1 Introduction

Common organic pollutants include dyes, pesticides, antibiotics and mycotoxins, some of which can pose a severe threat to the environment and human health [1,2,3]. Notably, mycotoxins produced as secondary metabolites by some fungi have multiple toxic effects on humans and livestock [4]. In addition, pesticides such as herbicides, insecticides and bactericides can also affect the environment and cause health problems [5]. Similarly, antibiotics are a class of drugs that treat infectious diseases in humans. However, the abuse of antibiotics is causing an increase in resistant bacteria [6], as well as metabolic and endocrine disruption. In addition, some dyes are also toxic and are the most commonly seen organic pollutants in the textile and leather industries. Therefore, there has been an increasing concern regarding purifying organic pollutants due to the various recalcitrant pollutants in wastewater from leather production.

Diverse methods have been involved in the removal of organic pollutants, such as physical removal, biodegradation and chemical degradation [7, 8]. Physical methods can separate organic pollutants from the environment and food matrix, but can cause secondary pollution [9]. Biodegradation based on single or combined cultures of bacteria, molds, yeasts, algae, or enzymes, is a popular remediation technique because it is highly economical and environmentally friendly [10, 11]. Although most organic pollutants can be removed by biodegradation [12, 13], some are highly recalcitrant and require additional chemical treatment. In chemical degradation, advanced oxidation processes (AOPs) have received significant attention, commonly using hydrogen peroxide (H2O2), hydroxyl radicals (•OH) and sulfate radicals (SO4•−), which can oxidize organic pollutants into less toxic products [14,15,16]. In addition, Fenton and Fenton-like processes, photocatalysis and SO4•−-based oxidative systems are also examples of AOPs.

Metal organic frameworks (MOFs), formed by the self-assembly of metal ions or metal clusters with organic ligands, have aroused wide attention [17, 18]. The large surface area, numerous adsorption sites [19] and variable functional groups have opened up potential applications of MOFs in catalysis [20], sensing [21], adsorption [22], conductivity [23] and drug delivery [24]. In addition, owing to their high surface areas, optimizable pore volume and pore size distributions, MOFs are becoming a promising class of adsorbents and enzyme immobilization carriers. Due to their tunable metal active sites, MOFs themselves show excellent catalytic performance. Therefore, MOFs have been developed as catalysts for the degradation of organic pollutants in advanced oxidation processes combined with enzyme catalysis and photocatalysis [25,26,27,28,29], while also promoting the adsorption of organic pollutants for effective removal. Several common types of organic pollutants that can be removed by MOFs are listed in Table 1. Consequently, MOFs have proven to be a promising platform for the removal of organic pollutants via physical adsorption, enzyme catalysis and chemical oxidation.

Table 1 Common organic pollutants that can be removed by MOFs catalysts

In this article, we provide a comprehensive review of recent findings and developments of MOFs for the removal of organic pollutants, including the fabrication strategies of MOFs and removal mechanisms. The role of MOF microstructures and properties in their catalytic degradation capability are discussed. Also, the strategies for enhancing the performance of pure MOFs to remove organic pollutants are summarized.

2 Synthesis of MOFs

MOFs were initially introduced by Omar Yaghi et al. by means of the combination of metal clusters and organic ligands [40]. After that, other types of MOFs such as Materials of Institute Lavoisier (MILs), Zeolitic imidazolate frameworks (ZIFs), University of Oslo (UiO), and PCN developed gradually [41]. MILs are a subclass of MOFs that are fabricated via the coordination of trivalent transition metal ions (such as Fe, Al, and Cr) and carboxylic acid ligands (Fig. 1a) [42, 43]. ZIFs can use transition metal ions to coordinate with imidazolate linkers through self-assembly (Fig. 1b) [44]. Among them, ZIF-8, a typical ZIFs composed of Zn (II) and 2-methylimidazole ligands, found its application as a catalyst in a variety of reactions [45]. Built with the metal center Zr4+ and dicarboxylic acid linkers, the UiO family has different ligand lengths but similar network topology, and the strong Zr-O bond coordination is conducive to its stability in various environmental conditions (Fig. 1c) [46]. PCN represents porous coordination network, Ma et al. first designed PCN-9 by a reaction between H3TATB and cobalt nitrate in DMSO [47]. And the reaction between H4adip and Cu (NO3)2 produced a new MOF designated PCN-14 by Ma and colleagues [48]. Subsequently, with meso-tetra(4-carboxyphenyl) porphyrin (TCPP) as the ligand, more porphyrinic MOFs have been prepared by researchers (Fig. 1d) [49, 50].

Fig. 1
figure 1

a Polyhedral structures of MILs [51] (with kind permission from Elsevier) b The crystal structure of ZIFs [52] (with kind permission from IOP Publishing Ltd) c Tolyhedral representation of the crystal structure of UiO-66 [53] (with kind permission from Elsevier) d Crystal structure of PCN-222 [54] (with kind permission from Wiley–VCH)

Different synthetic approaches were used for the preparation of MOFs, including solvothermal/hydrothermal methods, microwave synthesis, mechanochemical synthesis, sonochemical synthesis and electrochemical synthesis [55,56,57,58,59].

The most common approach is based on solvothermal/hydrothermal methods (Fig. 2a), which involve the execution of the reaction in an autoclave at a defined temperature [60]. The existing solvothermal methods are based on organic solvents, such as methanol, ethanol, acetone and N,N-dimethylformamide [61]. It is a relatively convenient and facile method, but it has some limitations, including a long reaction time and the high cost of the solvents [62, 63]. To accelerate crystallization and reduce liquid waste, alternative approaches, including electrochemical synthesis, microwave-assisted, mechanochemical and sonochemical, have been developed [64].

Fig. 2
figure 2

a Synthesis of MOFs by solvothermal/hydrothermal methods [71] (with kind permission from Wiley–VCH) b Microwave synthesis method [72] (with kind permission from Elsevier). c Mechanochemical synthesis method [73] (with kind permission from Elsevier) d Sonochemical synthesis method [74] (with kind permission from Wiley–VCH) e Electrochemical synthesis method [75] (with kind permission from Elsevier)

Microwave synthesis is a time-saving method, which uses microwave irradiation to heat reactant mixtures in domestic household microwave ovens (DMO) or similar commercially available instrumentation [65, 66]. Using microwave synthesis, various common MOFs such as MIL-101 [67], UiO-66 [68], ZIF-8 [69], PCN-134 [70] (PCN means porous coordination network) and others have been successfully prepared (Fig. 2b).

Mechanochemical synthesis is a green and eco-friendly approach (Fig. 2c), using methods such as ball-milling, screw extrusion, liquid-assisted resonant acoustic mixing and other approaches [76]. This synthesis method has attracted extensive attention because it uses little or no solvents, enables time-saving one pot synthesis, and generates minimal waste [77]. Liquid-assisted grinding (LAG) and ion- or liquid-assisted grinding (ILAG) are the most commonly used methods in mechanochemical synthesis. Compared with absolutely solvent-free approaches, these methods promote the dissolution of solid reagents and improve the formation of coordination bonds [78, 79]. Uzarevic et al. reported the first synthesis of zirconium MOFs (Zr-MOFs) using the LAG method [80]. The catalysis and porosity measurements showed that Zr-MOFs made by LAG had properties comparable to solvothermally synthesized materials.

Sonochemical synthesis uses ultrasound energy ranging from 20 to 1000 kHz, which enables the preparation of numerous MOFs with diverse crystal sizes and morphologies [81,82,83] (Fig. 2d). In general, the morphology and particle size of MOFs are affected by reaction time, temperature and ultrasonic power. Armstrong et al. optimized HKUST-1 crystals and revealed the crystallization mechanisms by modifying the reaction time and other parameters [84]. Compared with conventional solvothermal/ hydrothermal methods, this method can produce MOFs with homogeneous nucleation centers, while avoiding the need for high temperatures and pressures.

Electrochemical synthesis is a promising method that applies electrical current to chemical synthesis reactions [85] (Fig. 2e). This method can be divided into electrode superficial nucleation (ESN), indirect bipolar electrodeposition (IBED), and electrophoretic deposition (EPD) [86]. In this approach, there is no requirement for metal salts as precursors because the metal ions are generated by the electrodes [56].

3 Strategy for organic pollutants removal

3.1 Synergism of physical adsorption to catalytic degradation

Although MOFs are used for catalytic degradation, their adsorption capacity is also an important characteristic for the removal of organic pollutants. Due to their tunable porosities and large surface area, MOFs could remove organic pollutants by adsorption [87]. The removal of organic pollutants by MOFs is facilitated by multiple mechanisms including π-π stacking, electrostatic interactions, hydrogen bonding and acid–alkali interactions [88].

Gao et al. synthesized MIL-53(Cr), MIL-53(Fe) and MIL-53(Al) for sulfamethoxazole (SMZ) removal [89]. The maximum adsorption capacities predicted by a Langmuir model were 468.56, 450.83 and 75.53 mg·g−1 for MIL-53(Cr), MIL-53(Al), and MIL-53(Fe). The results revealed that metal nodes play an important role in SMZ removal. Zhao et al. reported PCN-222 for chloramphenicol (CAP) removal [90]. PCN-222 exhibited a large adsorption capacity of 370 mg g−1 and the adsorption equilibrium could be quickly reached after only 58 s. The large 1-D channels and the abundant hydroxyl groups of PCN-222 could improve the removal efficiency of CAP.

In addition, the introduction of functional groups such as -NH2, -NO2 or -SO4 and doping with metals such as Cu, Co, Mn and Ni could effectively improve the adsorption performance of MOFs [91]. Yu et al. fabricated a variety of porous MOFs, such as MIL-53(Fe), NH2-MIL-53(Fe), NO2-MIL-53(Fe) and Br-MIL-53(Fe) [92]. The maximum adsorption capacities of NH2-MIL-53(Fe), NO2-MIL-53(Fe) and Br-MIL-53(Fe) for the removal of tetracycline (TC) were 271.9 mg g−1, 272.6 mg g−1 and 309.6 mg g−1, which were respectively 9, 10 and 25% higher than the capacity of the pure MIL-53(Fe) (247.7 mg g−1) (Fig. 3a). Dehghan et al. compared the TC adsorption capacities of four MOFs (ZIF-67-NO3, ZIF-67-Cl, ZIF-67-SO4 and ZIF-67-OAC) with different chemical groups and four MOFs (ZIF-8-Octahedron, ZIF-8-Leaf, ZIF-8-Cuboid and ZIF-8-Cube) with different structures [93]. ZIF-67-Acetate exhibited the optimal performance (93.7%), showing 2.65 times higher removal efficiency than ZIF-67 (35.3%) (Fig. 3b). Yang et al. fabricated the Mn-doped UiO-66 (MnUiO-66) using solvothermal method for TC removal [94]. Doping with Mn added the active sites of UiO-66 (Fig. 3c). The maximum adsorption capacity of MnUiO-66 was 72.5 mg·g−1, almost six times higher than that of pure UiO-66. Jin et al. reported MIL-101 nanoparticles co-doped with Cu and Co, and used them as an adsorbent for efficient removal of TC [95]. Compared with pure MIL-101, the adsorption capacity of Cu-Co/MIL-101 was increased by 140% (Fig. 3d). In parallel, the outstanding adsorption performance can cooperate with the catalysis of MOFs for organic pollutants decontamination. For the elimination of dyes in wastewater, a novel Fe-loaded MOF-545(Fe) was synthesized by Zhang et al. [36]. The formed material not only showed high absorption capabilities, but also exhibited POD-like activity, which achieved removing MO and MB in a short period of time (about 2 h). In the case of photocatalysis, Jin et al. [96] synthesized MIL-101(Fe)@MIL-100(Fe) heterojunction to achieve 80% degradation of TC. Owing to the synergistic adsorption between the outer shell and nuclear layer, the Z-scheme heterojunction displayed a pore channel limited effect, which increased TC adsorption quantity and promoted TC photocatalytic properties. Li et al. [97] designed magnetic porous Fe3O4/carbon octahedra for Fenton-like catalytic removal of organic dye MB, the removal rate can reach nearly 100% in 30 min. The authors observed that the material exhibited excellent Fenton-like catalytic performance with MB molecules first adsorbed on the surface of catalysts, then diffused through mesoporous channels and sparked a Fenton-like reaction. Therefore, the synergistic impact of integrating physical adsorption and catalytic reactions may stimulate unique organic pollutant removal effects.

Fig. 3
figure 3

a The adsorption capacity of MIL-53(Fe)-based MOFs at different equilibrium concentrations [92] (with kind permission from Elsevier) b Removal efficiency and adsorption capacity of studied MOFs for TC [93] (with kind permission from Elsevier) c Adsorption performance of UiO-66 and MnUiO-66 for TC and Cr(VI) [94] (with kind permission from Elsevier) d Adsorption performance of pure MIL-101 and various MNPs/MIL-101 composites [95] (with kind permission from Elsevier)

3.2 Enzyme-immobilization carrier assisted enzymatic degradation

Although native enzymes are also used for the degradation of pollutants [98, 99], the fragile nature of native enzymes makes them susceptible to denaturation or instability in extreme environments which results in inactivation and an extremely high cost [100, 101]. Enzyme immobilization was an efficient strategy to improve the activity and stability of native enzymes. Owing to their high specific surface area, porous structure and good biocompatibility, MOFs hold great promise as enzyme immobilization carriers [102,103,104]. MOF-enzyme composites are also excellent catalysts for the degradation of organic pollutants.

MOF-enzyme composites show better catalytic activity, stability and reusability due to the protection of natural enzymes by MOFs [105]. Multifunctional groups on the surface of MOFs can contribute to improving the activity of immobilized enzymes. Furthermore, MOF nodes and linkers could offer numerous anchor sites for enzymes through chemical bonding, including coordinative bonding, covalent bonding and van der Waals forces [106, 107], which could prevent enzyme denaturation when exposed to extreme conditions and organic solvents. Liu et al. reported a hierarchically porous (HP) MOF HP-PCN-224(Fe) for glucose oxidase (GOx) immobilization [108]. Compared with free GOx, GOx@HP-PCN-224(Fe) displayed higher activity, pH and thermal stability. Park et al. found that Candida antarctica lipase B (CAL-B) conjugated on isoreticular MOF-3 (IRMOF-3) exhibited approximately 1000-fold higher activity than free CAL-B [109]. The immobilized enzyme showed higher thermal stability than the free enzyme and superior storage stability. Li et al. encapsulated organophosphorus acid anhydrase (OPAA) into zirconium MOF PCN-128y for the nerve agent simulant diisopropyl fluorophosphate (DFP) detoxication [110]. After three days of dry storage, OPAA@PCN-128y maintained 90% hydrolysis efficiency, while OPAA had only 30% hydrolysis efficiency. They further researched the catalytic performance of OPAA@PCN-128y for the real nerve agent, Soman, which indicated that the efficiency of OPAA@PCN-128y reached 90% in 60 min.

MOF-enzyme composites can remove organic pollutants through both the degradation activity of the encapsulated enzymes and the adsorption capacity of MOFs. Wang et al. found that encapsulation in Cu-MOF (HKUST-1) could enhance the catalytic activity of laccase [111] (Fig. 4a). The laccase/MOF system showed 50% higher degradation efficiency for bisphenol A (BPA) than free laccase. Jiang et al. synthesized the MIL-88A MOF and then used it to immobilize His-tagged organophosphohydrolase (OpdA) for degradation of organophosphorus pesticides (Ops) [112] (Fig. 4c-d). They used OpdA@MIL-88A for the degradation of OPs on grapes and cucumbers, which could achieve almost 100% removal efficiency and retain more than 66% and 61% of initial activity after 6 reuse cycles. Mo et al. encapsulated horseradish peroxidase (HRP) in the single-crystal ordered macroporous zeolitic imidazolate framework-8 (SOM-ZIF-8), which accelerated the degradation process of hazardous dyes (Fig. 4b) [113]. The HRP@SOM-ZIF-8 could rapidly remove congo red (CR) and rhodamine B (RB) by integrating the benefits of oxidative degradation by HRP with adsorption to the host material, exhibiting high removal efficiencies within 2 min. Wang et al. encapsulated the organophosphorus hydrolase (OPH) into Zn-doped Co-based ZIF (0.8CoZIF) for the effective detoxification of methyl parathion (MP) [114]. In the presence of 50 mM NaBH4, the OPH@0.8CoZIF completely converted 95 μM MP and produced nearly 100% 4-aminophenol within 15 min.

Fig. 4
figure 4

a Schematic illustration of the synthesis of laccase@HKUST-1 [111] (with kind permission from Elsevier) b Schematic diagram of the preparation of HRP embedded in SOM-ZIF-8 [113]. c Preparation of MIL-88A and OpdA@MIL-88A [112] (with kind permission from American Chemical Society) d Effect of adsorption time on the loading amount of His-OpdA at a 2:1 ratio (w/w) of MIL-88A to total protein; protein loading and activity of immobilized His-tagged OpdA at different ratios (w/w) of MIL-88A to total protein [112] (with kind permission from American Chemical Society)

3.3 MOF nanozyme-catalyzed degradation

By modulating the metal ion nodes and organic ligand, MOFs can be endowed with enzyme-like activities, including oxidase-, peroxidase-, and alkaline phosphatase-like activity, which can contribute to the removal of organic pollutants [115]. MOF-based nanozymes with oxidase-like activity can activate O2 to produce reactive oxygen species (ROS), which in turn can directly oxidize the pollutants [100]. In peroxidase-mimicking MOF nanozymes, MOFs can catalyze the reaction of H2O2 with other substrates [116]. For example, Zhou et al. reported PCN-222(Fe) containing Fe-TCPP as a heme-like ligand, mimicking the heme ligand of peroxidases [117]. In addition, MOFs can be used as hydrolase mimics which catalyze the hydrolysis of chemical bonds to achieve pollutants removal [100, 116]. Compared to natural enzymes, the MOF-based nanozymes show improved catalytic activity, better storage stability, and lower cost.

Luo et al. used three MOFs (MIL-100, MIL-53 and MIL-68) with peroxidase-like activity for aflatoxin B1 (AFB1) removal (Fig. 5a-b) [12]. The removal efficiency reached up to 100%, and animal feeding experiments confirmed that the hepatotoxicity of AFB1 can be neutralized by these peroxidase-like MOFs. Huang et al. reported hollow bimetallic Co-based nanocages (HNCs) (C-CoM-HNC, M = Ni, Mn, Cu, and Zn) for rhodamine B (RhB) degradation (Fig. 5c) [118]. In the strategy, the incorporation of secondary metal ions (Ni, Mn, Cu, and Zn) could provide new active sites and form synergistic active sites with Co. Meanwhile, C-CoM-HNC could mimic the oxidase enzyme and activate PMS, resulting in highly efficient RhB degradation. The C-CoCu-HNC had better oxidase activity than other HNCs and exhibited a promising catalytic performance. Therefore, C-CoCu-HNC was used for RhB degradation and the degradation efficiency could reach 93.41% after 60 min of reaction. They also reported peroxidase-mimicking NH2-MIL-88B(Fe) used for methylene blue (MB) degradation in water [119]. NH2-MIL-88B(Fe) could achieve 90% MB removal efficiency.

Fig. 5
figure 5

a MOF-loaded membranes for AFB1 removal [12] (with kind permission from American Chemical Society) b Schematic diagram of membrane preparation [12]. c Illustration of a general approach for C-CoM-HNC synthesis [118] (with kind permission from American Chemical Society) d Schematic illustration of the sensing platform for organic dye degradation with antibacterial activity based on the Au–Au/IrO2@Cu(PABA) cascade reactor [120] (with kind permission from American Chemical Society)

In addition, the current peroxide degradation systems are mainly based on the addition of H2O2 as a peroxidation agent. Zhao et al. developed an Au–Au/IrO2@Cu(p-aminobenzoic acid, PABA) catalytic reactor with tandem enzyme-like activity [120], which can exhibit excellent GOx- and peroxidase-like activity (Fig. 5d). More importantly, based on its GOx-like activity, the reactor can convert glucose into H2O2, which in turn can be used for the oxidation of organic dyes, avoiding the need to handle concentrated H2O2 with strong oxidizing and corrosive properties. In situ H2O2 generation therefore provides a promising direction for the development of novel MOF-based nanozymes.

Most studies focused on the oxidase- or peroxidase-like activity of MOF-based nanozymes. At present, there is little research on MOF nanozymes with hydrolase activity. However, Lin et al. reported a ZIF-90 nanozyme with organophosphate hydrolase (OPH) activity for the hydrolysis of MP (Fig. 6) [121]. The experimental results indicated that the mechanism for MP hydrolysis by ZIF-90 nanozyme could be ascribed to the synergistic effect of zinc and imidazole.

Fig. 6
figure 6

Schematic representation of the synthesis of ZIF-90 and the reaction mechanism of organophosphorus hydrolase and ZIF-90 [121] (with kind permission from American Chemical Society)

3.4 Photocatalytic degradation

As a green degradation technology based on solar energy, photocatalysis holds great promise for the degradation of pollutants [122, 123] (Table 2). To date, numerous photocatalysts have been used for the degradation of organic pollutants, including g-C3N4, TiO2, ZnO, CdS and their derivatives [124, 125]. Dionysiou et al. reported the use of nitrogen- and fluorine-doped titanium dioxide (NF-TiO2) for microcystin-LR (MC-LR) degradation [126]. Gong et al. found that g-C3N4/pyromellitic diimide (PDI)-g-C3N4 homojunction could photocatalytically degrade atrazine (ATZ) [127]. However, there are shortcomings in these photocatalysts, such as limited visible light utilization, improper band position and rapid recombination of charge carriers, which lead to low photocatalytic efficiency [128]. As porous coordination polymers consisting of tunable metal clusters and organic linkers, MOFs have photochemical properties. Considering the rich variety of possible MOF structures, using narrow gap semiconductors to construct MOF-based composites could overcome the above drawbacks and inherit the advantages of the individual MOFs or semiconductors [129,130,131]. For example, in order to make up for shortcomings such as the wide band gaps, insufficient light response and insufficient electric charge transfer rate, researchers introduced porous metal oxides, carbon materials, metal sulfides (MSs) and their heterostructures to form composites with improved photocatalytic performance [132, 133]. In an effort to even better support practical applications of MOFs in photocatalysts, the relevant theoretical calculation of MOF catalyst was elaborated by Hai and his colleague [134], guides the design and development of MOFs material. Kim et al. reported a novel MIL-125(Ti) modified with chemically reduced nitrogen-containing graphene oxide (CR–N-GO), named r-N-MIL [135]. With the incorporation of CR–N-GO, the pore size was increased from 2 nm to 2.8 nm, and the band gap of the semiconductor material was narrowed, which finally improved the photocatalytic activity of r-N-MIL. Wang et al. fabricated sulfur (S)-TiO2/UiO-66-NH2 to achieve Cr6+ reduction and BPA oxidation [136]. Wang et al. reported a ZnIn2S4@PCN-224 heterojunction that could degrade 99.9% tetracycline hydrochloride (TCH) within 60 min (Fig. 7) [137]. The increased photocatalytic performance of ZnIn2S4@PCN-224 compared to pure ZnIn2S4 was attributed to the construction of a hierarchical heterostructure between ZnIn2S4 and PCN-224.

Table 2 Summary of the photocatalytic degradation performance of functionalized MOFs as catalysts for the removal of organic pollutants
Fig. 7
figure 7

The mechanism of the electron/hole transfer and separation process of the ZIS@P20 composite under visible light irradiation [137] (with kind permission from Wiley–VCH)

The construction of heterojunctions is an ideal strategy for improving the photocatalytic efficiency of MOFs, which can inhibit the rapid recombination of photogenerated electrons and holes [147]. In recent years, type II composite heterojunctions combining MOFs with TiO2 have exhibited superior photocatalytic efficiency compared with pure MOFs or TiO2 [137]. Zhao et al. reported the photocatalyst PCN-222(Mn)-PW12/TiO2, which degraded 94.83% for ofloxacin in 120 min and 98.5% for RhB in 80 min [148]. The N2 adsorption–desorption and photoluminescence (PL) spectra indicated that the introduction of PCN-222(Mn) increased the number of active sites in PCN-222(Mn)-PW12/TiO2. The recombination rate of photoinduced electron–hole pairs was reduced, which in turn improved the photocatalytic efficiency. Liu et al. prepared 2D/1D core–shell heterostructures (ZnIn2S4@Fe3O4 and ZnIn2S4@α-Fe2O3) [149], and characterized them through a series of electrochemical measurements, including transient photocurrent responses, EIS Nyquist plots and polarization curves. The results showed that the 2D/1D core–shell heterostructures were beneficial for electron transfer, which facilitated the photodegradation of RhB. Lu et al. used g-C3N4/PDI@MOF heterojunctions as photocatalysts for the removal of TC, carbamazepine (CBZ), BPA and p-nitrophenol (PNP) [150].

As an emerging and effective method, these studies demonstrated the potential of functionalized MOFs in photocatalytic degradation for the removal of organic pollutants. Nevertheless, the research on MOF-based degradation of organic pollutants is still in its infancy. This research area remains challenging due to the complex reaction environments, such as photocatalytic reactors required for these catalysts and the relatively low degradation efficiency obtained. Moreover, it is a promising strategy for constructing MOF-based photocatalysts to remove pollutants and develop novel photocatalysts with high-efficiency optical and electronic properties. In addition, photocatalysis can also be combined with other reactions, including Fenton and SR-AOPs, which may increase the degradation efficiency.

3.5 MOF catalyst performance in the Fenton-like process

In Fenton and Fenton-like reactions, ferrous ion (FeII) acts as a catalyst on H2O2 to produce •OH, which in turn can attack even recalcitrant organic pollutants [151, 152]. However, conventional homogenous Fenton and Fenton-like reactions have various disadvantages, such as secondary pollution and the need for pH regulation [153, 154]. To avoid these disadvantages, the superior heterogeneous Fenton-like reactions, including photo-Fenton reaction, Hetero-electro-Fenton process and photo-electro-Fenton (PEF) reaction, have stood out.

Iron-based MOFs (Fe-MOFs) are a class of good photocatalysts and can be more efficiently coupled with Fenton reagents than other MOFs. Mei et al. reported benzimidazole (BIm)-modified Fe-MOFs, and the rod-like α-Fe2O3-x exhibited complete MB degradation [155]. The oxygen vacancies (OVs) and Fe2+ content determined the α-Fe2O3-x photo-Fenton-like catalytic activity, which was substantiated by the experimental data and density functional theory (DFT) calculations. Wang et al. reported a UiO-66-based MOF conjugated with an FeIII metalloporphyrin that could integrate photocatalysis and Fenton-like processes to degrade RhB [156] (Fig. 8a). Fe(III) tetra(4-carboxylphenyl)porphyrin chloride (FeIII-TCPPCl) not only played the role of a photosensitizer, but also acted as an iron-based catalyst that produced ·OH from H2O2, which could accelerate the Fenton-like process. Liu et al. used MIL-88A as a catalyst for the degradation of tris-(2-chloroisopropyl) phosphate (TCPP), a widely used organophosphorus flame retardant with adverse effects on the nervous system [157] (Fig. 8b). In the MIL-88A/H2O2/Vis system, the degradation efficiency of TCPP reached approximately 95% (Fig. 8c). These were ascribed to the Fe–O clusters in MIL-88A, which could activate H2O2 and then form ·OH radicals. Owing to the slow Fe(II)/Fe(III) cycle, the efficiency of Fenton and Fenton-like reactions is generally limited. Huang and co-workers reported a two-dimensional (2D) π-d conjugated MOF named Fe3(HITP)2 (HITP = 2,3,6,7,10,11-hexaiminotriphenylene) with high conductivity, which accelerated the Fe(III)/Fe(II) cycle to achieve 96.7% TC degradation within 30 min [158]. Furthermore, the efficiency of Fenton-like processes can be enhanced by introducing heterogeneous electro-Fenton catalysts. Electro-Fenton reactions realized catalytic degradation rely on in-situ generation of H2O2 on the cathode by O2 reduction and further conversion to ·OH [159]. Wu et al. have achieved electro-Fenton degradation of SMX through the use of Mn0.67Fe0.33-MOF-74. The SMX removal efficiency can reach 96% at pH 3 and 30 mA of current after 90 min [160]. Huang's team prepared a series of MnxCo3-x@C-GF with excellent catalytic performance by introducing Mn/Co MOFs derivatives into the graphite felt cathodes [161]. In their work, the CIP removal efficiency could achieve 99.8% in 60 min, which could be attributed to the electrochemically active metals that promoted the generation of active radicals (·OH). Another example of heterogeneous Fenton-like processes is the PEF reaction, which is an upgraded EF process. Ye and his colleagues [162] used 2,2′-bipyridine-5,5′-dicarboxylate (bpydc) as a linker to prepare heterogeneous PEF catalyst Fe − bpydc 2D MOF for bezafibrate treatment. Under UVA and visible light irradiation, bezafibrate in solution could be removed completely with a small amount (0.05 g L−1) Fe − bpydc 2D MOF as a catalyst. The experimental analysis and theoretical calculations revealed that newly developed MOFs have become highly efficient Fenton-like catalysts [163, 164].

Fig. 8
figure 8

a The proposed photocatalytic mechanism of FeIII-TCPPCl ⊂ UiO-66 in the co-catalytic Fenton-like reaction [160] (with kind permission from Elsevier) b Schematic diagram of the reaction mechanism of the MIL-88A/H2O2/Vis system [161] (with kind permission from Elsevier) c The efficiency of different systems in the removal of TCPP [161]

3.6 MOF catalytic degradation performance in SR-AOPs

SR-AOPs produce hydroxyl radicals (·OH) and sulfate radicals (SO4•−), which give them great potential for efficiently eliminating a variety of harmful pollutants. In SR-AOPs, SO4•− and singlet oxygen (1O2) are mainly produced, which contribute to the activation of persulfate (PS) and peroxymonosulfate (PMS) [155]. Electron paramagnetic resonance (EPR) analyses, and theoretical calculations based on DFT are used to explore the possible mechanisms in MOFs-based SR-AOPs [165], which demonstrated that PMS/PDS activation by MOFs-based materials for organic pollutants degradation is promising. PDS and PMS were activated by many strategies, including ultraviolet irradiation, chemical methods and other methods in which photo-activation deserves special attention [166]. In addition, partially coordinated metal ions in MOFs can activate PMS/PS to form SO4•− and ·OH, and thus degrade organic pollutants. Liu et al. embedded Co sites in a carbon nitride catalyst (CoCN), which were used for visible light-induced PMS activation [167]. The results of the radical quenching experiments and EPR analyses indicated the reaction mechanism of BPA degradation. The rate constant for the CoCN/Vis/PMS system (1.84 min−1) was 5.58 times higher than that of the CoCN/ PMS system (0.329 min−1), which indicated that visible light could help improve the activation performance of CoCN for PMS and produce more reactive free radicals to degrade BPA.

In addition, bimetallic MOFs exhibited several synergistic effects and enhanced properties compared with the monometallic MOFs. Roy et al. reported bimetallic MOF-based heterojunction MIL-53(Fe/Co)/CeO2 for atrazine degradation [168]. Visible light irradiation only achieved 24.3% ATZ degradation within 60 min, while the MIL-53(Fe/Co)/CeO2/PMS/Vis system could achieve 99.9% ATZ degradation. The results indicated that the Co and Fe sites in MIL-53(Fe/Co) could achieve simultaneous redox cycles and consequently activate PMS and generate the reactive species.

The above studies indicated that increasing metal sites could greatly improve PMS activation performance. This phenomenon can also be reflected in PS activation, which is based on MOFs. For example, Duan et al. fabricated Cu-MIL-101(Fe) andCo-MIL-101(Fe) to degrade Acid Orange 7 (AO7) via PS activation [169]. Compared with pristine MIL-101(Fe), the AO7 removal efficiency by 6 wt% Cu-MIL-101(Fe) and 6 wt% Co-MIL-101(Fe) has reached 92% and 98% within 150 min.

The SR-AOPs are still in their infancy, and more research is needed to understand their complex catalytic mechanisms.

3.7 Other MOF-based catalytic degradation

In addition to the above catalytic reactions, electrochemical catalytic degradation and ultrasonic reactions are important processes in the removal of pollutants. MOF-based materials are widely used as electrode active materials in the electrocatalytic degradation of organic pollutants. Arulpriya et al. [170] synthesized a MOF-modified electrode by introducing TiO2@Fe MOF for the simultaneous sensing and degradation of CPF. The TiO2@Fe MOF/SPE could degrade the chlorpyrifos with high efficiency due to its electrocatalytic activity. Xu et al. [171] introduced UiO-66 derived ZrO2-C nanoparticles into PbO2 electrode to construct a new type of ZrO2-C/PbO2 anode. The prepared functional electrode possessed advantageous electrochemical performance with the 2, 4, 6-trinitrophenol (TNP) removal efficiency of up to 94.48% in 140 min. In addition, MOF-derived nanomaterials can also be introduced into the electrode as photoelectrodes to enhance light utilization and improve degradation efficiency [172]. Jia et al. [173] used ZIF-8/NF-TiO2 nanocomposites as photoanodes for the degradation of sulfa antibiotics. Results showed that the hybrid photoelectrode can effectively improve light utilization and enhance electron–hole pairs, resulting in enhanced the photo-electrocatalytic degradation activity.

With respect to the ultrasound (US) process, coupled with photocatalysis, it could improve the degradation efficiency of organic compounds [174]. Mosleh et al. [175] prepared a novel photocatalyst by doping the Ce and Eu with HKUST-1 MOF for treating organophosphorus pesticide Malathion by sonophotocatalysis. More cavitation bubbles were generated and climaxed the mass transfer rate due to the ultrasonic field, resulting in a significantly improved Malathion degradation rate. Moreover, the sonolysis process has been presented for PS activation, which showed enhanced effects on the degradation of organic pollutants [176]. In Sajjadi and colleagues’ work, Fe3O4@MOF-2 was prepared as a heterogeneous nanocatalyst for PS activation under US irradiation to degrade diazinon [176]. In the presence of US irradiation, the Fe3O4@MOF-2/US/PS process accomplished the degradation of diazinon completely in 60 min, which could be attributed to the intensified generation of hydroxyl radicals by US. The hybrid systems integrating ultrasound and various AOPs are a low-cost and effective technique for organic contaminants treatment [177].

3.8 Reusability and safety of MOF catalysts

In practical applications, MOF catalysts are difficult to separate from the reaction solution for recycling. The fragile powder form of MOFs caused poor processing and recycling properties, which limit their practical applications [178, 179]. However, numerous strategies have been developed to overcome these defects. One is synthesizing magnetic MOF compounds. Niu et al. reported a CuCo/C catalyst which could degrade 90% CIP by activating PMS within 30 min [180]. The hysteresis curve revealed that CuCo/C could be easily separated using an external magnetic field (Fig. 9a). Cong et al. reported a yolk-shell Fe3O4@MOF-5 nanocomposite for MB removal [181]. In the yolk-shell Fe3O4@MOF-5, Fe3O4 was magnetic and catalytic, while the MOF-5 shell could effectively protect the Fe3O4 while providing numerous pores to accelerate the molecular transfer and improve the catalytic efficiency. Due to the Fe3O4 yolk, Fe3O4@MOF-5 could also be separated using an external magnetic field.

Fig. 9
figure 9

a Time-dependent adsorption (correlation curve was drawn using the kinetic parameters calculated from the pseudo-second-order model), pseudo-second-order plots (inset) and photographs of the contaminated aqueous solution before and after adsorption of RhB on MIL-100(Fe) (33.3wt.%) aerogel [182] (with kind permission from Wiley–VCH) b Magnetic properties of CuCo/C [180] (with kind permission from Elsevier) c Scheme of π-π and electrostatic interactions between the 3D ANF/ZIF-67 composite aerogel and dyes [183] (with kind permission from Elsevier)

Another effective recycling strategy is fixing MOFs on suitable carriers such as membranes or hydrogels/aerogels (Fig. 9b). Wang et al. used zeolitic imidazolate framework (ZIF)-67/PAN(polyacrylonitrile) composite nanofibers to activate PMS in the catalytic degradation of acid yellow-17 (AY) [184]. Different from dispersed MOF nanoparticle materials, the ZIF-67/PAN composite nanofibers could be easily separated from the reaction system due to their flexibility. He et al. reported Co@NCNT-MS catalysts with a favorable TC degradation capacity [185]. Researchers encapsulated the catalysts in graphene oxide (GO) by facile vacuum filtration to form a composite membrane with outstanding ease of separation. He et al. constructed a ZIF-8 photocatalyst membrane and its derived product (ZnS photocatalyst membrane) for the removal of MB under visible light irradiation [186]. The ZIF-8 photocatalyst membrane could easily achieve complete separation and avoid secondary contamination. Yao et al. reported hybrid aerogels made from ZIF-9 and ZIF-12 loaded onto cellulose aerogels [187]. The hybrid aerogels could remove about 90% of PNP in 60 min and could also be easily removed from the reaction system. Zhao et al. prepared a 3D aramid nanofiber (ANF)/ZIF-67 composite aerogel for the removal of organic dyes [183]. The 3D ANF aerogel served as a mechanical support to achieve the uniform assembly of ZIF-67 (Fig. 9c). Ren et al. designed and synthesized the MOF composite material copper-benzenedicarboxylate/cellulose aerogel (CuBDC/CA) [188], which could decompose more than 90% methylene blue in 240 min.

3.9 Functional MOF-based derivative for catalytic degradation

To achieve the superior performance of catalytic, pristine MOFs and MOF composites can be converted into derivatives by direct pyrolysis under appropriate conditions. MOF derivatives, synthesized via different pyrolysis strategies, are promising catalysts and absorbents for various reactions. For example, nanoporous metal-containing carbon (metal@C) catalysts are manufactured from MOFs via pyrolysis under an inert atmosphere (nitrogen or argon). Li and colleagues successfully synthesized a core/shell structured hollow Fe–Pd@C nanomaterial by carbonizing Fe-metal organic frameworks in the N2 atmosphere [189]. The as-synthesized nanomaterials show excellent performance as catalysts in strengthening homogeneous Fenton degradation of phenol. Moreover, porous regular-shaped metal oxide@C can be obtained when the MOFs are calcinated in reactive environments such as an oxygen atmosphere. Zhang and coworkers [190] constructed MOF-derived ZnFe2O4/Fe2O3 perforated nanotubes as catalysts for ciprofloxacin (CIP) photocatalytic removal. In their work, the magnetically recoverable Z-scheme photocatalysts were successfully synthesized by one-step calcination method using MOF as a precursor. Under light irradiation, with the help of the prominent photocatalytic performance of ZnFe2O4/Fe2O3 perforated nanotubes, the CIP removal percentage increased to 96.5% within 180 min.

In addition, the development of heterogeneous SR-AOPs based on MOF composites and their derivatives has drawn much attention. Pu et al. reported a MOF-derived novel magnetic Fe@C composite for PS activation and SMX degradation [191]. Fe@C-800, synthesized under the pyrolysis temperature of 800 °C, exhibits high catalytic capacities. Degradation efficiency for SMX is 98.3% and decomposition magnitude for PS is 93.6% after 90 min. Zhao and coworkers reported the CeO2/N-doped carbon/Ce-TCPP heterostructures that were converted from Ce-TCPP by performing a simple pyrolysis process at low temperature under N2 flow [192]. The prepared heterostructures as heterogeneous catalyst exhibit excellent photocatalytic activity of PMS under visible light, with the degradation rates of 98.6% and 94.4% for MB and MO within 60 min, respectively. The above works provided a possible degradation strategy for remediation of organic pollutants from the leather production.

3.10 MOF-functionalized products for pollutant removal

MOFs hold great promise as novel materials for the removal of pollutants, and integrating them into functionalized materials is very important from a practical point of view. Agrawal et al. fabricated MOF-functionalized fabrics ZIF-8@ carboxymethylated (CM) cotton and ZIF-67@CM cotton for the removal of volatile organic compounds (VOCs) from ambient air (Fig. 10) [193]. These fabrics have enormous potential for application in protective textiles, anti-odor clothing, air purification filters, and related products. Yoo et al. coated cotton with Zr-MOFs such as UiO-66, UiO-67 and DUT-52 and utilized the resulting composite for the removal of particulate matter PM from air, which could improve the performance of air filters [194]. Tahazadeh et al. synthesized biodegradable cellulose acetate/MOF-derived porous carbon (CA/MOFDPC) adsorptive membranes for MB removal [195]. These functionalized products have a number of potential commercial applications. Seo et al. prepared nanocellulose/MOF aerogel composites for effective detoxification of methyl paraoxon (MPO) in both static and dynamic continuous flow systems [196].

Fig. 10
figure 10

Schematic representation of MOF-based filters for integrated air purification [193] (with kind permission from Elsevier)

4 Conclusions and perspectives

In this article, we reviewed recent progress in treatment methods for removing organic pollutants, with a focus on catalytic degradation using MOFs. First, we described the synthesis methods of MOFs. Although solvothermal synthesis is the most popular strategy for the preparation of MOFs, other green strategies have been developed to avoid high energy consumption or the residue generation, which will gradually become the direction of the future. Then, recent advances in the applications of MOFs in the catalytic degradation of organic pollutants were systematically reviewed. MOF nanozymes have been found to have wide applications in pollutant removal due to their multiple enzyme-like activities, but they still face the challenge of further hoisting their stability in practical applications. The photocatalytic processes based on MOFs can degrade organic pollutants effectively in a sustainable way through the use of solar energy as energy sources. The heterogeneous Fenton-like reaction using solid catalysts removes organic contaminants by producing reactive species on the surface of the catalysts in an environmentally benign way. Meanwhile, research in this field is still in the early stages and some disadvantages, including aggregation and dispersibility, still exist. MOF-based materials exhibited high activity in the SR-AOPs process for the removal of organic pollutants due to their unique structural characteristics, and the catalytic performance in the field of SR-AOPs is often limited by activation time and temperature. Notably, MOFs can be combined with other nano- and functional materials to achieve synergistic effects. Currently, MOFs are gradually being developed from enzyme carriers and physical adsorbents into multifunctional materials with enzyme-mimetic catalytic activity and photocatalytic properties. With the progress of synthesis methods, multi-site transformation can be achieved, while the adsorption performance, enzyme catalysis, photocatalysis performance, and even auxiliary detection and removal functions can be achieved, resulting in truly multifunctional materials for the future. Finally, the heterogeneity of MOFs and their combination with hydrogels, aerogels or membranes will enable facile separation and recovery, which is very important for practical applications. We hope this review can provide useful information for researchers, and provide a reference for the removal of organic pollutants with MOF in the leather industry.

Further research on the removal of organic pollutants with MOFs should address the following issues:

  1. 1.

    Although the synthesis methods of most MOFs are simple, the costs remain high. We need to develop more cost-effective methods to fabricate MOFs, such as strategies to choose cheaper organic ligands.

  2. 2.

    Currently, experiments on the removal of pollutants using MOFs remain at the laboratory scale. In actual samples, there are many other substances which could disturb the removal process. Therefore, it is essential to evaluate these removal methods of organic pollutants for practical application.

  3. 3.

    Further studies should include specific experiments on the industrial utilization of MOFs. Although researchers have studied MOF-based catalysts in membrane reactors and MOF-based adsorbents as carriers, there are few studies on the removal of pollutants in large-scale industrial scenarios.

Given the above problems and challenges, researchers could synthesize new types of MOFs materials and optimize the methods to reduce the influence of the material itself on the experimental results, thus solving the problems of materials in practical applications. In summary, MOFs, as an excellent new porous nanomaterial, would be able to achieve large-scale industrial production and practical application in the near future with the joint efforts of researchers.