Next Article in Journal
Supramolecular Self-Assembly Strategy towards Fabricating Mesoporous Nitrogen-Rich Carbon for Efficient Electro-Fenton Degradation of Persistent Organic Pollutants
Next Article in Special Issue
Electrochemical Performance of Layer-Structured Ni0.8Co0.1Mn0.1O2 Cathode Active Materials Synthesized by Carbonate Co-Precipitation
Previous Article in Journal
Effect of Ionic Liquids in the Elaboration of Nanofibers of Cellulose Bagasse from Agave tequilana Weber var. azul by Electrospinning Technique
Previous Article in Special Issue
Interface Engineered V-Zn Hybrids: Electrocatalytic and Photocatalytic CO2 Reductions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress of Metal-Organic Frameworks and Metal-Organic Frameworks-Based Heterostructures as Photocatalysts

by
Mohammad Mansoob Khan
*,
Ashmalina Rahman
and
Shaidatul Najihah Matussin
Chemical Sciences, Faculty of Science, Universiti Brunei Darussalam, Jalan Tungku Link, Gadong BE 1410, Brunei
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(16), 2820; https://doi.org/10.3390/nano12162820
Submission received: 23 July 2022 / Revised: 13 August 2022 / Accepted: 14 August 2022 / Published: 17 August 2022

Abstract

:
In the field of photocatalysis, metal-organic frameworks (MOFs) have drawn a lot of attention. MOFs have a number of advantages over conventional semiconductors, including high specific surface area, large number of active sites, and an easily tunable porous structure. In this perspective review, different synthesis methods used to prepare MOFs and MOFs-based heterostructures have been discussed. Apart from this, the application of MOFs and MOFs-based heterostructures as photocatalysts for photocatalytic degradation of different types of pollutants have been compiled. This paper also highlights the different strategies that have been developed to modify and regulate pristine MOFs for improved photocatalytic performance. The MOFs modifications may result in better visible light absorption, effective photo-generated charge carriers (e/h+), separation and transfer as well as improved recyclability. Despite that, there are still many obstacles and challenges that need to be addressed. In order to meet the requirements of using MOFs and MOFs-based heterostructures in photocatalysis for low-cost practical applications, future development and prospects have also been discussed.

1. Introduction

Photocatalysis is a chemical reaction that takes place in presence of a photocatalyst and a suitable light. In short, it is the acceleration of a photo-reaction in the presence of a catalyst. Photocatalysis over semiconductors occurs when light with an energy greater than the band gap of semiconductors is absorbed generating photo-excited electrons (e) and holes (h+) for subsequent reduction and oxidation processes by e and h+, respectively [1]. Metal oxides and chalcogenides are two semiconductors that have been extensively studied. Metal oxides, such as TiO2 [2], ZnO [3], SnO2 [4], and CeO2 [5,6], are commonly used as photocatalysts. Despite their stability, the main drawback of using metal oxides as photocatalysts is their wide band gap energy. The wide band gap only enables them to absorb ultraviolet light and this restricts their practical usage [7]. In addition, carbon-based materials, such as graphene and g-C3N4, have also attracted remarkable research interest in recent years, owing to their ultrathin, two-dimensional nature, and unique properties including high thermal conductivity and excellent mechanical stiffness [8,9]. However, the fast recombination of the charge carriers is the major disadvantage of using graphene as a photocatalyst material. To overcome this problem, it requires modification with nanoparticles or light-sensitive materials. In contrast, chalcogenides exhibit narrow band gap energy. However, it has the disadvantage of having a high rate of recombination of photo-generated ē and h+ as well as the photo-corrosion phenomena [10,11].
Recently, metal-organic frameworks (MOFs) have emerged as novel photocatalysts due to their intrinsic structural properties by having a high surface area and a well-ordered porous structure [12,13]. MOFs are porous crystalline hybrid materials formed by the linkage of metal centers (cluster) and organic ligands (organic linkers). These MOFs are known as porous coordination polymers [14]. MOFs exhibit reticular design in which two designs approach: supermolecular building layer and supermolecular building block demonstrate the influential capability of the MOFs as precursor materials with unique structure. In general, metal-enclosed inorganic compounds and organic ligands (moieties) are linked through coordination bonds, thus developing MOFs. Therefore, a three-dimensional interconnected network is formed creating open frameworks with high permanent porosity [15].
Properties such as regular channels, high specific area, and abundant active sites make MOFs as promising materials [16]. MOFs are considered low electron conductivity materials due to the poor overlap between frontier orbits and the electronic states on ligands and metal ions. However, the electronic properties of a metal can be changed by changing its coordination environment to enforce the desired geometries around the metal [17]. Figure 1 shows some of the tunable properties of MOFs [18]. In the last decade, MOFs have been employed for water treatment [19], chemical hydrogen storage [20], photochemical and electrochemical energy storage and conversion [21,22], removal of heavy metals [23], solar energy conversion [24,25], fuel production [26], gaseous fuel storage [27], water harvester [28], water splitting for hydrogen evolution reaction [29], and oxygen evolution reaction [30]. The organic linkers tend to have large π-conjugated systems in which electron transfer between ligands and metals is possible. Interestingly, if the MOFs have a metal-ligand with photocatalytic properties, then ligands themselves can generate photo-generated electrons for photocatalytic reactions [31].
Therefore, the advantages of using MOFs as photocatalytic materials are as follows [26]:
  • The porous structure of MOFs exposes more active sites and facilitates;
  • Ordered crystal structure of MOFs can separate electron-hole pairs;
  • MOFs structure can be tailored;
  • MOFs are easy to separate and recycle from the reaction system;
  • MOFs have the tendency for electron transfer between ligands and metals due to the large π-conjugated systems in the organic linkers;
  • A long-wavelength absorbing organic bridging ligand can enhance the optical absorption of MOFs.
MOFs are widely researched due to their potential applications such as photocatalytic of degradation of organic and inorganic pollutants [32,33], photocatalytic of antibiotics [34], photocatalytic antimicrobials [35], bio-imaging [36], drug delivery [37], photocatalytic organic transformations [36,38], photocatalytic production of hydrogen (H2) [39], and photocatalytic reduction of carbon dioxide (CO2) [40], as illustrated in Figure 2.
However, recombination of photogenerated electron-hole pairs has become the main challenge that limits the improvement of photocatalysis [41]. By constructing heterostructure, the charge separation can be promoted due to difference in the interface potential that generates a built-in electric field in which it helps in inhibiting recombination [42]. A typical strategy includes the introduction of metal nanoparticles to improve the separation of photogenerated electron-hole pairs [41]. Another strategy is to integrate MOFs and inorganic semiconductors/organic polymers. The interaction between the inorganic-organic complex is versatile due to the strong charge interaction between components [41]. In addition, the organic linkers of MOFs could also be further modified which have a great potential to construct heterostructures [43]. Therefore, MOFs-based heterostructure photocatalysts are mainly composed by different components.
In summary, this article emphasizes the recent progress and developments in the photocatalytic activities of MOFs and MOFs-based heterostructures. Various synthesis methods to fabricate MOFs and their respective heterostructures have been discussed in-depth. Moreover, the possible mechanisms of the photocatalytic performance of MOFs and MOFs-based heterostructures photocatalysts have also been illustrated and discussed. Despite the current progress of MOFs and MOFs-based heterostructures, there are still limitations and gaps that hinder the potential of the materials in different applications. Therefore, the future prospects and conclusions have also been highlighted and discussed.

2. Different Synthesis Methods of MOFs and MOFs-Based Heterostructures

2.1. Synthesis of MOFs

The main goal in MOFs synthesis is to establish the synthesis conditions that lead to defined inorganic building blocks without the decomposition of the organic linker. The understanding of typical metal coordination environments or the formation conditions of typical inorganic building blocks is crucial for direct synthesis efforts [44]. Higher surface area (ultrahigh porosity) of MOFs provides extraordinary properties. This requires an increase in storage space per weight of the material. The pore size and topology of the framework can be finely tuned by selecting appropriate linkers and metal nodes. Longer organic linkers provide larger storage space and a greater number of adsorption sites. Moreover, MOF topology has shown some effects on catalytic performance [45]. However, a unique spatial distribution of active sites can only be achieved within a specific MOF structure. Therefore, the topological diversity of MOFs can be used to change the position and orientation of MOF catalytic sites which can be located at the nodes or the linkers [46]. Ideally, it is important to keep the pore diameter below 2 nm in order to maximize the surface area of the framework [47]. Apart from that, MOFs are well-known for their ability to impart functionality using a judicious selection of linkers and metal nodes. They undergo post-synthetic modification to further tune properties through swapping or altering linker or node components in the framework. Furthermore, select crystallographic phases and crystallite size/morphology can be controlled to modify the surface chemistry of the MOF [18]. In the last decade, various synthesis methods have been used to design MOFs, such as hydrothermal [48], solvothermal [49], precipitation [50], and microwave-assisted synthesis [51].

2.1.1. Microwave-Assisted Synthesis

Microwave synthesis [48,49,50,51] is one of the most common synthesis methods. It uses microwave irradiation, which facilitates the nucleation and crystal growth in the synthesis of MOFs synthesis [51]. For instance, mesoporous iron trimesate nano MOF MIL-100(Fe) was synthesized using microwave-assisted hydrothermal reaction [52]. In general, due to the high-frequency motion generated by microwave heating, the heating is uniform without the need for heat conduction. Apart from the temperature, the pressure of the reaction process can also be controlled. This method has a significant improvement in saving reaction time and reducing energy consumption [53]. This method obtained a homogenous and chemically stable coating of polyethylene glycol (PEG) with enhanced colloidal and chemical robustness. Moreover, it was found that the material induced a low immune response. [RE(L)(H2O)(DMF)n] where RE = Dy, Gd, Ho, Pr, and Sm have been synthesized through the hydrothermal method [54]. Structural analyses reveal that the ligands are linked by five different rare-earth centers forming iso-structural nanoporous frameworks. Other work by Goswami et al. reported on the synthesis of 2D-MOFs with redox active metal centers (Ni(II), Co(II), and Cu(II)) through hydrothermal synthesis [55]. The metal centers contain two types of electron-rich linkers such as bis(5-azabenzimidazole, Linear L1, and angular L2) and aromatic dicarboxylates. In another study, Zr-based metal-organic frameworks were prepared using microwave-assisted synthesis [51]. The porosity of the material was based on the quantity of the modulator, reaction time, and temperature. It was found the optimum amount of modulator was in the range of 40 to 185 mol and the optimum temperature was between 80 to 120 °C.

2.1.2. Solvothermal Method

MOFs have also been synthesized using the solvothermal method as it is considered one of the most popular methods used to prepare MOFs. The solvothermal method offers controllable geometry of the synthesized product with a large yield compared with other methods [56]. Solvothermal reactions are carried out in the liquid phase and crystals are precipitated through self-assembly between organic ligands and metal salts to obtain target MOFs [53]. Denisov et al. reported the synthesis of [Zn4O(BDS)3] via the solvothermal method [49]. The Zn4O6+ tetrahedral are bound to each other into a three-dimensional cubic framework by the BDC anions to form large spherical pores with a volume of 1838.78 Å3 accessible for the solvent. A facile and template-free solvothermal method has been used to prepare Cu-based MOFs nanosheets as reported by Zhang et al [57]. The starting coordination complexes of different copper (II)-ligand was found to control the growth and morphology of MOF crystal. Duan et al. synthesized Cu-BTC using a room temperature-template strategy [50]. The synthesized material exhibited a hierarchically porous structure and excellent thermostability. Moreover, various organic solvents, such as dimethyl sulfoxide, N, N-dimethylformamide, and toluene, have been widely used to fabricate MOFs and MOFs-based heterostructures to date. Among the above-mentioned organic solvents, dimethylformamide is most often selected for the synthesis of MOF and its modifications, due to its high boiling point and solubility. Although the solvothermal method has been widely used, the toxicity and environmental unfriendliness of the organic solvents severely restrict their use.

2.1.3. Vapor Assisted Synthesis

In Virmani’s work, UiO-66 MOFs were obtained from a vapor-assisted conversion synthesis [58]. A mixed solution of ZrOCl2, dicarboxylic acid, and acetic acid were prepared and deposited. The deposited materials were stored in an oven and dried under pressure subsequently. The thickness of the MOF films was found to be 200 nm. A copper dicarboxylate metal-organic framework film was prepared by chemical vapor deposition (CVD) as stated by Stassin et al. [59]. The MOF-CVD method involved two steps. Firstly, the substrate was changed into a Cu precursor layer by physical vapor deposition and secondly, the precursor layer reacted with the vapor of the dicarboxylic acid linker. The authors extended this process to a thin film of copper dicarboxylate MOFs, which was the first crystalline and oriented porous material to be obtained entirely through vapor deposition. In another work, high quality Mg-MOF thin films were prepared using vapor assisted synthesis [60]. A mixture of H2O, acetic acid, and dimethylformamide (DMF) was used as the vapor source and filled into a sealed glass container. The precursor, Mg(CH3COO)2·4H2O was dissolved in DMF and 2,5-dihydroxy-benzenedicarboxylic (DOBDC) acid in DMF. AA was added into the precursor solution to prevent a rapid precipitation of the precursors when the Mg salt and the DOBDC mixed.

2.1.4. Sonochemical Method

Ultrasonic irradiation has also been used in the synthesis of MOFs. The energy of high frequency in the liquid phase induces the generation, growth, and rupture of bubbles which make the local instantaneous temperature and pressure of the reactant increase [53]. The immense energy and high pressure created during the sonication process help to create a stable structure. Thus, the sonication technique has a great potential for the fabrication of MOFs in large quantities and at a low cost. For instance, Zn-based MOFs were synthesized by the aforementioned method [61]. DMF and triethylamine were included in the reaction and were mixed with the precursor vigorously until the solution was clear. An ultrasound rod was then placed into the reaction reactor. The obtained product was filtered and obtained uniform nucleation and high production yield. Table 1 shows the various synthesis of MOF materials.
There are several works relied on low-temperature routes to synthesize MOFs. A precipitation reaction method is well-known to grow simple molecular or ionic crystals since it is possible to tune the reaction conditions [62]. Vapor phase synthesis has been used as it can gently and efficiently produce oriented MOF films [63]. Since high orientation and high crystallinity MOFs are produced, they are suitable for applications in sensor and electronic devices. Meanwhile, microwave-assisted provides an efficient and sustainable route to synthesize MOF materials. This is due to the microwave energy directly penetrating into the substance as opposed to just relying on the heat conduction of the medium for heating. The microwave-assisted method usually takes only a short amount of time to complete the entire heating process which can effectively reduce the temperature and time [64]. The sonochemical method on the other hand can improve the crystallization and growth rate during the growth process [65].

2.2. Synthesis Method of MOFs-Based Heterostructures

In addition to the synthesis of MOFs as mentioned earlier, the synthesis of MOFs-based heterostructures is also widely studied in the field of heterogeneous catalysis due to their large specific surface area, modifiable pore structure, and high exposed active sites [66]. The construction of heterojunction is an effective method to improve the photocatalytic property of photocatalyst by improving the photogenerated carrier separation efficiency [67,68]. Moreover, the porous nature of MOFs facilitates the migration and separation of photo-generated charge carriers. The following shows the different modifications employed to synthesize MOFs-based heterostructures.

2.2.1. Addition of Linkers

The MOF structures may be easily altered by changing the substituent groups in linkers and/or linker itself, as well as exchanging or doping metal ions in the framework nodes [42,69]. As a result, the band gap energy of MOFs can be easily controlled to improve photocatalytic efficiency. For instance, Dan-Hardi et al. have synthesized highly porous titanium-based MOFs using terephthalic acid (H2BDC) precursor and they found out that MIL-125(Ti) synthesized with unmodified H2BDC can only respond to UV light. Another study, reported by Fu and co-workers introduced a NH2 group into the H2BDC precursor and successfully synthesized isostructural NH2-MIL-125(Ti) in the same way. The authors reported that the synthesized materials displayed yellow color and their light absorption ability extended to the visible region. The band gap dramatically decreased from 3.60 to 2.46 eV. Moreover, this shows that the incorporation of BDC derivatives could tune the optical properties of the MOFs [70].
The organic linkers are used in some modifications because they can function as an antenna for light sensitizing and charge transfer to the inorganic cluster by exploiting ligand-to-metal charge transfer (LMCT). The LMCT feature of MOFs is ideal due to the efficient charge separation. Moreover, alteration of the organic linker could introduce additional substituents (using mixed linkers or even capping additional metal ions) affecting the oxidative power of these MOFs. A similar concept may be used for the reduction; this is because metal orbitals in MOFs define the position of the lowest unoccupied crystal orbital (LUCO), and the reductive power can be changed by choosing metal ions with suitable orbitals [71]. For instance, Wang et al. have successfully fabricated porphyrin-MOFs to degrade bisphenol F under irradiation of visible light [72]. The organic linkers of porphyrin can harvest light by playing the role of antennas, and further activate the metal clusters via a linker to a cluster charge transition that could generate enough reactive oxygen species (ROS). They also reported that the prepared photocatalyst was quite stable in a saline environment and could achieve over 78% bisphenol F removal (with/without salt) after 8 cycles.

2.2.2. Modification of its Metal Clusters

Alternatively, the optical response can be altered by tuning the cluster-forming metal or by employing mixed metal clusters. The latter approach has been utilized to create mid-gap metal-centered states that result in the MOFs with LMCT transition which are clusters that cannot be fabricated via direct synthesis. The second method is to use MOFs as a passive container for encapsulating a light-absorbing photocatalyst. These active species are often homogeneous catalysts based on precious metals, and leaching has been effectively prevented by encapsulating or covalently bonding them to the framework [73,74]. Table 2 shows some of the MOFs-based heterostructures that have been used in various photocatalytic activities.
MOFs have a number of advantages in photocatalysis due to their inherent structural features which include large surface area and porous structure and a tunable combination between metallic nodes and organic linkers. The photocatalytic efficiency is, however, still insufficient to meet the actual needs. As discussed above, a series of strategies have been introduced to enhance the photocatalytic properties of MOFs and their applications including ligand functionalization [87], mixed ligands [88], mixed metal centers [89,90], metal ions immobilization [91], the addition of magnetic materials, MOF or COF coupling [92,93], semiconductors coupling [94,95], surface decoration using carbon-based materials [95] or metal nanoparticles [96], and dye sensitization [40]. These strategies have also been employed to extend visible light absorption of MOFs, more efficient generation, and separation and transfer of charge carriers, as well as its good recyclability. Owing to its outstanding properties, all of the above-mentioned MOFs-based heterostructures can effectively use light energy for various photocatalytic activities, and most of them can be recycled several times. Due to their extraordinary porosity and large surface area, both the metal nodes and organic linkers can be easily tuned; thus, MOFs-based heterostructures exhibited promising photocatalytic performance [97]. However, the stability of MOFs during the photocatalytic reaction is an important issue needed to be further considered, thus searching MOFs with optimized coordination between metal clusters and organic linkers for high stability is of great importance.

3. Applications of MOFs and MOFs-Based Heterostructures

The use of MOFs as photocatalysts for photocatalytic activities has been reported in previous literature and discussed below. Table 3 summarizes the results of photocatalytic degradation of various waste reported in the different studies.

3.1. Photocatalytic Evolution of H2

Non-metal g-C3N4 is a well-known photocatalyst due to its appealing electronic structure, low cost, and high stability. The band gap of g-C3N4 was approximately 2.7 eV, indicating the light absorption edge into the visible region (up to 450 nm). Thus, coupling g-C3N4 with MOFs may lead to an enhanced visible light response. For instance, Zhao et al. investigated the photocatalytic evolution of H2 using protonated g-C3N4 co-operated with Sm-doped Co-MOF to form a 2D/2D heterojunction via electrostatic self-assembly [108]. The H2 evolution rate of 5% Sm-doped exhibited the highest activity has reached 73.42 μmol h−1 within 5 h under simulated sunlight. The improved activity may be attributed to the presence of a small amount of Sm ions in the trivalent state which increases the chaos of coordination in the Co-MOF. Moreover, it further facilitates the separation of photo-excited charge carriers. Moreover, Li et al. have fabricated Co-MOF via the solvothermal method for photocatalytic evolution of H2 under irradiation of visible light [39]. The experiment results revealed that the photocatalytic H2 evolution rate of the synthesized material reaches a high value of 1.44 mmolg−1 h−1 in H2O solution with Na2S and Na2SO3 as a sacrificial agent without co-catalyst.

3.2. Photocatalytic Degradation of Organic Pollutants

Chen et al. have used hybrid modification strategies to construct Co/Ni-MOFs@BiOI catalyst based using a simple hydrothermal synthesis method [117]. In this study, different photoactive semiconductors were also coupled with MOFs constructing multi-heterojunctions for more efficient charge separation. The authors found that the combination of Co/Ni-MOFs with BiOI has significantly improved the capability of absorbing visible light which results in enhancing photocatalytic degradation of methylene blue in aqueous solutions without any chemical additives. They reported that the 20 mg/L of Co/[email protected]% exhibited the highest methylene blue removal of 81.3% within 240 min under irradiation of visible light and showed almost four times higher photocatalytic activity than that of Co/Ni-MOFs. This could be attributed to the absorption wavelength to visible light, the acceleration of charge-carrier separation efficiency as well as a synergetic effect.
In another study, Govindaraju and co-workers successfully used Zn-MOF decorated bio-activated carbon for photocatalytic degradation of methyl orange (86.4%) and brilliant green (78.8%) within 90 min under irradiation of UV light [118]. Hussain et al. showed the highest photocatalytic degradation activity with 99.7% methylene blue removal in 3 h under visible light using N-O-TiO2/C composite derived from NH2-MIL-125(Ti) [119]. The high photocatalytic performance may be due to the presence of carboxylate functional groups (-COOH) on the surface of the carbon matrix in the composite because these functional groups increase the hydrophilicity of the composite and provide additional active sites which facilitate better accommodation of methylene blue molecules. It may also have attributed to the optimal anatase/rutile phase junctions together with the formation of photo-active oxygen rich N-O like interstitial/intraband states above the valence band of TiO2.
Zhang et al. have successfully prepared magnetically recoverable Z-scheme ZnFe2O4/Fe2O3 perforated nanotube using a one-step MOF-derived calcination method by using MIL-88B/Zn core/shell nanorod as a precursor [99]. The authors reported that the MIL-88B core reacted with O2 to form hollow Fe2O3 and the Zn2+ shell underwent an in situ solid-state reaction with Fe2O3 to cause volume shrinkage and formation of ZnFe2O4 and finally turned into ZnFe2O4/Fe2O3 perforated nanotube in the calcination process. They found that ZFF-2 exhibited the highest photocatalytic ciprofloxacin degradation performance with a degradation percentage of 96.5% and a TOC removal percentage up to 89% under light irradiation for 180 min. The synthesized material showed no significant decrease in photocatalytic degradation of ciprofloxacin even after five successive recycling experiments. The enhanced photocatalytic performance may be due to the perforated nanotube and Z-scheme transfer pathway, high light absorption ability, and high interfacial separation of photo-generated charge.
In a different study, Hou et al. showed outstanding photocatalytic degradation of perfluorooctanoic acid up to 89.6% within 3 h under optimal conditions using lignin/PVA/bimetallic MOFs composite [108]. Perfluorooctanoic acid is used as an industrial surfactant and as a material feedstock. In addition, the lignin/PVA/bimetallic MOFs composite retained 77% catalytic capacity after 4 cycles.

3.3. Photocatalytic Reduction of CO2

Photocatalytic reduction of CO2 to value added products, such as methane, ethanol, methanol, CO, methanoic acid, etc., has been widely reported. MOFs has recently surfaced as a better candidate for CO2 reduction. This is owing to the presence of organic linkers and metal center nodes in MOFs that enable easy absorption of light energy and desirable photoactive catalytic properties for CO2 reduction. For instance, Idris and co-authors have demonstrated the photocatalytic reduction of CO2 using dye-sensitized 2D Fe-MOF nanosheets [40]. The dye//Fe-MOF nanosheets catalytic system exhibited a good photocatalytic CO2-to-CO activity of 1120 μmol g−1 h−1 under visible light irradiation. Moreover, the photocatalytic CO production was further enhanced by regulating the electronic structure of the 2D Fe-MOF nanosheets by doping with Co ions, achieving a remarkable photocatalytic activity of 1637 μmolg−1 h−1. In a different study, Xu and co-workers have employed CdS/Ni-MOF for selective photocatalytic reduction of CO2 to CO. 130Among the prepared materials, 20%-CdS/Ni-MOF showed the best photocatalytic reduction performance, the yield of CO reached 7.47 μmol g in the 4th hour, which was nearly 16 times and 7 times that of Ni-MOF and CdS, respectively. In another study, Sonowal et al. have reported the reduction of CO2 to methanol using g-C3N4@Zr-based MOF under visible light irradiation [120]. The maximum yield of methanol obtained is 386 µmol g−1 h−1.
In summary, as a novel organic-inorganic hybrid structure, MOFs can function as photocatalysts because they can combine photo-sensitizer and catalytic capabilities in one structure. Photo-active MOFs can function as improved photocatalysts by minimizing contamination, and easily be recovered and reused due to their solid nature. Various factors including tunable energy levels, metal-to-ligand or ligand-to-metal charge transfer transitions effect, and orientation of organic linkers should be further studied and optimized in order to develop stable and high functionality structure and components of the MOF photocatalysts.

4. Photocatalytic Mechanisms of MOFs and MOFs-Based Heterostructures

Light absorption in MOFs and MOFs-based heterostructures can occur either through the inorganic metal atom or through the organic linker [121]. MOFs often contain few absorption bands in the UV-Vis region due to the chemical nature of the chromophore centers originate from the inorganic metal atom or organic linkers [122]. In addition, distinct π–π * transitions of the aromatic units in the organic linker or to metal-to-ligand or ligand-to-metal charge transfer transitions often contribute the absorption bands [123,124]. As a result, the semiconductors band theory is not appropriate to describe the light absorption and subsequent transitions observed in MOF materials.

4.1. Photocatalysis of Pristine MOFs Structures

According to the principle of the traditional photocatalytic process of semiconductors, an incident light with an energy greater than the band gap energy of a semiconductor can directly excite a photocatalyst. Upon light illumination, charge transfer from ligand (organic site) to metal (inorganic site) occurs in the MOF [125]. Figure 3 illustrated the general photocatalytic mechanism of the MOFs. In the pristine MOFs, e are excited from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) leaving behind h+ in the HOMO [126,127]. The HOMO/LUMO functions similarly to the VB/CB in semiconductors in this context [128]. In the LUMO (i.e., CB in a typical semiconductor), the photo-generated e can be transferred to O2, resulting in the generation of O2 radicals. Meanwhile, the surface hydroxyl group/water can be oxidized by h+ in the HOMO (i.e., VB in a typical semiconductor), resulting in the formation of OH. The presence of these reactive species can then be used to degrade different types of pollutants [97]. In a review reported by Ikreedeegh et al., the authors have explained in-depth the redox reactions of the MOF [129]. In brief, the reduction and oxidation of acceptors (CO2) and donor species (H2O) by the e/h+ pair, respectively. However, for MOF semiconductors with small band gap energy, generally, they generate a reverse process and recombine immediately to release unproductive energy in the form of heat.
For instance, Sun et al. have successfully oxidized amines to imines using NH2-MIL-125(Ti) under exposure to visible light [130]. The photo-generated Ti3+ and O2 which were formed via the reaction between Ti3+ and O2 were reported to be involved in the transformation process based on the experimental observations and their electron spin resonance result. Chen et al. reported an enhanced photocatalytic CO2 reduction performance using NH2-MIL-125(Ti) [131]. This is due to the enhanced optical absorption and CO2 uptake ability with the introduction of NH2BDC. However, it was further mentioned that after the introduction of ZnTCPP, it further enhanced the CO2 uptake ability and, hence, significantly increased the photoelectric conversion ability (Figure 4).

4.2. Photocatalysis of MOFs-Based Heterostructures

Researchers have used various methods to fabricate visible light responsive MOFs including altering metal nodes, functionalizing organic linkers, employing hybrid linkers, and developing composite photocatalysts in order to maximize visible light harvesting in the solar spectrum. Figure 5 shows a general photocatalytic mechanism of MOFs-based heterostructures. For example, Qiu and co-workers have reported the remarkable photocatalytic degradation of an oral antibiotic, sulfamethoxazole using Ag@Zr-based MOFs [115]. They also reported the low leaching amount of Ag and Zr under different pH values which indicated the excellent stability of the synthesized materials. Under exposure to light, the e would be enriched on the surface of Ag nanoparticles (Figure 5). Subsequently, the photo-excited e would transfer from Ag nanoparticles to Zr-based MOFs to further react with O2 to produce O2. On the other hand, the photo-excited h+ could react with H2O to generate O2. In addition, some generated O2 could be further transferred to react with OH radicals. All of the oxygen-active species took part in the degradation of sulfamethoxazole.
In a different study, Li et al. fabricated 2D porphyrin Co-TCPP MOF decorated 2D B-TiO2−X nanosheet (Co-TCPP MOF@B-TiO2−X) photocatalysts via the hydrothermal method [114]. The prepared photocatalysts exhibited outstanding photocatalytic performance for the degradation of bisphenol A up to 97% within 120 min irradiation. The degradation rate of the composite was 7.3 and 19.3 times higher than Co-TCPP MOF and B-TiO2−X, respectively. They reported that both B-TiO2−X and Co-TCPP MOF could concurrently produce photogenerated e/h+ pairs upon irradiation of visible light. Based on the conventional type II heterojunction, the photogenerated e are transferred from the LUMO of Co-TCPP MOF to the CB of B-TiO2−X due to the more negative LUMO position of Co-TCPP MOF. Meanwhile, due to the more positive VB position of B-TiO2−X, the photogenerated h+ were transferred from the VB of B-TiO2−X to the HOMO of Co-TCPP MOF indicating that these heterostructures are effective in suppressing the recombination of photogenerated carriers and boosting the photocatalytic performance.
However, according to a recent literature review by Behera et al., there are four types of possible mechanisms by various MOFs based photocatalysts: (i) Heterojunction based charge dynamics, (ii) Z-scheme based charge dynamics, (iii) Cocatalyst mediated charge dynamics, and (iv) Elemental doping based charge dynamics [132]. For instance, MOF-derived ZnFe2O4/Fe2O3 nanotube showed a direct Z-scheme rather than a typical heterojunction photocatalyst [99]. Based on the optical and band properties determined by DRS and Mott-Schottky, as well as fluorescence analysis, a corresponding photocatalytic CIP degradation mechanism has been proposed. The photogenerated e on the CB of Fe2O3 will pass through the Fe2O3/ZnFe2O4 interface which, therefore, will recombine with the photogenerated holes on the VB of ZnFe2O4. The heterojunction facilitates the separation of photogenerated e/h+ pairs and maintains the strong oxidizing holes on the VB of Fe2O3 and highly reductive electrons on the CB of ZnFe2O4. Therefore, in the presence of visible light and ZnFe2O4/Fe2O3 composites, O2 and OH are thermodynamically generated in the photocatalytic CIP degradation system (Figure 6). Other studies that proposed the Z-scheme have been reported by Qiao et al. for hydrogen evolution from photocatalytic water splitting in the MOF/MoS2 photocatalyst and by Wang et al. for photocatalytic removal of alkylphenol ethoxylate using CuO@Cu-based MOFs composite [101,133]. Photocatalytic reduction of Cr(VI) using NiFe2O4/MOF-808 composite was reported by Khosroshahi et al. [134]. The mechanism of photocatalytic reduction of Cr(VI) has been proposed using the Z-scheme pathway as described earlier.
On the other hand, Cao et al. reported an S-scheme heterojunction for photocatalytic hydrogen evolution using Cu-MOFs modifies Mn0.05Cd0.95S NPs [135]. The charge transfer mode has been explained using Z-scheme and S-scheme (Figure 7) [136]. For Z-scheme, semiconductor II (SC II) with a higher Fermi level to ensure its electron flow to semiconductor I (SC I) through the interface of SC II-SC I until both Fermi levels of SC I and SC II are eventually at the same level [136]. In addition, the positive and negative charges gather at interfaces near SC II and SC I, respectively, which produces an internal electric field (IEF). Then, from CB of SC I, photogenerated e are transferred to SC II VB with IEF action. For the S-scheme principle, the difference in Fermi level increases of SC II band edge bending for e supply layer formation. This leads to the decrease of the SC I band edge for electron acquisition. The difference in Fermi levels between SC I and SC II results in the transfer of the photogenerated e from CB of SC I to VB of SC II through the bending charge transfer channel within IEF. Therefore, in this case, Cao et al. mentioned that under visible light irradiation, the photogenerated e from the CB of the Cu-MOFs and holes from the VB of the MCS are recombined under the internal electric field, band edge bending, and coulomb interaction. The S-scheme pathway shows strong redox capacity, thus providing the powerful driving force for photocatalytic H2 evolution [135].
In summary, MOFs and MOFs-based heterostructures can absorb light either through the inorganic metal atom or through the organic linker. Under light exposure, e are photoexcited from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) leaving behind h+ in the HOMO in pure MOFs materials. However, to maximize visible light harvesting in the solar spectrum, the properties of MOFs have been altered by altering metal nodes, functionalizing organic linkers, employing hybrid linkers, and developing composite photocatalysts.

5. Future Outlook

To summarize, MOFs are emerging as the new promising materials in the field of photocatalysis. However, they still have some limitations. The following future recommendations should be considered to address all these issues and challenges:
  • Firstly, most MOF synthesis methods require high production costs which may restrict their applications in various fields. Therefore, a simple and cost-effective synthesis method to prepare MOFs should be introduced;
  • Recyclability, reusability, and stability of MOFs have not been discussed widely. These qualities are crucial for preparing MOFs with excellent properties;
  • In-depth study of the physicochemical properties of these materials, such as light absorption in the visible region, electronic structure, crystallographic properties, porosity, and surface area, is highly required with the aid of advanced characterization techniques;
  • MOFs usually exist in powder form which leads to poor stability and reusability. Recycling issue of MOFs should be addressed properly. For example, MOFs fixed to wood channels can solve the recycling problems as stated by Ma et al. [137,138];
  • More detailed studies to improve the stability of MOFs by in-situ growth of MOFs on wood should be carried out;
  • Huge efforts should be exerted by carrying out more intensive research works on the determination of the MOFs photocatalytic mechanisms such as the Z-schemes and S-schemes since it is considered to be the key to developing and engineering photocatalysts with efficient charge separation and transport.

6. Conclusions

MOFs have emerged as novel photocatalysts owing to their structural characteristics of large surface area, abundant active sites as well as well-ordered porous structure. However, fast recombination of electron-hole pairs has become one of the main challenges to provide an effective photocatalyst. Therefore, in this review, an overview of synthesis methods used in the fabrication of MOFs and MOFs-based heterostructures for various potential photocatalytic applications were comprehensively compiled and discussed. Mechanisms of photocatalytic performance using MOFs and MOFs-based heterostructures, such as generation of the charge carriers, separation of e/h+ pairs, and the degradation/removal pathways of representative pollutants have been illustrated in detail. However, an improvement in the overall characteristics of MOFs and MOFs-based heterostructures as visible light active materials should be the focus in a future study. Hence, research gaps regarding the development of MOFs and MOFs-based heterostructures were highlighted in the future prospects at the end of this review.

Author Contributions

A.R.: writing—original draft preparation, S.N.M.: writing—original draft preparation, M.M.K.: writing—review and editing, supervision, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universiti Brunei Darussalam FIC block grant number UBD/RSCH/1.4/FICBF(b)/2021/035. This is an invited review and the APC was waived off.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Khan, M.M.; Pradhan, D.; Sohn, Y. Nanocomposites for Visible Light-Induced Photocatalysis. In Springer Series on Polymer and Composite Materials; Khan, M.M., Pradhan, D., Sohn, Y., Eds.; Springer International Publishing: Cham, Switzerland, 2017. [Google Scholar] [CrossRef]
  2. Khan, M.E.; Khan, M.M.; Min, B.K.; Cho, M.H. Microbial fuel cell assisted band gap narrowed TiO2 for visible light induced photocatalytic activities and power generation. Sci. Rep. 2018, 8, 1723. [Google Scholar] [CrossRef] [PubMed]
  3. Rahman, A.; Harunsani, M.H.; Tan, A.L.; Khan, M.M. Zinc Oxide and Zinc Oxide-Based Nanostructures. In Biogenic and Phytogenic Synthesis, Properties and Applications; Springer: Berlin/Heidelberg, Germany, 2021; Volume 44, pp. 1333–1372. [Google Scholar] [CrossRef]
  4. Matussin, S.; Harunsani, M.H.; Tan, A.L.; Khan, M.M. Plant-extract-mediated SnO2 nanoparticles: Synthesis and applications. ACS Sustain. Chem. Eng. 2020, 8, 3040–3054. [Google Scholar] [CrossRef]
  5. Naidi, S.N.; Harunsani, M.H.; Tan, A.L.; Khan, M.M. Green-Synthesized CeO2 Nanoparticles for photocatalytic, antimicrobial, antioxidant and cytotoxicity activities. J. Mater. Chem. B 2021, 9, 5599–5620. [Google Scholar] [CrossRef] [PubMed]
  6. Matussin, S.N.; Rahman, A.; Khan, M.M. Role of Anions in the Synthesis and Crystal Growth of Selected Semiconductors. Front. Chem. 2022, 10, 881518. [Google Scholar] [CrossRef]
  7. Khan, M.M.; Adil, S.F.; Al-Mayouf, A. Metal oxides as photocatalysts. J. Saudi Chem. Soc. 2015, 19, 462–464. [Google Scholar] [CrossRef]
  8. Khan, M.E.; Khan, M.M.; Cho, M.H. Green synthesis, photocatalytic and photoelectrochemical performance of an Au–Graphene nanocomposite. RSC Adv. 2015, 5, 26897–26904. [Google Scholar] [CrossRef]
  9. Mohammad, A.; Karim, M.R.; Khan, M.E.; Khan, M.M.; Cho, M.H. Biofilm-assisted fabrication of Ag@SnO2-g-C3N4 nanostructures for visible light-induced photocatalysis and photoelectrochemical performance. J. Phys. Chem. C 2019, 123, 20936–20948. [Google Scholar] [CrossRef]
  10. Rahman, A.; Khan, M.M. Chalcogenides as photocatalysts. New J. Chem. 2021, 45, 19622–19635. [Google Scholar] [CrossRef]
  11. Khan, M.M. Chalcogenide-Based Nanomaterials as Photocatalysts, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  12. Kolobov, N.; Goesten, M.G.; Gascon, J. Metal-organic frameworks: Molecules or semiconductors in photocatalysis? Angew. Chem. Int. Ed. 2021, 60, 26038–26052. [Google Scholar] [CrossRef]
  13. Hao, Y.-C.; Chen, L.-W.; Li, J.; Guo, Y.; Su, X.; Shu, M.; Zhang, Q.; Gao, W.-Y.; Li, S.; Yu, Z.-L.; et al. Metal-organic framework membranes with single-atomic centers for photocatalytic CO2 and O2 reduction. Nat. Commun. 2021, 12, 2682. [Google Scholar] [CrossRef]
  14. Chuhadiya, S.; Suthar, D.; Patel, S.L.; Dhaka, M.S. Metal organic frameworks as hybrid porous materials for energy storage and conversion devices: A review. Coord. Chem. Rev. 2021, 446, 214115. [Google Scholar] [CrossRef]
  15. Dutta, A.; Pan, Y.; Liu, J.-Q.; Kumar, A. Multicomponent isoreticular metal-organic frameworks: Principles, current status and challenges. Coord. Chem. Rev. 2021, 445, 214074. [Google Scholar] [CrossRef]
  16. Nemiwal, M.; Gosu, V.; Zhang, T.C.; Kumar, D. Metal organic frameworks as electrocatalysts: Hydrogen evolution reactions and overall water splitting. Int. J. Hydrog. Energy 2021, 46, 10216–10238. [Google Scholar] [CrossRef]
  17. Nordin, N.A.; Mohamed, M.A.; Salehmin, M.N.I.; Mohd Yusoff, S.F. Photocatalytic active metal-organic Framework and its derivatives for solar-driven environmental remediation and renewable energy. Coord. Chem. Rev. 2022, 468, 214639. [Google Scholar] [CrossRef]
  18. Baumann, A.E.; Burns, D.A.; Liu, B.; Thoi, V.S. Metal-organic framework functionalization and design strategies for advanced electrochemical energy storage devices. Commun. Chem. 2019, 2, 86. [Google Scholar] [CrossRef]
  19. Li, X.; Wu, D.; Hua, T.; Lan, X.; Han, S.; Cheng, J.; Du, K.-S.; Hu, Y.; Chen, Y. Micro/macrostructure and multicomponent design of catalysts by MOF-derived strategy: Opportunities for the application of nanomaterials-based advanced oxidation processes in wastewater treatment. Sci. Total Environ. 2022, 804, 150096. [Google Scholar] [CrossRef]
  20. Suresh, K.; Aulakh, D.; Purewal, J.; Siegel, D.J.; Veenstra, M.; Matzger, A.J. Optimizing hydrogen storage in MOFs through engineering of crystal morphology and control of crystal size. J. Am. Chem. Soc. 2021, 143, 10727–10734. [Google Scholar] [CrossRef]
  21. Qiu, T.; Liang, Z.; Guo, W.; Tabassum, H.; Gao, S.; Zou, R. Metal-organic framework-based materials for energy conversion and storage. ACS Energy Lett. 2020, 5, 520–532. [Google Scholar] [CrossRef]
  22. Morshedy, A.S.; Abd El Salam, H.M.; El Naggar, A.M.A.; Zaki, T. Hydrogen production and in situ storage through process of water splitting using mono/binary metal-organic framework (MOF) structures as new chief photocatalysts. Energy Fuels 2020, 34, 11660–11669. [Google Scholar] [CrossRef]
  23. Zhang, X.; Tong, S.; Huang, D.; Liu, Z.; Shao, B.; Liang, Q.; Wu, T.; Pan, Y.; Huang, J.; Liu, Y.; et al. Recent advances of Zr based metal organic frameworks photocatalysis: Energy production and environmental remediation. Coord. Chem. Rev. 2021, 448, 214177. [Google Scholar] [CrossRef]
  24. Liu, M.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in core—Shell metal organic frame-based photocatalysts for solar energy conversion. Coord. Chem. Rev. 2021, 446, 214123. [Google Scholar] [CrossRef]
  25. Wang, Y.-C.; Liu, X.-Y.; Wang, X.-X.; Cao, M.-S. Metal-organic frameworks based photocatalysts: Architecture strategies for efficient solar energy conversion. Chem. Eng. J. 2021, 419, 129459. [Google Scholar] [CrossRef]
  26. Zhao, X.; Li, J.; Li, X.; Huo, P.; Shi, W. Design of Metal-organic frameworks (MOFs)-based photocatalyst for solar fuel production and photo-degradation of pollutants. Chin. J. Catal. 2021, 42, 872–903. [Google Scholar] [CrossRef]
  27. Li, H.; Li, L.; Lin, R.-B.; Zhou, W.; Zhang, Z.; Xiang, S.; Chen, B. Porous metal-organic frameworks for gas storage and separation: Status and challenges. EnergyChem 2019, 1, 100006. [Google Scholar] [CrossRef]
  28. Hanikel, N.; Prévot, M.S.; Fathieh, F.; Kapustin, E.A.; Lyu, H.; Wang, H.; Diercks, N.J.; Glover, T.G.; Yaghi, O.M. Rapid cycling and exceptional yield in a metal-organic framework water harvester. ACS Cent. Sci. 2019, 5, 1699–1706. [Google Scholar] [CrossRef]
  29. Wu, X.-Q.; Huang, D.-D.; Wu, Y.-P.; Zhao, J.; Liu, X.; Dong, W.-W.; Li, S.; Li, D.-S.; Li, J.-R. In situ synthesis of nano CuS-embedded MOF hierarchical structures and application in dye adsorption and hydrogen evolution reaction. ACS Appl. Energy Mater. 2019, 2, 5698–5706. [Google Scholar] [CrossRef]
  30. Zou, Z.; Wang, T.; Zhao, X.; Jiang, W.-J.; Pan, H.; Gao, D.; Xu, C. Expediting in-situ electrochemical activation of two-dimensional metal-organic frameworks for enhanced OER intrinsic activity by iron incorporation. ACS Catal. 2019, 9, 7356–7364. [Google Scholar] [CrossRef]
  31. Al-Rowaili, F.N.; Zahid, U.; Onaizi, S.; Khaled, M.; Jamal, A.; AL-Mutairi, E.M. A review for metal-organic frameworks (MOFs) utilization in capture and conversion of carbon dioxide into valuable products. J. CO2 Util. 2021, 53, 101715. [Google Scholar] [CrossRef]
  32. 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]
  33. Oladoye, P.O.; Adegboyega, S.A.; Giwa, A.-R.A. Remediation potentials of composite metal-organic frameworks (MOFs) for dyes as water contaminants: A comprehensive review of recent literatures. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100568. [Google Scholar] [CrossRef]
  34. Liu, S.; Jiang, X.; Waterhouse, G.I.N.; Zhang, Z.-M.; Yu, L. Construction of Z-scheme titanium-MOF/plasmonic silver nanoparticle/NiFe layered double hydroxide photocatalysts with enhanced dye and antibiotic degradation activity under visible light. Sep. Purif. Technol. 2022, 278, 119525. [Google Scholar] [CrossRef]
  35. Deng, Z.; Li, M.; Hu, Y.; He, Y.; Tao, B.; Yuan, Z.; Wang, R.; Chen, M.; Luo, Z.; Cai, K. Injectable biomimetic hydrogels encapsulating gold/metal-organic frameworks nanocomposites for enhanced antibacterial and wound healing activity under visible light actuation. Chem. Eng. J. 2021, 420, 129668. [Google Scholar] [CrossRef]
  36. Ghasempour, H.; ZareKarizi, F.; Morsali, A.; Yan, X.-W. Development of a highly porous Fe-based MOF using symmetrically incompatible building blocks: Selective oxidation of benzyl alcohols. Appl. Mater. Today 2021, 24, 101157. [Google Scholar] [CrossRef]
  37. Mallakpour, S.; Nikkhoo, E.; Hussain, C.M. Application of MOF materials as drug delivery systems for cancer therapy and dermal treatment. Coord. Chem. Rev. 2022, 451, 214262. [Google Scholar] [CrossRef]
  38. Lu, G.; Chu, F.; Huang, X.; Li, Y.; Liang, K.; Wang, G. Recent advances in metal-organic frameworks-based materials for photocatalytic selective oxidation. Coord. Chem. Rev. 2022, 450, 214240. [Google Scholar] [CrossRef]
  39. Li, T.-T.; Dang, L.-L.; Zhao, C.-C.; Lv, Z.-Y.; Yang, X.-G.; Zhao, Y.; Zhang, S.-H. A self-sensitized Co (II)-MOF for efficient visible-light-driven hydrogen evolution without additional cocatalysts. J. Solid State Chem. 2021, 304, 122609. [Google Scholar] [CrossRef]
  40. Mahmoud Idris, A.; Jiang, X.; Tan, J.; Cai, Z.; Lou, X.; Wang, J.; Li, Z. Dye-sensitized Fe-MOF nanosheets as visible-light driven photocatalyst for high efficient photocatalytic CO2 reduction. J. Colloid Interface Sci. 2022, 607, 1180–1188. [Google Scholar] [CrossRef]
  41. Tang, H.L.; Sun, X.J.; Zhang, F.M. Development of MOF-based heterostructures for photocatalytic hydrogen evolution. Dalton Trans. 2020, 49, 12136–12144. [Google Scholar] [CrossRef]
  42. Xiao, Y.; Guo, X.; Yang, N.; Zhang, F. Heterostructured MOFs photocatalysts for water splitting to produce hydrogen. J. Energy Chem. 2021, 58, 508–522. [Google Scholar] [CrossRef]
  43. Zhang, F.-M.; Dong, H.; Zhang, X.; Sun, X.-J.; Liu, M.; Yang, D.-D.; Liu, X.; Wei, J.-Z. Postsynthetic modification of ZIF-90 for potential targeted codelivery of two anticancer drugs. ACS Appl. Mater. Interfaces 2017, 9, 27332–27337. [Google Scholar] [CrossRef]
  44. Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): Routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933–969. [Google Scholar] [CrossRef] [PubMed]
  45. Gomez-Gualdron, D.A.; Gutov, O.V.; Krungleviciute, V.; Borah, B.; Mondloch, J.E.; Hupp, J.T.; Yildirim, T.; Farha, O.K.; Snurr, R.Q. Computational design of metal-organic frameworks based on stable zirconium building units for storage and delivery of methane. Chem. Mater. 2014, 26, 5632–5639. [Google Scholar] [CrossRef]
  46. Deria, P.; Gómez-Gualdrón, D.A.; Hod, I.; Snurr, R.Q.; Hupp, J.T.; Farha, O.K. Framework-topology-dependent catalytic activity of zirconium-based (porphinato)zinc(II) MOFs. J. Am. Chem. Soc. 2016, 138, 14449–14457. [Google Scholar] [CrossRef] [PubMed]
  47. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef]
  48. Chen, W.; Du, L.; Wu, C. Hydrothermal Synthesis of MOFs; Woodhead Publishing: Sawston, UK, 2020; pp. 141–157. [Google Scholar] [CrossRef]
  49. Denisov, G.L.; Primakov, P.V.; Korlyukov, A.A.; Novikov, V.V.; Nelyubina, Y.V. Solvothermal synthesis of the metal-organic framework MOF-5 in autoclaves prepared by 3D printing. Russ. J. Coord. Chem. Koord. Khimiya 2019, 45, 836–842. [Google Scholar] [CrossRef]
  50. Duan, C.; Zhang, Y.; Li, J.; Kang, L.; Xie, Y.; Qiao, W.; Zhu, C.; Luo, H. Rapid room-temperature preparation of hierarchically porous metal-organic frameworks for efficient uranium removal from aqueous solutions. Nanomaterials 2020, 10, 1539. [Google Scholar] [CrossRef]
  51. Vakili, R.; Xu, S.; Al-Janabi, N.; Gorgojo, P.; Holmes, S.M.; Fan, X. Microwave-assisted synthesis of zirconium-based metal organic frameworks (MOFs): Optimization and gas adsorption. Microporous Mesoporous Mater. 2018, 260, 45–53. [Google Scholar] [CrossRef]
  52. Giménez-Marqués, M.; Bellido, E.; Berthelot, T.; Simón-Yarza, T.; Hidalgo, T.; Simón-Vázquez, R.; González-Fernández, Á.; Avila, J.; Asensio, M.C.; Gref, R.; et al. GraftFast surface engineering to improve MOF nanoparticles furtiveness. Small 2018, 14, 1801900. [Google Scholar] [CrossRef]
  53. Xia, T.; Lin, Y.; Li, W.; Ju, M. Photocatalytic degradation of organic pollutants by MOFs based materials: A review. Chin. Chem. Lett. 2021, 32, 2975–2984. [Google Scholar] [CrossRef]
  54. Wang, X.R.; Huang, Z.; Du, J.; Wang, X.Z.; Gu, N.; Tian, X.; Li, Y.; Liu, Y.Y.; Huo, J.Z.; Ding, B. Hydrothermal preparation of five rare-earth (Re = Dy, Gd, Ho, Pr, and Sm) luminescent cluster-based coordination materials: The first MOFs-based ratiometric fluorescent sensor for lysine and bifunctional sensing platform for insulin and Al3+. Inorg. Chem. 2018, 57, 12885–12899. [Google Scholar] [CrossRef] [PubMed]
  55. Goswami, A.; Ghosh, D.; Chernyshev, V.V.; Dey, A.; Pradhan, D.; Biradha, K. 2D MOFs with Ni(II), Cu(II), and Co(II) as efficient oxygen evolution electrocatalysts: Rationalization of catalytic performance vs. structure of the MOFs and potential of the redox couples. ACS Appl. Mater. Interfaces 2020, 12, 33679–33689. [Google Scholar] [CrossRef] [PubMed]
  56. Remya, V.R.; Kurian, M. Synthesis and catalytic applications of metal-organic frameworks: A review on recent literature. Int. Nano Lett. 2019, 9, 17–29. [Google Scholar] [CrossRef]
  57. Zhang, B.; Luo, Y.; Kanyuck, K.; Saenz, N.; Reed, K.; Zavalij, P.; Mowery, J.; Bauchan, G. Facile and template-free solvothermal synthesis of mesoporous/macroporous metal-organic framework nanosheets. RSC Adv. 2018, 8, 33059–33064. [Google Scholar] [CrossRef] [PubMed]
  58. Virmani, E.; Rotter, J.M.; Mähringer, A.; von Zons, T.; Godt, A.; Bein, T.; Wuttke, S.; Medina, D.D. On-surface synthesis of highly oriented thin metal-organic framework films through vapor-assisted conversion. J. Am. Chem. Soc. 2018, 140, 4812–4819. [Google Scholar] [CrossRef]
  59. Stassin, T.; Rodríguez-Hermida, S.; Schrode, B.; Cruz, A.J.; Carraro, F.; Kravchenko, D.; Creemers, V.; Stassen, I.; Hauffman, T.; De Vos, D.; et al. Vapour-phase deposition of oriented copper dicarboxylate metal-organic framework thin films. Chem. Commun. 2019, 55, 10056–10059. [Google Scholar] [CrossRef]
  60. Kim, K.J.; Culp, J.T.; Ohodnicki, P.R.; Thallapally, P.K.; Tao, J. Synthesis of high-quality Mg-MOF-74 thin films via vapor-assisted crystallization. ACS Appl. Mater. Interfaces 2021, 13, 35223–35231. [Google Scholar] [CrossRef]
  61. Cho, H.-Y.; Kim, J.; Kim, S.-N.; Ahn, W.-S. High yield 1-L scale synthesis of ZIF-8 via a sonochemical route. Microporous Mesoporous Mater. 2013, 169, 180–184. [Google Scholar] [CrossRef]
  62. Hoskins, B.F.; Robson, R. Design and construction of a new class of scaffolding-like materials comprising infinite polymeric frameworks of 3D-linked molecular rods. A reappraisal of the zinc cyanide and cadmium cyanide structures and the synthesis and structure of the diamond-rela. J. Am. Chem. Soc. 1990, 112, 1546–1554. [Google Scholar] [CrossRef]
  63. Reif, B.; Somboonvong, J.; Hartmann, M.; Kaspereit, M.; Schwieger, W. Synthesis of ZIF-11 membranes: The influence of preparation technique and support type. Membranes 2021, 11, 523. [Google Scholar] [CrossRef]
  64. Liu, J.; Wöll, C. Surface-supported metal-organic framework thin films: Fabrication methods, applications, and challenges. Chem. Soc. Rev. 2017, 46, 5730–5770. [Google Scholar] [CrossRef]
  65. Kim, J.; Yang, S.-T.; Choi, S.B.; Sim, J.; Kim, J.; Ahn, W.-S. Control of catenation in CuTATB-n metal-organic frameworks by sonochemical synthesis and its effect on CO2 adsorption. J. Mater. Chem. 2011, 21, 3070. [Google Scholar] [CrossRef]
  66. Mahendran, N.; Praveen, K. BiPO4/Fe-metal organic framework composite: A promising photocatalyst toward the abatement of tetracycline hydrochloride, indigo carmine and reduction of 4-nitrophenol. J. Ind. Eng. Chem. 2021, 100, 220–232. [Google Scholar] [CrossRef]
  67. Zhang, Y.; Xu, J.; Zhou, J.; Wang, L. Metal-organic framework-derived multifunctional photocatalysts. Chin. J. Catal. 2022, 43, 971–1000. [Google Scholar] [CrossRef]
  68. Yang, H. A short review on heterojunction photocatalysts: Carrier transfer behavior and photocatalytic mechanisms. Mater. Res. Bull. 2021, 142, 111406. [Google Scholar] [CrossRef]
  69. Shi, Y.; Yang, A.-F.; Cao, C.-S.; Zhao, B. Applications of MOFs: Recent advances in photocatalytic hydrogen production from water. Coord. Chem. Rev. 2019, 390, 50–75. [Google Scholar] [CrossRef]
  70. Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium metal-organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem. Int. Ed. 2012, 51, 3364–3367. [Google Scholar] [CrossRef]
  71. Santaclara, J.G.; Kapteijn, F.; Gascon, J.; van der Veen, M.A. Understanding metal-organic frameworks for photocatalytic solar fuel production. CrystEngComm 2017, 19, 4118–4125. [Google Scholar] [CrossRef]
  72. Wang, Z.; Li, Q.; Su, R.; Lv, G.; Wang, Z.; Gao, B.; Zhou, W. Enhanced degradation of bisphenol F in a porphyrin-MOF based visible-light system under high salinity conditions. Chem. Eng. J. 2022, 428, 132106. [Google Scholar] [CrossRef]
  73. Li, Z.; Wang, L.; Qin, L.; Lai, C.; Wang, Z.; Zhou, M.; Xiao, L.; Liu, S.; Zhang, M. Recent advances in the application of water-stable metal-organic frameworks: Adsorption and photocatalytic reduction of heavy metal in water. Chemosphere 2021, 285, 131432. [Google Scholar] [CrossRef]
  74. Qiu, J.; Zhang, X.; Feng, Y.; Zhang, X.; Wang, H.; Yao, J. Modified metal-organic frameworks as photocatalysts. Appl. Catal. B Environ. 2018, 231, 317–342. [Google Scholar] [CrossRef]
  75. Zhang, Y.; Li, G.; Guo, Q. CdSe QDs@ Fe-based metal organic framework composites for improved photocatalytic RhB degradation under visible Light. Microporous Mesoporous Mater. 2021, 324, 111291. [Google Scholar] [CrossRef]
  76. Xiao, L.; Zhang, Q.; Chen, P.; Chen, L.; Ding, F.; Tang, J.; Li, Y.; Au, C.-T.; Yin, S.-F. Copper-mediated metal-organic framework as efficient photocatalyst for the partial oxidation of aromatic alcohols under visible-light irradiation: Synergism of plasmonic effect and schottky junction. Appl. Catal. B Environ. 2019, 248, 380–387. [Google Scholar] [CrossRef]
  77. Kamandi, R.; Mahmoodi, N.M.; Kazemeini, M. Graphitic carbon nitride nanosheet/metal-organic framework heterostructure: Synthesis and pollutant degradation using visible light. Mater. Chem. Phys. 2021, 269, 124726. [Google Scholar] [CrossRef]
  78. Hou, W.; Chen, M.; Chen, C.; Wang, Y.; Xu, Y. Increased production of H2 under visible light by packing CdS in a Ti, Zr-based metal organic framework. J. Colloid Interface Sci. 2021, 604, 310–318. [Google Scholar] [CrossRef] [PubMed]
  79. Zhai, B.Y.; Chen, Y.; Liang, Y.N.; Gao, Y.H.; Shi, J.L.; Zhang, H.L.; Li, Y. Modifying Ag3VO4 with metal-organic frameworks for enhanced photocatalytic activity under visible light. Mater. Chem. Phys. 2020, 239, 122078. [Google Scholar] [CrossRef]
  80. Huang, W.; Liu, N.; Zhang, X.; Wu, M.; Tang, L. Metal organic framework G-C3N4/MIL-53(Fe) heterojunctions with enhanced photocatalytic activity for Cr(VI) reduction under visible light. Appl. Surf. Sci. 2017, 425, 107–116. [Google Scholar] [CrossRef]
  81. He, Y.; Wang, D.; Li, X.; Fu, Q.; Yin, L.; Yang, Q.; Chen, H. Photocatalytic degradation of tetracycline by metal-organic frameworks modified with Bi2WO6 nanosheet under direct sunlight. Chemosphere 2021, 284, 131386. [Google Scholar] [CrossRef]
  82. Zhang, C.; Guo, D.; Shen, T.; Hou, X.; Zhu, M.; Liu, S.; Hu, Q. Titanium dioxide/magnetic metal-organic framework preparation for organic pollutants removal from water under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2020, 589, 124484. [Google Scholar] [CrossRef]
  83. Xia, Q.; Huang, B.; Yuan, X.; Wang, H.; Wu, Z.; Jiang, L.; Xiong, T.; Zhang, J.; Zeng, G.; Wang, H. Modified stannous sulfide nanoparticles with metal-organic framework: Toward efficient and enhanced photocatalytic reduction of chromium (VI) under visible light. J. Colloid Interface Sci. 2018, 530, 481–492. [Google Scholar] [CrossRef]
  84. Wang, Y.L.; Zhang, S.; Zhao, Y.F.; Bedia, J.; Rodriguez, J.J.; Belver, C. UiO-66-based metal organic frameworks for the photodegradation of acetaminophen under simulated solar irradiation. J. Environ. Chem. Eng. 2021, 9, 106087. [Google Scholar] [CrossRef]
  85. Parnicka, P.; Lisowski, W.; Klimczuk, T.; Łuczak, J.; Żak, A.; Zaleska-Medynska, A. Visible-light-driven lanthanide-organic-frameworks modified TiO2 photocatalysts utilizing up-conversion effect. Appl. Catal. B Environ. 2021, 291, 120056. [Google Scholar] [CrossRef]
  86. Tang, T.; Jin, X.; Tao, X.; Huang, L.; Shang, S. Low-crystalline Ce-based bimetallic MOFs synthesized via DBD plasma for excellent visible photocatalytic performance. J. Alloy. Compd. 2022, 895, 162452. [Google Scholar] [CrossRef]
  87. Chen, H.; Liu, C.; Guo, W.; Wang, Z.; Shi, Y.; Yu, Y.; Wu, L. Functionalized UiO-66(Ce) for photocatalytic organic transformation: The role of active sites modulated by ligand functionalization. Catal. Sci. Technol. 2022, 12, 1812–1823. [Google Scholar] [CrossRef]
  88. Somnath; Ahmad, M.; Siddiqui, K.A. Synthesis of mixed ligand 3D cobalt MOF: Smart responsiveness towards photocatalytic dye degradation in environmental contaminants. J. Mol. Struct. 2022, 1265, 133399. [Google Scholar] [CrossRef]
  89. Shan, C.; Zhang, X.; Ma, S.; Xia, X.; Shi, Y.; Yang, J. Preparation and application of bimetallic mixed ligand MOF photocatalytic materials. Colloids Surf. A Physicochem. Eng. Asp. 2022, 636, 128108. [Google Scholar] [CrossRef]
  90. Li, M.; Yuan, J.; Wang, G.; Yang, L.; Shao, J.; Li, H.; Lu, J. One-step construction of Ti-in bimetallic MOFs to improve synergistic effect of adsorption and photocatalytic degradation of bisphenol A. Sep. Purif. Technol. 2022, 298, 121658. [Google Scholar] [CrossRef]
  91. Han, W.; Shao, L.-H.; Sun, X.-J.; Liu, Y.-H.; Zhang, F.-M.; Wang, Y.; Dong, P.-Y.; Zhang, G.-L. Constructing Cu ion sites in MOF/COF heterostructure for noble-metal-free photoredox catalysis. Appl. Catal. B Environ. 2022, 317, 121710. [Google Scholar] [CrossRef]
  92. Ye, Z.; Feng, S.; Wu, W.; Zhou, Y.; Wang, Y.; Dai, X.; Cao, X. Synthesis of double MOFs composite material for visible light photocatalytic degradation of tetracycline. Solid State Sci. 2022, 127, 106842. [Google Scholar] [CrossRef]
  93. Lv, S.-W.; Liu, J.-M.; Li, C.-Y.; Zhao, N.; Wang, Z.-H.; Wang, S. Two novel MOFs@COFs hybrid-based photocatalytic platforms coupling with sulfate radical-involved advanced oxidation processes for enhanced degradation of bisphenol A. Chemosphere 2020, 243, 125378. [Google Scholar] [CrossRef]
  94. Guo, J.; Liang, Y.; Liu, L.; Hu, J.; Wang, H.; An, W.; Cui, W. Noble-metal-free CdS/Ni-MOF composites with highly efficient charge separation for photocatalytic H2 evolution. Appl. Surf. Sci. 2020, 522, 146356. [Google Scholar] [CrossRef]
  95. Liu, X.; Zhang, L.; Li, Y.; Xu, X.; Du, Y.; Jiang, Y.; Lin, K. A novel heterostructure coupling MOF-derived fluffy porous indium oxide with g-C3N4 for enhanced photocatalytic activity. Mater. Res. Bull. 2021, 133, 111078. [Google Scholar] [CrossRef]
  96. Lu, W.; Duan, C.; Liu, C.; Zhang, Y.; Meng, X.; Dai, L.; Wang, W.; Yu, H.; Ni, Y. A self-cleaning and photocatalytic cellulose-fiber-supported “Ag@AgCl@MOF-Cloth” membrane for complex wastewater remediation. Carbohydr. Polym. 2020, 247, 116691. [Google Scholar] [CrossRef]
  97. Wang, Q.; Gao, Q.; Al-Enizi, A.M.; Nafady, A.; Ma, S. Recent advances in MOF-based photocatalysis: Environmental remediation under visible light. Inorg. Chem. Front. 2020, 7, 300–339. [Google Scholar] [CrossRef]
  98. Zhao, S.; Xu, J.; Mao, M.; Li, L.; Li, X. Protonated G-C3N4 cooperated with Co-MOF doped with Sm to construct 2D/2D heterojunction for integrated dye-sensitized photocatalytic H2 evolution. J. Colloid Interface Sci. 2021, 583, 435–447. [Google Scholar] [CrossRef] [PubMed]
  99. Zhang, X.; Lin, B.; Li, X.; Wang, X.; Huang, K.; Chen, Z. MOF-derived magnetically recoverable Z-scheme ZnFe2O4/Fe2O3 perforated nanotube for efficient photocatalytic ciprofloxacin removal. Chem. Eng. J. 2022, 430, 132728. [Google Scholar] [CrossRef]
  100. Xu, G.; Zhang, Y.; Peng, D.; Sheng, D.; Zhang, Y.; Tian, Y.; Ma, D. MOF derived carbon modified porous TiO2 mixed-phase junction with efficient visible-light photocatalysis for cyclohexane oxidation. Mater. Res. Bull. 2021, 146, 111602. [Google Scholar] [CrossRef]
  101. Xia, Z.; Shi, B.; Zhu, W.; Lü, C. Temperature-responsive polymer-tethered Zr-porphyrin MOFs encapsulated carbon dot nanohybrids with boosted visible-light photodegradation for organic contaminants in water. Chem. Eng. J. 2021, 426, 131794. [Google Scholar] [CrossRef]
  102. Wang, D.; Zhu, R.; Lou, J.; Yuan, J.; Xu, J.; Fan, X. Novel marigold-like CuO@Cu-Based MOFs composite photocatalyst for high-performance removal of alkylphenol ethoxylate under visible light. J. Environ. Chem. Eng. 2021, 9, 106434. [Google Scholar] [CrossRef]
  103. Prakash Tripathy, S.; Subudhi, S.; Das, S.; Kumar Ghosh, M.; Das, M.; Acharya, R.; Acharya, R.; Parida, K. Hydrolytically stable citrate capped Fe3O4@UiO-66-NH2 MOF: A hetero-structure composite with enhanced activity towards Cr (VI) adsorption and photocatalytic H2 evolution. J. Colloid Interface Sci. 2022, 606, 353–366. [Google Scholar] [CrossRef]
  104. Thi, Q.V.; Tamboli, M.S.; Thanh Hoai Ta, Q.; Kolekar, G.B.; Sohn, D. A nanostructured MOF/reduced graphene oxide hybrid for enhanced photocatalytic efficiency under solar Light. Mater. Sci. Eng. B 2020, 261, 114678. [Google Scholar] [CrossRef]
  105. Tang, B.; Dai, Y.; Sun, Y.; Chen, H.; Wang, Z. Graphene and MOFs co-modified composites for high adsorption capacity and photocatalytic performance to remove pollutant under both UV- and visible-light irradiation. J. Solid State Chem. 2020, 284, 121215. [Google Scholar] [CrossRef]
  106. Qian, J.-F.; Yue, H.-D.; Qiu, P.-X.; Liang, Q.; Hang, M.-T.; He, M.-Y.; Bu, Y.-F.; Chen, Q.; Zhang, Z.-H. Anions mediated amino-type Cd-MOFs catalysts for efficient photocatalytic hydrogen evolution. J. Solid State Chem. 2021, 304, 122632. [Google Scholar] [CrossRef]
  107. Peña-Velasco, G.; Hinojosa-Reyes, L.; Morán-Quintanilla, G.A.; Hernández-Ramírez, A.; Villanueva-Rodríguez, M.; Guzmán-Mar, J.L. Synthesis of heterostructured catalyst coupling MOF derived Fe2O3 with TiO2 for enhanced photocatalytic activity in anti-inflammatory drugs mixture degradation. Ceram. Int. 2021, 47, 24632–24640. [Google Scholar] [CrossRef]
  108. Hou, C.; Chen, W.; Fu, L.; Zhang, S.; Liang, C.; Wang, Y. Efficient degradation of perfluorooctanoic acid by electrospun lignin-based bimetallic MOFs nanofibers composite membranes with peroxymonosulfate under solar light irradiation. Int. J. Biol. Macromol. 2021, 174, 319–329. [Google Scholar] [CrossRef] [PubMed]
  109. 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]
  110. Zhu, Z.; Zhu, C.; Hu, C.; Liu, B. Facile fabrication of BiOIO3/MIL-88B heterostructured photocatalysts for removal of pollutants under visible light irradiation. J. Colloid Interface Sci. 2022, 607, 595–606. [Google Scholar] [CrossRef]
  111. Pattappan, D.; Kavya, K.V.; Vargheese, S.; Kumar, R.T.R.; Haldorai, Y. Graphitic carbon Nitride/NH2-MIL-101(Fe) composite for environmental remediation: Visible-light-assisted photocatalytic degradation of acetaminophen and reduction of hexavalent chromium. Chemosphere 2022, 286, 131875. [Google Scholar] [CrossRef]
  112. 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]
  113. Abdollahi, B.; Najafidoust, A.; Abbasi Asl, E.; Sillanpaa, M. Fabrication of ZiF-8 metal organic framework (MOFs)-based CuO-ZnO photocatalyst with enhanced solar-light-driven property for degradation of organic dyes. Arab. J. Chem. 2021, 14, 103444. [Google Scholar] [CrossRef]
  114. Li, J.; Li, X.; Wu, G.; Guo, J.; Yin, X.; Mu, M. Construction of 2D Co-TCPP MOF decorated on B-TiO2−X nanosheets: Oxygen vacancy and 2D-2D heterojunctions for enhancing visible Light-driven photocatalytic degradation of bisphenol A. J. Environ. Chem. Eng. 2021, 9, 106723. [Google Scholar] [CrossRef]
  115. Qiu, J.-L.; Su, J.; Muhammad, N.; Zheng, W.-T.; Yue, C.-L.; Liu, F.-Q.; Zuo, J.-L.; Ding, Z.-J. Facile encapsulating Ag nanoparticles into a tetrathiafulvalene-based Zr-MOF for enhanced photocatalysis. Chem. Eng. J. 2022, 427, 131970. [Google Scholar] [CrossRef]
  116. Xu, M.; Sun, C.; Zhao, X.; Jiang, H.; Wang, H.; Huo, P. Fabricated hierarchical CdS/Ni-MOF heterostructure for promoting photocatalytic reduction of CO2. Appl. Surf. Sci. 2022, 576, 151792. [Google Scholar] [CrossRef]
  117. Chen, L.; Ren, X.; Alharbi, N.S.; Chen, C. Fabrication of a novel Co/Ni-MOFs@BiOI composite with boosting photocatalytic degradation of methylene blue under visible light. J. Environ. Chem. Eng. 2021, 9, 106194. [Google Scholar] [CrossRef]
  118. Govindaraju, S.; Arumugasamy, S.K.; Chellasamy, G.; Yun, K. Zn-MOF decorated bio activated carbon for photocatalytic degradation, oxygen evolution and reduction catalysis. J. Hazard. Mater. 2022, 421, 126720. [Google Scholar] [CrossRef] [PubMed]
  119. Hussain, M.Z.; Yang, Z.; Khalil, A.M.E.; Hussain, S.; Awan, S.U.; Jia, Q.; Fischer, R.A.; Zhu, Y.; Xia, Y. Metal-organic framework derived multi-functionalized and co-doped TiO2/C nanocomposites for excellent visible-light photocatalysis. J. Mater. Sci. Technol. 2022, 101, 49–59. [Google Scholar] [CrossRef]
  120. Sonowal, K.; Nandal, N.; Basyach, P.; Kalita, L.; Jain, S.L.; Saikia, L. Photocatalytic reduction of CO2 to methanol using Zr (IV)-based MOF composite with g-C3N4 quantum dots under visible light irradiation. J. CO2 Util. 2022, 57, 101905. [Google Scholar] [CrossRef]
  121. Kumar, V.; Singh, V.; Kim, K.-H.; Kwon, E.E.; Younis, S.A. Metal-organic frameworks for photocatalytic detoxification of chromium and uranium in water. Coord. Chem. Rev. 2021, 447, 214148. [Google Scholar] [CrossRef]
  122. Pan, Y.; Abazari, R.; Yao, J.; Gao, J. Recent progress in 2D metal-organic framework photocatalysts: Synthesis, photocatalytic mechanism and applications. J. Phys. Energy 2021, 3, 032010. [Google Scholar] [CrossRef]
  123. Yu, J.; Park, J.; Van Wyk, A.; Rumbles, G.; Deria, P. Excited-state electronic properties in Zr-based metal-organic frameworks as a function of a topological network. J. Am. Chem. Soc. 2018, 140, 10488–10496. [Google Scholar] [CrossRef]
  124. Du, C.; Zhang, Z.; Yu, G.; Wu, H.; Chen, H.; Zhou, L.; Zhang, Y.; Su, Y.; Tan, S.; Yang, L.; et al. A review of metal organic framework (MOFs)-based materials for antibiotics removal via adsorption and photocatalysis. Chemosphere 2021, 272, 129501. [Google Scholar] [CrossRef]
  125. Mukherjee, D.; van der Bruggen, B.; Mandal, B. Advancements in visible light responsive MOF composites for photocatalytic decontamination of textile wastewater: A review. Chemosphere 2022, 295, 133835. [Google Scholar] [CrossRef] [PubMed]
  126. Xiao, J.-D.; Jiang, H.-L. Metal-organic frameworks for photocatalysis and photothermal catalysis. Acc. Chem. Res. 2019, 52, 356–366. [Google Scholar] [CrossRef] [PubMed]
  127. Younis, S.A.; Kwon, E.E.; Qasim, M.; Kim, K.-H.; Kim, T.; Kukkar, D.; Dou, X.; Ali, I. Metal-organic framework as a photocatalyst: Progress in modulation strategies and environmental/energy applications. Prog. Energy Combust. Sci. 2020, 81, 100870. [Google Scholar] [CrossRef]
  128. Ramyashree, M.S.; Shanmuga Priya, S.; Freudenberg, N.C.; Sudhakar, K.; Tahir, M. Metal-organic framework-based photocatalysts for carbon dioxide reduction to methanol: A review on progress and application. J. CO2 Util. 2021, 43, 101374. [Google Scholar] [CrossRef]
  129. Ikreedeegh, R.R.; Tahir, M. A Critical review in recent developments of metal-organic-frameworks (MOFs) with band engineering alteration for photocatalytic CO2 reduction to solar fuels. J. CO2 Util. 2021, 43, 101381. [Google Scholar] [CrossRef]
  130. Sun, D.; Ye, L.; Li, Z. Visible-light-assisted aerobic photocatalytic oxidation of amines to imines over NH2-MIL-125(Ti). Appl. Catal. B Environ. 2015, 164, 428–432. [Google Scholar] [CrossRef]
  131. Chen, S.; Yang, F.; Gao, H.; Wang, J.; Chen, X.; Zhang, X.; Li, J.; Li, A. Construction of dual ligand Ti-based MOFs with enhanced photocatalytic CO2 reduction performance. J. CO2 Util. 2021, 48, 101528. [Google Scholar] [CrossRef]
  132. Behera, P.; Subudhi, S.; Tripathy, S.P.; Parida, K. MOF derived nano-materials: A recent progress in strategic fabrication, characterization and mechanistic insight towards divergent photocatalytic applications. Coord. Chem. Rev. 2022, 456, 214392. [Google Scholar] [CrossRef]
  133. Qiao, Z.; Wang, W.; Liu, N.; Huang, H.T.; Karuppasamy, L.; Yang, H.J.; Liu, C.H.; Wu, J.J. Synthesis of MOF/MoS2 composite photocatalysts with enhanced photocatalytic performance for hydrogen evolution from water splitting. Int. J. Hydrog. Energy, 2021; in press. [Google Scholar] [CrossRef]
  134. Khosroshahi, N.; Bakhtian, M.; Safarifard, V. Mechanochemical synthesis of ferrite/MOF nanocomposite: Efficient photocatalyst for the removal of meropenem and hexavalent chromium from water. J. Photochem. Photobiol. A Chem. 2022, 431, 114033. [Google Scholar] [CrossRef]
  135. Cao, Y.; Wang, G.; Liu, H.; Li, Y.; Jin, Z.; Ma, Q. Regular octahedron Cu-MOFs modifies Mn0.05Cd0.95S nanoparticles to form a S-scheme heterojunction for photocatalytic hydrogen evolution. Int. J. Hydrog. Energy 2021, 46, 7230–7240. [Google Scholar] [CrossRef]
  136. Bao, Y.; Song, S.; Yao, G.; Jiang, S. S-scheme photocatalytic systems. Solar RRL 2021, 5, 2100118. [Google Scholar] [CrossRef]
  137. Ma, X.; Xiong, Y.; Liu, Y.; Han, J.; Duan, G.; Chen, Y.; He, S.; Mei, C.; Jiang, S.; Zhang, K. When MOFs meet wood: From opportunities toward applications. Chem, 2022; in press. [Google Scholar] [CrossRef]
  138. 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]
Figure 1. Tunable properties of MOF during a synthesis.
Figure 1. Tunable properties of MOF during a synthesis.
Nanomaterials 12 02820 g001
Figure 2. Different applications of MOFs and MOFs-based heterostructures.
Figure 2. Different applications of MOFs and MOFs-based heterostructures.
Nanomaterials 12 02820 g002
Figure 3. The mechanism of photocatalysis of pristine MOFs.
Figure 3. The mechanism of photocatalysis of pristine MOFs.
Nanomaterials 12 02820 g003
Figure 4. The reaction mechanism of photocatalytic CO2 reduction using D-TiMOF. Reprinted with permission from Ref. [131]. Copyright 2022 Elsevier.
Figure 4. The reaction mechanism of photocatalytic CO2 reduction using D-TiMOF. Reprinted with permission from Ref. [131]. Copyright 2022 Elsevier.
Nanomaterials 12 02820 g004
Figure 5. Example of photocatalysis mechanism of MOFs-based heterostructures.
Figure 5. Example of photocatalysis mechanism of MOFs-based heterostructures.
Nanomaterials 12 02820 g005
Figure 6. Z-scheme of photocatalytic CIP degradation system. Reprinted with permission from Ref. [99]. Copyright 2022 Elsevier.
Figure 6. Z-scheme of photocatalytic CIP degradation system. Reprinted with permission from Ref. [99]. Copyright 2022 Elsevier.
Nanomaterials 12 02820 g006
Figure 7. (ac) Z-scheme and (df) S-scheme photocatalytic systems. Reprinted with permission from Ref. [136]. Copyright 2022 John Wiley and Sons.
Figure 7. (ac) Z-scheme and (df) S-scheme photocatalytic systems. Reprinted with permission from Ref. [136]. Copyright 2022 John Wiley and Sons.
Nanomaterials 12 02820 g007
Table 1. Synthesis of MOF materials using various methods.
Table 1. Synthesis of MOF materials using various methods.
No.MaterialsSynthesis MethodsAverage Particle SizeRef.
1PEGylated iron trimesate (materials institute Lavoisier) MIL-100 (Fe)Hydrothermal129–174 nm[52]
2Rare-earth 3D cluster-based MOFsHydrothermal0.4–2.7 μm[54]
32D-MOFs with Ni(II), Cu(II), and Co(II)Hydrothermal-[55]
4Zn4O(BDS)3 MOFsSolvothermal807 nm[49]
5Cu(II)-based MOFsSolvothermal50 nm[57]
6Zr-based MOFsMicrowaveLess than 100 nm[51]
7Cu-BTC MOFsRoom temperature template method-[50]
8UiO-66 MOFsVapor-assisted200 nm[58]
9Copper dicarboxylate MOFsChemical vapor deposition-[59]
10Zn-based MOFsSonochemical-[61]
Table 2. Different modifications of MOFs.
Table 2. Different modifications of MOFs.
No.MOFs-Based HeterostructuresSynthesis MethodBand Gap
Energy
Photocatalytic ActivitiesPhotocatalytic PerformanceRef.
1BiPO4/Fe-metal
organic framework composite
Solvothermal methodPristine NH2-MIL-53(Fe): 2.11 eV
BiPO4: 3.75 eV
Photocatalytic degradation of tetracycline hydrochloride,
Indigo Carmine and reduction of 4-nitrophenol
Tetracycline hydrochloride: degraded 80% within 120 min
Indigo Carmine: degraded 94% within 120 min
4-nitrophenol: 95% reduction efficiency within 12 min
[66]
2CdSe QDs@ Fe-based metal-organic framework compositesHeated to reflux in oil bathCdSe: 1.97 eV
Fe-BDC: 2.71 eV
Photocatalytic degradation of rhodamine BAlmost completely degraded ~99.8% after 240 min[75]
3Copper-mediated metal-organic frameworkAdvanced double-solvent approach followed by one-step reduction-Photocatalytic oxidation of aromatic alcoholsUp to 69.4% conversion[76]
4Graphitic carbon nitride nanosheet/metal-organic frameworkHydrothermalg-C3N4: 2.72 eV
MIL-101(Fe): 2.55 eV
g-C3N4/MIL-101(Fe):
2.05 eV
Photocatalytic degradation of rhodamine BMIL-101(Fe)/g-C3N4: 99.3% Bare g-C3N4: 40%
Bare MIL-101(Fe): 73%
[77]
5CdS in a Ti, Zr-based metal-organic frameworkStepwise precipitationP415: 3.47 eV
P415-NH2:
2.30 eV
CdS: 2.41 eV
Photocatalytic production of H2The rate of H2 production:
5.8 μmol min−1g−1
[78]
6Hydrogels encapsulating Gold/metal-organic
frameworks nanocomposites
Schiff base reaction and radical polymerizationZIF-8: 3.19 eVPhotocatalytic antibacterial and wound healing
activities
Au@ZIF-8@GCOA: up to 99.1% for E. coli and 99.6% for S. aureus after 20 min of irradiation[35]
7Modified Ag3VO4 with metal-organic frameworksFacile two-step methodAg3VO4: 2.0 eV 20%AZ: 2.11 eV
40%AZ: 2.21 eV
60%AZ: 2.25 eV
80%AZ: 2.83 eV
ZIF-8: 5.06 eV
Photocatalytic degradation of rhodamine B60%AZ composites exhibited the highest, about 4.2 times
of pure Ag3VO and 22.8 times of bare ZIF-8, respectively
[79]
8Metal-organic framework
g-C3N4/MIL-53(Fe) heterojunctions
Solvothermal methodMIL-53(Fe):
2.72 eV
CMFe-3:
2.51 eV
Photocatalytic reduction of Cr(VI)g-C3N4/MIL-53(Fe) showed about 2.1 and 2.0 times higher photocatalytic efficiency for the reduction of Cr(VI) in comparison to pure g-C3N4 and MIL-53(Fe), respectively.[80]
9Metal-organic frameworks
modified with Bi2WO6 nanosheet
Hydrothermal method-Photocatalytic degradation of tetracycline12%MIL/BWO achieved
the highest removal efficiency of about 92.4% within 120 min
[81]
10Titanium dioxide/magnetic metal-organic frameworkHydrothermal methodTiO2: 3.1 eV
TiO2/magnetic MIL-101(Cr): 1.61 eV
Photocatalytic degradation of
acid red 1
TiO2/magnetic MIL-101(Cr) showed 90% degradation of acid red 1[82]
11Modified stannous sulfide nanoparticles with a metal-organic frameworkDeposition method at room temperaturepure MIL-53(Fe): 1.6 eV
SnS: 0.87 eV
Photocatalytic reduction of chromium
(VI)
71.3% of Cr(VI) removal is achieved after 60 min[83]
12UiO-66-based metal-organic frameworksSolvothermal methodUiO-66:
3.82 eV
UiO-66-NH2: 2.63 eV
UiO-66-(OH)2: 2.50 eV
Photocatalytic degradation of acetaminophen90% after 6 h[84]
13Lanthanide-organic-frameworks modified TiO2Solvothermal methodTiO2: ~3.2 eV
Ln(ndc) MOF: ~2.9 eV
Photocatalytic degradation of phenol87.5% after 60 min[85]
14Fe/Ce-based bimetallic MOFDielectric barrier discharge plasmaCe-MOF: 3.01 eV
Fe/Ce-MOF-1: 1.90 eV
Fe/Ce-MOF-2: 1.97 eV
Fe/Ce-MOF-3: 1.75 eV
Photocatalytic degradation of methyl orangeFe/Ce-MOF-2 could degrade 93% methyl orange in 30 min under visible light[86]
Table 3. Different photocatalytic applications of MOFs and MOFs-based heterostructures.
Table 3. Different photocatalytic applications of MOFs and MOFs-based heterostructures.
No.Synthesized
Materials
Band Gap
Energy
ApplicationsSource of LightEfficiencyRef.
1Sm doped 2D Co-MOF/
protonated-g-C3N4
g-C3N4:
2.82 eV
Protonated-g-C3N4:
2.70 eV
2D Co-MOF:
3.12 eV
Photocatalytic evolution of hydrogen300 W Xe lamp with an AM 1.5G filterThe H2 evolution rate of 5% Sm doped with the highest activity has reached 73.42 μmolh−1.[98]
2MOF-derived magnetically recoverable ZnFe2O4/Fe2O3
perforated nanotube
Fe2O3: 2.10 eV
ZnFe2O4: 2.24 eV
Photocatalytic removal of ciprofloxacin 300 W Xe lamp with AM 1.5 filterZFF-2 exhibits the best photocatalytic ciprofloxacin degradation performance with
a degradation percentage of 96.5% and a TOC removal percentage up to 89% under light irradiation for 180 min
[99]
3MOF derived carbon modified porous TiO2C-TiO2-3: 2.45 eV
C-TiO2-4: 2.85 eV
C-TiO2-5: 2.94 eV
C-TiO2-6: 2.37 eV
C-TiO2-7: 2.93 eV
Photocatalytic oxidation of cyclohexane 100 W mercury lamp
with a UV cutoff filter (λ ≥ 420 nm)
C-TiO2-4 catalyst, the yield of cyclohexanol and cyclohexanone could reach 13.18 and 107.07 μmol respectively,
and the selectivity to cyclohexanone reaches 89.0%.
[100]
4Polymer-tethered Zr-porphyrin MOFs
encapsulated carbon dot nanohybrids
PCN-222: 1.75 eV
Citric acid derived QDs@PCN-222:
1.64 eV
Nitrogen-doped CDs@PCN-222:
1.59 eV
Sulfur- doped CDs@PCN-222:
1.66 eV
Photocatalytic degradation of rhodamine B and tetracycline300 W UV
lamp with a 420 filter
The removal efficiency of rhodamine B and tetracycline
reached almost 100% and 90.93%, respectively under 20 min visible light irradiation
[101]
5Marigold-like CuO@Cu-based MOFs compositeCuO: 1.41 eV
Cu-H3BTC MOF:
2.0 eV
Photocatalytic removal of alkylphenol ethoxylate300 W Xe lamp with
a 420 nm cut-off filter
CuO@Cu-H3BTC MOF composite exhibited up to 79% removal of alkylphenol polyethoxylate within 120 min of visible light irradiation[102]
6Citrate capped Fe3O4@
UiO-66-NH2 MOF
Citrate capped Fe3O4:
1.45 eV
UiO-66-NH2:
2.67 eV
Adsorption of Cr (VI) and photocatalytic evolution of H2300 W Xe lampMU-2 showed a maximum monolayer adsorption capacity of 743 mg g−1 which followed pseudo-second-order kinetics.
In addition, the synthesized composite material displayed enhanced activity towards photocatalytic H2 evolution with a maximum evolution rate of 417 μmolh−1
with an apparent conversion efficiency of 3.12%.
[103]
7MOF/
reduced graphene oxide hybrid
Not calculatedPhotocatalytic degradation of methylene blue, rhodamine B and methyl orangeSolar simulator
with wavelengths in the range of 300–1900 nm
MOF-5@rGO photocatalytic degradation efficiency reached 93% after 20 min illumination[104]
8ZIF-8,
UiO-66 and Cu-BTC are adopted to combine with
rGO and TiO2
Uio-66-rGO/TiO2:
2.6 eV
Cu-BTC-rGO/TiO2:
2.80 eV
ZIF-8-rGO/TiO2:
2.90 eV
Photocatalytic degradation of rhodamine B500 W Xe lamp and a cutoff filter
(λ > 400 nm)
Degradation rate constants of rhodamine B for UiO-66-RGO/TiO2 reached
1.65 × 10−1min−1 and
1.12 × 10−1 min−1 under UV- and visible-light irradiation, respectively.
[105]
9Cd-MOFsNot calculatedPhotocatalytic evolution of H2300 W Xe
lamp
The highest H2 production rate reached 17,242 μmolg−1 h−1[106]
10Co/Ni-MOFs@
BiOI composite
Not calculatedPhotocatalytic degradation of crystal violet, rhodamine 6G, malachite green, Congo red and methyl orange300 W PLS-
SXE with a cut-off filter (λ = 420 nm)
The removal efficiency of Co/[email protected]%
followed the order of MG > MB > MO > CR > CV > R6G, and the degradation rate constants (k, min−1) were 0.04705, 0.00627, 0.00681,
0.00695, 0.00347, 0.00066, respectively.
[91]
11MIL-101(Fe) derived Fe2O3 with TiO2Fe2O3: 1.7–2.0 eV
TiO2: 3.17 eV
FeTi-3%: 2.94 eV
Photocatalytic degradation of a mixture of nonsteroidal anti-inflammatory drugs namely ibuprofen and naproxen Philips 25 W/m2,
λ = 365–700 nm
The material exhibited up to 91% and 100% degradation of ibuprofen and naproxen within 240 and 15 min of reaction, respectively.[107]
12Ti-MOF/ plasmonic Ag nanoparticle/
NiFe layered double hydroxide
Ti-MOF:
2.53 eV
NiFeLDH:
2.19 eV
Ti-MOF/
NiFeLDH: 2.41 eV
Ti-MOF/Ag/
NiFeLDH-1:
2.17 eV
Ti-MOF/Ag/
NiFeLDH-2:
2.09 eV
Ti-MOF/Ag/
NiFeLDH-3:
2.12 eV Ti-MOF/Ag/
NiFeLDH-4:
2.15 eV
Photocatalytic degradation of antibiotics,
levofloxacin and rhodamine B dye
300 W Xe
With a cut-off filter of
λ > 420 nm
Ti-MOF/Ag/NiFeLDH composite displayed outstanding photocatalytic degradation up to 95% for rhodamine B removal within 50 min and 92% levofloxacin degradation efficiency in
70 min
[34]
13Dye sensitized Fe-MOF nanosheetsNot calculatedPhotocatalytic reduction of CO2300 W Xe lamp with a 420 nm
cut-off filter
The synthesized material exhibited a significant photocatalytic CO production rate of 1120 μmol g−1 h−1[40]
14Lignin-based bimetallic MOFs nanofibers composite membranes with
peroxymonosulfate
2.66 eVPhotocatalytic degradation of perfluorooctanoic acid9 W UV lamp with the wavelength of photoexcitation is 185 nm and under solar lightLignin/PVA/bimetallic-MOFs showed outstanding performance up to 89.6% degradation of perfluorooctanoic acid
within 3 h
[108]
15Pt photo
deposited on MIL-53(Fe)
Not calculatedPhotocatalytic evolution of H2300 W Xe lampDeposition of Pt nanoparticles on Fe-MOFs can lower the overpotential for H2 evolution toward further enhanced photocatalytic activity.[109]
16BiOIO3/
MIL-88B
BiOIO3: 2.88 eV
MIL-88B: 2.31 eV
BMIL-5: 2.48 eV
Photocatalytic degradation of Reactive Blue 19 and tetracycline hydrochloride300 W Xe lampThe BiOIO3/MIL-88B composites exhibited an excellent removal rate for Reactive Blue 19 and tetracycline hydrochloride under visible light irradiation, which was approximately 3.28 and 4.22 times higher than the pristine BiOIO3, respectively.[110]
17Porphyrin-
Zr MOF
Not calculatedPhotocatalytic degradation of bisphenol F500 W Xe lamp with a 420 nm cut off filterThe material could achieve over 78% BPF removal (with/without salt) after
8 cycles.
[83]
18Graphitic carbon nitride/NH2-MIL-101(Fe)Fe-MOF: 1.64 eV
g-C3N4: 2.77 eV
Composite: 1.90 eV
Photocatalytic degradation of
acetaminophen and reduction of Cr(VI)
Solar light (60,000 lux)The composite showed the highest degradation of acetaminophen 94% at pH 7 and Cr(VI) reduction efficiency of 91% at pH 2[111]
19MIL-100 (Fe) MOF/MOX homojunctionNot calculatedPhotocatalytic oxidation of gaseous benzene, toluene, xylenes and styrene250 W Xe
lamp
MIL-100(Fe) MOF/MOX homojunction showed up to 23%, 41%, 82%, 79% and 83% photocatalytic oxidation of benzene, toluene, p-xylene, m-xylene and styrene, respectively[112]
20CuO-ZnO/
ZiF-8 MOF
CuO-ZnO: 2.89 eV
CuO-ZnO/ZiF-8(20): 1.96 eV
ZiF-8: 5.34 eV
Photocatalytic degradation of acid orange 7400 W halogen lampCuO-ZnO/ZiF-8(20) showed the highest photocatalytic degradation of 98.1% acid orange 7 in 100 min [113]
21Co-TCPP MOF@B-
TiO2−X
pure B-TiO2−x:
2.71 eV
Co-TCPP MOF:
2.55 eV
BTC-10%:
2.24 eV
Photocatalytic degradation of bisphenol AA 300W Xe lamp with a 420nm cut-
off filter
Co-TCPP MOF@B-TiO2−X exhibited remarkable photocatalytic degradation of
bisphenol A up to 97% within 120 min irradiation
[114]
22Ag@Tetra
thiafulvalene-based
Zr-MOF
Not calculatedPhotocatalytic
degradation of sulfamethoxazole
500 W Xe lampAg NPs@Zr-TTFTB were found to efficiently remove sulfamethoxazole
(k = 0.009 min−1)
[115]
23CdS/Ni-MOF heterostructureCdS: 2.0 eV
Ni-MOF: 3.3 eV
Photocatalytic reduction of CO2300 W Xe lamp20%-CdS/Ni-MOF showed the best photocatalytic reduction
performance, the yield of CO reached 7.47 μmol/g in the 4th hour, which was nearly 16 times and 7 times that of Ni-MOF and CdS
[116]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, M.M.; Rahman, A.; Matussin, S.N. Recent Progress of Metal-Organic Frameworks and Metal-Organic Frameworks-Based Heterostructures as Photocatalysts. Nanomaterials 2022, 12, 2820. https://doi.org/10.3390/nano12162820

AMA Style

Khan MM, Rahman A, Matussin SN. Recent Progress of Metal-Organic Frameworks and Metal-Organic Frameworks-Based Heterostructures as Photocatalysts. Nanomaterials. 2022; 12(16):2820. https://doi.org/10.3390/nano12162820

Chicago/Turabian Style

Khan, Mohammad Mansoob, Ashmalina Rahman, and Shaidatul Najihah Matussin. 2022. "Recent Progress of Metal-Organic Frameworks and Metal-Organic Frameworks-Based Heterostructures as Photocatalysts" Nanomaterials 12, no. 16: 2820. https://doi.org/10.3390/nano12162820

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

Article Metrics

Back to TopTop