] [CrossRef]
  51. Hu, W.; Song, Z.; Sun, L.; Zhang, L.; Zhang, Q.; Ren, X.; Li, Y. TiO2/CoOx heterostructure decorated MIL-100(Fe) by atomic layer deposition for enhanced photocatalytic oxygen production. Dalton Trans. 2025, 54, 3467–3477. [Google Scholar] [CrossRef] [PubMed]
  52. Li, S.; Cui, J.; Wu, X.; Zhang, X.; Hu, Q.; Hou, X. Rapid in situ microwave synthesis of Fe3O4@MIL-100(Fe) for aqueous diclofenac sodium removal through integrated adsorption and photodegradation. J. Hazard. Mater. 2019, 373, 408–416. [Google Scholar] [CrossRef] [PubMed]
  53. Ramos Corona, A.; Rodríguez López, J.; Rangel Segura, R.; Martínez Garcia, M.M.; Flores, E.; Rodríguez Gattorno, G.; Alvarado Gil, J.J. Microwave-Assisted Synthesis of CdS-MOF MIL-101 (Fe) Composite: Characterization and Photocatalytic Performance. Inorg. Chem. 2024, 63, 19536–19552. [Google Scholar] [CrossRef] [PubMed]
  54. Zhu, X.; Zhang, C.; Li, Y.; Lu, Y.; Huang, N.; Wang, D. Degradation of perfluorooctanoic acid by inductively heated Fenton-like process over the Fe3O4/MIL-101 composite. Chin. Chem. Lett. 2024, 35, 109753. [Google Scholar] [CrossRef]
  55. Solís, R.R.; Gómez-Avilés, A.; Belver, C.; Rodriguez, J.J.; Bedia, J. Microwave-assisted synthesis of NH2-MIL-125(Ti) for the solar photocatalytic degradation of aqueous emerging pollutants in batch and continuous tests. J. Environ. Chem. Eng. 2021, 9, 106230. [Google Scholar] [CrossRef]
  56. Guo, Z.; Chen, H.; Feng, D.; Guo, F. Enhanced peroxymonosulfate-activated antibiotic degradation via mechanochemically synthesized cobalt-doped MIL-53(Al). J. Environ. Chem. Eng. 2025, 13, 115356. [Google Scholar] [CrossRef]
  57. Yang, Y.; Chen, J.; Fu, D.; Chang, Y.; Liu, K.; Cui, S.; Li, Y.; Li, B.; Wei, S.; Li, D.; et al. Enhanced nonradical oxidation of thiocyanate by nano γ-Fe2O3 and Fe3C-modified 2D hierarchical porous carbon nanosheets. J. Environ. Chem. Eng. 2025, 13, 115810. [Google Scholar] [CrossRef]
  58. Hidayat, D.; Lestari, W.W.; Dendy, D.; Khoerunnisa, F.; Handayani, M.; Sanjaya, E.H.; Gunawan, T. Adsorption Studies of Anionic and Cationic Dyes on MIL-100(Cr) Synthesized Using Facile and Green Mechanochemical Method. J. Inorg. Organomet. Polym. Mater. 2023, 33, 1548–1561. [Google Scholar] [CrossRef]
  59. Souza, B.E.; Tan, J.-C. Mechanochemical approaches towards the in situ confinement of 5-FU anti-cancer drug within MIL-100 (Fe) metal–organic framework. CrystEngComm 2020, 22, 4526–4530. [Google Scholar] [CrossRef]
  60. Kuzharov, A.A.; Gritsai, M.A.; Butova, V.V.; Soldatov, M.A.; Polyakov, V.A.; Rud, P.A.; Rusalev, Y.V.; Kubrin, S.P.; Roldugin, V.A.; Trigub, A.L.; et al. One-step electrochemical synthesis of γ-Fe2O3@MIL-88a magnetic composite for heterogeneous Fenton-like catalysis. Ceram. Int. 2022, 48, 34864–34876. [Google Scholar] [CrossRef]
  61. Antonio, A.M.; Rosenthal, J.; Bloch, E.D. Electrochemically Mediated Syntheses of Titanium(III)-Based Metal–Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 11383–11387. [Google Scholar] [CrossRef] [PubMed]
  62. Larasati, L.; Dendy, D.; Lestari, W.W.; Suharbiansah, R.S.R.; Firdaus, M.; Masykur, A.; Wibowo, F.R. Rapid and Facile Electrochemical Synthesis of MIL-101(Fe)-NH2 and Its Curcumin Loading and Release Studies. J. Inorg. Organomet. Polym. Mater. 2024, 34, 4039–4049. [Google Scholar] [CrossRef]
  63. Pangestu, A.; Lestari, W.W.; Wibowo, F.R.; Larasati, L. Green Electro-Synthesized MIL-101(Fe) and Its Aspirin Detoxification Performance Compared to MOF-808. J. Inorg. Organomet. Polym. Mater. 2022, 32, 1828–1839. [Google Scholar] [CrossRef]
  64. Wu, W.; Decker, G.E.; Weaver, A.E.; Arnoff, A.I.; Bloch, E.D.; Rosenthal, J. Facile and Rapid Room-Temperature Electrosynthesis and Controlled Surface Growth of Fe-MIL-101 and Fe-MIL-101-NH2. ACS Cent. Sci. 2021, 7, 1427–1433. [Google Scholar] [CrossRef]
  65. Xuemin, W.; Longxue, L.; Ling, L.; Zihan, Y.; Hongxu, G.; Chen, Z. Ultrasound Assisted Synthesis of Nanoscale NH2-MIL-53(Fe) for the Adsorption of Dye. Chin. J. Struct. Chem. 2021, 40, 42–46. [Google Scholar] [CrossRef]
  66. Abbasian, M.; Khayyatalimohammadi, M. Ultrasound-assisted synthesis of MIL-88(Fe) conjugated starch-Fe3O4 nanocomposite: A safe antibacterial carrier for controlled release of tetracycline. Int. J. Biol. Macromol. 2023, 234, 123665. [Google Scholar] [CrossRef]
  67. Shiri, P.; Cui, H.; Zhang, L. Sustainable process intensification: Ultrasonic preparation of MIL-88A for benzoxazole synthesis. Chem. Eng. Process.-Process Intensif. 2024, 201, 109797. [Google Scholar] [CrossRef]
  68. Le, B.T.; La, D.D.; Nguyen, P.T.H. Ultrasonic-Assisted Fabrication of MIL-100(Fe) Metal–Organic Frameworks as a Carrier for the Controlled Delivery of the Chloroquine Drug. ACS Omega 2022, 8, 1262–1270. [Google Scholar] [CrossRef]
  69. Mirzaei, D.; Zabardasti, A.; Sadeghi, M.; Yekta, S. Synthesis of the Novel ZSM-5/NiO/MIL-101(Cr) Zeolite Catalyst Nanocomposite and Its Performance for the Sonodegradation of Organic Dyes in Aqueous Solutions. J. Inorg. Organomet. Polym. Mater. 2021, 31, 960–977. [Google Scholar] [CrossRef]
  70. Al Amery, N.; Abid, H.R.; Al-Saadi, S.; Wang, S.; Liu, S. Facile directions for synthesis, modification and activation of MOFs. Mater. Today Chem. 2020, 17, 100343. [Google Scholar] [CrossRef]
  71. Valekar, A.H.; Lee, S.-K.; Kim, Y.K.; Cho, K.H.; Jo, D.; Hwang, Y.K.; Yoon, J.W.; Lee, U.H. Facile shaping of flexible MIL-53(Al) for effective separation of propylene over propane. Chem. Eng. J. 2024, 480, 147872. [Google Scholar] [CrossRef]
  72. Wang, Y.; Wang, L.; Yu, X.; Shen, X.; Xu, L.; Zhang, Y.; Shi, J.; Gan, L. Alginate/MIL-88B(Fe) derived magnetic biochar bead for non-radical degradation of organic pollutant with low peroxymonosulfate consumption. Chem. Eng. J. 2024, 502, 157892. [Google Scholar] [CrossRef]
  73. Dapaah, M.F.; Niu, Q.; Yu, Y.-y.; Jia, H.; Cheng, L. Tunning MIL-101(Cr) MOF polymorphism towards efficient adsorption and localized Fenton-like degradation of para-nitroaniline by Fe incorporation. Chem. Eng. J. 2024, 486, 150162. [Google Scholar] [CrossRef]
  74. An, Y.; Lv, X.; Jiang, W.; Wang, L.; Shi, Y.; Hang, X.; Pang, H. The stability of MOFs in aqueous solutions—Research progress and prospects. Green Chem. Eng. 2024, 5, 187–204. [Google Scholar] [CrossRef]
  75. Szliszka, E.; Czuba, Z.P.; Domino, M.; Mazur, B.; Zydowicz, G.; Krol, W. Ethanolic Extract of Propolis (EEP) Enhances the Apoptosis- Inducing Potential of TRAIL in Cancer Cells. Molecules 2009, 14, 738–754. [Google Scholar] [CrossRef] [PubMed]
  76. Tao, Y.; Teng, C.; Musho, T.D.; van de Burgt, L.; Lochner, E.; Heller, W.T.; Strouse, G.F.; Dudley, G.B.; Stiegman, A.E. Direct Measurement of the Selective Microwave-Induced Heating of Agglomerates of Dipolar Molecules: The Origin of and Parameters Controlling a Microwave Specific Superheating Effect. J. Phys. Chem. B 2021, 125, 2146–2156. [Google Scholar] [CrossRef] [PubMed]
  77. Fang, X.; Choi, J.Y.; Lu, C.; Reichert, E.; Pham, H.T.B.; Park, J. From 0D to 2D: Microwave-assisted Synthesis of Electrically Conductive Metal-Organic Frameworks with Controlled Morphologies. Chem. Sci. 2025, 16, 3168–3172. [Google Scholar] [CrossRef]
  78. Wang, H.-C.; Liu, X.; Ma, J.-G.; Cheng, P. Rapid and high-throughput synthesis of diverse MOFs with centrifuge tube grinding strategy. Green Chem. 2024, 26, 7312–7319. [Google Scholar] [CrossRef]
  79. Wang, Y.; Liu, Z.; Huang, P.; Lei, B.; Qiao, L.; Li, T.; Lin, K.-Y.A.; Wang, H. Mechanochemical synthesis of biochar encapsulated FeMn nanoparticles with strong metal–carbon interactions for efficient degradation of tetracycline via activating peroxymonosulfate. Chem. Eng. J. 2024, 479, 147525. [Google Scholar] [CrossRef]
  80. Zelenka, T.; Baláž, M.; Férová, M.; Diko, P.; Bednarčík, J.; Királyová, A.; Zauška, Ľ.; Bureš, R.; Sharda, P.; Király, N. The influence of HKUST-1 and MOF-76 hand grinding/mechanical activation on stability, particle size, textural properties and carbon dioxide sorption. Sci. Rep. 2024, 14, 15386. [Google Scholar] [CrossRef]
  81. Varsha, M.V.; Nageswaran, G. Review—Direct Electrochemical Synthesis of Metal Organic Frameworks. J. Electrochem. Soc. 2020, 167, 155527. [Google Scholar] [CrossRef]
  82. Cao, Y.-J.; Chen, Z.-Y.; Peng, H.-L.; Chen, H.-P.; Xie, C.-F.; Fan, J.-P. One-step electrochemical synthesis of Zn-MOFs/NF composite electrodes for efficient electrosorption of organic dyes. Sep. Purif. Technol. 2025, 360, 131020. [Google Scholar] [CrossRef]
  83. Sachdeva, S.; Pustovarenko, A.; Sudhölter, E.J.R.; Kapteijn, F.; de Smet, L.C.P.M.; Gascon, J. Control of interpenetration of copper-based MOFs on supported surfaces by electrochemical synthesis. CrystEngComm 2016, 18, 4018–4022. [Google Scholar] [CrossRef]
  84. Campagnol, N.; Van Assche, T.; Boudewijns, T.; Denayer, J.; Binnemans, K.; De Vos, D.; Fransaer, J. High pressure, high temperature electrochemical synthesis of metal–organic frameworks: Films of MIL-100 (Fe) and HKUST-1 in different morphologies. J. Mater. Chem. A 2013, 1, 5827–5830. [Google Scholar] [CrossRef]
  85. He, G.; Dakhchoune, M.; Zhao, J.; Huang, S.; Agrawal, K.V. Electrophoretic Nuclei Assembly for Crystallization of High-Performance Membranes on Unmodified Supports. Adv. Funct. Mater. 2018, 28, 1707427. [Google Scholar] [CrossRef]
  86. Al-Kutubi, H.; Gascon, J.; Sudhölter, E.J.R.; Rassaei, L. Electrosynthesis of Metal–Organic Frameworks: Challenges and Opportunities. ChemElectroChem 2015, 2, 462–474. [Google Scholar] [CrossRef]
  87. Ansari, K.R.; Singh, A.; Younas, M.; Ali, I.H.; Lin, Y. Progress in metal-organic frameworks (MOFs) as multifunctional material: Design, synthesis and anticorrosion performance techniques. Coord. Chem. Rev. 2025, 523, 216294. [Google Scholar] [CrossRef]
  88. Matyjaszewski, K.; Davis, T.P. Handbook of Radical Polymerization; Wiley-Interscience: Hoboken, NJ, USA, 2013. [Google Scholar] [CrossRef]
  89. Safarifard, V.; Morsali, A. Applications of ultrasound to the synthesis of nanoscale metal–organic coordination polymers. Coord. Chem. Rev. 2015, 292, 1–14. [Google Scholar] [CrossRef]
  90. Safarifard, V.; Morsali, A. Facile preparation of nanocubes zinc-based metal-organic framework by an ultrasound-assisted synthesis method; precursor for the fabrication of zinc oxide octahedral nanostructures. Ultrason. Sonochemistry 2018, 40, 921–928. [Google Scholar] [CrossRef]
  91. Suslick, K.S. The sonochemical hot spot. J. Acoust. Soc. Am. 2005, 89, 1885–1886. [Google Scholar] [CrossRef]
  92. Yan, X.-W.; Hakimifar, A.; Bigdeli, F.; Hanifehpour, Y.; Wang, S.-J.; Liu, K.-G.; Morsali, A.; Joo, S.W. Rapid and Selective Sensing of 2,4,6-Trinitrophenol via a Nano-Plate Zn(II)-Based MOF Synthesized by Ultrasound Irradiation. Crystals 2023, 13, 1344. [Google Scholar] [CrossRef]
  93. Liang, Z.; Liang, Y.; Yu, P.; Wang, X. Ultrasonic-assisted in situ synthesis of MOF-199 on the surface of carboxylated cellulose fibers for efficient adsorption of methylene blue. RSC Adv. 2024, 14, 15095–15105. [Google Scholar] [CrossRef] [PubMed]
  94. Khan, N.A.; Jhung, S.H. Synthesis of metal-organic frameworks (MOFs) with microwave or ultrasound: Rapid reaction, phase-selectivity, and size reduction. Coord. Chem. Rev. 2015, 285, 11–23. [Google Scholar] [CrossRef]
  95. Baluk, M.A.; Pieczyńska, A.; Kroczewska, M.; Łuczak, J.; Matus, K.; Nikiforow, K.; Zaleska-Medynska, A. Efficient method for octahedral NH2-MIL-125 (Ti) synthesis: Fast and mild conditions. Chem. Eng. J. 2024, 492, 152313. [Google Scholar] [CrossRef]
  96. Liu, N.; Huang, W.; Zhang, X.; Tang, L.; Wang, L.; Wang, Y.; Wu, M. Ultrathin graphene oxide encapsulated in uniform MIL-88A(Fe) for enhanced visible light-driven photodegradation of RhB. Appl. Catal. B Environ. 2018, 221, 119–128. [Google Scholar] [CrossRef]
  97. Liu, F.; Cao, J.; Yang, Z.; Xiong, W.; Xu, Z.; Song, P.; Jia, M.; Sun, S.; Zhang, Y.; Zhong, X. Heterogeneous activation of peroxymonosulfate by cobalt-doped MIL-53(Al) for efficient tetracycline degradation in water: Coexistence of radical and non-radical reactions. J. Colloid Interface Sci. 2021, 581, 195–204. [Google Scholar] [CrossRef]
  98. Yan, J.; Wu, Y.; Huang, M.; Cheng, L.; Pan, Y.; Wu, C.-C.; Yeh, C.-H.; Li, J.-L.; Lin, Y.-D.; Chi, Y.; et al. Iridium(III) Carbene Complexes Featuring Either Metal-to-Ligand Charge Transfer (MLCT) or Through-Space Charge Transfer (TSCT) Blue Luminescence. Angew. Chem. Int. Ed. 2025, 64, e202424694. [Google Scholar] [CrossRef]
  99. Wang, C.; Wang, S.; Ping, Y.; Zhao, Z.; Guo, D.; Wang, D.; Bu, X. Ru@MIL-125/MnOx metal-organic-framework-based cocatalysts for photocatalytic nitrogen fixation. Appl. Catal. B Environ. Energy 2024, 347, 123781. [Google Scholar] [CrossRef]
  100. Du, C.; Lv, Y.; Cao, J.; Zhu, H.; Zhang, Y.; Zou, Y.; Peng, H.; Dong, W.; Zhou, L.; Yu, G.; et al. Removal of oxytetracycline from water by S-doped MIL-53(Fe): Synergistic effect of surface adsorption and persulfate activation. Environ. Res. 2023, 239, 116842. [Google Scholar] [CrossRef]
  101. George, P.; Chaudhari, K.; Chowdhury, P. Influence of cation doping (Li+, Na+, K+) on photocatalytic activity of MIL-53(Fe). J. Mater. Sci. 2018, 53, 11694–11714. [Google Scholar] [CrossRef]
  102. Jiang, X.; Su, S.; Ren, B.; Qiu, Y.; Wang, S.; Yang, X. Lanthanum-Doped iron MOFs: A sustainable solution for Arsenic(V) and phosphate pollution in water. Sep. Purif. Technol. 2025, 354, 129098. [Google Scholar] [CrossRef]
  103. Wan, Z.; Xu, X.; Bi, Z.; Jiajia, D.; Li, Y.; Chen, M.; Huang, Z. Gadolinium doping-induced electronic structure optimization of MIL-101-NH2: Efficient adsorption of arsenic (V) and phosphorus and electrochemical regeneration. Sep. Purif. Technol. 2025, 357, 130133. [Google Scholar] [CrossRef]
  104. Mo, G.; Wang, L.; Luo, J. Controlled thermal treatment of NH2-MIL-125(Ti) for drastically enhanced photocatalytic reduction of Cr(VI). Sep. Purif. Technol. 2021, 277, 119643. [Google Scholar] [CrossRef]
  105. Thi Nguyen, Q.; Na, J.; Lee, Y.-R.; Baek, K.-Y. Boosting catalytic performance for CO2 cycloaddition under mild condition via amine grafting on MIL-101-SO3H catalyst. J. Environ. Chem. Eng. 2024, 12, 111852. [Google Scholar] [CrossRef]
  106. Zhao, X.; Han, X.; Li, Z.; Huang, H.; Liu, D.; Zhong, C. Enhanced removal of iodide from water induced by a metal-incorporated porous metal–organic framework. Appl. Surf. Sci. 2015, 351, 760–764. [Google Scholar] [CrossRef]
  107. Inchongkol, Y.; Saothayanun, T.K.; Adpakpang, K.; Phongsuk, N.; Impeng, S.; Kosasang, S.; Ma, N.; Horike, S.; Bureekaew, S. Tuning Electronic and Proton Transfer Properties on Amino-Functionalized Co-Based MOF for Efficient Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2024, 16, 64638–64645. [Google Scholar] [CrossRef] [PubMed]
  108. Yu, T.; Gao, P.; Du, H.; Dong, L. Interfacial Co–O–Fe bonding in novel amorphous NiCo-ZIF@MIL-100 as efficient active sites enabling electrocatalytic water oxidation. J. Mater. Chem. A 2024, 12, 12712–12720. [Google Scholar] [CrossRef]
  109. Bhuyan, A.; Ahmaruzzaman, M. Design and Development of p-toluenesulfonic acid Functionalized Dual-functional Fe-MOF@SBC Heterogeneous Catalyst for Efficient Biodiesel Production and Photocatalytic Degradation of Metronidazole. J. Inorg. Organomet. Polym. Mater. 2024, 34, 4842–4862. [Google Scholar] [CrossRef]
  110. Huang, B.; Wang, J.; Xie, D.; Huang, Q.; Wen, D.; Zeng, X.; Lin, D.; Guo, W.; Sun, H.; Xie, F. Surface reconstruction of defect-engineered MIL-88@Fe2O3 p-n heterojunction for enhanced electrocatalytic water and urea oxidation. Chem. Eng. J. 2024, 498, 155006. [Google Scholar] [CrossRef]
  111. Yao, J.; Zhang, M.; Ma, X.; Xu, L.; Gao, F.; Xiao, J.; Gao, H. Interfacial electronic modulation of CoP-CoO p-p type heterojunction for enhancing oxygen evolution reaction. J. Colloid Interface Sci. 2022, 607, 1343–1352. [Google Scholar] [CrossRef]
  112. Yang, J.; Liu, T.; Zhou, H.; Cao, W.; Chen, C.; He, X.; Jiang, C.; Li, Y.; Wang, Y. In situ conversion of typical type-I MIL-125(Ti)/BiOBr into type-II heterostructure photocatalyst via MOF self-sacrifice: Photocatalytic mechanism and theoretical study. J. Alloys Compd. 2022, 900, 163440. [Google Scholar] [CrossRef]
  113. Humayun, M.; Shu, L.; Pi, W.; Xia, H.; Khan, A.; Zheng, Z.; Fu, Q.; Tian, Y.; Luo, W. Vertically grown CeO2 and TiO2 nanoparticles over the MIL53Fe MOF as proper band alignments for efficient H2 generation and 2,4-DCP degradation. Environ. Sci. Pollut. Res. 2022, 29, 34861–34873. [Google Scholar] [CrossRef]
  114. Li, R.; Nie, J.-H.; Xian, J.-J.; Zhou, J.-W.; Lu, Y.; Miao, M.-P.; Zhang, W.-H.; Fu, Y.-S. Planar Heterojunction of Ultrathin CrTe3 and CrTe2 van der Waals Magnet. ACS Nano 2022, 16, 4348–4356. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, X.; Zhou, L.; Cai, X.; Zheng, X.; Liu, W.; Wang, G.; Tan, D.; Luo, X.; Dong, M. Heterogeneous structures and morphological transitions of composite materials and its applications. Adv. Compos. Hybrid Mater. 2024, 7, 251. [Google Scholar] [CrossRef]
  116. Lin, Z.-F.; Wang, T.-H.; Venkatesan, P.; Doong, R.-A. Indirect Z-scheme MXene@g-C3N4/MIL-101(Fe) heterojunction for the enhanced visible-light-responsive enrofloxacin photodegradation. Chem. Eng. J. 2025, 505, 159411. [Google Scholar] [CrossRef]
  117. Wang, Q.; Zhou, Q.; Wang, P.; Wang, E.; Chen, S.; Shi, Y.; Deng, H.; Liu, A.; Du, H.; Li, Z.; et al. S-scheme POM/MIL-101(Fe) heterojunction for photocatalytic decontamination of Cr(VI). Sep. Purif. Technol. 2025, 361, 131498. [Google Scholar] [CrossRef]
  118. Xue, B.; Yang, C.; Guo, P.; Wang, T.; Li, G.; Liu, D. The type II MIL-100/BiOBr heterojunctions promote efficient photo-Fenton degradation of tetracycline hydrochloride. Appl. Surf. Sci. 2025, 689, 162552. [Google Scholar] [CrossRef]
  119. Lu, H.; Deng, C.; Yu, Z.; Zhang, D.; Li, W.; Huang, J.; Bao, T.; Liu, X. Synergistic degradation of fluorene in soil by dielectric barrier discharge plasma combined with P25/NH2-MIL-125(Ti). Chemosphere 2022, 296, 133950. [Google Scholar] [CrossRef]
  120. Hou, L.; Li, W.; Wu, Z.; Wei, Q.; Yang, H.; Jiang, Y.; Wang, T.; Wang, Y.; He, Q. Embedding ZnCdS@ZnIn2S4 into thiazole-modified g-C3N4 by electrostatic self-assembly to build dual Z-scheme heterojunction with spatially separated active centers for photocatalytic H2 evolution and ofloxacin degradation. Sep. Purif. Technol. 2022, 290, 120858. [Google Scholar] [CrossRef]
  121. Li, Q.; Deng, M.; Gao, J.; Liu, A.; Wang, Q.; Wang, E.; Li, Z.; Zhu, H.; Tang, X. Rational design of a novel MIL-100(Fe)/TpPa-1 COF direct Z-scheme heterojunction for photo-self-Fenton removal of antibiotics: Performance and ecotoxicity assessment. Sep. Purif. Technol. 2025, 362, 131722. [Google Scholar] [CrossRef]
  122. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-Scheme Photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  123. Dong, W.; Du, C.; Zhang, Y.; Cao, J.; Jiang, J.; Zhou, L.; Yu, G.; Zou, Y.; Peng, H.; Yan, R.; et al. Enhanced carrier separation and photocatalytic degradation of oxytetracycline via S-scheme MIL-53(Fe)/FeOCl heterojunction composites with peroxydisulfate activation. J. Environ. Chem. Eng. 2025, 13, 115024. [Google Scholar] [CrossRef]
  124. Li, Q.; Zhou, Q.; Wu, Y.; Shi, Y.; Liu, Y.; Deng, H.; Chen, S.; Li, Z.; Wang, E.; Zhu, H.; et al. Dual S-scheme heterojunction via MOF-on-MOF strategy for efficient photoelectrocatalytic removal of organic contaminants: Detoxification and mechanism. J. Environ. Sci. 2025, 155, 111–126. [Google Scholar] [CrossRef]
  125. Du, J.; Ahmad, I.; Ashraf, I.M.; Ahmed, F.B.M.; Aslam, A.; Ali, I.; Mohammad, A.; Khasawneh, M.A. Advancements in dual S-scheme heterojunction systems for photocatalytic applications: A mini-review. Int. J. Hydrogen Energy 2025, 100, 1361–1384. [Google Scholar] [CrossRef]
  126. Patial, S.; Sudhaik, A.; Sonu; Thakur, S.; Van Le, Q.; Ahamad, T.; Singh, P.; Huang, C.-W.; Nguyen, V.-H.; Raizada, P. Synergistic interface engineering in n-p-n type heterojunction Co3O4/MIL/Mn-STO with dual S-scheme multi-charge migration to enhance visible-light photocatalytic degradation of antibiotics. Environ. Res. 2024, 240, 117481. [Google Scholar] [CrossRef] [PubMed]
  127. Zhao, H.; Xing, Z.; Su, S.; Song, S.; Li, Z.; Zhou, W. Gear-shaped mesoporous NH2-MIL-53(Al)/CdS P-N heterojunctions as efficient visible-light-driven photocatalysts. Appl. Catal. B Environ. 2021, 291, 120106. [Google Scholar] [CrossRef]
  128. Liu, H.; Jiang, L.; Huang, S.; Niu, J.; Zhang, Y.; Liao, J.; Dong, G.; Song, D.; Zhou, Q. Metal-organic framework functionalized magnetic Nb2CTX for high enrichment of polychlorinated biphenyls in water prior to gas chromatography tandem mass spectrometry. J. Chromatogr. A 2025, 1740, 465560. [Google Scholar] [CrossRef]
  129. Su, P.; Zhang, C.; Liu, Y.; Zhang, J.; Djellabi, R.; Wang, R.; Guo, J.; Zhang, R.; Guo, H.; Ding, X.; et al. Boosting PFOA photocatalytic removal from water using highly adsorptive and sunlight-responsive ZIF67/MIL-100(Fe) modified C3N4. J. Environ. Chem. Eng. 2023, 11, 110765. [Google Scholar] [CrossRef]
  130. Ismail, M.A.; Taki, A.G.; Kumar, S.; Sammen, S.S.; Amari, A.; Bongale, A.; Kisi, O.; Salem, A. Effectiveness of waste-derived MIL type MOFs in removing PFOA and PFAS pollutants for environmental remediation. Sci. Rep. 2025, 15, 9439. [Google Scholar] [CrossRef]
  131. Alrefaee, S.H.; Al-bonayan, A.M.; Alsharief, H.H.; Aljohani, M.; Alshammari, K.F.; Saad, F.A.; Abumelha, H.M.; El-Metwaly, N.M. Efficient removal of carbofuran by sono-photo active CdS@MIL based Ti Framework. Surf. Interfaces 2023, 40, 103133. [Google Scholar] [CrossRef]
  132. Wen, Y.; Rentería-Gómez, Á.; Day, G.S.; Smith, M.F.; Yan, T.-H.; Ozdemir, R.O.K.; Gutierrez, O.; Sharma, V.K.; Ma, X.; Zhou, H.-C. Integrated Photocatalytic Reduction and Oxidation of Perfluorooctanoic Acid by Metal–Organic Frameworks: Key Insights into the Degradation Mechanisms. J. Am. Chem. Soc. 2022, 144, 11840–11850. [Google Scholar] [CrossRef] [PubMed]
  133. Sharmin, A.; Bhuiyan, M.A.; Pramanik, B.K. A hybrid LMO MOF catalytic membrane with PMS activation for efficient degradation of pharmaceutical micropollutants and nanoplastics removal. Sep. Purif. Technol. 2025, 360, 130961. [Google Scholar] [CrossRef]
  134. Nguyen, V.-T.; Tseng, Y.-S.; Nguyen, T.-B.; Chen, C.-W.; Bui, X.-T.; Dong, C.-D. Z-scheme TiO2 derived from MIL-125(Ti)/g-C3N5 heterojunction for enhanced photocatalytic degradation of sulfadiazine via peroxymonosulfate activation. J. Environ. Chem. Eng. 2025, 13, 115910. [Google Scholar] [CrossRef]
  135. Jiang, Z.; Wang, T.; Wang, J.; Yu, T.; Kong, C.; Yang, Z.; Zhu, H. Oxygen vacancy-rich BiVO4 modified with mesoporous MIL-88A(Fe) Z-scheme heterojunction for enhanced photocatalytic formaldehyde degradation. Sep. Purif. Technol. 2025, 353, 128581. [Google Scholar] [CrossRef]
  136. Sun, Z.; Wang, M.; Fan, J.; Feng, R.; Zhou, Y.; Zhang, L. TiO2@MIL-101(Cr) nanocomposites as an efficient photocatalyst for degradation of toluene. Adv. Compos. Hybrid Mater. 2021, 4, 1322–1329. [Google Scholar] [CrossRef]
  137. Zhu, P.; Hu, Z.; Chen, S. Praseodymium-Doped Cr2O3 Prepared by In Situ Pyrolysis of MIL-101(Cr) for Highly Efficient Catalytic Oxidation of 1,2-Dichloroethane. Molecules 2024, 29, 3417. [Google Scholar] [CrossRef]
  138. Zhang, X.; Yue, K.; Rao, R.; Chen, J.; Liu, Q.; Yang, Y.; Bi, F.; Wang, Y.; Xu, J.; Liu, N. Synthesis of acidic MIL-125 from plastic waste: Significant contribution of N orbital for efficient photocatalytic degradation of chlorobenzene and toluene. Appl. Catal. B Environ. 2022, 310, 121300. [Google Scholar] [CrossRef]
  139. Zhao, C.; Wang, Z.; Chen, X.; Chu, H.; Fu, H.; Wang, C.-C. Robust photocatalytic benzene degradation using mesoporous disk-like N-TiO2 derived from MIL-125(Ti). Chin. J. Catal. 2020, 41, 1186–1197. [Google Scholar] [CrossRef]
  140. Yang, S.; Wang, X.; Song, G.; Lu, G.; Shi, G.; Wang, Y.; Xie, X.; Sun, J. Ti-O-Mo bond-bridged PMA@MIL-125-NH2 photocatalyst for gas acetone photocatalytic degradation. Appl. Catal. B Environ. Energy 2025, 367, 125112. [Google Scholar] [CrossRef]
  141. Liu, H.; Feng, J.; Wang, X.; Xu, M.; Deng, Y.; Li, G.; Yu, Y.; An, T. Improving photocatalytic activity and chlorine resistance of carbon nanolayer-wrapped TiO2 nanocomposite catalysts for dichloromethane purification. Environ. Sci. Nano 2025, 12, 2486–2494. [Google Scholar] [CrossRef]
  142. He, L.; Xu, Y.; Yang, Z.; Lu, X.; Yao, X.; Li, C.; Xu, D.; Wu, C.; Yao, Z. Copper-decorated strategy based on defect-rich NH2-MIL-125(Ti) boosts efficient photocatalytic degradation of methyl mercaptan under sunlight. Environ. Pollut. 2024, 344, 123341. [Google Scholar] [CrossRef]
  143. Liu, Q.; Chen, Y.; Chen, Z.; Xie, Y.; Yu, H.; Yuan, S.; Guo, Y.; Cheng, Y.; Qian, H.; Yao, W. Rapid magnetization and removal of microplastics from environment and food based on magnetic metal-organic framework Fe3O4@SiO2@MIL-53(Al). Environ. Sci. Pollut. Res. 2023, 30, 117373–117389. [Google Scholar] [CrossRef] [PubMed]
  144. Feng, X.; Long, R.; Liu, C.; Liu, X. Visible-light-driven removal of tetracycline hydrochloride and microplastics (HDPE) by nano flower hybrid heterojunction NH2-MIL-88B(Fe)/MoS2 via enhanced electron-transfer. Sep. Purif. Technol. 2022, 302, 122138. [Google Scholar] [CrossRef]
  145. Rojas-Guerrero, C.A.; Villanueva-Rodríguez, M.; Guzmán-Mar, J.L.; Hernández-Ramírez, A.; Cedillo-González, E.I.; Longoria Rodríguez, F.E.; Hinojosa-Reyes, L. Solar photocatalytic degradation of polyethylene terephthalate nanoplastics: Evaluation of the applicability of the TiO2/MIL-100(Fe) composite material. J. Environ. Chem. Eng. 2023, 11, 110415. [Google Scholar] [CrossRef]
  146. Gu, X.; Li, L.; Wu, Y.; Dong, W. Enhancement of microplastics degradation with MIL-101 modified BiOI photocatalyst under light and dark alternated system. J. Environ. Chem. Eng. 2024, 12, 112958. [Google Scholar] [CrossRef]
  147. Yang, F.; Li, J.; Dong, J.; Chen, S.; Hu, W.; Zhang, Y.; Wang, H.; Li, Z.; Wang, Z. MX@MIL-125(Ti)-mediated sonocatalytic degradation for the dyes and microplastics. Sep. Purif. Technol. 2024, 337, 126488. [Google Scholar] [CrossRef]
  148. Su, X.; Dong, Y.; Zhu, Y.; Shi, H. MIL-125-NH2/BNQDs persistent photocatalyst enhanced peroxymonosulfate activation for efficient PET plastics removal. Chem. Eng. J. 2024, 501, 157764. [Google Scholar] [CrossRef]
  149. Wang, Z.; Adu-Kumi, S.; Diamond, M.L.; Guardans, R.; Harner, T.; Harte, A.; Kajiwara, N.; Klánová, J.; Liu, J.; Moreira, E.G.; et al. Enhancing Scientific Support for the Stockholm Convention’s Implementation: An Analysis of Policy Needs for Scientific Evidence. Environ. Sci. Technol. 2022, 56, 2936–2949. [Google Scholar] [CrossRef]
  150. Tian, Y.; Liu, Q.; Lin, S.; Liang, X.; Khan, S.; Yang, X.; Wang, X. Magnetic Z-scheme CuFe2O4/MIL-101(Fe) toward chlorpyrifos degradation: Photocatalytic mechanism, degradation pathways, and intermediates toxicity evaluation. J. Environ. Chem. Eng. 2023, 11, 110054. [Google Scholar] [CrossRef]
  151. Madaj, R.; Sobiecka, E.; Kalinowska, H. Lindane, kepone and pentachlorobenzene: Chloropesticides banned by Stockholm convention. Int. J. Environ. Sci. Technol. 2018, 15, 471–480. [Google Scholar] [CrossRef]
  152. Keshta, B.E.; Yu, H.; Wang, L. MIL series-based MOFs as effective adsorbents for removing hazardous organic pollutants from water. Sep. Purif. Technol. 2023, 322, 124301. [Google Scholar] [CrossRef]
  153. Abdelhameed, R.M.; Darwesh, O.M.; El-Shahat, M. Titanium-based metal-organic framework capsulated with magnetic nanoparticles: Antimicrobial and photocatalytic degradation of pesticides. Microporous Mesoporous Mater. 2023, 354, 112543. [Google Scholar] [CrossRef]
  154. Vigneshwaran, S.; Sirajudheen, P.; Karthikeyan, P.; Nikitha, M.; Ramkumar, K.; Meenakshi, S. Immobilization of MIL-88(Fe) anchored TiO2-chitosan(2D/2D) hybrid nanocomposite for the degradation of organophosphate pesticide: Characterization, mechanism and degradation intermediates. J. Hazard. Mater. 2021, 406, 124728. [Google Scholar] [CrossRef]
  155. Saini, K.; Singh, J.; Malik, S.; Saharan, Y.; Goyat, R.; Umar, A.; Akbar, S.; Ibrahim, A.A.; Baskoutas, S. Metal-Organic Frameworks: A promising solution for efficient removal of heavy metal ions and organic pollutants from industrial wastewater. J. Mol. Liq. 2024, 399, 124365. [Google Scholar] [CrossRef]
  156. Schlummer, M.; Sölch, C.; Meisel, T.; Still, M.; Gruber, L.; Wolz, G. Emission of perfluoroalkyl carboxylic acids (PFCA) from heated surfaces made of polytetrafluoroethylene (PTFE) applied in food contact materials and consumer products. Chemosphere 2015, 129, 46–53. [Google Scholar] [CrossRef] [PubMed]
  157. Li, W.; Cao, J.; Xiong, W.; Yang, Z.; Sun, S.; Jia, M.; Xu, Z. In-situ growing of metal-organic frameworks on three-dimensional iron network as an efficient adsorbent for antibiotics removal. Chem. Eng. J. 2020, 392, 124844. [Google Scholar] [CrossRef]
  158. Kong, Z.; Lu, L.; Zhu, C.; Xu, J.; Fang, Q.; Liu, R.; Shen, Y. Enhanced adsorption and photocatalytic removal of PFOA from water by F-functionalized MOF with in-situ-growth TiO2: Regulation of electron density and bandgap. Sep. Purif. Technol. 2022, 297, 121449. [Google Scholar] [CrossRef]
  159. Cui, J.; Gao, P.; Deng, Y. Destruction of Per- and Polyfluoroalkyl Substances (PFAS) with Advanced Reduction Processes (ARPs): A Critical Review. Environ. Sci. Technol. 2020, 54, 3752–3766. [Google Scholar] [CrossRef]
  160. Eberle, D.; Ball, R.; Boving, T.B. Impact of ISCO treatment on PFAA co-contaminants at a former fire training area. Environ. Sci. Technol. 2017, 51, 5127–5136. [Google Scholar] [CrossRef]
  161. Babalar, M.; Siddiqua, S.; McIntyre, L.; Ellenor, D.; Usakiewicz, J. Nanoporous dopamine/β-cyclodextrin PES-PMACZ/MOF modified membrane for high-efficiency, low-fouling extraction of microplastics and PCB 209 from synthetic landfill leachate. J. Hazard. Mater. Adv. 2025, 18, 100637. [Google Scholar] [CrossRef]
  162. Shi, X.; Ren, B.; Jin, X.; Wang, X.C.; Jin, P. Metabolic hazards of pharmaceuticals and personal care products (PPCPs) in sewers. J. Hazard. Mater. 2022, 432, 128539. [Google Scholar] [CrossRef]
  163. Zhichang, D.; Chunyan, S.; Xiangrui, Z.; Wensong, L.; Xue-Rong, S. The Application of Metal-Organic Frameworks in the Adsorptive Removal of Harmful Species from Aqueous Solutions. Mini-Rev. Org. Chem. 2023, 20, 227–239. [Google Scholar] [CrossRef]
  164. Preeyanghaa, M.; Vinesh, V.; Sabarikirishwaran, P.; Rajkamal, A.; Ashokkumar, M.; Neppolian, B. Investigating the role of ultrasound in improving the photocatalytic ability of CQD decorated boron-doped g-C3N4 for tetracycline degradation and first-principles study of nitrogen-vacancy formation. Carbon 2022, 192, 405–417. [Google Scholar] [CrossRef]
  165. Huang, C.-Y.; Song, M.; Gu, Z.-Y.; Wang, H.-F.; Yan, X.-P. Probing the Adsorption Characteristic of Metal–Organic Framework MIL-101 for Volatile Organic Compounds by Quartz Crystal Microbalance. Environ. Sci. Technol. 2011, 45, 4490–4496. [Google Scholar] [CrossRef] [PubMed]
  166. Zhao, Z.; Nie, J.; Huang, H.; He, F.; Wang, K.; Yang, P. Experimental Study on the Adsorption Performance of Metal–Organic Framework MIL-101 (Cr) for Indoor Toluene. Buildings 2025, 15, 2506. [Google Scholar] [CrossRef]
  167. Montano, L.; Baldini, G.M.; Piscopo, M.; Liguori, G.; Lombardi, R.; Ricciardi, M.; Esposito, G.; Pinto, G.; Fontanarosa, C.; Spinelli, M.; et al. Polycyclic Aromatic Hydrocarbons (PAHs) in the Environment: Occupational Exposure, Health Risks and Fertility Implications. Toxics 2025, 13, 151. [Google Scholar] [CrossRef]
  168. Gao, Z.; Wang, J.; Muhammad, Y.; Hu, P.; Hu, Y.; Chu, Z.; Zhao, Z.; Zhao, Z. Hydrophobic shell structured NH2-MIL(Ti)-125@mesoporous carbon composite via confined growth strategy for ultra-high selective adsorption of toluene under highly humid environment. Chem. Eng. J. 2022, 432, 134340. [Google Scholar] [CrossRef]
  169. Tu, T.N.; Pham, T.M.; Nguyen, Q.H.; Tran, N.T.; Le, V.N.; Ngo, L.H.; Chang, K.; Kim, J. Metal–organic frameworks for aromatic-based VOC capture. Sep. Purif. Technol. 2024, 333, 125883. [Google Scholar] [CrossRef]
  170. Yu, J.; Wang, X.; Lu, G.; Xie, H.; Xie, X.; Sun, J. Improving the removal efficiency of oxygenated volatile organic compounds by defective UiO-66 regulated with water. J. Hazard. Mater. 2024, 469, 134055. [Google Scholar] [CrossRef]
  171. Yu, X.; Hu, Y.; Luan, X.; Jalil Shah, S.; Liu, L.; Li, C.; Ren, Y.; Zhou, L.; Li, J.; Deng, J.; et al. Microwave-assisted construction of MXene/MOF aerogel via N-metal bonds for efficient photodegradation of vapor acetone under high humidity. Chem. Eng. J. 2023, 476, 146878. [Google Scholar] [CrossRef]
  172. Zhang, Y.; Zhu, Z.; Wang, W.-N.; Chen, S.-C. A Novel Sustainable Semiconductor/Metal-organic Framework Coated Electret Filter for Simultaneous Removal of PM2.5 and VOCs. Aerosol Air Qual. Res. 2023, 23, 220445. [Google Scholar] [CrossRef]
  173. Tian, T.; Zhang, J.; Ge, S.; Tian, L. Enhanced gaseous benzene degradation by bimetallic MIL-101(Fe, Cu) activated persulfate system: Efficiency and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2025, 706, 135785. [Google Scholar] [CrossRef]
  174. Lin, H.; Liu, Y.; Deng, J.; Jing, L.; Wang, Z.; Wei, L.; Wei, Z.; Hou, Z.; Tao, J.; Dai, H. The Advancement of Supported Bimetallic Catalysts for the Elimination of Chlorinated Volatile Organic Compounds. Catalysts 2024, 14, 531. [Google Scholar] [CrossRef]
  175. Fan, Y.; Zhang, H.; Ren, M.; Zhang, Y.; Li, Y.; Wang, L.; Chen, J. Low-temperature catalytic degradation of chlorinated aromatic hydrocarbons over bimetallic Ce-Zr/UiO-66 catalysts. Chem. Eng. J. 2021, 414, 128782. [Google Scholar] [CrossRef]
  176. Wang, Y.; Guo, X.; Zhao, Y.; Xu, Z.; Fan, S.; Long, H.; Fan, Q.; Duan, Z.; Zhang, J.; Jin, Z. The anodic chlorine ion repelling mechanisms of Fe/Co/Ni-based nanocatalysts for seawater electrolytic hydrogen production. Nano Energy 2025, 135, 110662. [Google Scholar] [CrossRef]
  177. Ding, S.; Wu, S.; Wang, P.; Fang, N.; Zhang, Q.; Li, S.; Chu, Y. Structure-selectivity relevance of multiple-component catalysts for CVOCs’ complete oxidation: State-of-the-art and perspectives. Sep. Purif. Technol. 2025, 354, 128964. [Google Scholar] [CrossRef]
  178. Zhou, Z.; Li, Q.; Su, G.; Pang, J.; Sun, B.; Meng, J.; Shi, B. Catalytic degradation of chlorinated volatile organic compounds (CVOCs) over Ce-Mn-Ti composite oxide catalysts. J. Environ. Sci. 2024, 138, 326–338. [Google Scholar] [CrossRef]
  179. Starkova, J.E.; Borisov, R.S.; Kanateva, A.Y. Recent Advances in the Detection of Sulfur Compounds in Crude Oil and Petroleum Products. J. Anal. Chem. 2024, 79, 2005–2022. [Google Scholar] [CrossRef]
  180. Yao, X.; Zhou, J.; Liu, Z. Study on adsorption of low-concentration methyl mercaptan by starch-based activated carbon. Chemosphere 2022, 302, 134901. [Google Scholar] [CrossRef]
  181. McNamara, N.D.; Neumann, G.T.; Masko, E.T.; Urban, J.A.; Hicks, J.C. Catalytic performance and stability of (V) MIL-47 and (Ti) MIL-125 in the oxidative desulfurization of heterocyclic aromatic sulfur compounds. J. Catal. 2013, 305, 217–226. [Google Scholar] [CrossRef]
  182. Lv, M.; Jiang, B.; Xing, Y.; Ya, H.; Zhang, T.; Wang, X. Recent advances in the breakdown of microplastics: Strategies and future prospectives. Environ. Sci. Pollut. Res. 2022, 29, 65887–65903. [Google Scholar] [CrossRef]
  183. Zhang, D.; Xing, Y.; Wang, X.; Li, W.; Guo, Y.; Tang, Y.; Zhang, H.; Chen, J.; Jiang, B. The effect of polyvinyl chloride microplastics on soil properties, greenhouse gas emission, and element cycling-related genes: Roles of soil bacterial communities and correlation analysis. J. Hazard. Mater. 2024, 480, 136248. [Google Scholar] [CrossRef]
  184. Peller, J.R.; Mezyk, S.P.; Shidler, S.; Castleman, J.; Kaiser, S.; Faulkner, R.F.; Pilgrim, C.D.; Wilson, A.; Martens, S.; Horne, G.P. Facile nanoplastics formation from macro and microplastics in aqueous media. Environ. Pollut. 2022, 313, 120171. [Google Scholar] [CrossRef]
  185. Wang, J.; Zhu, Z.; Sun, L.; Sun, Y.; Yang, S.; Qin, Q.; Xue, Y. The bridging role of soil organic carbon in regulating bacterial community by microplastic pollution: Evidence from different microplastic additions. J. Hazard. Mater. 2025, 490, 137761. [Google Scholar] [CrossRef]
  186. Sharma, V.K.; Ma, X.; Lichtfouse, E.; Robert, D. Nanoplastics are potentially more dangerous than microplastics. Environ. Chem. Lett. 2023, 21, 1933–1936. [Google Scholar] [CrossRef]
  187. Morreale, M.; La Mantia, F.P. Current Concerns about Microplastics and Nanoplastics: A Brief Overview. Polymers 2024, 16, 1525. [Google Scholar] [CrossRef]
Figure 1. (a) Number of publications and (b) number of citations from Web of Science searches using the keywords “MIL” and “emerging contaminants” (data for 2021–2025).
Figure 1. (a) Number of publications and (b) number of citations from Web of Science searches using the keywords “MIL” and “emerging contaminants” (data for 2021–2025).
Inorganics 13 00324 g001
Figure 2. Five different ways to synthesize MIL. (a) Hydrothermal synthesis; (b) microwave-assisted synthesis; (c) mechanical grinding synthesis; (d) electrochemical synthesis; (e) ultrasound-assisted synthesis.
Figure 2. Five different ways to synthesize MIL. (a) Hydrothermal synthesis; (b) microwave-assisted synthesis; (c) mechanical grinding synthesis; (d) electrochemical synthesis; (e) ultrasound-assisted synthesis.
Inorganics 13 00324 g002
Figure 3. (a) SEM image of MIL-53(Fe). (bf) SEM images of S-MIL-53(Fe). (gk) Elemental mapping images of S-MIL-53(Fe). Reprinted from [100].
Figure 3. (a) SEM image of MIL-53(Fe). (bf) SEM images of S-MIL-53(Fe). (gk) Elemental mapping images of S-MIL-53(Fe). Reprinted from [100].
Inorganics 13 00324 g003
Figure 4. SEM images of (a) NH2-MIL-53, (b) La-BDC-NH2, (c) La0.25Fe1.0-MOF, (d) La0.5Fe1.0-MOF, (e) La0.75Fe1.0-MOF, (f) La1.0Fe1.0-MOF, and (g) elemental mapping of La0.75Fe1.0-MOF. Reprinted from [102].
Figure 4. SEM images of (a) NH2-MIL-53, (b) La-BDC-NH2, (c) La0.25Fe1.0-MOF, (d) La0.5Fe1.0-MOF, (e) La0.75Fe1.0-MOF, (f) La1.0Fe1.0-MOF, and (g) elemental mapping of La0.75Fe1.0-MOF. Reprinted from [102].
Inorganics 13 00324 g004
Figure 5. (a,b) Structural model of TpPa-1 COF and MIL-100(Fe); (c) side view of the charge density difference over the heterojunction; (d) SEM images of (d) MT2; (e,f) EDS mapping images of MT2. Reprinted from [121].
Figure 5. (a,b) Structural model of TpPa-1 COF and MIL-100(Fe); (c) side view of the charge density difference over the heterojunction; (d) SEM images of (d) MT2; (e,f) EDS mapping images of MT2. Reprinted from [121].
Inorganics 13 00324 g005
Figure 6. (a) XRD patterns and (b) FTIR patterns of Co3O4, MIL−88A, SrTiO3, Mn−SrTiO3, MIL/Mn−STO, and Co3O4/MIL/Mn−STO composite. Reprinted from [126].
Figure 6. (a) XRD patterns and (b) FTIR patterns of Co3O4, MIL−88A, SrTiO3, Mn−SrTiO3, MIL/Mn−STO, and Co3O4/MIL/Mn−STO composite. Reprinted from [126].
Inorganics 13 00324 g006
Figure 7. (a) Cycling stability of D-MIL-53(Fe/Ce)/400@GF for amoxicillin removal. Reprinted from [43]. (b) Levofloxacin by DNMB-5 in four consecutive runs. Reprinted from [44]. (c) The cycling tests of the MTO for TC degradation under visible light irradiation. Reprinted from [45]. (d) MIL-100(Fe), MIL-100(Fe)/TiO2, and MIL-100(Fe)/TiO2/CoOx with Co deposition cycles of 5, 10, 20, 30, and 40 under 300 W Xe lamp irradiation. Reprinted from [51].
Figure 7. (a) Cycling stability of D-MIL-53(Fe/Ce)/400@GF for amoxicillin removal. Reprinted from [43]. (b) Levofloxacin by DNMB-5 in four consecutive runs. Reprinted from [44]. (c) The cycling tests of the MTO for TC degradation under visible light irradiation. Reprinted from [45]. (d) MIL-100(Fe), MIL-100(Fe)/TiO2, and MIL-100(Fe)/TiO2/CoOx with Co deposition cycles of 5, 10, 20, 30, and 40 under 300 W Xe lamp irradiation. Reprinted from [51].
Inorganics 13 00324 g007
Figure 8. Photocatalytic radical generation and driven degradation of emerging contaminants.
Figure 8. Photocatalytic radical generation and driven degradation of emerging contaminants.
Inorganics 13 00324 g008
Figure 9. (a) Cycling stability of TCS@MOF photocatalyst after five experimental cycles; (b) XRD patterns of TCS@MOF used after four successive cycles for the degradation of MCP. Reprinted from [154].
Figure 9. (a) Cycling stability of TCS@MOF photocatalyst after five experimental cycles; (b) XRD patterns of TCS@MOF used after four successive cycles for the degradation of MCP. Reprinted from [154].
Inorganics 13 00324 g009
Figure 10. Main mechanisms of photocatalytic degradation of PFOA.
Figure 10. Main mechanisms of photocatalytic degradation of PFOA.
Inorganics 13 00324 g010
Figure 11. Main mechanisms of photocatalytic degradation of TC.
Figure 11. Main mechanisms of photocatalytic degradation of TC.
Inorganics 13 00324 g011
Figure 12. Main mechanisms of photocatalytic degradation of acetone.
Figure 12. Main mechanisms of photocatalytic degradation of acetone.
Inorganics 13 00324 g012
Figure 13. (a) Photocatalytic degradation of CH3SH over composite materials with different ratios of NH2-BDC; (b) different activation temperatures; (c) and different introduced Cu contents; (d) recycling tests for 0.2Cu/80%NH2-MIL-125 (300 °C). Reprinted from [142].
Figure 13. (a) Photocatalytic degradation of CH3SH over composite materials with different ratios of NH2-BDC; (b) different activation temperatures; (c) and different introduced Cu contents; (d) recycling tests for 0.2Cu/80%NH2-MIL-125 (300 °C). Reprinted from [142].
Inorganics 13 00324 g013
Figure 14. FESEM images of HDPE microplastics with different degrees of alteration. HDPE-M-X represents the photocatalytic process of HDPE microplastics in NH2-MIL-88B(Fe)/MoS2 for X hours. Reprinted from [144].
Figure 14. FESEM images of HDPE microplastics with different degrees of alteration. HDPE-M-X represents the photocatalytic process of HDPE microplastics in NH2-MIL-88B(Fe)/MoS2 for X hours. Reprinted from [144].
Inorganics 13 00324 g014
Figure 15. Main mechanisms of photocatalytic degradation of PE.
Figure 15. Main mechanisms of photocatalytic degradation of PE.
Inorganics 13 00324 g015
Table 1. Overview of the three-dimensional structure and properties of MILs.
Table 1. Overview of the three-dimensional structure and properties of MILs.
MOFThe 3D StructureLigandCentral Metal AtomTypical ShapeRef.
MIL-53Inorganics 13 00324 i001Terephthalic acidFe, Al, CrRod-shaped, rectangular cone-like[21]
MIL-88Inorganics 13 00324 i002Fumarate, terephthalic acidFeFusiform, hexagonal rod-shaped[22]
MIL-100Inorganics 13 00324 i003Tricarboxylic benzene, trimesic acidFe, Al, CrOctahedron[23]
MIL-101Inorganics 13 00324 i004Terephthalic acidCr, FeOctahedron[24]
MIL-125Inorganics 13 00324 i005Terephthalic acidTiDisc/cake, octahedron[25]
Table 2. MIL preparation methods, synthesis strategies, and properties.
Table 2. MIL preparation methods, synthesis strategies, and properties.
Synthesis MethodMILMetal SourceLigandSynthesis ConditionsMaterial PropertiesNumber of Experimental CyclesModification MethodRef
Solvent synthesis/Hydrothermal synthesisMIL-53Cr(NO3)3·9H2OH2bdc220 °C, 72 h, acidification by adding 40% HFSpecific surface area of 1050 m2/gFive times/[42]
MIL-53FeCl3·6H2O, Ce(NO3)3·6H2OH2bdc140 °C, 24 h, 10% citric acid instead of DMFSpecific surface area > 1000 m2/gEight timesCarbon-based material loading[43]
MIL-88BFeCl3·6H2OH2bdc100 °C, 12 h, add NaOH solution to adjust the Ph of the solutionSpecific surface area of 35.05 m2/gFour timesAmino functionalization[44]
MIL-88AFeCl3·6H2OFumarate65 °C, 12 h/Five timesZ-type heterojunction[45]
MIL-100FeCl3·6H2OH3BTC95 °C, 18 h/Five timesS-type heterojunction[46]
MIL-101Cr(NO3)2H2bdc210 °C, 10 h//Postsynthetic modifier[47]
MIL-125C16H36O4TiH2bdc150 °C, 20 hSpecific surface area of 943.9 m2/gFive timesZ-type heterojunction[48]
Microwave-assisted synthesisMIL-53FeCl3·6H2OH2bdc150 °C, 30 min, 300 WSpecific surface area of 96.33 m2/gFive timesCarbon-based material load[49]
MIL-88BFeCl3·6H2OH2bdc150 °C, 10 min, 800 WSpecific surface area of 47 m2/gFour times/[50]
MIL-100FeCl3·6H2OH3BTC150 °C, 15 min, 800 WSpecific surface area of 487 m2/gTwenty timesZ-type heterojunction[51]
MIL-100FeCl3·6H2OH3BTC150 °C, 30 min, 300 WSpecific surface area of 1244.62 m2/g/Composite modified[52]
MIL-101FeCl3·6H2OH2bdc200 °C, 50 min, 400 W/Four timesType I heterozygous junction[53]
MIL-101FeCl3·6H2OH2bdc200 °C, 50 min, 600 WSpecific surface area of 960.3 m2/gFour timesComposite modified[54]
MIL-125C12H28O4TiH2bdc200 °C, 15 min, 600 WSpecific surface area of 1030 m2/gFour timesAmino functionalization[55]
Mechanical grinding synthesisMIL-53Al2(SO4)3·18H2OH2bdcBall milling frequency 30 Hz, ball milling time 1 hSpecific surface area of 1143 m2/g//[21]
MIL-53AlCl3·6H2OH2bdcBall milling frequency 30 Hz, ball milling time 1 hSpecific surface area of 638.96 m2/gFour timesMetal ion doping[56]
MIL-88AFeCl3·6H2OFumarateBall milling at 200 °C for 5 h at a rate of 5 °C/min.Specific surface area of 723.17 m2/gThree times/[57]
MIL-100CrCl3·6H2OH3BTCGrinding at room temperature for 40 min, heating at 220 °C for 15 hSpecific surface area of 1557 m2/g//[58]
MIL-100Fe(NO3)3·9H2OH3BTCContinuous grinding at room temperature for 3–5 hSpecific surface area of 753 m2/g//[59]
Electrochemical synthesisMIL-88AIron plate electrodeFumarateMixed solution of 50% H2O and 50% ethanol as electrolyte 15 V voltage, 30 minSpecific surface area of 128.5 m2/gThree timesMagnetic material load[60]
MIL-100TiCl4H3TATB−1.20 V potential, 120 °C, 18 hSpecific surface area of 5842 m2/g//[61]
MIL-101TiCl4H2BPDC−1.20 V potential, 120 °C, 18 hSpecific surface area of 3263 m2/g//[61]
MIL-101FeCl3·6H2OH2BDC-NH2TBATFB electrolyte, 15 V, 30 minSpecific surface area of 139 m2/g/Amino functionalization[62]
MIL-101Iron plate electrodeH2BDCTBATFB electrolyte, 15 V, 30 minSpecific surface area of 131 m2/g//[63]
MIL-101FeCl2H2BPDC+0.75 V potential, room temperature, 8 hSpecific surface area of 1918 m2/g/Amino functionalization[64]
Ultrasound-assisted synthesisMIL-53Fe(NO3)3·9H2OH2BDC-NH280 °C, 1 h, 305 WSpecific surface area of 179.9 m2/g/Amino functionalization[65]
MIL-88Fe(NO3)3·9H2OH2bdcUltrasonic treatment 0.5 h, heating at 85 °C for 12 h//Magnetic material load[66]
MIL-88AFeCl3·6H2OFumarateRoom temperature, 4 h, 70% power/Three times/[67]
MIL-100FeCl3·6H2OH3BTC150 °C, 10 min, 1080 W, 20.5 kHzSpecific surface area of 1033 m2/g//[68]
MIL-101Cr(NO3)3·9H2OH2BDCRoom temperature, 1 h, 100 W, 30 kHz/Four times/[69]
Table 3. Summary of MIL photocatalytic removal of emerging contaminants.
Table 3. Summary of MIL photocatalytic removal of emerging contaminants.
Classification of ContaminantsMIL NamesSynthesis MethodModification MethodTarget
Contaminants
Initial
Contaminant
Concentration
Catalytic Degradation
Times
Catalytic
Degradation
Temperature
EfficiencyEnvironmental
Adaptation
Cycling StabilityRef.
POPsNH2-MIL-53(Al)/CdSHydrothermal synthesisAmino functionalization, heterojunctionTrichlorophenol10 mg/L180 minRoom temperature98.85%Stable at 410 °CFive times[127]
Fe3O4@Nb2CTX@NH2–MIL-88Solvothermal synthesisAmino functionalization, heterojunctionPolychlorinated biphenyl0.005–50 μg/L60 min25 °C87.60%//[128]
MIL-100(Fe)@C3N4Hydrothermal synthesisHeterojunctionPerfluorooctanoic acid10–150 μg/L6 hRoom temperature60.50%//[129]
NH2-MIL-101(Cr)Solvothermal synthesisAmino functionalizationPerfluorooctanoic acid10–50 mg/L//698.4 mg/g/Seven times[130]
CdS@NH2-MIL-125Solvothermal synthesisAmino functionalization, heterojunctionCarbofuran10 g/L90 minRoom temperature98.40%/Five times[131]
NH2-MIL-125Hydrothermal synthesisAmino functionalizationPerfluorooctanoic acid100 μg/L24 h20 °C98.9%/Three times[132]
PPCPsMIL-53(Fe)Microwave-assisted synthesisCarbon material loadsCiprofloxacin50–150 mg/L60 min25 °C98.53%The effect is less in conditions where inorganic ions are presentFive times[49]
Co-MIL-53(Al)Mechanical grinding synthesisDoping of elementsTetracycline hydrochloride10–120 mg/L30 min25 °C96.10%Less effect in the presence of inorganic ionsFour times[56]
MIL-88AUltrasound-assisted synthesis/Benzoxazole/1 h25 °C84%~96%//[67]
MIL-88A(Fe)/Ti3C2/MoO3Solvothermal synthesisHeterojunctionTetracycline hydrochloride100 mg/L120 min25 °C94.90%//[45]
Fe3O4@MIL-100(Fe)Microwave-assisted synthesisMagnetic material loadsDiclofenac sodium60 mg/L3 h25 °C99.40%//Five times[52]
MIL-100(Fe)/TpPa-1 COFSolvothermal synthesisHeterojunctionTetracycline hydrochloride20 mg/L120 minRoom temperature91%36 h of continuous catalyticFive times[121]
g-C3N4/MIL-101(Fe)Hydrothermal synthesisHeterojunctionEnrofloxacin10 mg/L60 min25 °C100%Degradation efficiency above 80% at pH 3 and 9Five times[116]
MIL-100(Fe)/CoFeLMO/PEGHydrothermal synthesisHeterojunctionRanitidine1–10 mg/L1 minRoom temperature99.50%Extreme pH conditions maintain 99.5% degradation efficiencyTwenty times[133]
MIL-100(Fe)/TpPa-1 COFSolvothermal synthesisHeterojunctionTetracycline hydrochloride20 mg/L120 minRoom temperature91%36 h of continuous catalyticFive times[121]
MIL-125(Ti)/g-C3N5Solvothermal synthesisHeterojunctionSulfadiazine0.04 mM15 min25 °C95.60%Degradation rate of 64.5% at pH 3Five times[134]
VOCsMIL-88A(Fe)@BiVO4Hydrothermal synthesisHeterojunctionFormaldehyde50 ppm120 min25 °C91%Slight decrease in degradation efficiency at higher humidity levelsFive times[135]
TiO2@MIL-101(Cr)Hydrothermal synthesisHeterojunctionToluene/120 minRoom temperature39.60%//[136]
Pr/MIL-101(Cr)Hydrothermal synthesisDoping of elements1,2-dichloroethane1000 ppm/290 °C95%Stabilized at 290 °C for 120 h/[137]
NH2-MIL-125Solvothermal synthesisAmino functionalizationMethylbenzene2000 ppm20 min25 °C96%High degradation efficiency in the presence of water vaporFour times[138]
N/MIL-125(Ti)Solvothermal synthesisDoping of elementsBenzene0.007%480 min25 °C99.10%/Ten times[139]
PMA@NH2-MIL-125Solvothermal synthesisAmino functionalization, compound modificationAcetone400 ppm/Room temperature78%/Five times[140]
NH2-MIL-125Solvothermal synthesisCarbon material loadsDichloromethane70 ± 1 ppmv5 hRoom temperature85%Stabilizes degradation at high humidity/[141]
Cu/NH2-MIL-125Solvothermal synthesisDoping of elementsMethyl mercaptan100 ppm25 minRoom temperature100%/Five times[142]
MicroplasticsFe3O4@SiO2@MIL-53(Al)Hydrothermal synthesisHeterojunctionPolyvinyl chloride1 mg/mL10 hRoom temperature93.17%/Five times[143]
NH2-MIL-88B(Fe)/MoS2Solvothermal synthesisAmino functionalization, heterojunctionPolyethylene10 g/L48 hRoom temperature/High degradation efficiency over the full pH rangeSeven times[144]
TiO2/MIL-100(Fe)Microwave-assisted synthesisHeterojunctionPolyethylene terephthalate100 mg/L5 hRoom temperatureCI = 0.99The degradation efficiency was higher at pH 3/[145]
BiOI@MIL-101Solvothermal synthesisHeterojunctionPolyethylene1.0 g/L6 hRoom temperature//Five times[146]
Ti3C2 MXene @MIL-125(Ti)Hydrothermal synthesisHeterojunctionPolyethylene0.5 g/L4 hRoom temperature78%Better removal at low pHFive times[147]
MIL-125-NH2/BNQDsSolvothermal synthesisAmino functionalization, heterojunctionPolyethylene terephthalate/5 hRoom temperature95.71%The degradation efficiency was higher at pH 3Four times[148]
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

Chen, Y.; Jiang, Y.; Li, W.; Su, W.; Xing, Y.; Yu, S.; Li, W.; Guo, Y.; Zhang, D.; Wang, S.; et al. MIL Series in MOFs for the Removal of Emerging Contaminants: Application and Mechanisms. Inorganics 2025, 13, 324. https://doi.org/10.3390/inorganics13100324

AMA Style

Chen Y, Jiang Y, Li W, Su W, Xing Y, Yu S, Li W, Guo Y, Zhang D, Wang S, et al. MIL Series in MOFs for the Removal of Emerging Contaminants: Application and Mechanisms. Inorganics. 2025; 13(10):324. https://doi.org/10.3390/inorganics13100324

Chicago/Turabian Style

Chen, Yixiang, Yusheng Jiang, Weiping Li, Wei Su, Yi Xing, Shuyan Yu, Wenxin Li, Ying Guo, Duo Zhang, Shanqing Wang, and et al. 2025. "MIL Series in MOFs for the Removal of Emerging Contaminants: Application and Mechanisms" Inorganics 13, no. 10: 324. https://doi.org/10.3390/inorganics13100324

APA Style

Chen, Y., Jiang, Y., Li, W., Su, W., Xing, Y., Yu, S., Li, W., Guo, Y., Zhang, D., Wang, S., Qian, Z., Hong, C., & Jiang, B. (2025). MIL Series in MOFs for the Removal of Emerging Contaminants: Application and Mechanisms. Inorganics, 13(10), 324. https://doi.org/10.3390/inorganics13100324

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