Next Article in Journal
Advancements in Preprocessing and Analysis of Nitrite and Nitrate since 2010 in Biological Samples: A Review
Previous Article in Journal
A Functional Biological Molecule Restores the PbI2 Residue-Induced Defects in Two-Step Fabricated Perovskites
Previous Article in Special Issue
Recent Advances in Metal-Organic Framework (MOF)-Based Photocatalysts: Design Strategies and Applications in Heavy Metal Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Fe-MOFs in Photodegradation and Removal of Air and Water Pollutants: A Review

1
National Joint Engineering Research Center of Energy Saving and Environmental Protection Technology in Metallurgy and Chemical Engineering Industry, Kunming University of Science and Technology, Kunming 650093, China
2
Kunming Electric Power Design Institute Limited Liability Company, Kunming 650034, China
3
Yunnan Hubai Environmental Protection Technology Co., Ltd., Kunming 650034, China
4
Zhejiang Ecological and Environmental Monitoring Center, Hangzhou 310012, China
5
State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, Kunming University of Science and Technology, Kunming 650093, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(20), 7121; https://doi.org/10.3390/molecules28207121
Submission received: 31 August 2023 / Revised: 8 October 2023 / Accepted: 14 October 2023 / Published: 17 October 2023
(This article belongs to the Special Issue Design and Synthesis of Functional Materials for Photocatalysis)

Abstract

:
Photocatalytic technology has received increasing attention in recent years. A pivotal facet of photocatalytic technology lies in the development of photocatalysts. Porous metal–organic framework (MOF) materials, distinguished by their unique properties and structural characteristics, have emerged as a focal point of research in the field, finding widespread application in the photo-treatment and conversion of various substances. Fe-based MOFs have attained particular prominence. This review explores recent advances in the photocatalytic degradation of aqueous and gaseous substances. Furthermore, it delves into the interaction between the active sites of Fe-MOFs and pollutants, offering deeper insights into their mechanism of action. Fe-MOFs, as photocatalysts, predominantly facilitate pollutant removal through redox processes, interaction with acid sites, the formation of complexes with composite metal elements, binding to unsaturated metal ligands (CUSs), and hydrogen bonding to modulate their respiratory behavior. This review also highlights the focal points of future research, elucidating the challenges and opportunities that lie ahead in harnessing the characteristics and advantages of Fe-MOF composite catalysts. In essence, this review provides a comprehensive summary of research progress on Fe-MOF-based catalysts, aiming to serve as a guiding reference for other catalytic processes.

1. Introduction

In the wake of the continuous development of the global economy and the rapid expansion of industry and agriculture activities, a variety of environmental and energy challenges has grown increasingly severe. These environmental pollutants, encompassing waterborne and atmospheric contaminants, pose significant threats to public health [1,2]. These pollutants can be broadly categorized into two main types: natural pollutants and anthropogenic pollutants, with the latter being the primary culprits behind public health hazards. Water pollutants encompass a diverse array, including antibiotics, dye wastewater, heavy metals, pesticides, polycyclic aromatic hydrocarbons, pharmaceuticals, and PPCPS, while air pollutants include carbon oxides, nitrogen oxides, hydrocarbons, and VOCs [3,4,5,6,7,8,9,10,11,12,13]. Conventional treatment methods, such as physical adsorption, chemical oxidation, and biological treatment, are unable to meet the requirements of efficient, economical, and environmentally friendly treatment of such pollutants. Therefore, the green, cost-effective, and efficient photocatalytic approach has emerged as a standout solution. Photocatalytic oxidation, relying on semiconductor materials, leverages the abundant solar energy available in nature as its driving force to decompose pollutants. The chief advantage of this technology lies in its ability to achieve complete pollutant degradation within environmental media at ambient temperature and pressure. As a result, pollutants are rendered relatively environmentally friendly and more amenable to subsequent treatment [14,15,16,17,18,19,20,21].
Metal–organic frameworks (MOFs) are polycrystalline coordination polymers featuring a multi-dimensional stereo-structure, formed by amalgamation of metal ions (or clusters of metal oxides) with organic ligand molecules containing elements such as carbon, nitrogen, and oxygen through coordination bonds [22,23,24,25,26,27,28]. In comparison to other commonly used photocatalysts, MOF materials boast a series of distinctive advantages: (1) Both metal nodes and organic linkers within MOFs play pivotal roles in photocatalytic reactions. It has been reported that metal nodes and organic linkers can be regarded as isolated semiconductor quantum dots and light-absorbing antennae, respectively. Under light irradiation, metal nodes can be directly excited or activated by organic linkers; (2) MOFs exhibit high porosity and a significant specific surface area, facilitating the exposure of a greater number of reaction sites and promoting the transport of substrates and products; (3) the structure of MOFs is adjustable, allowing for the expansion of the optical response range, and the theoretical bandgap of MOFs falls within the range of 1–5.5 eV; (4) the porous structure of MOFs creates shorter pathways for charge carriers to reach the system and engage with substrates. This aids in enhancing the separation of photogenerated electrons from holes; (5) metal cations exhibit diverse coordination chemistry, and there exists a multitude of organic connectors. Reasonable selection of metal ions and organic ligands can modify their light absorption properties, enabling the efficient utilization of sunlight, among other advantages [29,30,31,32].
In metal–organic skeleton composites, transition metal oxides, such as Fe-MOFs, Co-MOFs, along with various other composite materials, are commonly used. Among these, Fe-MOFs have attracted substantial interest due to their potential applications in catalysis. This fascination arises from the cost-effectiveness, low toxicity, variable valence, biocompatibility, and environmental friendliness associated with the iron (Fe) element [33,34,35]. Fe-MOFs not only inherit several characteristics from conventional MOF materials, such as a large pore volume and a high specific surface area, but they also possess abundant Lewis acid centers, Brønsted acid centers, and unsaturated metal sites, rendering them ideal catalytic active centers [36]. Furthermore, the incorporation of metal ions with redox activity into Fe-MOFs can further enhance their catalytic activity, thanks to the coordination effect. Additionally, the formation of hydrogen bonds between μ2-OH or μ2-O in the inorganic chains of Fe-MOFs and guest molecules facilitates the efficient removal of pollutants. In comparison to MOFs in the ZIF series and other transition metals like Cr, Ni, Cu, and Co, Fe-MOFs exhibit superior stability, environmental friendliness, and affordability. Prominent examples of Fe-MOFs include MIL-53(Fe), MIL-100(Fe), MIL-88A/B(Fe), MIL-101(Fe), and various functionalized MOFs, often featuring ligands such as Schiff base and pyrazine. As shown in Figure 1, Fe-MOFs are widely used in the study of water pollutants and gaseous pollutants. Their remarkable ability to separate charge carriers contributes to outstanding light absorption and photovoltaic properties. For instance, the Fe33-O node-centered cluster structure in the MIL series of Fe-MOFs exhibits a low band gap (Eg) (2.01 eV for MIL-100 (Fe) and 2.62 eV for MIL-53 (Fe)), making it susceptible to the excitation of visible light and the generation of photogenerated electrons–holes. Consequently, Fe-MOFs are widely used in photocatalytic treatment applications [37,38,39,40,41,42]. In addition, it is worth highlighting that Fe-MOFs possess an exceptionally large specific surface area and porosity [43]. When coupled with their ligand-unsaturated sites, such as Lewis acid sites within the structure, this attribute leads to enhanced interactions with guest molecules and reaction intermediates, making them exceptionally well-suited for photocatalytic applications. Li et al. [44] covalently modified NH2-MIL-101 (Fe) with 2-anthraquinone sulfonic acid (AQS) and used it as a redox mediator to augment the degradation of bisphenol A through persulfate activation. The result showed that AQS-NH2-MIL-101 (Fe) led to an impressive 97.7% degradation of BPA.
This review primarily focuses on the utilization of Fe-based MOFs in the photocatalytic adsorptive removal of harmful water pollutants and gaseous substances commonly found in the environment. The principles governing the removal of water pollutants and large gaseous substances with Fe-MOFs are summarized and discussed. The main reaction mechanism is shown in Figure 1. In addition, their strengths and weaknesses in the context of gaseous pollutant removal are elucidated, while the current research status of Fe-MOFs in the realm of water pollution and air substance photo-treatment is summarized.

2. The Application of Fe-MOFs in Photocatalytic Removal of Organic Pollutants

Organic dyes represent a prominent category of pollutants frequently encountered in dyeing and finishing wastewater. These pollutants, most notably aromatic ring structures like rhodamine B, methyl orange, methylene blue, and malachite green, pose a formidable challenge due to their resistance to complete degradation and removal through conventional treatments [45,46]. Furthermore, the environmental presence of antibiotics, including tetracycline, ciprofloxacin, and chlortetracycline, has garnered significant attention as a major concern. These antibiotic contaminants, resulting from intensive usage and occasional misuse, have led to substantial antibiotic residues in aquatic environments, creating an alarming antibiotic selection pressure. For example, antibiotics are reported to cause the development of resistance in aquatic microorganisms, a phenomenon with direct implications for human health (e.g., superbugs, etc.). Increased resistance escalates infection rates and mortality, posing a serious threat to both human well-being and sustainable development. In addition, commonly used nonsteroidal anti-inflammatory drugs (NSAIDs), such as ibuprofen (IBP), are associated with genotoxicity, aquatic environmental toxicity, and the potential for endocrine disruption in organisms. Organophosphorus pesticides (OPPs), by irreversibly inhibiting important enzymes such as acetylcholinesterase (AChE) in the body, present another environmental challenge. Meanwhile, the plastic monomer and plasticizer bisphenol A (BPA) have been linked to increased macrophage activity, reduced antibody production, and immune system disorders, among other adverse effects [47,48]. In light of these pressing concerns, there is a compelling imperative to explore green methods for the removal of organic pollutants. Photocatalytic degradation emerges as a highly promising approach. Fe-MOFs have played a central role in advancing this field, with notable materials, including MIL-53(Fe), MIL-68(Fe), MIL-88A(Fe), MIL-100(Fe), MIL-101(Fe), the ZIF series, and their composites, forming the mainstay of research in the photodegradation of organic pollutants [49,50,51,52,53,54,55,56].

2.1. Dyestuffs Treatment

To harness a broader spectrum of light for the effective utilization of Fe-MOFs, the incorporation of more metal active sites is essential to activate reactants. Designing catalysts with an increased number of active sites stands out as an effective strategy for enhancing photocatalytic performance [57,58,59]. Jin et al. [60] devised a novel dual-ligand Fe-MOF using dielectric barrier discharge (DBD). In this innovative approach, the N ligand served to regulate the active site, while the addition of N enhanced conductivity and provided additional sites for anchoring Fe. This ingenious design ultimately led to the highly efficient degradation of methyl orange. Furthermore, D-Fe-MOFs (dual-ligand Fe-based MOFs) with a greater specific surface area and pore volume were synthesized using DBD [61]. Under simulated sunlight, these materials achieved an impressive 97% degradation of MO (20 mg/L) within 48 min. It was also found that O2, ·OH, and h+ played a crucial role in the system. During the degradation of MO, the photocatalytic efficiency witnessed a notable decline upon the addition of AO and BQ scavengers, underscoring the significance of holes (h+) and superoxide radical anions (O2) as the main active species in the photodegradation of methyl orange dye. In contrast, the addition of TBA had a negligible impact on the photodegradation of methyl orange, implying that ·OH played a relatively minor role in the process. Additionally, it was observed that the molar ratios of different organic ligands had exerted a profound influence on the photocatalytic performance of the catalysts. The formation of these ligands resulted in an increased oxygen vacancy and Fe2+ content, which in turn facilitated the transport of photogenerated electrons and holes, thus promoting the degradation of MO. Wang et al. [62] employed the BDB plasma method to prepare a composite material consisting of Fe-MOFs@Fe2O3 using waste PET. Remarkably, this composite material achieved an outstanding degradation rate of 99.3% for malachite green (MG) within 30 min. This innovative process not only demonstrated high-performance MG removal but also contributed to PET plastics recycling.
Doping and the formation of complexes represent common strategies employed to improve the photocatalytic performance of Fe-MOFs [63,64]. Metal doping serves to suppress electron–hole recombination to the fullest extent, thereby amplifying the photocatalytic activity of Fe-MOFs. Furthermore, the creation of heterojunction through composites with semiconductor materials extends the spectral response range of Fe-MOFs while also impeding the recombination of photogenerated carriers, prolonging their lifetimes, and consequently enhancing their photocatalytic activity. A significant synergistic effect between sonication and the Fenton reaction was elucidated by Geng [65] et al. They achieved high-performance removal of tetracycline hydrochloride using Fe-MOFs synthesized via an ultrasound/H2O2/MOF system (Figure 2a,b). The best catalytic performance was observed with MIL-88B, primarily attributable to its higher number of Lewis acid sites. It was experimentally confirmed that ·OH played a pivotal role in the removal of tetracycline hydrochloride, as indicated by EPR testing and quenching experiments. An important synergistic effect between ultrasound and the Fenton reaction was unveiled through a power-factor assessment of the ultrasound. During ultrasound irradiation, microcurrents are generated, promoting mixing at the solid–liquid interface. This mixing enhances the cleansing of the catalyst’s surface, exposing more active sites and accelerating the reaction between H2O2 and MIL-88B, thereby facilitating the degradation of contaminants. The main reaction pathway in the entire system involves the adsorption and degradation of H2O2 by unsaturated iron sites on Fe-MOFs. Cao et al. [66] introduced PS into the MIL-53(Fe)/LED visible light photocatalytic degradation system for organic pollutants. PS was employed to facilitate the separation of photogenerated electrons and holes, thereby enhancing the photocatalytic activity of MIL-53(Fe). Under the same experimental conditions, the degradation rate of Acid Orange 7 in the MIL-53(Fe)/PS/LED system reached an impressive 100%, in stark contrast to the MIL-53(Fe)/LED system, which achieved only 24% degradation. This significant enhancement was attributed to PS’s capacity to increase the generation of ·OH in the MIL-53(Fe)/LED system. Photoluminescence spectroscopy (PL) and electron cis-resonance (EPR) analyses confirmed the ability of PS to accept photogenerated electrons, thus promoting the degradation of the dye Acid Orange 7 (Figure 2c). Tran et al. [67], among others, achieved efficient photocatalytic degradation of RhB using bimetallic Mn/Fe-MOF materials obtained by doping Mn2+ ions. The activity of these bimetallic centers played a crucial role in shaping unique structures and compositions, ultimately enhancing their photocatalytic ability. Through interfacial conjugation engineering, TiO2@NH2-MIL-101(Fe) was successfully synthesized to enhance the electron transfer capacity from MOFs to TiO2 and induce the band-tail state to obtain a narrower energy band for the composite catalyst. This innovation led to the achievement of high-performance photocatalytic degradation of methylene blue [68]. Brahmi et al. [69] devised MIL-100 (Fe)/polymer and MIL-88A (Fe)/polymer composites using a photopolymerization process, successfully achieving efficient degradation of acidic black pollutants. Tran et al. [70] synthesized M/Fe-MOF (M = Co, Cu, and Mg) materials through bimetallic modification, leading to the efficient removal of rhodamine B. The M/Fe-MOF materials were also modified with M, Cu, and Mg. It was confirmed that after M modification, all metal ions were inserted inside the structure of Fe-MOF material without disturbing their crystal structure. Mousavi et al. [71] successfully synthesized α-Fe2O3@C@SiO2/TiO2 with carboxylic acid surface functional groups (α-Fe2O3@C-COOH) via a two-step dry process. This innovative material demonstrated high-performance removal of dye RY145 while maintaining good recyclability. Zhao et al. [72] found that the presence of amino groups enhances the adsorption capacity of NH2-MILs. This enhanced adsorption capability arises from hydrogen bonding interactions and π–π stacking between the amino groups of the dye molecules and the amino-functionalized MOFs. Gong et al. [73] synthesized a core-shell structure, Fe3O4@GO@MIL-100 (Fe), achieving high-performance degradation of 2,4 dichlorophenol with excellent stability, as well as exo-magnetic field recovery of the catalyst.
Fe-MOFs also play a key role in the dye-sensitized hydrogen production system, where organic dyes function as photosensitizers, expanding the light response range and facilitating the photocatalytic decomposition of water for hydrogen production within visible light. Li et al. [74] used dye sensitization to enable photocatalytic hydrogen production from Fe-MOFs. Their finding indicated that dye sensitization effectively injects high-energy electrons into Fe-MOFs, enhancing their Fermi energy levels under light. This process helps overcome the surface overpotential for H2 production from water. Zhang et al. [75] utilized the coordination and valence mixing of heteroatoms in iron clusters to modulate the electronic energy band structure and charge transfer ability of metal–organic frameworks. They achieved a remarkable twenty-fivefold increase in the hydrogen production rate compared to a pristine Fe-MOF by incorporating Pt as a co-catalyst. ZnIn2S4@NH2-MIL-53(Fe/Co0.75), prepared by Dai et al. [76], achieved an impressive hydrogen production rate of 16,1724.8 μmol/g over a 6 h period. Li et al. [77] attained exceptional photocatalytic hydrogen evolution performance in a dye-sensitized system by anchoring earth-rich copper species to NH2-MIL-101 (Fe). The activity achieved, at 5770.96 μmol/g·h, was attributed to a transient CuII/CuI center, rendering it a noble metal-free synergistic catalyst.

2.2. Antibiotics Removal

He et al. [78] successfully synthesized Fe3O4@MIL-100 (Fe) hetero-complexes using an in situ growth method, achieving an increase in the removal efficiency of levofloxacin from spiked wastewater from 77.9% to 85.5%. Moreover, the catalyst could be magnetically recycled, thus reducing further environmental pollution. Fu et al. [79] developed the novel MIL-101(Fe)-1-4((ethyl)phenyl)urea (MIL-101(Fe)-EPU) by grafting a phenylethyl side chain onto NH2-MIL-101(Fe). This modification imparted superior hydrophobicity and water stability to the material, resulting in high degradation efficiency for tetrabromobisphenol A. The mechanism was attributed to the oxidation of hydroxyl groups, enabling the degradation of pollutants. Yi et al. [80] achieved excellent degradation of chloroquine phosphate using PDINH/ MIL-88A (Fe) composites prepared through a ball milling strategy. The proposed mechanism is shown in Figure 3a. Yan et al. [81] synthesized CNT@MIL-101 (Fe) with outstanding visible light absorption, enabling efficient removal of ciprofloxacin under visible light irradiation. Du et al. [82] accomplished the efficient removal of oxytetracycline (OTC) utilizing a magnetic nanocomposite, MIL-101(Fe)/γ-Fe2O3, synthesized by a hydrothermal method. This catalyst could be recovered magnetically, enhancing its recyclability (Figure 3b). By modifying the ligand structure and fine-tuning the coordination of 2-MI to control the crystal size and morphology of Fe-MOFs, Cheng [83] et al. achieved the first one-pot hydrogenation and N-alkylation reaction of benzopropionitrile with alcohols, demonstrating excellent performance. Moreover, visible light irradiation was found to significantly enhance the activation efficiency compared to heating through in situ infrared. Liu et al. [84] obtained a magnetic Fe3O4@MIL-53(Fe) photocatalyst by calcining pristine MIL-53(Fe). This material exhibited high-performance degradation of ibuprofen under visible light irradiation.

2.3. Photo-Fenton System

Metal–organic frameworks (MOFs), particularly iron-based MOFs (Fe-MOFs), are emerging as promising Fenton-like catalysts due to their well-developed pores and available active sites [85,86,87]. However, the strong coordination between metal ions and organic ligands often reduces the exposure of active sites in most iron-based MOFs, limiting Lewis acid site availability and thus hampering the activation efficiency of H2O2. In addition, the lower cycling efficiency of Fe3+/Fe2+ further limits the activation efficiency of H2O2. Therefore, there is a significant focus on improving the cycling efficiency of Fe2+/Fe3+. Horiuchi et al. [88] discovered that MIL-101 (Fe) promotes photocatalytic water oxidation under visible light irradiation, producing oxygen from a silver nitrate solution. The finely dispersed Fe-O clusters had a very effective positive impact on the reaction. Huang et al. [89] synthesized a surface-coated Fe-MOF with recrystallized PDI (perylene-3,4,9,10-tetracarboxylic diimide), where the Fe-MOF in the northern part of the coating remained intact due to the stabilized H-type π–π stacking of the PDI under acidic conditions. The resulting Fe-MOF@PDI photocatalyst exhibited increased acid resistance, resulting in enhanced Fenton activity and stability. A series of modulations of α-Fe2O3 at the nanoscale using surface engineering was carried out by Xu et al. [90], resulting in catalysts with various shapes, such as spherical, octahedral, spindle, rod-shaped, and controllable oxygen vacancies (Ovs) of α-Fe2O3-x. Based on the photocatalytic results, rod-shaped α-Fe2O3-x exhibited the best performance in the complete degradation of methylene blue (MB). This performance was attributed to the synergistic effect of a large specific surface area, a high concentration of Fe2+, and Ovs. Using Fe-MOFs, Wu et al. [91] demonstrated, for the first time, that visible light can accelerate the Fe2+/Fe3+ cycling in a photo-Fenton system based on Fe-O clusters within the framework. This acceleration promoted the degradation of tetracycline hydrochloride. Utilizing a solid acid catalyst, MIL-101 effectively extended the pH range up to 10.2 while also demonstrating good reusability and stability (Figure 4a,b). The proposed mechanism is presented in Figure 4c.
Guo et al. [92] accomplished the degradation of tetracycline hydrochloride (HT-HCl) under visible light by hydrothermally synthesizing a series of mixed-valence coordinated unsaturated metal sites (CUSs) with varying Fe2+/Fe3+ ratios in MIL-100 (Fe). This is due to the fact that visible light irradiation accelerated the cycling of Fe2+/Fe3+, resulting in enhanced catalytic performance. The synergistic relationship between the porous structure of Fe-MOFs and the exposure of active sites for APO catalysis was intensively investigated by Hu et al. [93] (Figure 5a). It was found that the ozone-catalyzed reaction activity of MIL-53(Fe) was regulated by the evolution of crystal structures over time. Six evolutionary steps of MIL-53(Fe) crystals under the action of intramolecular hydrogen bonding were revealed, leading to a volcano plot depicting the relationship between rhodamine degradation-catalyzed ozone reactivity and crystallization time (Figure 5b–f).
Yang et al. [94] modified the structure of the Fe-O node in MIL-53(Fe) by substituting Fe2+ to obtain a catalyst with mixed valence coexistence. This modification led to excellent catalytic degradation of 4-nitrophenol. The substitution of Fe2+ introduced higher active sites than the original Fe3+ center (Fe2+→Fe3+ half-reaction) rapidly activating H2O2 and efficiently destroying 4-nitrophenol. It was also found that the accelerated cycle of Fe2+ and Fe3+ coupling promoted increasing ·OH generation and allowed more 4-NP to be degraded. Using Fe-MOF with more Lewis acid sites (LAS), Crittrnden et al. [95] obtained a synergistic effect that enhanced photogenerated carrier separation when ozone served as the electron acceptor. It was also demonstrated that ozone not only reduced the photogenerated carrier recombination but also decomposed on the LAS of MIL-88A (Fe) to generate more reactive oxygen species, further promoting the degradation of 4-nitrophenol (Figure 6a–c). Fe-BDC1, synthesized by Wu et al. [96] under simple stirring conditions, achieved high-performance degradation of Rhodamine B. The degradation of Rhodamine B was achieved by Fe-BDC1. Iron coordination unsaturated sites (Fe CUSs) and ·OH were also identified as active species for efficient degradation of the pollutant (Figure 6d,e). Tang et al. [97] accomplished the degradation of ibuprofen by anchoring ReS2 nanoparticles in a heterojunction formed in MIL-88B (Fe) under visible light and persulfate (PS). Song et al. [98] obtained mixed-valent Cu@MIL-88B(Fe) after processing the Fe-oxo node structure in MIL-88B(Fe) by the Cu+ species substitution method. Both experimental and computational results showed that Cu+ played a shuttling role in promoting the transfer between Fe2+/Fe3+ and inducing the formation of a large number of stable Fe2+ sites, as well as suppressing the leaching of iron ions during the catalytic process. These modifications significantly improved the performance of the Fenton-like reaction (Figure 6f).

2.4. Hexavalent Chromium Removal

Cr (VI) is a common highly toxic water pollutant originating from various sources like leather tanning, cooling tower discharge, electroplating, and anodizing baths. It poses harmful effects by penetrating the cell wall, unlike the less toxic Cr (III), which is almost insoluble at neutral pH. Therefore, reducing Cr (VI) to Cr (III) is a key process for the effective removal of Cr (VI) from water and wastewater [99,100].
The reduction of Fe3+ to Fe2+ in the linker-to-metal clusters of Fe-MOFs is crucial for charge transfer. Wang et al. [101] constructed a series of MIL-53(Fe) and Bi12O17Cl2 (MBx) composites by ball milling, demonstrating superior Cr(VI) degradation performance under visible light irradiation. They also confirmed that Fe-μ3-oxo clusters in MIL-53(Fe) triggered the Cr(VI) reduction process through linker-to-metal-cluster charge transfer (LCCT), which reduced Fe3+ to Fe2+, facilitating Cr(VI) removal. Graphite-like phase C3N4, a metal-free semiconductor material known for its π–π conjugated electronic structure, narrow bandgap (~2.7 eV), excellent chemical stability, good thermal stability, and low cost, is increasingly used for Fe-MOF modification (Figure 7a). Zhang [102] et al. successfully synthesized four novel iron-based metal–organic frameworks (MOFs) through ligand modulation to achieve improved Cr(VI) reduction properties. MTBDC-TPT-Fe exhibited the best performance due to enhanced electron push–pull effects between the iron–oxygen clusters and the organic ligands. The introduction of the -SCH3 group enhanced light absorption of the MOFs and provided electrons for the Fe centers, while TPT enhanced the photogenerated charge carriers’ separation and transfer. Wang et al. [103] introduced p-type carbon nitride (CN75) nanoparticles into NH2-MIL-53(Fe) through a solvothermal method to form p-n heterojunctions. This promoted charge carrier separation and migration, broadened visible light response, and achieved efficient hexavalent chromium photoreduction. At the same time, the sample also exhibited excellent cyclic stability and structural stability. Gao [104] et al. synthesized Fe-2MI compositions with visible light response using 2-methylimidazole (2MI) as an organic ligand. This led to the formation of complexes with Fe2+ and Fe3+ where Fe2+ formed a more stable complex with a higher coordination ratio with 2MI. Fe2+-2MI had suitable conduction and valence band positions for the reduction of Cr (VI) and organic matter oxidation, demonstrated in the mechanism study (Figure 7b). From the right side (A) of Figure 7b, it can be seen that compared to other samples, Fe (II)-2MI exhibited the best activity for Cr (VI) reduction. In the comparative experiment of different organic ligands (2MI, H2BDC-NH2-H2BDC, and FA), it can be seen that Fe (II)-2MI still exhibited the best performance (Figure 7b, right (B)). Meanwhile, He et al. [105] synthesized Fe-BDC/Fe-2MI heterojunctions via a one-pot method. They observed the presence of MIL-88B (Fe) in the Fe-BDC/Fe-2MI complexes, unlike MIL-53 (Fe) in bare Fe-BDCs, indicating the modulating effect of 2MI on the structure of Fe-BDCs.

2.5. Other Water Pollutants

In addition to dyes, antibiotics, and heavy metals, Fe-MOF photocatalysts have shown promise in the removal of various other water pollutants. Oladipo et al. [106] achieved the decomposition of two organophosphorus pesticides, methylmalathion (MP) and chlorpyrifos (CP), under sunlight irradiation using AgIO3/MIL-53(Fe). The individual degradation rates of CP and MP were 78% and 90%, respectively, under 60 min of sunlight irradiation in both tap water and distilled water. The individual degradation rates of CP and MP in tap water and distilled water were 78% and 90%, respectively, under 60 min sunlight irradiation. Zhong et al. [107] prepared Cu2O/MIL(Fe/Cu) composites through an in situ copper bridge strategy to enhance visible light absorption and achieve efficient degradation of thiacloprid (TCL). It was demonstrated that the Cu bridge promoted the charge interfacial transfer and Fe2+/Fe3+ redox reaction between Cu2O and MIL(Fe/Cu). Ahmad et al. [108] successfully synthesized stable mesoporous MIL-100(Fe)-loaded ZnO NS using an in situ self-assembly method. This composite achieved high-performance removal of chlorpyrifos and malathion–methyl. It was also found that the catalytic activity of mesoporous MIL-100(Fe) was significantly enhanced thanks to the mesoporous structure, which improved the diffusion and accessibility of molecules within the MOF channel.

2.6. Simultaneous Removal of Multiple Water Pollutants

In recent years, researchers have been exploring the simultaneous removal of heavy metals and the byproduction of valuable products during the process of removing organic pollutants [109]. Wang’s group prepared Z-Scheme WO3/MIL-100(Fe)(MxWy) composites using a ball milling strategy [110]. This composite exhibited high performance in the removal of both hexavalent chromium and BPA. The successful synthesis of the Z-Scheme structure was confirmed with a series of characterizations, and it was found to contribute positively to the efficient removal of Cr (VI) and BPA (Figure 8a,b).
In addition to introducing an oxygen source with strong electron trapping ability into the photocatalytic system of MOFs, another strategy to enhance the photocatalytic activity of MOFs is to prepare heterostructured materials by compositing with photoresponsive materials (e.g., TiO2, g-C3N4)/strong electron transporting materials (e.g., (oxidized) graphene). Similarly, the role of such strategies ultimately is to promote photogenerated electron and hole separation [32,52,111]. Liu et al. [112] prepared a sandwich-like heterostructure, TiO2NS@MIL-100(Fe), by combining MIL-100(Fe) with two-dimensional TiO2 nanosheets via a self-assembly method for the photocatalytic degradation of methylene blue. The interface between MIL-100(Fe) and TiO2 nanosheets facilitated the rapid photogenerated electron transfer and conversion and enhanced the photocatalytic activity of TiO2NS@MIL-I00(Fe). The g-C3N4/NH2-MIL-101 (Fe) prepared via the in situ solvothermal method by Liu et al. [113] achieved efficient photocatalytic removal of Cr (VI) and MO. The pH was found to be crucial for the reduction of Cr (VI). It was also found by infrared spectroscopy that as the content of g-C3N4 increased, the corresponding absorption became stronger, indicating the successful incorporation of g-C3N4 into the composite. The Z-Scheme CBO/MIL-88A(Fe)p-n heterojunction constructed by Li et al. [114] achieved nearly 100% removal of hexavalent chromium within 30 min under visible light irradiation. In terms of antimicrobial activity, complete inactivation was also achieved within 60 min. Notably, the leaching of iron ions was minimal after several cycles (Figure 8c).
The heterostructure of Fe2O3@Ag-ZnO@C of Z-Scheme designed by Li et al. [115] demonstrated efficient degradation of tetracycline and methylene blue in wastewater. Through a combination of experimental and computational results, it was found that the Fe2O3@ZnO nano-interface and the carbon sheath worked together to enhance the efficiency of photogenerated carrier transfer and absorption capacity. Additionally, the treated wastewater showed good biocompatibility (Figure 8d). Nguyen et al. [116] utilized a microwave-assisted solvothermal method to synthesize a series of M/Fe-MOFs (M = Ni, Mg, and Sn) for the high-performance and sustainable removal of different dyes, including rhodamine B (RhB), crystal violet (CV), methyl orange (MO), and methylene blue (MB).
Figure 8. (a) The diagram of photocatalytic Cr(VI) sequestration and photo-Fenton BPA decomposition mechanism of M80W120; (b) HRTEM images of photo-deposited PbO2 over M80W120. Reprinted with permission from Ref. [110]. Copyright 2021, Elsevier. (c) Proposed charge transfer mechanism toward Cr(VI) reduction and antibacterial applications by CBO/MIL-88(Fe). Reprinted with permission from Ref. [114]. Copyright 2023, Elsevier. (d) Proposed Z-scheme photocatalytic system of FAZ. Reprinted with permission from Ref. [115]. Copyright 2022, American Chemical Society.
Figure 8. (a) The diagram of photocatalytic Cr(VI) sequestration and photo-Fenton BPA decomposition mechanism of M80W120; (b) HRTEM images of photo-deposited PbO2 over M80W120. Reprinted with permission from Ref. [110]. Copyright 2021, Elsevier. (c) Proposed charge transfer mechanism toward Cr(VI) reduction and antibacterial applications by CBO/MIL-88(Fe). Reprinted with permission from Ref. [114]. Copyright 2023, Elsevier. (d) Proposed Z-scheme photocatalytic system of FAZ. Reprinted with permission from Ref. [115]. Copyright 2022, American Chemical Society.
Molecules 28 07121 g008

3. The Application of Fe-MOF Photocatalytic Removal and Conversion of Gaseous Substances

The distinctive structural characteristics and presence of unsaturated metal sites in Fe-MOFs have led to numerous applications in the field of the photocatalytic removal of gaseous pollutants. These applications predominantly revolve around the photoreduction of CO2, the elimination of NOx, nitrogen fixation, and the abatement of VOCs.

3.1. CO2 Removal

The imperative to mitigate CO2 emissions has emerged as a pressing environmental concern. At the same time, combating climate change through controlling CO2 emissions has become an international consensus. CO2 capture methods primarily encompass absorption, adsorption, biofixation, membrane separation, and photocatalysis. Among them, photocatalysis has been widely used due to its environmentally friendly and low-carbon attributes. The utilization of MOF materials in photocatalytic CO2 removal represents a burgeoning research area in recent years, with Fe-based MOFs occupying an indispensable role [117,118,119,120,121,122,123,124,125,126].
Subsequently, a series of NH2-MIL-101 (Fe) catalysts prepared by Dao et al. [127] demonstrated an efficient and highly selective reduction in CO2 at the gas–solid interface. NH2-MIL-101(Fe) obtained by Wang et al. [128] via growing varying metal ions in situ onto covalent frameworks containing triazines exhibited remarkable CO2 reduction to CO and CH4. Dong et al. [129] synthesized MOF-PCN-250-Fe2M, featuring Fe2+/Fe3+ metal cluster nodes and open metal sites, and further enhanced its CO2 photoreduction performance through modulation of MII metal ions in the metal clusters (M = Mn, Zn, Ni, Co) (Figure 9a). Zhang et al. [130] synthesized Fe-soc-MOF (Fe-soc-O) with eight {111} crystalline facets exposed and Fe-soc-MOF (Fe-soc-M) with eight {111} crystalline facets and six {100} crystalline facets, achieving high-performance CO2 photoreduction. Notably, the surface-active sites of Fe-soc-MOF were found to be closely related to the crystal faces, with Fe-soc-O exhibiting the highest CO generation and Fe-soc-M displaying superior CO selectivity (Figure 9b,c).

3.2. NOx Removal and Nitrogen Fixation

Nitrogen oxides, inherently toxic and detrimental to the ozone layer, contribute to the formation of photochemical smog, acid deposition, and eutrophication of water bodies. Excessive levels also pose direct threats to human respiratory health, presenting a significant risk to humanity’s well-being. While widely used catalysts like V2O5-WO3 (MoO3)/TiO2 are effective, the presence of toxic vanadium can compromise their selectivity for nitrogen, leading to the production of N2O or NO at high temperatures [131,132,133]. It has been shown that Fe-MOFs can form a large number of coordination unsaturated metal centers (CUSs) serving as Lewis acid centers in the NH3-SCR catalytic system [134]. N2 desorbs into the gas phase, forming Brønsted acid sites on the backbone, where NH3 adsorption is more favorable, facilitating the reaction. Therefore, Fe-MOFs find extensive use in this industry. Li et al. [132] observed exceptional photocatalytic nitrogen fixation activity in a series of Fe-based MOFs (MIL-101 (Fe), MIL-100 (Fe), and MIL-88 (Fe)). Photocatalytic nitrogen fixation activity was compared between MIL-101 (Fe) and MIL-101 (Cr), with the former exhibiting an activity of 50.355 μmol/h, while the latter remained inactive. This led to the conclusion that the source of nitrogen fixation activity was the iron catalytic center. Given the similarity of the outer electronic structures of Cr and Fe, it is hypothesized that the two extra multi-site electrons of elemental Fe are the fundamental reason for the excellent photocatalytic nitrogen fixation activity. Nguyen et al. [135] synthesized MIL-101 (Fe) using a facile microwave solvothermal method and employed it as a photocatalyst for NOx degradation. The results showed that under the optimal reaction conditions (alumina matrix, initial NO concentration of 350 ppb, and relative humidity of 40–50%), photocatalytic degradation of NOx under sunlight irradiation of 0.1 g of the material reached an efficiency of up to 77%, offering optimal conditions for efficient photocatalytic degradation of NOx. MIL-101, when exposed to appropriate light, generates charges at the Fe33-OXO cluster or BDC (benzene dicarboxylic acid ligand) junction, which then oxidize and reduce H2O and O2 to ·OH and ·O2 for catalytic removal of pollutants. During irradiation, the competition between the guest molecules (H2O, NO, and O2) and the Fe active center significantly impacts denitrification performance. Chen et al. [136] utilized the H4abtc ligand attached to an iron cluster to create the stable metal–organic framework Fe-abtc, achieving nitrogen fixation under visible light. The stable structure formed by linking the tetracarboxylic acid to the metal cluster and the presence of N@N in the ligand is crucial to the sustainability of the photocatalytic reaction. Huang et al. [137] greatly promoted the photocatalytic nitrogen fixation reaction using oxygen-deficient MO@Fe/Ce-MOFs synthesized via dielectric barrier discharge (DBD) plasma. Their work provided a detailed description of the critical role played by oxygen defects in nitrogen fixation reactions (Figure 10a,b). Li et al. [132] demonstrated outstanding photocatalytic nitrogen fixation activity in a series of Fe-MOFs, highlighting the iron catalytic center and its two extra divalent electrons as the key factors behind this exceptional performance.

3.3. VOCs Removal

VOCs are known for their strong irritant and toxic properties, with some even being teratogenic and carcinogenic. Common treatment technologies for VOCs include adsorption, absorption, condensation, membrane separation, catalytic combustion, thermal incineration, biodegradation, photocatalytic methods, and plasma technology. MOFs, characterized by their substantial specific surface area and strong catalytic activity, hold significant promise for VOC treatment applications. They have been extensively studied and applied as catalysts for photocatalytic degradation of VOCs [120], including applications involving Fe-based MOFs. Li et al. [138] proposed an efficient and stable Fe-based metal-organic skeleton (Fe-MOF) in which adsorption and photodegradation regeneration alternated to achieve repetitive and efficient removal of VOCs. The Fe-MOF could be regenerated by converting adsorbed gases into CO2 under 1 solar (100 mw/cm2) irradiation, demonstrating excellent durability over 100 cycles of repeated adsorption and degradation. Jiang et al. [139] synthesized Pd/Fe MOFs nanocomposites through coordination engineering, achieving the conversion of various aldehydes, alcohols, and toluene into amide compounds and amines using only flue gas (or air) as an oxidant, with water as a byproduct. Chen et al. [140] synthesized MIL-100 (Fe) by adjusting the coordination of Fe3+ ions/α-Fe2O3, achieving efficient removal of ortho xylene. This study also confirms the critical role played by the reversible conversion of Fe3+ and Fe2+ in the oxidation of ortho-xylene and the effective generation of active radicals (Figure 11a). Subsequently, the team utilized MIL-100 (Fe)/MOX synjunction to efficiently capture BTXS (benzene, toluene, xylene, and styrene) using coordination of unsaturated acidic Fe3-O sites [141], participating in the PCO process through the conversion between Fe3+ and Fe2+ (Figure 11b). Qin et al. [142] prepared a Fe MOF derivative (M-300) using MIL-100 (Fe) as the precursor, exhibiting excellent performance in degrading VOCs and eliminating Escherichia coli under visible light. At the same time, it achieved high-performance visible light degradation of acetaldehyde. This was attributed to the exposure of unsaturated Fe2+ active sites, which promoted the transfer of photogenerated carriers and redox processes, thereby improving the overall photocatalytic performance of the system (Figure 11c).
The synthesis methods of relevant catalysts and the removal performance of pollutants are shown in Table 1.

4. Strengths and Challenges

In the field of photocatalysis, most conventional catalysts consist of semiconductor materials. However, Fe-MOFs offer significant advantages by substantially enhancing the porosity of the materials and finely regulating inherent Lewis acidity through the incorporation of Fe as a node. Moreover, the combination of Fe ion nodes with organic ligands as bridge compositions renders Fe-MOFs a promising catalyst for various applications. Fe-MOFs can also be effectively integrated with functional materials, such as metal oxides, carbon-based materials, and metal nanoparticles, to construct new composites.
The advantages of Fe-MOFs can be summarized as follows:
(1)
Fe-MOFs exhibit a wide spectral response, enabling them to catalyze chemical reactions within the visible light spectrum. This characteristic enhances their applicability. Fe-MOFs find applications in photocatalytic processes focused on degrading organic pollutants, converting CO2 through photocatalysis, nitrogen fixation, and degrading of gaseous organic pollutants.
(2)
Featuring uniform, stable, and infinitely elongated structural composition units, Fe-MOFs allow for precise structural regulation of composite materials by adjusting the structural composition of MOFs.
(3)
Fe-MOFs possess high porosity, an intricate pore structure, and an extremely large internal specific surface area. Additionally, the pore size is controllable, providing ideal conditions for introducing other precursors into the pore structure.
(4)
Fe-MOFs exhibit a soft nature and can be modified by altering external conditions during the synthesis process. Factors such as temperature, pressure, acidity, and alkalinity can be adjusted to tailor the material’s properties as needed.
However, to advance the application of Fe-MOFs in catalytic water pollution purification and atmospheric substance treatment, several key improvements are necessary:
(1)
While Fe-MOF photocatalytic materials have found applications in water pollutants and atmospheric treatment, the current study focus has primarily been on novel MOF photocatalysts and the effect of ligand variations on the photocatalytic activity of MOFs. There exists a notable gap in systematic research concerning the photocatalytic mechanism of MOF materials and the relationship between photocatalytic performance and material structure. Future research efforts should prioritize strengthening and enhancing our understanding in these areas.
(2)
Fe-MOFs encounter challenges related to stability, high production cost, complex preparation processes, and limited photocatalytic activity. While some auxiliary methods have shown promise in shortening the synthesis time of Fe-MOF photocatalytic materials and improving their catalytic activity, many of these studies remain in the experimental stage. It is crucial to further develop and scale up these techniques to facilitate their industrial application.
(3)
Current research has placed relatively less emphasis on the effective methods of separating used catalysts from the reactants. To enhance the practical industrial application of these materials, it is essential to explore approaches such as imparting magnetic properties to the materials or converting them into loaded catalysts. How to combine the good photocatalytic degradation function with recyclability and low cost remains a challenging yet crucial issue to address.
(4)
Utilizing advanced fine structure analysis and density functional theory calculations, researchers can delve deeper into the conformational relationship between the microstructure and macroscopic performance of the catalysts. This approach allows for a more profound understanding of the interfacial electron transfer mechanism, guiding the design and regulation of catalysts with improved catalytic activity.
(5)
The synergistic removal of various pollutants in complex environment settings should be further emphasized to maximize the excellent catalytic performance of Fe-MOF composite catalysts when addressing real-world challenges.

5. Conclusions and Outlook

In summary, this review highlights recent advancements in the utilization of Fe-MOFs-based catalysts for highly efficient photocatalytic degradation of water and atmospheric pollutants. It has been emphasized throughout this review that functionalized Fe-based MOFs, as well as their derivatives and complexes, have successfully addressed challenges related to photogenerated carrier susceptibility to complexation, Fe2+/Fe3+ transfer, and insufficient exposure of the active site compared to pristine Fe-based MOFs. These innovations have significantly expanded the range of applications for Fe-MOFs and have shown promising results in both water pollution and air substances remediation. This review provides a comprehensive overview of the current state of research on Fe-MOFs and their composites, highlighting advanced materials used in photocatalytic technologies aimed at achieving the complete removal and mineralization of pollutants present in water and the atmosphere.
Despite the existing challenges, Fe-MOFs hold a promising outlook in the removal of environmental pollutants. With the continued dedication of scholars from various fields, it is anticipated that Fe-MOFs will eventually find practical applications in water and atmospheric purification. In the future, the development trajectory of Fe-MOF photocatalytic materials can be explored through efforts aimed at reducing production time, achieving large-scale production, reducing the cost, simplifying the preparation processes, and enhancing the catalytic activity. As science and technology continue to advance, future Fe-MOF materials will usher in a more vigorous development in the field of photocatalysis.

Author Contributions

J.C.: Writing-original draft, funding acquisition. Y.P.: Conceptualization. Y.J.: Methodology and analysis. L.L.: Investigation and writing—review and editing. H.W.: Supervision and resources. K.L.: Resources. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the project funded by the China Post-doctoral Science Foundation (2023MD734182).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationship that could have appeared to influence the work reported in this paper.

Sample Availability

Not applicable.

References

  1. He, M.; Xu, Z.; Hou, D.; Gao, B.; Cao, X.; Ok, Y.S.; Rinklebe, J.; Bolan, N.S.; Tsang, D.C.W. Waste-derived biochar for water pollution control and sustainable development. Nat. Rev. Earth Environ. 2022, 3, 444–460. [Google Scholar] [CrossRef]
  2. Lin, L.; Yang, H.; Xu, X. Effects of Water Pollution on Human Health and Disease Heterogeneity: A Review. Front. Environ. Sci. 2022, 10, 880246. [Google Scholar] [CrossRef]
  3. Pandey, S.; Makhado, E.; Kim, S.; Kang, M. Recent developments of polysaccharide based superabsorbent nanocomposite for organic dye contamination removal from wastewater—A review. Environ. Res. 2023, 217, 114909. [Google Scholar] [CrossRef]
  4. Ahmadijokani, F.; Molavi, H.; Rezakazemi, M.; Tajahmadi, S.; Bahi, A.; Ko, F.; Aminabhavi, T.M.; Li, J.-R.; Arjmand, M. UiO-66 metal–organic frameworks in water treatment: A critical review. Prog. Mater. Sci. 2022, 125, 100904. [Google Scholar] [CrossRef]
  5. Choi, W.S.; Lee, H.J. Nanostructured Materials for Water Purification: Adsorption of Heavy Metal Ions and Organic Dyes. Polymers 2022, 14, 2183. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, Y.; Cui, X.; Zhang, P.; Wang, Y.; Lu, W. Synthesis of porphyrin porous organic polymers and their application of water pollution treatment: A review. Environ. Technol. Innov. 2023, 29, 102972. [Google Scholar] [CrossRef]
  7. Sun, P.; Wang, Z.; An, S.; Zhao, J.; Yan, Y.; Zhang, D.; Wu, Z.; Shen, B.; Lyu, H. Biochar-supported nZVI for the removal of Cr(VI) from soil and water: Advances in experimental research and engineering applications. J. Environ. Manag. 2022, 316, 115211. [Google Scholar] [CrossRef]
  8. Jaison, A.; Mohan, A.; Lee, Y.-C. Recent Developments in Photocatalytic Nanotechnology for Purifying Air Polluted with Volatile Organic Compounds: Effect of Operating Parameters and Catalyst Deactivation. Catalysts 2023, 13, 407. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Liu, H.; Gao, F.; Tan, X.; Cai, Y.; Hu, B.; Huang, Q.; Fang, M.; Wang, X. Application of MOFs and COFs for photocatalysis in CO2 reduction, H2 generation, and environmental treatment. EnergyChem 2022, 4, 100078. [Google Scholar] [CrossRef]
  10. Ajmal, Z.; Haq, M.U.; Naciri, Y.; Djellabi, R.; Hassan, N.; Zaman, S.; Murtaza, A.; Kumar, A.; Al-Sehemi, A.G.; Algarni, H.; et al. Recent advancement in conjugated polymers based photocatalytic technology for air pollutants abatement: Cases of CO2, NOx, and VOCs. Chemosphere 2022, 308, 136358. [Google Scholar] [CrossRef]
  11. Sharma, V.K.; Feng, M. Water depollution using metal-organic frameworks-catalyzed advanced oxidation processes: A review. J. Hazard. Mater. 2019, 372, 3–16. [Google Scholar] [CrossRef]
  12. Hu, J.; Song, J.; Han, X.; Wen, Q.; Yang, W.; Pan, W.; Jian, S.; Jiang, S. Fabrication of Ce-La-MOFs for defluoridation in aquatic systems: A kinetics, thermodynamics and mechanisms study. Sep. Purif. Technol. 2023, 314, 123562. [Google Scholar] [CrossRef]
  13. Song, J.; Yang, W.; Han, X.; Jiang, S.; Zhang, C.; Pan, W.; Jian, S.; Hu, J. Performance of Rod-Shaped Ce Metal–Organic Frameworks for Defluoridation. Molecules 2023, 28, 3492. [Google Scholar] [CrossRef] [PubMed]
  14. Musadiq Anis, S.; Habibullah Hashemi, S.; Nasri, A.; Sajjadi, M.; Eslamipanah, M.; Jaleh, B. Decorated ZrO2 by Au nanoparticles as a potential nanocatalyst for the reduction of organic dyes in water. Inorg. Chem. Commun. 2022, 141, 109489. [Google Scholar] [CrossRef]
  15. Nisa, M.U.; Abid, A.G.; Gouadria, S.; Munawar, T.; Alrowaili, Z.A.; Abdullah, M.; Al-Buriahi, M.S.; Iqbal, F.; Ehsan, M.F.; Ashiq, M.N. Boosted electron-transfer/separation of SnO2/CdSe/Bi2S3 heterostructure for excellent photocatalytic degradation of organic dye pollutants under visible light. Surf. Interfaces 2022, 31, 102012. [Google Scholar] [CrossRef]
  16. Liu, M.; Zheng, J.; Wang, L.; Hu, Z.; Rao, W.; Liu, Y.; Xie, Y.; Yu, C. Amino-modified magnetic glucose-based carbon composites for efficient Cr(VI) removal. J. Taiwan Inst. Chem. Eng. 2022, 136, 104419. [Google Scholar] [CrossRef]
  17. Almaie, S.; Vatanpour, V.; Rasoulifard, M.H.; Koyuncu, I. Volatile organic compounds (VOCs) removal by photocatalysts: A review. Chemosphere 2022, 306, 135655. [Google Scholar] [CrossRef]
  18. Yang, R.-G.; Fu, Y.-M.; Wang, H.-N.; Zhang, D.-P.; Zhou, Z.; Cheng, Y.-Z.; Meng, X.; He, Y.-O.; Su, Z.-M. ZIF-8/covalent organic framework for enhanced CO2 photocatalytic reduction in gas-solid system. Chem. Eng. J. 2022, 450, 138040. [Google Scholar] [CrossRef]
  19. Li, S.; Shan, S.; Chen, S.; Li, H.; Li, Z.; Liang, Y.; Fei, J.; Xie, L.; Li, J. Photocatalytic degradation of hazardous organic pollutants in water by Fe-MOFs and their composites: A review. J. Environ. Chem. Eng. 2021, 9, 105967. [Google Scholar] [CrossRef]
  20. Jian, S.; Cheng, Y.; Ma, X.; Guo, H.; Hu, J.; Zhang, K.; Jiang, S.; Yang, W.; Duan, G. Excellent fluoride removal performance by electrospun La–Mn bimetal oxide nanofibers. New J. Chem. 2022, 46, 490–497. [Google Scholar] [CrossRef]
  21. Wei, Q.; Xue, S.; Wu, W.; Liu, S.; Li, S.; Zhang, C.; Jiang, S. Plasma Meets MOFs: Synthesis, Modifications, and Functionalities. Chem. Rec. 2023, 23, e202200263. [Google Scholar] [CrossRef]
  22. Ren, X.; Gao, Z.; Wu, G. Tunable nano-effect of Cu clusters derived from MOF-on-MOF hybrids for electromagnetic wave absorption. Compos. Commun. 2022, 35, 101292. [Google Scholar] [CrossRef]
  23. Peng, X.; Chen, L.; Li, Y. Ordered macroporous MOF-based materials for catalysis. Mol. Catal. 2022, 529, 112568. [Google Scholar] [CrossRef]
  24. Jo, Y.M.; Jo, Y.K.; Lee, J.H.; Jang, H.W.; Hwang, I.S.; Yoo, D.J. MOF-Based Chemiresistive Gas Sensors: Toward New Functionalities. Adv. Mater. 2022, 2206842. [Google Scholar] [CrossRef]
  25. Luo, Y.; Bag, S.; Zaremba, O.; Cierpka, A.; Andreo, J.; Wuttke, S.; Friederich, P.; Tsotsalas, M. MOF Synthesis Prediction Enabled by Automatic Data Mining and Machine Learning. Angew. Chem. Int. Ed. Engl. 2022, 61, e202200242. [Google Scholar] [CrossRef] [PubMed]
  26. Yao, Y.; Wang, C.; Na, J.; Hossain, M.S.A.; Yan, X.; Zhang, H.; Amin, M.A.; Qi, J.; Yamauchi, Y.; Li, J. Macroscopic MOF Architectures: Effective Strategies for Practical Application in Water Treatment. Small 2022, 18, e2104387. [Google Scholar] [CrossRef] [PubMed]
  27. Tao, Y.; Yang, B.; Wang, F.; Yan, Y.; Hong, X.; Xu, H.; Xia, M.; Wang, F. Green synthesis of MOF-808 with modulation of particle sizes and defects for efficient phosphate sequestration. Sep. Purif. Technol. 2022, 300, 121825. [Google Scholar] [CrossRef]
  28. Jian, S.; Chen, Y.; Shi, F.; Liu, Y.; Jiang, W.; Hu, J.; Han, X.; Jiang, S.; Yang, W. Template-Free Synthesis of Magnetic La-Mn-Fe Tri-Metal Oxide Nanofibers for Efficient Fluoride Remediation: Kinetics, Isotherms, Thermodynamics and Reusability. Polymers 2022, 14, 5417. [Google Scholar] [CrossRef]
  29. Alikhani, N.; Hekmati, M.; Karmakar, B.; Veisi, H. Green synthesis of gold nanoparticles (Au NPs) using Rosa canina fruit extractand evaluation of its catalytic activity in the degradation of organic dye pollutants of water. Inorg. Chem. Commun. 2022, 139, 109351. [Google Scholar] [CrossRef]
  30. Beydaghdari, M.; Saboor, F.H.; Babapoor, A.; Asgari, M. Recent Progress in Adsorptive Removal of Water Pollutants by Metal-Organic Frameworks. ChemNanoMat 2021, 8, e202100400. [Google Scholar] [CrossRef]
  31. He, D.; Jiang, L.; Yuan, K.; Zhang, J.; Zhang, J.; Fan, Q.; Liu, H.; Song, S. Synthesis and study of low-cost nitrogen-rich porous organic polyaminals for efficient adsorption of iodine and organic dye. Chem. Eng. J. 2022, 446, 137119. [Google Scholar] [CrossRef]
  32. Ibrahim, I.; Belessiotis, G.V.; Antoniadou, M.; Kaltzoglou, A.; Sakellis, E.; Katsaros, F.; Sygellou, L.; Arfanis, M.K.; Salama, T.M.; Falaras, P. Silver decorated TiO2/g-C3N4 bifunctional nanocomposites for photocatalytic elimination of water pollutants under UV and artificial solar light. Results Eng. 2022, 14, 100470. [Google Scholar] [CrossRef]
  33. Du, C.; Zhang, Y.; Zhang, Z.; Zhou, L.; Yu, G.; Wen, X.; Chi, T.; Wang, G.; Su, Y.; Deng, F.; et al. Fe-based metal organic frameworks (Fe-MOFs) for organic pollutants removal via photo-Fenton: A review. Chem. Eng. J. 2022, 431, 133932. [Google Scholar] [CrossRef]
  34. Zhu, R.; Cai, M.; Fu, T.; Yin, D.; Peng, H.; Liao, S.; Du, Y.; Kong, J.; Ni, J.; Yin, X. Fe-Based Metal Organic Frameworks (Fe-MOFs) for Bio-Related Applications. Pharmaceutics 2023, 15, 1599. [Google Scholar] [CrossRef]
  35. Dong, J.; Wen, L.; Yang, H.; Zhao, J.; He, C.; Hu, Z.; Peng, L.; Hou, C.; Huo, D. Catalytic Hairpin Assembly-Driven Ratiometric Dual-Signal Electrochemical Biosensor for Ultrasensitive Detection of MicroRNA Based on the Ratios of Fe-MOFs and MB-GA-UiO-66-NH2. Anal. Chem. 2022, 94, 5846–5855. [Google Scholar] [CrossRef] [PubMed]
  36. Ma, X.; Zhao, S.; Tian, Z.; Duan, G.; Pan, H.; Yue, Y.; Li, S.; Jian, S.; Yang, W.; Liu, K.; et al. MOFs meet wood: Reusable magnetic hydrophilic composites toward efficient water treatment with super-high dye adsorption capacity at high dye concentration. Chem. Eng. J. 2022, 446, 136851. [Google Scholar] [CrossRef]
  37. Acharya, D.; Pathak, I.; Muthurasu, A.; Bhattarai, R.M.; Kim, T.; Ko, T.H.; Saidin, S.; Chhetri, K.; Kim, H.Y. In situ transmogrification of nanoarchitectured Fe-MOFs decorated porous carbon nanofibers into efficient positrode for asymmetric supercapacitor application. J. Energy Storage 2023, 63, 106992. [Google Scholar] [CrossRef]
  38. Liang, X.; Wang, C.; Yu, M.; Yao, Z.; Zhang, Y. Fe-MOFs derived porous Fe4N@carbon composites with excellent broadband electromagnetic wave absorption properties. J. Alloys Compd. 2022, 910, 164844. [Google Scholar] [CrossRef]
  39. Gallareta-Olivares, G.; Rivas-Sanchez, A.; Cruz-Cruz, A.; Hussain, S.M.; Gonzalez-Gonzalez, R.B.; Cardenas-Alcaide, M.F.; Iqbal, H.M.N.; Parra-Saldivar, R. Metal-doped carbon dots as robust nanomaterials for the monitoring and degradation of water pollutants. Chemosphere 2023, 312, 137190. [Google Scholar] [CrossRef]
  40. Saghir, S.; Pu, C.; Fu, E.; Wang, Y.; Xiao, Z. Synthesis of high surface area porous biochar obtained from pistachio shells for the efficient adsorption of organic dyes from polluted water. Surf. Interfaces 2022, 34, 102357. [Google Scholar] [CrossRef]
  41. Yuan, B.; Wang, X.; Zhou, X.; Xiao, J.; Li, Z. Novel room-temperature synthesis of MIL-100(Fe) and its excellent adsorption performances for separation of light hydrocarbons. Chem. Eng. J. 2019, 355, 679–686. [Google Scholar] [CrossRef]
  42. Cai, J.; Li, D.; Jiang, L.; Yuan, J.; Li, Z.; Li, K. Review on CeO2-Based Photocatalysts for Photocatalytic Reduction of CO2: Progresses and Perspectives. Energy Fuels 2023, 37, 4878–4897. [Google Scholar] [CrossRef]
  43. Wang, J.; Sun, Y.; Zhao, X.; Chen, L.; Peng, S.; Ma, C.; Duan, G.; Liu, Z.; Wang, H.; Yuan, Y.; et al. A poly(amidoxime)-modified MOF macroporous membrane for high-efficient uranium extraction from seawater. e-Polymers 2022, 22, 399–410. [Google Scholar] [CrossRef]
  44. Li, X.; Guo, W.; Liu, Z.; Wang, R.; Liu, H. Quinone-modified NH2-MIL-101(Fe) composite as a redox mediator for improved degradation of bisphenol A. J. Hazard. Mater. 2017, 324, 665–672. [Google Scholar] [CrossRef] [PubMed]
  45. Madima, N.; Kefeni, K.K.; Mishra, S.B.; Mishra, A.K. TiO2-modified g-C3N4 nanocomposite for photocatalytic degradation of organic dyes in aqueous solution. Heliyon 2022, 8, e10683. [Google Scholar] [CrossRef] [PubMed]
  46. Zhu, S.; Xu, J.; Wang, B.; Xie, J.; Ying, G.; Li, J.; Cheng, Z.; Li, J.; Chen, K. Highly efficient and rapid purification of organic dye wastewater using lignin-derived hierarchical porous carbon. J. Colloid. Interface Sci. 2022, 625, 158–168. [Google Scholar] [CrossRef]
  47. Sharma, P.; Singh, S.P.; Parakh, S.K.; Tong, Y.W. Health hazards of hexavalent chromium (Cr (VI)) and its microbial reduction. Bioengineered 2022, 13, 4923–4938. [Google Scholar] [CrossRef]
  48. Joseph, J.; Iftekhar, S.; Srivastava, V.; Fallah, Z.; Zare, E.N.; Sillanpaa, M. Iron-based metal-organic framework: Synthesis, structure and current technologies for water reclamation with deep insight into framework integrity. Chemosphere 2021, 284, 131171. [Google Scholar] [CrossRef] [PubMed]
  49. Jin, Y.; Zhang, T.; Pan, N.; Wang, S.; Zhang, B.; Zhu, X.; Hao, Y.; Wang, X.; Song, L.; Zhang, M. Surface functionalization of carbon cloth with conductive Ni/Fe-MOFs for highly efficient oxygen evolution. Surf. Interfaces 2022, 33, 102294. [Google Scholar] [CrossRef]
  50. Ronda-Leal, M.; Osman, S.M.; Won Jang, H.; Shokouhimehr, M.; Romero, A.A.; Luque, R. Selective hydrogenation of furfural using TiO2-Fe2O3/C from Ti-Fe-MOFs as sacrificial template: Microwave vs Continuous flow experiments. Fuel 2023, 333, 126221. [Google Scholar] [CrossRef]
  51. Sun, C.; Li, C.; Guo, M.; Yang, X.; Luo, Y.; Chen, L.; Zheng, H.; Zhao, S.; Li, F. Fabrication and optimization of paper chips from calcinated Fe-MOFs for rapid and in situ visual detection of tetracyclines in water environments. J. Hazard. Mater. 2023, 458, 131946. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.; Zhao, G.; Zhang, G.; Zhang, Y.; Wang, H.; Cao, W.; Li, T.; Wei, Q. An electrochemical aptasensor based on gold-modified MoS2/rGO nanocomposite and gold-palladium-modified Fe-MOFs for sensitive detection of lead ions. Sens. Actuators B Chem. 2020, 319, 128313. [Google Scholar] [CrossRef]
  53. Wan, Y.; Wan, J.; Ma, Y.; Wang, Y.; Luo, T. Sustainable synthesis of modulated Fe-MOFs with enhanced catalyst performance for persulfate to degrade organic pollutants. Sci. Total Environ. 2020, 701, 134806. [Google Scholar] [CrossRef] [PubMed]
  54. Xia, Q.; Wang, H.; Huang, B.; Yuan, X.; Zhang, J.; Zhang, J.; Jiang, L.; Xiong, T.; Zeng, G. State-of-the-Art Advances and Challenges of Iron-Based Metal Organic Frameworks from Attractive Features, Synthesis to Multifunctional Applications. Small 2019, 15, e1803088. [Google Scholar] [CrossRef]
  55. Xie, A.D.; Hu, M.-G.; Luo, Y.-H.; Zhu, X.-G.; Wang, Z.-H.; Geng, W.-Y.; Zhang, H.; Zhang, D.-E.; Zhang, H. Synthesis of a stable iron(iii)–organic framework for a visible light induced simultaneous photocatalytic reduction of Cr(vi) and the degradation of organic dyes in water. New J. Chem. 2021, 45, 13406–13414. [Google Scholar] [CrossRef]
  56. Yang, M.; Zhou, Y.-N.; Cao, Y.-N.; Tong, Z.; Dong, B.; Chai, Y.-M. Advances and Challenges of Fe-MOFs Based Materials as Electrocatalysts for Water Splitting. Appl. Mater. Today 2020, 20, 100692. [Google Scholar] [CrossRef]
  57. Wang, K.; Huang, X.; Zhou, T.; Wang, H.; Xie, H.; Ren, Y. Boosted electrochemical properties of porous Li2FeSiO4/C based on Fe-MOFs precursor for lithium ion batteries. Vacuum 2020, 171, 108997. [Google Scholar] [CrossRef]
  58. Zhang, M.; Cao, A.; Zhang, H.; Zhao, Y.; Su, X.; Wang, L.; Wu, R.; Yang, C. Urchin-like hybrid nanostructures of CuOx/Fe2O3 from Cu-mediated pyrolysis of Fe-MOFs for catalytic reduction of organic pollutants. Nanoscale 2022, 14, 1826–1833. [Google Scholar] [CrossRef]
  59. Cheng, H.; Zang, C.; Bian, F.; Jiang, Y.; Yang, L.; Dong, F.; Jiang, H. Boosting free radical type photocatalysis over Pd/Fe-MOFs by coordination structure engineering. Catal. Sci. Technol. 2021, 11, 5543–5552. [Google Scholar] [CrossRef]
  60. Jin, X.; Tang, T.; Tao, X.; Huang, L.; Shang, S. Regulating N content to anchor Fe in Fe-MOFs: Obtaining multiple active sites as efficient photocatalysts. J. Taiwan Inst. Chem. Eng. 2022, 132, 104133. [Google Scholar] [CrossRef]
  61. Jin, X.; Tang, T.; Tao, X.; Huang, L.; Xu, D. A novel dual-ligand Fe-based MOFs synthesized with dielectric barrier discharge (DBD) plasma as efficient photocatalysts. J. Mol. Liq. 2021, 340, 117290. [Google Scholar] [CrossRef]
  62. Wang, C.; Wang, L.; Tao, X.; Huang, L.; Yang, Z.; Guo, Q. Novel dielectric barrier discharge (DBD) plasma method for preparation of Fe-MOFs@Fe2O3 composites to treat waste polyethylene terephthalate (PET). J. Environ. Chem. Eng. 2023, 11, 109281. [Google Scholar] [CrossRef]
  63. Yu, Y.; Yu, C.; Niu, Y.; Chen, J.; Zhao, Y.; Zhang, Y.; Gao, R.; He, J. Target triggered cleavage effect of DNAzyme: Relying on Pd-Pt alloys functionalized Fe-MOFs for amplified detection of Pb2+. Biosens. Bioelectron. 2018, 101, 297–303. [Google Scholar] [CrossRef]
  64. Yao, J.; Xie, Z.; Zeng, X.; Wang, L.; Yue, T. Bimetallic Eu/Fe-MOFs ratiometric fluorescent nanoenzyme for selective cholesterol detection in biological serum: Synthesis, characterization, mechanism and DFT calculations. Sens. Actuators B Chem. 2022, 354, 130760. [Google Scholar] [CrossRef]
  65. Geng, N.; Chen, W.; Xu, H.; Ding, M.; Lin, T.; Wu, Q.; Zhang, L. Insights into the novel application of Fe-MOFs in ultrasound-assisted heterogeneous Fenton system: Efficiency, kinetics and mechanism. Ultrason. Sonochem. 2021, 72, 105411. [Google Scholar] [CrossRef] [PubMed]
  66. Gao, Y.; Li, S.; Li, Y.; Yao, L.; Zhang, H. Accelerated photocatalytic degradation of organic pollutant over metal-organic framework MIL-53(Fe) under visible LED light mediated by persulfate. Appl. Catal. B Environ. 2017, 202, 165–174. [Google Scholar] [CrossRef]
  67. Tran, K.N.T.; Tran, T.B.; Do, S.T.; Nguyen, K.O.T.; Lam, T.V. Synthesis of Mn-Doped Fe-MOFs with Different Ratios and Its Application for Photocatalytic Degradation of Rhodamine B Dye. Indones. J. Chem. 2022, 22, 96–104. [Google Scholar] [CrossRef]
  68. Ma, Y.; Lu, Y.; Hai, G.; Dong, W.; Li, R.; Liu, J.; Wang, G. Bidentate carboxylate linked TiO2 with NH2-MIL-101(Fe) photocatalyst: A conjugation effect platform for high photocatalytic activity under visible light irradiation. Sci. Bull. 2020, 65, 658–669. [Google Scholar] [CrossRef]
  69. Brahmi, C.; Benltifa, M.; Vaulot, C.; Michelin, L.; Dumur, F.; Gkaniatsou, E.; Sicard, C.; Airoudj, A.; Morlet-Savary, F.; Bousselmi, L.; et al. New Hybrid Fe-based MOFs/Polymer Composites for the Photodegradation of Organic Dyes. ChemistrySelect 2021, 6, 8120–8132. [Google Scholar] [CrossRef]
  70. Ngan Tran, T.K.; Ho, H.L.; Nguyen, H.V.; Tran, B.T.; Nguyen, T.T.; Thi Bui, P.Q.; Bach, L.G. Photocatalytic degradation of Rhodamine B in aqueous phase by bimetallic metal-organic framework M/Fe-MOF (M = Co, Cu, and Mg). Open Chem. 2022, 20, 52–60. [Google Scholar] [CrossRef]
  71. Mousavi, S.E.; Younesi, H.; Bahramifar, N.; Tamunaidu, P.; Karimi-Maleh, H. A novel route to the synthesis of alpha-Fe2O3@C@SiO2/TiO2 nanocomposite from the metal-organic framework as a photocatalyst for water treatment. Chemosphere 2022, 297, 133992. [Google Scholar] [CrossRef]
  72. Zhao, Q.; Zhang, L.; Wang, X.; Jia, X.; Xu, P.; Zhao, M.; Dai, R. Simultaneous efficient adsorption and photocatalytic degradation of methylene blue over iron(III)-based metal–organic frameworks: A comparative study. Transit. Met. Chem. 2019, 44, 789–797. [Google Scholar] [CrossRef]
  73. Gong, Q.; Liu, Y.; Dang, Z. Core-shell structured Fe3O4@GO@MIL-100(Fe) magnetic nanoparticles as heterogeneous photo-Fenton catalyst for 2,4-dichlorophenol degradation under visible light. J. Hazard. Mater. 2019, 371, 677–686. [Google Scholar] [CrossRef]
  74. Li, S.; Wu, F.; Lin, R.; Wang, J.; Li, C.; Li, Z.; Jiang, J.; Xiong, Y. Enabling photocatalytic hydrogen production over Fe-based MOFs by refining band structure with dye sensitization. Chem. Eng. J. 2022, 429, 132217. [Google Scholar] [CrossRef]
  75. Zhang, X.; Ma, X.; Ye, Y.; Guo, C.; Xu, X.; Zhou, J.; Wang, B. Enhanced photocatalytic hydrogen evolution with a Mixed-Valence iron Metal-Organic framework. Chem. Eng. J. 2023, 456, 140939. [Google Scholar] [CrossRef]
  76. Dai, F.; Wang, Y.; Zhou, X.; Zhao, R.; Han, J.; Wang, L. ZnIn2S4 decorated Co-doped NH2-MIL-53(Fe) nanocomposites for efficient photocatalytic hydrogen production. Appl. Surf. Sci. 2020, 517, 146161. [Google Scholar] [CrossRef]
  77. Li, H.; Liu, X.; Song, M.; Feng, H.; Yu, J.; Liu, B.; Ren, Q.; Yang, Z.; Xu, M.; He, Y. Non-noble copper ion anchored on NH2-MIL-101(Fe) as a novel cocatalyst with transient metal centers for efficient photocatalytic water splitting. J. Alloys Compd. 2022, 905, 164153. [Google Scholar] [CrossRef]
  78. He, W.; Li, Z.; Lv, S.; Niu, M.; Zhou, W.; Li, J.; Lu, R.; Gao, H.; Pan, C.; Zhang, S. Facile synthesis of Fe3O4@MIL-100(Fe) towards enhancing photo-Fenton like degradation of levofloxacin via a synergistic effect between Fe3O4 and MIL-100(Fe). Chem. Eng. J. 2021, 409, 128274. [Google Scholar] [CrossRef]
  79. Fu, J.; Wang, L.; Chen, Y.; Yan, D.; Ou, H. Enhancement of aqueous stability of NH2-MIL-101(Fe) by hydrophobic grafting post-synthetic modification. Environ. Sci. Pollut. Res. Int. 2021, 28, 68560–68571. [Google Scholar] [CrossRef]
  80. Yi, X.-H.; Ji, H.; Wang, C.-C.; Li, Y.; Li, Y.-H.; Zhao, C.; Wang, A.; Fu, H.; Wang, P.; Zhao, X.; et al. Photocatalysis-activated SR-AOP over PDINH/MIL-88A(Fe) composites for boosted chloroquine phosphate degradation: Performance, mechanism, pathway and DFT calculations. Appl. Catal. B Environ. 2021, 293, 120229. [Google Scholar] [CrossRef]
  81. Yan, D.; Hu, H.; Gao, N.; Ye, J.; Ou, H. Fabrication of carbon nanotube functionalized MIL-101(Fe) for enhanced visible-light photocatalysis of ciprofloxacin in aqueous solution. Appl. Surf. Sci. 2019, 498, 143836. [Google Scholar] [CrossRef]
  82. Du, C.; Zhang, Y.; Zhang, Z.; Song, D.; Cao, J.; Yu, H.; Yu, G.; Zhou, L.; Su, Y.; Lv, Y.; et al. Highly efficient removal of oxytetracycline using activated magnetic MIL-101(Fe)/gamma-Fe2O3 heterojunction catalyst. J. Environ. Manag. 2022, 317, 115327. [Google Scholar] [CrossRef] [PubMed]
  83. Cheng, H.; Long, X.; Bian, F.; Yang, C.; Liu, X.; Jiang, H. Efficient photocatalytic one-pot hydrogenation and N-alkylation of nitrobenzenes/benzonitriles with alcohols over Pd/MOFs: Effect of the crystal morphology & “quasi-MOF” structure. J. Catal. 2020, 389, 121–131. [Google Scholar] [CrossRef]
  84. Liu, N.; Wang, J.; Wu, J.; Li, Z.; Huang, W.; Zheng, Y.; Lei, J.; Zhang, X.; Tang, L. Magnetic Fe3O4@MIL-53(Fe) nanocomposites derived from MIL-53(Fe) for the photocatalytic degradation of ibuprofen under visible light irradiation. Mater. Res. Bull. 2020, 132, 111000. [Google Scholar] [CrossRef]
  85. Wang, Z.; Cheng, Y.; Wang, C.; Guo, R.; You, J.; Zhang, H. Optimizing the performance of Fe-based metal-organic frameworks in photo-Fenton processes: Mechanisms, strategies and prospects. Chemosphere 2023, 339, 139673. [Google Scholar] [CrossRef] [PubMed]
  86. Li, S.; Luo, P.; Wu, H.; Wei, C.; Hu, Y.; Qiu, G. Strategies for Improving the Performance and Application of MOFs Photocatalysts. ChemCatChem 2019, 11, 2978–2993. [Google Scholar] [CrossRef]
  87. Yi, J.; Wu, X.; Wu, H.; Guo, J.; Wu, K.; Zhang, L. Facile synthesis of novel NH2-MIL-53(Fe)/AgSCN heterojunction composites as a highly efficient photocatalyst for ciprofloxacin degradation and H2 production under visible-light irradiation. React. Chem. Eng. 2022, 7, 84–100. [Google Scholar] [CrossRef]
  88. Horiuchi, Y.; Toyao, T.; Miyahara, K.; Zakary, L.; Van, D.D.; Kamata, Y.; Kim, T.H.; Lee, S.W.; Matsuoka, M. Visible-light-driven photocatalytic water oxidation catalysed by iron-based metal-organic frameworks. Chem. Commun. 2016, 52, 5190–5193. [Google Scholar] [CrossRef]
  89. Huang, W.; Shao, H.; Song, M.; Yang, Z.; Li, G.; Liao, X. Perylene diimides coated Fe-MOFs as acid-tolerant photo-Fenton catalyst for phenol removal. Appl. Surf. Sci. 2021, 547, 149222. [Google Scholar] [CrossRef]
  90. Xu, W.; Xue, W.; Huang, H.; Wang, J.; Zhong, C.; Mei, D. Morphology controlled synthesis of α-Fe2O3-x with benzimidazole-modified Fe-MOFs for enhanced photo-Fenton-like catalysis. Appl. Catal. B Environ. 2021, 291, 120129. [Google Scholar] [CrossRef]
  91. Wu, Q.; Yang, H.; Kang, L.; Gao, Z.; Ren, F. Fe-based metal-organic frameworks as Fenton-like catalysts for highly efficient degradation of tetracycline hydrochloride over a wide pH range: Acceleration of Fe(II)/ Fe(III) cycle under visible light irradiation. Appl. Catal. B Environ. 2020, 263, 118282. [Google Scholar] [CrossRef]
  92. Guo, J.; Jia, H.; Zhang, A.; Pei, Z.; Luo, M.; Xue, J.; Shen, Q.; Liu, X.; Xu, B. MIL-100 (Fe) with mix-valence coordinatively unsaturated metal site as Fenton-like catalyst for efficiently removing tetracycline hydrochloride: Boosting Fe(III)/Fe(II) cycle by photoreduction. Sep. Purif. Technol. 2021, 262, 118334. [Google Scholar] [CrossRef]
  93. Hu, Q.; Zhang, M.; Xu, L.; Wang, S.; Yang, T.; Wu, M.; Lu, W.; Li, Y.; Yu, D. Unraveling timescale-dependent Fe-MOFs crystal evolution for catalytic ozonation reactivity modulation. J. Hazard. Mater. 2022, 431, 128575. [Google Scholar] [CrossRef]
  94. Yang, T.; Yu, D.; Wang, D.; Yang, T.; Li, Z.; Wu, M.; Petru, M.; Crittenden, J. Accelerating Fe(III)/Fe(II) cycle via Fe(II) substitution for enhancing Fenton-like performance of Fe-MOFs. Appl. Catal. B Environ. 2021, 286, 119859. [Google Scholar] [CrossRef]
  95. Yu, D.; Li, L.; Wu, M.; Crittenden, J.C. Enhanced photocatalytic ozonation of organic pollutants using an iron-based metal-organic framework. Appl. Catal. B Environ. 2019, 251, 66–75. [Google Scholar] [CrossRef]
  96. Wu, Q.; Siddique, M.S.; Yang, Y.; Wu, M.; Kang, L.; Yang, H. Facile and scalable synthesis of Fe-based metal organic frameworks for highly efficient photo-Fenton degradation of organic contaminants. J. Clean. Prod. 2022, 374, 134033. [Google Scholar] [CrossRef]
  97. Liu, N.; Dai, W.; Fei, F.; Xu, H.; Lei, J.; Quan, G.; Zheng, Y.; Zhang, X.; Tang, L. Insights into the photocatalytic activation persulfate by visible light over ReS2/MIL-88B(Fe) for highly efficient degradation of ibuprofen: Combination of experimental and theoretical study. Sep. Purif. Technol. 2022, 297, 121545. [Google Scholar] [CrossRef]
  98. Song, M.; Han, J.; Wang, Y.; Chen, L.; Chen, Y.; Liao, X. Effects and Mechanisms of Cu Species in Fe-MOFs on Fenton-Like Catalytic Activity and Stability. ACS Appl. Mater. Interfaces 2023, 15, 36201–36213. [Google Scholar] [CrossRef]
  99. Wei, F.H.; Chen, D.; Liang, Z.; Zhao, S.Q.; Luo, Y. Preparation of Fe-MOFs by microwave-assisted ball milling for reducing Cr(vi) in wastewater. Dalton. Trans. 2017, 46, 16525–16531. [Google Scholar] [CrossRef]
  100. Garazhian, Z.; Farrokhi, A.; Rezaeifard, A.; Jafarpour, M.; Khani, R. The enhanced visible-light-induced photocatalytic activities of bimetallic Mn-Fe MOFs for the highly efficient reductive removal of Cr(vi). RSC Adv. 2021, 11, 21127–21136. [Google Scholar] [CrossRef]
  101. Li, H.; Zhao, C.; Li, X.; Fu, H.; Wang, Z.; Wang, C.-C. Boosted photocatalytic Cr(VI) reduction over Z-scheme MIL-53(Fe)/Bi12O17Cl2 composites under white light. J. Alloy. Compd. 2020, 844, 156147. [Google Scholar] [CrossRef]
  102. Zhang, H.; Luo, Y.H.; Chen, F.Y.; Geng, W.Y.; Lu, X.X.; Zhang, D.E. Enhancing the spatial separation of photogenerated charges on Fe-based MOFs via structural regulation for highly-efficient photocatalytic Cr(VI) reduction. J. Hazard. Mater. 2023, 441, 129875. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, K.; Fu, Y.; Hao, D.; Guo, J.; Ni, B.-J.; Jiang, B.; Xu, L.; Wang, Q. Fabrication of CN75/NH2-MIL-53(Fe) p-n heterojunction with wide spectral response for efficiently photocatalytic Cr(VI) reduction. J. Alloys Compd. 2022, 891, 161994. [Google Scholar] [CrossRef]
  104. Gao, Q.; Lin, D.; Fan, Y.; He, Q.; Wang, Q. Visible light induced photocatalytic reduction of Cr(VI) by self-assembled and amorphous Fe-2MI. Chem. Eng. J. 2019, 374, 10–19. [Google Scholar] [CrossRef]
  105. He, Q.; Fu, Y.; Ge, X.; Al-Enizi, A.M.; Nafady, A.; Wang, Q.; Ma, S. Facile fabrication of Fe-BDC/Fe-2MI heterojunction with boosted photocatalytic activity for Cr(VI) reduction. J. Environ. Chem. Eng. 2021, 9, 105961. [Google Scholar] [CrossRef]
  106. Oladipo, A.A.; Vaziri, R.; Abureesh, M.A. Highly robust AgIO3/MIL-53 (Fe) nanohybrid composites for degradation of organophosphorus pesticides in single and binary systems: Application of artificial neural networks modelling. J. Taiwan Inst. Chem. Eng. 2018, 83, 133–142. [Google Scholar] [CrossRef]
  107. Zhong, Z.; Li, M.; Fu, J.; Wang, Y.; Muhammad, Y.; Li, S.; Wang, J.; Zhao, Z.; Zhao, Z. Construction of Cu-bridged Cu2O/MIL(Fe/Cu) catalyst with enhanced interfacial contact for the synergistic photo-Fenton degradation of thiacloprid. Chem. Eng. J. 2020, 395, 125184. [Google Scholar] [CrossRef]
  108. Ahmad, M.; Chen, S.; Ye, F.; Quan, X.; Afzal, S.; Yu, H.; Zhao, X. Efficient photo-Fenton activity in mesoporous MIL-100(Fe) decorated with ZnO nanosphere for pollutants degradation. Appl. Catal. B Environ. 2019, 245, 428–438. [Google Scholar] [CrossRef]
  109. Yan, J.; Wen, X.; Yin, L.; Wang, Y.; Li, H.; Tu, Y. An electrochemiluminescence aptasensor for amyloid-β protein with signal enhancement from AuNPs/Fe-MOFs nanocomposite. J. Electroanal. Chem. 2023, 933, 117293. [Google Scholar] [CrossRef]
  110. Wang, J.-W.; Qiu, F.-G.; Wang, P.; Ge, C.; Wang, C.-C. Boosted bisphenol A and Cr(VI) cleanup over Z-scheme WO3/MIL-100(Fe) composites under visible light. J. Clean. Prod. 2021, 279, 123408. [Google Scholar] [CrossRef]
  111. Wang, B.-Y.; Yuan, J.; Guo, J.; Zhang, F.-Q. Synthesis of Fe-MOFs/h-CeO2 hollow micro-spheres and their highly efficient photocatalytic degradation of RhB. J. Dispers. Sci. Technol. 2022, 44, 1479–1489. [Google Scholar] [CrossRef]
  112. Liu, X.; Dang, R.; Dong, W.; Huang, X.; Tang, J.; Gao, H.; Wang, G. A sandwich-like heterostructure of TiO2 nanosheets with MIL-100(Fe): A platform for efficient visible-light-driven photocatalysis. Appl. Catal. B Environ. 2017, 209, 506–513. [Google Scholar] [CrossRef]
  113. Liu, B.; Wu, Y.; Han, X.; Lv, J.; Zhang, J.; Shi, H. Facile synthesis of g-C3N4/amine-functionalized MIL-101(Fe) composites with efficient photocatalytic activities under visible light irradiation. J. Mater. Sci. Mater. Electron. 2018, 29, 17591–17601. [Google Scholar] [CrossRef]
  114. Li, N.; Du, H.; Tan, M.; Yang, L.; Xue, B.; Zheng, S.; Wang, Q. Construction of Z-scheme CuBi2O4/MIL-88A(Fe) heterojunctions with enhanced LED light driven photocatalytic Cr(VI) reduction and antibacterial performance. Appl. Surf. Sci. 2023, 614, 156249. [Google Scholar] [CrossRef]
  115. Li, Y.; Xia, Y.; Liu, K.; Ye, K.; Wang, Q.; Zhang, S.; Huang, Y.; Liu, H. Constructing Fe-MOF-Derived Z-Scheme Photocatalysts with Enhanced Charge Transport: Nanointerface and Carbon Sheath Synergistic Effect. ACS Appl. Mater. Interfaces 2020, 12, 25494–25502. [Google Scholar] [CrossRef] [PubMed]
  116. Nguyen, H.-T.T.; Tran, K.-N.T.; Van Tan, L.; Tran, V.A.; Doan, V.-D.; Lee, T.; Nguyen, T.D. Microwave-assisted solvothermal synthesis of bimetallic metal-organic framework for efficient photodegradation of organic dyes. Mater. Chem. Phys. 2021, 272, 125040. [Google Scholar] [CrossRef]
  117. Bai, S.; Qiu, H.; Song, M.; He, G.; Wang, F.; Liu, Y.; Guo, L. Porous fixed-bed photoreactor for boosting C–C coupling in photocatalytic CO2 reduction. eScience 2022, 2, 428–437. [Google Scholar] [CrossRef]
  118. Wang, W.; Wang, L.; Su, W.; Xing, Y. Photocatalytic CO2 reduction over copper-based materials: A review. J. CO2 Util. 2022, 61, 102056. [Google Scholar] [CrossRef]
  119. Wu, Q.J.; Liang, J.; Huang, Y.B.; Cao, R. Thermo-, Electro-, and Photocatalytic CO2 Conversion to Value-Added Products over Porous Metal/Covalent Organic Frameworks. Acc. Chem. Res. 2022, 55, 2978–2997. [Google Scholar] [CrossRef]
  120. Zheng, Y.; Fu, K.; Yu, Z.; Su, Y.; Han, R.; Liu, Q. Oxygen vacancies in a catalyst for VOCs oxidation: Synthesis, characterization, and catalytic effects. J. Mater. Chem. A 2022, 10, 14171–14186. [Google Scholar] [CrossRef]
  121. Jiang, L.; Li, Z.; Wang, D.; Guo, T.; Hu, Y. In-situ growth of p-type Ag2O on n-type Bi2O2S with intimate interfacial contact for NIR light-driven photocatalytic CO2 reduction. Appl. Surf. Sci. 2022, 601, 154185. [Google Scholar] [CrossRef]
  122. Wang, J.; Guo, R.T.; Bi, Z.X.; Chen, X.; Hu, X.; Pan, W.G. A review on TiO2-x-based materials for photocatalytic CO2 reduction. Nanoscale 2022, 14, 11512–11528. [Google Scholar] [CrossRef] [PubMed]
  123. Zhang, S.; Zhang, Y.; Baig, F.; Liu, T.-F. Synthesis and Applications of Stable Iron-Based Metal–Organic Framework Materials. Cryst. Growth Des. 2021, 21, 3100–3122. [Google Scholar] [CrossRef]
  124. Lou, W.; Yang, J.; Li, L.; Li, J. Adsorption and separation of CO2 on Fe(II)-MOF-74: Effect of the open metal coordination site. J. Solid State Chem. 2014, 213, 224–228. [Google Scholar] [CrossRef]
  125. Yang, W.; Wang, Y.; Wang, Q.; Wu, J.; Duan, G.; Xu, W.; Jian, S. Magnetically separable and recyclable Fe3O4@PDA covalent grafted by l-cysteine core-shell nanoparticles toward efficient removal of Pb2+. Vacuum 2021, 189, 110229. [Google Scholar] [CrossRef]
  126. Jian, S.; Tian, Z.; Zhang, K.; Duan, G.; Yang, W.; Jiang, S. Hydrothermal Synthesis of Ce-doped ZnO Heterojunction Supported on Carbon Nanofibers with High Visible Light Photocatalytic Activity. Chem. Res. Chin. Univ. 2021, 37, 565–570. [Google Scholar] [CrossRef]
  127. Dao, X.Y.; Guo, J.H.; Wei, Y.P.; Guo, F.; Liu, Y.; Sun, W.Y. Solvent-Free Photoreduction of CO2 to CO Catalyzed by Fe-MOFs with Superior Selectivity. Inorg. Chem. 2019, 58, 8517–8524. [Google Scholar] [CrossRef] [PubMed]
  128. Wang, L.; Mao, J.; Huang, G.; Zhang, Y.; Huang, J.; She, H.; Liu, C.; Liu, H.; Wang, Q. Configuration of hetero-framework via integrating MOF and triazine-containing COF for charge-transfer promotion in photocatalytic CO2 reduction. Chem. Eng. J. 2022, 446, 137011. [Google Scholar] [CrossRef]
  129. Dong, H.; Zhang, X.; Lu, Y.; Yang, Y.; Zhang, Y.-P.; Tang, H.-L.; Zhang, F.-M.; Yang, Z.-D.; Sun, X.; Feng, Y. Regulation of metal ions in smart metal-cluster nodes of metal-organic frameworks with open metal sites for improved photocatalytic CO2 reduction reaction. Appl. Catal. B Environ. 2020, 276, 119173. [Google Scholar] [CrossRef]
  130. Zhang, X.Y.; Wang, P.; Zhang, Y.; Cheng, X.M.; Sun, W.Y. Facet-Dependent Photocatalytic Behavior of Fe-soc-MOF for Carbon Dioxide Reduction. ACS Appl. Mater. Interfaces 2023, 15, 3348–3356. [Google Scholar] [CrossRef]
  131. Xie, X.; Liu, Y.; Li, Y.; Tao, J.; Liu, C.; Feng, J.; Feng, L.; Shan, Y.; Yang, S.; Xu, K. Nitrogen-doped Fe-MOFs derived carbon as PMS activator for efficient degradation of tetracycline. J. Taiwan Inst. Chem. Eng. 2023, 146, 104891. [Google Scholar] [CrossRef]
  132. Li, G.; Li, F.; Liu, J.; Fan, C. Fe-based MOFs for photocatalytic N2 reduction: Key role of transition metal iron in nitrogen activation. J. Solid State Chem. 2020, 285, 121245. [Google Scholar] [CrossRef]
  133. Lasek, J.; Yu, Y.-H.; Wu, J.C.S. Removal of NOx by photocatalytic processes. J. Photochem. Photobiol. C Photochem. Rev. 2013, 14, 29–52. [Google Scholar] [CrossRef]
  134. Yong, S.; Xiao-yu, S.; Wang, Z.; Chun-yan, L.; Qidong, Z.; Jinsuo, G.; Xinyong, L. Synthesis and characterization of Cu+/MIL-100(Fe) with enhanced catalytic performance for NH3-SCR. Ferroelectrics 2018, 522, 20–28. [Google Scholar] [CrossRef]
  135. Nguyen, H.P.; Kim, T.H.; Lee, S.W. Influence of operational parameters on the photocatalytic performance of DE-NOx process Via MIL-101(Fe). Prog. Nat. Sci. Mater. Int. 2018, 28, 689–695. [Google Scholar] [CrossRef]
  136. Chen, L.; Chen, Y.; Tu, X.; Zhu, S.; Sun, C.; Zhang, L.; Han, W.; Duan, X.; Sun, Q.; Zheng, H. Fe/Zr-MOFs constructed by a sunlight-responsive ligand for efficient photocatalytic nitrogen fixation under ambient condition. J. Colloid. Interface Sci. 2023, 633, 703–711. [Google Scholar] [CrossRef]
  137. Huang, L.; Kong, X.; Chang, K.; Yu, Z.; Tao, X.; Tang, T.; Xu, Y. DBD plasma assisted synthesis of MO@Fe/Ce-MOFs with rich oxygen defects as efficient photocatalysts for nitrogen fixation. J. Environ. Chem. Eng. 2023, 11, 109836. [Google Scholar] [CrossRef]
  138. Li, P.; Kim, S.; Jin, J.; Do, H.C.; Park, J.H. Efficient photodegradation of volatile organic compounds by iron-based metal-organic frameworks with high adsorption capacity. Appl. Catal. B Environ. 2020, 263, 118284. [Google Scholar] [CrossRef]
  139. Jiang, H.; Cheng, H.; Zang, C.; Tan, J.; Sun, B.; Bian, F. Photocatalytic aldehydes/alcohols/toluenes oxidative amidation over bifunctional Pd/MOFs: Effect of Fe-O clusters and Lewis acid sites. J. Catal. 2021, 401, 279–287. [Google Scholar] [CrossRef]
  140. Chen, L.; Wang, X.; Rao, Z.; Tang, Z.; Wang, Y.; Shi, G.; Lu, G.; Xie, X.; Chen, D.; Sun, J. In-situ synthesis of Z-Scheme MIL-100(Fe)/α-Fe2O3 heterojunction for enhanced adsorption and Visible-light photocatalytic oxidation of O-xylene. Chem. Eng. J. 2021, 416, 129112. [Google Scholar] [CrossRef]
  141. Chen, L.; Wang, X.; Rao, Z.; Tang, Z.; Shi, G.; Wang, Y.; Lu, G.; Xie, X.; Chen, D.; Sun, J. One-pot Synthesis of the MIL-100 (Fe) MOF/MOX Homojunctions with Tunable Hierarchical Pores for the Photocatalytic Removal of BTXS. Appl. Catal. B Environ. 2022, 303, 120885. [Google Scholar] [CrossRef]
  142. Qin, J.; Pei, Y.; Zheng, Y.; Ye, D.; Hu, Y. Fe-MOF derivative photocatalyst with advanced oxygen reduction capacity for indoor pollutants removal. Appl. Catal. B Environ. 2023, 325, 122346. [Google Scholar] [CrossRef]
Figure 1. Fe-MOFs for water and gas pollutant photodegradation (A and B represent two different Fe-MOFs materials or Fe-MOFs composites).
Figure 1. Fe-MOFs for water and gas pollutant photodegradation (A and B represent two different Fe-MOFs materials or Fe-MOFs composites).
Molecules 28 07121 g001
Figure 2. (a) Schematic diagram of the proposed mechanism of US/Fenton system for TC-HCl removal; (b) EPR spectra of DMPO/·OH and DMPO/·O2. Reprinted with permission from Ref. [57]. Copyright 2021, Elsevier. (c) Plausible mechanism of photocatalytic degradation of AO7 in the MIL-53(Fe)/PS/Vis process. Reprinted with permission from Ref. [58]. Copyright 2017, Elsevier.
Figure 2. (a) Schematic diagram of the proposed mechanism of US/Fenton system for TC-HCl removal; (b) EPR spectra of DMPO/·OH and DMPO/·O2. Reprinted with permission from Ref. [57]. Copyright 2021, Elsevier. (c) Plausible mechanism of photocatalytic degradation of AO7 in the MIL-53(Fe)/PS/Vis process. Reprinted with permission from Ref. [58]. Copyright 2017, Elsevier.
Molecules 28 07121 g002
Figure 3. (a) Illustration of plausible mechanism of photocatalysis-activated SR-AOP oxidation of CQ over P25M175 under LED visible light. Reprinted with permission from Ref. [80]. Copyright 2021, Elsevier. (b) Schematic diagram of the proposed photocatalytic mechanism of OTC degradation in γMF-10/light/PDS. Reprinted with permission from Ref. [82]. Copyright 2022, Elsevier.
Figure 3. (a) Illustration of plausible mechanism of photocatalysis-activated SR-AOP oxidation of CQ over P25M175 under LED visible light. Reprinted with permission from Ref. [80]. Copyright 2021, Elsevier. (b) Schematic diagram of the proposed photocatalytic mechanism of OTC degradation in γMF-10/light/PDS. Reprinted with permission from Ref. [82]. Copyright 2022, Elsevier.
Molecules 28 07121 g003
Figure 4. (a,b) Photo-Fenton degradation of TC-HCl and the corresponding kinetics with various Fenton-like catalysts. Experimental conditions: [TC-HCl] = 50 mg/L, [Catalyst] = 0.10 g/L, [H2O2] = 10 mL/L, pH = 4.1, T = 4.1, T = 287 K, visible light. (c) Illustration of the proposed reaction mechanism for TC-HCl removal in MIL-101/H2O2/visible light system. Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Figure 4. (a,b) Photo-Fenton degradation of TC-HCl and the corresponding kinetics with various Fenton-like catalysts. Experimental conditions: [TC-HCl] = 50 mg/L, [Catalyst] = 0.10 g/L, [H2O2] = 10 mL/L, pH = 4.1, T = 4.1, T = 287 K, visible light. (c) Illustration of the proposed reaction mechanism for TC-HCl removal in MIL-101/H2O2/visible light system. Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Molecules 28 07121 g004
Figure 5. (a) Schematic illustration of the morphology evolution progress of MIL-53(Fe)s as crystallization time increases. (b,c) Evolution of normalized RhB concentration versus reaction time as a function of the screened MIL-53(Fe)-XH (X = 2, 4, 6, 8, 10, 12, 14, 16, 18, 20, 22, 24, 48, 96, 120, and 144) over HCO process. (d,e) Fitting results of RhB destruction by a series of MIL-53(Fe)s using a pseudo-first-order kinetic model. (f) Pseudo-first-order rate constants for probe reactions and the volcano plot of probe destruction rate constants over the solvothermal reaction time for MIL-53(Fe)s synthesis. Experimental conditions: [RhB]0 = 100 mg/L, [caralyst] = 0.1 g/L, [O3]in = 30 mg/L, gas flow rate = 60 mL/min, T = 25 °C. Reprinted with permission from Ref. [93]. Copyright 2022, Elsevier.
Figure 5. (a) Schematic illustration of the morphology evolution progress of MIL-53(Fe)s as crystallization time increases. (b,c) Evolution of normalized RhB concentration versus reaction time as a function of the screened MIL-53(Fe)-XH (X = 2, 4, 6, 8, 10, 12, 14, 16, 18, 20, 22, 24, 48, 96, 120, and 144) over HCO process. (d,e) Fitting results of RhB destruction by a series of MIL-53(Fe)s using a pseudo-first-order kinetic model. (f) Pseudo-first-order rate constants for probe reactions and the volcano plot of probe destruction rate constants over the solvothermal reaction time for MIL-53(Fe)s synthesis. Experimental conditions: [RhB]0 = 100 mg/L, [caralyst] = 0.1 g/L, [O3]in = 30 mg/L, gas flow rate = 60 mL/min, T = 25 °C. Reprinted with permission from Ref. [93]. Copyright 2022, Elsevier.
Molecules 28 07121 g005
Figure 6. (a) EIS Nyquist impedance plots and (b) PL spectra of MIL-88A(Fe) in the presence or absence of ozone. (c) Proposed PCO mechanism of MIL-88A(Fe) for 4-NP degradation and mineralization. Reprinted with permission from Ref. [95]. Copyright 2019, Elsevier. (d) EPR spectra with DMPO (·OH) as trapping agents in different systems. (e) Fe 2p spectra of Fe-BDC1 and Fe-BDC2. Reprinted with permission from Ref. [96]. Copyright 2022, Elsevier. (f) Model of H2O2 dissociation process in MIL-88B and Cu@MIL-88B. Reprinted with permission from Ref. [98]. Copyright 2023, Elsevier.
Figure 6. (a) EIS Nyquist impedance plots and (b) PL spectra of MIL-88A(Fe) in the presence or absence of ozone. (c) Proposed PCO mechanism of MIL-88A(Fe) for 4-NP degradation and mineralization. Reprinted with permission from Ref. [95]. Copyright 2019, Elsevier. (d) EPR spectra with DMPO (·OH) as trapping agents in different systems. (e) Fe 2p spectra of Fe-BDC1 and Fe-BDC2. Reprinted with permission from Ref. [96]. Copyright 2022, Elsevier. (f) Model of H2O2 dissociation process in MIL-88B and Cu@MIL-88B. Reprinted with permission from Ref. [98]. Copyright 2023, Elsevier.
Molecules 28 07121 g006
Figure 7. (a) A schematic diagram of Ce(VI) photoreduction over the MB100 under white light illumination. Reprinted with permission from Ref. [101]. Copyright 2020, Elsevier. (b) Self-assembled and amorphous Fe-2MI displayed fascinating photocatalytic activity for Cr(VI) reduction. Reprinted with permission from Ref. [104]. Copyright 2019, Elsevier.
Figure 7. (a) A schematic diagram of Ce(VI) photoreduction over the MB100 under white light illumination. Reprinted with permission from Ref. [101]. Copyright 2020, Elsevier. (b) Self-assembled and amorphous Fe-2MI displayed fascinating photocatalytic activity for Cr(VI) reduction. Reprinted with permission from Ref. [104]. Copyright 2019, Elsevier.
Molecules 28 07121 g007
Figure 9. (a) A scheme of photocatalytic routes for visible-light-driven CO2 reduction in PCN-250-Fe2M (or PCN-250-Fe3). Reprinted with permission from Ref. [129]. Copyright 2020, Elsevier. (b) Illustration of the Band Structure and Photoinduced Electron Transfer Progress. (c) Illustration of the Intermediate during CO2-to-CO Conversion over Fe3O Clusters. Reprinted with permission from Ref. [130]. Copyright 2022, American Chemical Society.
Figure 9. (a) A scheme of photocatalytic routes for visible-light-driven CO2 reduction in PCN-250-Fe2M (or PCN-250-Fe3). Reprinted with permission from Ref. [129]. Copyright 2020, Elsevier. (b) Illustration of the Band Structure and Photoinduced Electron Transfer Progress. (c) Illustration of the Intermediate during CO2-to-CO Conversion over Fe3O Clusters. Reprinted with permission from Ref. [130]. Copyright 2022, American Chemical Society.
Molecules 28 07121 g009
Figure 10. (a) O 1s spectra of Ce-MOFs-D and MO@Fe/Ce-MOFs. (b) Possible mechanism of photocatalytic nitrogen fixation. Reprinted with permission from Ref. [137]. Copyright 2023, Elsevier.
Figure 10. (a) O 1s spectra of Ce-MOFs-D and MO@Fe/Ce-MOFs. (b) Possible mechanism of photocatalytic nitrogen fixation. Reprinted with permission from Ref. [137]. Copyright 2023, Elsevier.
Molecules 28 07121 g010
Figure 11. (a) Direct Z-scheme mechanism in MIL-100(Fe)/α-Fe2O3 hybrid. Reprinted with permission from Ref. [140]. Copyright 2021, Elsevier. (b) The mechanism of photocatalytic degradation of BTXS on the MIL-100(Fe) MOF/MOX homojunctions. Reprinted with permission from Ref. [141]. Copyright 2022, Elsevier. (c) The mechanism for photocatalytic oxidation on M-300 catalyst under visible light. Reprinted with permission from Ref. [142]. Copyright 2023, Elsevier.
Figure 11. (a) Direct Z-scheme mechanism in MIL-100(Fe)/α-Fe2O3 hybrid. Reprinted with permission from Ref. [140]. Copyright 2021, Elsevier. (b) The mechanism of photocatalytic degradation of BTXS on the MIL-100(Fe) MOF/MOX homojunctions. Reprinted with permission from Ref. [141]. Copyright 2022, Elsevier. (c) The mechanism for photocatalytic oxidation on M-300 catalyst under visible light. Reprinted with permission from Ref. [142]. Copyright 2023, Elsevier.
Molecules 28 07121 g011
Table 1. The degradation performance of Fe-based MOFs and their composites.
Table 1. The degradation performance of Fe-based MOFs and their composites.
CatalystsSynthesis MethodTarget SubstancesConcentration(mg·L−1)/Volume (mL)Catalyst (g/L)/
(g)
Reaction Time (min)Efficiency (%)/EvolutionLight SourceRef.
N-Fe-MOFsDBD plasma MO200.804897.00500 W xenon lamp[60]
D-Fe-MOFsDBD plasmaMO200.804897.00500 W xenon lamp[61]
Fe-MOFs@Fe2O3DBD plasmaMG150.302099.30500 W xenon lamp[62]
MIL-88B(Fe)SolvothermalTC-HCl100.20783.30/[65]
MIL-53(Fe)SolvothermalAO70.05 mM0.6090Almost 100LED lamps[66]
Mn/Fe-MOFsSolvothermalRhB3 × 10−5 M0.1012091.7840 W LED lamps[67]
TiO2@NH2-MIL-101(Fe)Self-assemblyMB500.20 3096.00300 W xenon lamp[68]
MIL-100(Fe)/polymerPhotopolymerizationAcid Black15 ppm/3095.20UV–Visible Light[69]
M/Fe-MOF (M = Co, Cu, Mg)SolvothermalRhB3 × 10−5 M0.2512092.0040 W LED [70]
α-Fe2O3@C@SiO2/TiO2SolvothermalRY145 dye50–2500.30901008 W 12in T5 TUV[71]
Fe-MOFsOil bathMB200.2518076.16500 W mercury lamp[72]
Fe3O4@GO@MIL-100(Fe)Hydrothermal2,4-DCP50/60Almost 100500 W xenon lamp[73]
Fe3O4@MIL-100(Fe)In-suit growthLevofloxacin2000.3318093.40PLS-SXE300/300UV[78]
NH2-MIL-101(Fe)Post-synthetic modificationTBBPA1.84 mM0.50120Almost 100/[79]
PDINH/MIL-88A(Fe)Facile ball-millingCQ100.403094.60PCX50C[80]
CNT@MIL-101(Fe)HydrothermalCiprofloxacin3.02 μM0.5060Almost 100White light LEDs[81]
MIL-101(Fe)/γ-Fe2O3HydrothermalOTC25/6091.20Visible light[82]
Fe3O4@MIL-53(Fe)CalcinationIBP100.406099500 W xenon lamp[84]
NH2-MIL-53(Fe)SolvothermalCO2500.0025 hCO/87.6 μmol·g−1 (6 h)300 W xenon lamp[127]
NH2-MIL-101(Fe)HydrothermalCO2500.108 hCO/25 μmol·g−1·h−1
CH4/11.67 μmol·g−1·h−1
300 W xenon lamp[128]
PCN-250-Fe2M (M = Mn, Zn, Ni, Co)SolvothermalCO2//4 hCO/21.51 mmol·g−1 (4 h)300 W xenon lamp[129]
Fe-soc-MOFsSolvothermalCO2500.055 hCO/1804 μmol·g−1·h−1Visible light[130]
MIL-101(Fe)SolvothermalN2800.0560Nitrogen fixation activity 50.35 μmol·L−1·h−1300 W xenon lamp[132]
MIL-101(Fe)Microwave-solvothermalNOx0.1 ppm0.106077150 W xenon lamp[135]
Fe/Zr-MOFsSolvothermalN2400.0025 hNitrogen fixation activity 49.8 μmol·L−1·h−1300 W xenon lamp[136]
MO@Fe/Ce-MOFsCalcinationN25000.00252 hNitrogen fixation activity 299 μmol·L−1·h−1500 W xenon lamp[137]
Fe-MOFsSolvothermalVOCs10000.202 hCO2/460 ppmINNOVA 1412i[138]
MIL-100(Fe)/α-Fe2O3HydrothermalVOCs1200.095200100250 W xenon lamp[139]
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

Cai, J.; Peng, Y.; Jiang, Y.; Li, L.; Wang, H.; Li, K. Application of Fe-MOFs in Photodegradation and Removal of Air and Water Pollutants: A Review. Molecules 2023, 28, 7121. https://doi.org/10.3390/molecules28207121

AMA Style

Cai J, Peng Y, Jiang Y, Li L, Wang H, Li K. Application of Fe-MOFs in Photodegradation and Removal of Air and Water Pollutants: A Review. Molecules. 2023; 28(20):7121. https://doi.org/10.3390/molecules28207121

Chicago/Turabian Style

Cai, Jun, Yang Peng, Yanxin Jiang, Li Li, Hua Wang, and Kongzhai Li. 2023. "Application of Fe-MOFs in Photodegradation and Removal of Air and Water Pollutants: A Review" Molecules 28, no. 20: 7121. https://doi.org/10.3390/molecules28207121

Article Metrics

Back to TopTop