Next Article in Journal
Synthesis of High Near-Infrared Reflective Black Pigment Based on YMn2O5
Previous Article in Journal
Squaric Acid Core Substituted Unsymmetrical Squaraine Dyes for Dye-Sensitized Solar Cells: Effect of Electron Acceptors on Their Photovoltaic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Revolutionizing Wastewater Treatment: Harnessing Metal–Organic Frameworks for Exceptional Photocatalytic Degradation of Azo-Type Dyes

by
Jose Manuel Barrera-Andrade
1,
Natali de la Fuente-Maldonado
1,
Ricardo Lopez-Medina
2,
Ana Marisela Maubert-Franco
3 and
Elizabeth Rojas-Garcia
1,4,*
1
Lab. Catálisis y Materiales, ESIQIE-Instituto Politécnico Nacional Zacatenco, Mexico City 07738, Mexico
2
Área de Procesos de la Industria Química, Departamento de Energía, Universidad Autónoma Metropolitana-Unidad Azcapotzalco, Mexico City 02200, Mexico
3
Área de Química de Materiales, Departamento de Ciencias Básicas e Ingeniería, Universidad Autónoma Metropolitana-Azcapotzalco, Mexico City 02200, Mexico
4
Área de Ingeniería Química, Departamento de Ciencias Básicas e Ingeniería, Universidad Autónoma Metropolitana-Iztapalapa, Mexico City 09340, Mexico
*
Author to whom correspondence should be addressed.
Colorants 2023, 2(4), 674-704; https://doi.org/10.3390/colorants2040035
Submission received: 9 September 2023 / Revised: 5 November 2023 / Accepted: 8 November 2023 / Published: 13 November 2023

Abstract

:
Due to the high stability of azo-type dyes, conventional treatment processes such as adsorption, flocculation, and activated sludge are not efficient for decolorizing wastewater effluents. An alternative to traditional wastewater treatment is photocatalysis, which has gained significant interest because research has shown it to be a viable and cost-effective process that uses sunlight as an inexhaustible energy source. In heterogeneous photocatalysis, a photocatalyst is required, such as TiO2, ZnO, composite materials, and, more recently, metal–organic frameworks (MOFs). MOFs, also known as “coordination polymers”, exhibit photocatalytic properties and have been proven to be promising materials in the photocatalytic degradation of dyes. This study presents recent advances in using MOFs as photocatalysts to degrade recalcitrant contaminants like azo-type dyes. Recent advancements in developing photocatalysts based on MOFs are focused on two strategies. Firstly, the development of new MOFs composed of complex ligands or a mixed ligand system, and secondly, the synthesis of composite materials based on MOFs and metal oxides, metals, sulfides, nitrides, etc. Both strategies have significantly contributed to the search for new semiconductors to degrade some recalcitrate contaminants in wastewater.

1. Introduction

1.1. Colorants

Color has a long history with humans and is valuable and essential in our daily lives. Colorants are classified into dyes and pigments and are crucial to adding color to various materials (Figure 1) [1]. The key distinction between both is that dyes are soluble in the substrate, dispersing at a molecular level, while pigments are insoluble and are applied as particles [2]. There are two main types of dyes: synthetic and natural [3]. Synthetic dyes are derived from petroleum compounds, while natural dyes are obtained from plants, animals, and minerals [4]. Synthetic dyes consist of organic molecules with three essential components: the chromophore, the auxochrome, and the matrix [2,3]. The chromophore is the part of the molecule responsible for absorbing specific wavelengths of light and giving the colorant its color. The common chromophores include the azo (-N=N-) and non-azo (nitro (-NO2), nitroso (-N=O), carbonyl (-C=O), and thiocarbonyl (-C=S)) group. On the other hand, an auxochrome is a group of atoms or ions that can alter or enhance the color of compounds when attached to a chromophore group. They are often responsible for the different colors we observe in organic compounds, such as dyes. Common auxochromes include the hydroxyl (OH) group, amino (NH2), aldehyde (CHO), and methyl mercaptan (SCH3). Finally, the remaining atoms in the molecule make up the matrix, the dye’s third component. The matrix includes the solubilizing groups, stabilizing groups, and counterions. Azo dyes are mainly used by textile, paper, plastics, and food industries and are thus estimated to account for approximately 70% of the world’s dye production [5,6].
Azo-type dyes find various industrial applications and can be categorized into several classes, including acid, basic, neutral, reactive, vat, mordant, and disperse, as shown in Figure 1 [4]. Table 1 shows the chemical structure of any azo-type organic dyes according to their classification. Acid dyes are anionic compounds that are generally water-soluble. They are used for dyeing protein fibers under acidic conditions [7]. A distinctive characteristic of acid dyes for protein and polyamide fibers is the presence of one or more sulfonate (-SO3) groups, usually in sodium salt (-Na+). These groups serve a dual purpose: ensuring solubility in water and imparting a negative charge to the dyes.
On the other hand, basic dyes are cationic compounds carrying a positive charge, typically originating from the ammonium cation. They adhere well to negatively charged fibers and are employed to dye materials like acrylic, nylon, and other synthetic fibers. Basic dyes also have applications in paper and ink dyeing [8]. Direct or substantive dyes constitute a significant proportion of anionic azo dyes and are used to dye natural or regenerated cellulose [9]. The essential requirement for the classification of a dye in this group is its substantivity, i.e., its absorption from an aqueous salt-containing solution onto cellulosic materials.

1.2. Main Degradation Methods of Dyes in Wastewater

The advent of the Industrial Revolution marked a profound transformation in human life as machines were harnessed for mass production, creating a substantial demand for various dyes to color large quantities of fabrics. During this era, the available dyes were primarily of plant and animal origin, which proved impractical in meeting the growing demand. Consequently, humans began synthesizing various dyes without considering these compounds’ potential degradation and toxicity. As time passed, these compounds were recklessly disposed of into water bodies and sewers, leading to severe environmental issues [10].
Since most azo-type dyes are of anthropogenic origin, their natural degradation is minimal, if not virtually non-existent. As a result, they persist in the environment for extended periods [11]. For this reason, in recent years, significant efforts have been made to improve conventional treatments for the degradation of azo-type dyes, such as adsorption, oxidation methods, biological methods, and photocatalysis [12]. All these dye treatment methods are depicted in Figure 2.
Table 2 shows the advantages and disadvantages of any of the convectional and advanced treatments used for the degradation of azo-type dyes in wastewater.
One initial method for treating azo dyes was adsorption onto a substrate [19,20,21,22,23]. This approach was also necessitated by early water treatment stages involving sand, gravel, or clay filters that partially adsorbed these dyes [24,25]. However, this initial option was specialized, prompting the search for more efficient treatment alternatives since these filters needed frequent replacement, incurring higher costs (Table 2). Thus, pre-treatment stages were developed where adsorbents with high adsorption capacities, such as activated carbon, were utilized [26]. Activated carbon cannot only be used as an adsorbent in the removal of dyes [8,27,28]; it has other applications such as air and gas purification, gold recovery, chemical catalyst support, solvent recovery, gas masks and protective gear, cosmetics and beauty products, and other [29,30,31,32]. However, it is important to mention that the adsorption method often involves transferring dyes from one medium to another, i.e., from water to a solid substrate (Table 2). Hence, the appropriate treatment method must be sought to prevent future contamination [33]. For example, chemical oxidation treatments aim to oxidize dyes using reactive agents, but in some cases, these dyes contain halogens that produce even more toxic byproducts, exacerbating environmental issues (Table 2) [34].
Other methods for the degradation of azo dyes have been explored via electrochemistry, which can effectively eliminate dyes but often generate sludge requiring further treatment [35]. The biological treatment has gained traction as bacteria have evolved to tolerate more toxic environments and break down complex dye molecules as a food source, facilitating partial or complete azo dye degradation [36,37]. However, this approach has not been effective for all dyes. Another option involves using UV-C light for dye oxidation; its high energy can break chemical bonds, resulting in partial dye degradation [38]. However, a high consumption of electricity is necessary. Combining multiple degradation techniques has been essential in many instances to break down the dyes fully.
Recent research has delved into a novel treatment method using a semiconductor, a light source, and a dye, where these components interact to achieve complete dye mineralization; this process is called photocatalysis [18]. The concept of using semiconductor materials such as TiO2 in photocatalytic treatments emerged in the 1990s. Researchers discovered that titanium dioxide could generate electrons (e−) and holes (h+) when exposed to UV light. These two photo-species enter the surrounding medium, where the electrons can interact with dissolved oxygen in water, forming superoxide (·O2). Simultaneously, the photogenerated holes (h+) interact with water molecules, producing hydroxyl radicals (·OH) capable of oxidizing organic matter [39].
Over the years, this application has evolved by using solar radiation as an environmentally friendly light source. Additionally, semiconductor materials have improved via enhanced synthesis methods, modifications, or doping [40,41,42,43]. In some cases, entirely new semiconductors have been developed, expanding the possibilities for treating dye-contaminated effluents.

1.3. Fundamentals of Metal–Organic Frameworks (MOFs) as Semiconductors

Metal–organic frameworks (MOFs) are three-dimensional crystalline materials characterized by metallic clusters connected by organic ligands [44]. Beyond separation and gas storage applications, MOFs have found use in various fields, including catalysis (e.g., photo-CO2 reduction, and photocatalytic dye degradation), drug delivery and transport (enabled by mesopores for substance storage), optical sensors due to their semiconductor properties or as photocatalyst, and more. Currently, MOFs are of great interest because some have photocatalytic properties such as NH2-MIL-125 (Ti) [45,46,47,48,49], UIO-66 (Zr) [50], NH2-UIO-66 (Zr) [51,52,53], MIL-53 (Al) [54,55], NH2-MIL-101 (Fe) [47,56], and new MOFs with a complex ligand or mixed ligand [57,58,59], among others. Figure 3 shows how research into MOFs as photocatalysts has been growing in recent years, as well as their use as photocatalysts for the degradation of dyes.
When a beam of light interacts with the MOF, it allows the formation of charge carriers (holes (h+) and electrons (e−)) that are produced by the absorption of energy equal to or greater than the bandgap (Eg) [44]. However, these materials, due to their particular structure and organic nature, must take into account some additional and complementary considerations to common metallic semiconductors, such as (a) light absorption and photoexcitation: MOFs with photosensitive properties can play an active role since organic ligands and metal centers can absorb photons and convert a photogenerated excited state and transfer electrons from the ligand to the metal center [60,61,62]. This is possible due to the multiple possibilities that MOFs must be modified by replacing their organic linkers or metal sites, introducing molecules and nanoparticles within their structure, creating defects, or forming composites with other materials. In MOFs, charge transfer can occur via various mechanisms, leading to distinct charge transfer complexes (CTCs) with unique characteristics. Table 3 shows some common types of charge transfer in MOFs and their characteristics.
The characteristics and applications of these charge transfer types can vary depending on the specific MOF structure, the metal ions, the ligands involved, and the environmental conditions. The ability of MOFs to facilitate these types of charge transfer processes upon exposure to light makes them promising photocatalysts for a wide range of applications, including solar energy conversion, environmental remediation, and sustainable chemistry, and b) about the efficiency of these processes: these depend mainly on the type of material, its crystallinity and particle size. For example, high charge mobility is commonly associated with highly crystalline semiconductor materials in which oriented charge mobility is observed. Semiconducting nanocrystals with small particle sizes generally possess satisfactory charge transfer performance; However, if the particles are too small, recombination may occur due to enhanced surface defects [76]. Charge carrier recombination is a process in which photogenerated electrons and holes will recombine and, as a result, release energy obtained during excitation in the form of fluorescence or heat. In general, this process can be avoided mainly by decreasing the particle size of the photocatalyst.
In addition, this charge carrier recombination process can also reduce doping/impregnation/support in MOFs of metals (Au, Ag, Pt, Pd, etc.) [47,50,77], oxides, or mixtures of metal oxides. (TiO2, CoFe2O4, ZnFe2O4, Ag3VO4, etc. [45,47,55], graphene oxide [78,79], quantum dots [80,81], and g-C3N4 [82,83], among others. The synthesis of these composites makes it possible to reduce the recombination of charge carriers by separating the photogenerated electrons and sending them to a metal, metal oxide, graphene oxide, or another MOF with a reduction potential less than the activated one. Figure 4 provides a scheme of the charge transfer pathway in the photocatalytic degradation of dyes employing a MOF and a semiconductor; a MOF (support), a semiconductor and a metal; and a MOF and a semiconductor with Z-scheme.
Figure 4a shows charge transfer mechanism proposed when a MOF and a semiconductor (SC) interact, where the MOF is activated. After, the photogenerated electrons migrate to the conduction band of the semiconductor, generating reactive species that oxidize of dye molecules and the holes in the valence band, also they generate (·OH) hydroxyl radicals that oxidize the dye molecules. While Figure 4b exemplifies a composite based on a MOF, a semiconductor, and a metal. In this case, the semiconductor is activated, generating the charge carriers, where the electrons migrate to the metal, and in this, the active species responsible for the oxidation of the dye are formed. In addition, water oxidation reactions take place in the valence band, creating hydroxyl radicals responsible for the oxidation/degradation of the dye molecule. Finally, Figure 4c shows a composite with a Z-scheme, where the MOF is activated, generating charge carriers, the electrons from the conduction band migrate to the valence band of the semiconductor, these electrons migrate again to the conduction band of the semiconductor, generating in this site the reactive species responsible for the oxidation of the dye.

1.4. Paragraph in the Degradation of Dyes Using MOFs

In recent years, there has been a growing interest in improving MOFs for use as semiconductors, particularly in photocatalytic applications [84,85,86,87]. These improvement strategies can be categorized into two fundamental axes, as depicted in Figure 5: (i) the synthesis of new MOFs with simple or complex ligands or mixtures of complex ligands and (ii) the synthesis of composite materials incorporating metals, metal oxides, carbon materials, or other MOFs, often combined with quantum dots (QDs) of carbon and metal oxides. Almost in all cases, this last axis allows us to improve the physicochemical, structural, optical, and textural properties of conventional MOFs, their thermal stability, and in aqueous environments, a key challenge when water is involved in MOF applications.
In the case of thermal stability, Healy et al. [88] identified several factors that affect the thermal stability of MOFs, such as the nature and position of functional groups, metal hardness, and the presence of coordinated solvent molecules. Also, they mentioned that thermal stability was found to have a relatively modest impact on the synthesis of MOFs by isoreticular expansion and interpenetration, where mixed ligands are generally present. They also proposed a possible thermal decomposition mechanism of MOFs, which they divided into two categories depending on whether the decomposition is ligand-centered, inorganic node-centered, or SBUs. However, ligand-centered decomposition is the dominant factor in determining the decomposition temperature of MOFs.
This section will review recent advancements along these two axes, analyzing strategies to enhance semiconductor efficiency in the photocatalytic degradation of dyes found in wastewater via heterogeneous photocatalysis.

1.5. Possible Degradation Mechanism of Dyes

Several studies have shown that the main reactive species in the photocatalytic degradation of dyes can be superoxide radicals (·O2−), hydroxyl radicals (·OH), and holes (h+) [89,90]. These species are responsible for oxidizing the dye molecule until their mineralization (CO2 + H2O) or degradation of these into molecules that are sometimes more toxic than the dye itself. To achieve the above, the dye molecule, in interaction with the reactive species form reaction intermediates, which are finally mineralized into inorganic products. For example, Shi et al. [89] showed a possible reaction mechanism for the degradation of methyl blue (MB) dye (Figure 6). They mention that during the degradation, both S and N in the molecular structure of MB underwent redox reactions to form amines and sulfonic acid groups as intermediates, which are finally further mineralized into inorganic products.
Other authors also proposed a possible degradation mechanism of Rhodamine B (RhB) [51,90], methyl violet (MV) [91], and azure I [92].
Li et al. [90] determined using liquid chromatography–mass spectrometry (LC-MS) the possible intermediates of the reaction and therefore a possible degradation mechanism of Rhodamine B (RhB) (Figure 7). As shown in Figure 7, the proposed reaction mechanism for the degradation of Rhodamine B could begin with the formation of B1 (m/z = 443.19), which is characteristic of the protonated form (C28H31N2O+3) of RhB (mechanism 1), or by the methyl oxidation of the latter to B2 (m/z = 458.69) (mechanism 2). Via mechanism 1, the protonated form of RhB could undergo a series of N-de-ethylation reactions to form macromolecular intermediates B3–9 (m/z = 415.16, 387.15, 362.30, 330.90), which subsequently, via de-amination, de-alkylation, de-carboxylation, and cleavage of the chromophore reactions allow the formation of smaller molecular intermediate B10–22 (m/z = 318.27, 310.07, 279.11, 274.27, 258.19, 254.17, 244.03, 230.24, 213.07, 212.1 5, 167.03, and 122.00), which are finally further mineralized into inorganic products. Meanwhile, via mechanism 2, the direct formation of macromolecular intermediates B3–9 is possible, which also forms small molecular intermediates B10–22 via de-amination, de-alkylation, de-carboxylation, and cleavage of the chromophore reactions, where finally the latter are mineralized to inorganic products.
However, despite the efforts made in the search for a reaction mechanism for the photocatalytic degradation of dyes, it is still necessary to use more sophisticated techniques, such as in situ and even theoretical studies, which allow for a more accurate determine the possible reaction intermediates, and thus be able to establish a more realistic reaction mechanism.

2. Recent Advances in the Photocatalytic Degradation of Dyes According to Classification

2.1. Acid Dyes

Table 4 presents recent advances in using MOFs as photocatalysts for degrading various dyes, including acid, reactive, direct, and mordant. Methyl Orange (MO) (I.C. Acid Orange 52), an azo-type dye (Figure 8), has been a critical acid dye used to evaluate the performance of MOFs in photocatalysis. Dyes are commonly employed in characterizing semiconductors due to their convenient monitoring via spectrophotometry. The degradation process is also visually observable, as the reduction in coloration can be detected with the naked eye. Recent research on MOFs in photocatalysis for dye degradation has explored several avenues, including synthesizing new MOFs, using multiple ligands, and doping MOFs with other semiconductors.
In recent years, the utilization of pristine MOFs has declined due to their inherent limitations, primarily related to their stability both thermally and in aqueous media. These MOFs rapidly decompose upon contact with water molecules. In the context of acid dyes, novel MOFs have been introduced, such as [Cu3(L)2(4′-bipy)] [89], designed to enhance stability in aqueous media.
Another noteworthy example is the MOF La-PTC, composed of perylene-3,4,9,10-tetracarboxylate ligand and lanthanum metal cluster. This MOF achieved a remarkable 99.6% degradation of MO at 240 min, with the addition of hydrogen peroxide facilitating the formation of hydroxyl radicals, leading to the complete degradation of the dye. MOF La-PTC was activated using visible light [93], wherein the dye facilitated light absorption; this is activated, generating electrons that transitioned to the valence band of the MOF, ultimately yielding superoxide radicals (Figure 9).
The combination of ligands in the synthesis of MOFs has been applied in the last four years, providing better characteristics to the MOF and increasing its photocatalytic activity. For instance, MIL-88A catalysts with mixed C4H6N2–fumaric acid ligands were introduced, featuring small, highly crystalline particles. This MOF efficiently degraded to MO at a concentration of 100 ppm for 45 min [40]. In another case, a new 8-fold interpenetrated MOF was synthesized via a mixed-ligand system of benzene-p-dicarboxylic acid and tris(4-(4H-1,2,4-triazol-4-yl) phenyl)amine under hydrothermal conditions, demonstrating excellent water stability. Figure 10 illustrates the structure of the MOF with two ligands, highlighting the band positions. This MOF eliminated nearly 91% of MO due to the mixed ligands, which reduced the band gap of the MOF (Figure 10d). Remarkably, it could be reused five times under the same conditions, achieving over 90% elimination of MO [94].
The mixed ligands also contributed to more excellent thermal stability, allowing these modified D-Fe-MOFs to eliminate 97% of MO in 48 min [95]. In a related study, identical ligands were applied to obtain Fe-MOFs, which exhibited a larger specific surface area, smaller particle size, a band gap of 2.55 eV, and a degradation rate of 97% within 48 min for MO (20 mg/L) under visible light [96].
The third configuration involves modifying MOFs by doping them with another semiconductor, forming a heterojunction between the two semiconductors. This configuration enhances the separation of generated photo species, increases the surface area, and improves light absorption in the visible region. Various semiconductors, such as MoS2 [49], GO [41], and TiO2 [42,43,97], have been employed, consistently achieving excellent results with over 90% degradation of MO. Moreover, all doped MOFs were activated using visible light. For instance, when MOFs were modified or doped with Ag2CrO4 [98,99], AO7 dye was effectively degraded in both cases, surpassing 90% degradation.

2.2. Basic-Dyes

There are several types of basic dyes, but in the literature, basic organic dyes like RhB, MV, and MB have been extensively investigated due to their common occurrence in wastewater, as evident in Table 4. Moreover, these dyes serve as target molecules in photocatalytic degradation reactions to assess the photocatalytic properties of semiconductors, including TiO2, ZnO, g-C3N4, Fe2O3, composite materials, and more recently, metal–organic frameworks (MOFs) and their composites. Simultaneously, another objective has been focused on the search for highly active photocatalysts for complete dye mineralization.
Several semiconductors have been studied in the photocatalytic degradation of dyes, such as ZnO, g-C3N4, CeO2, TiO2, etc. However, TiO2 is one of the most used semiconductors in the photocatalytic process. Nevertheless, it presents several disadvantages, such as a wide gap band, high recombination of the photogenerated charge carriers (electron-hole), low mechanical stability, etc. Hence, exploring alternative semiconductors for dye degradation reactions, like MOFs, is imperative. MOFs find extensive use as semiconductors in the photocatalytic degradation of dyes due to their exceptional properties, including a high specific surface area, large pore volumes, and exposed metal centers that facilitate catalytic and photocatalytic reactions. Additionally, MOFs can adsorb substantial quantities of dyes due to their abundant active sites [20].
In the degradation of basic dyes, it is noteworthy that the most researched MOFs include UiO-66, NH2-UiO-66, MIL-53 (Al), MIL-53 (Fe), NH2-MIL-53 (Fe), NH2-MIL-125 (Ti), MIL-88 (Fe), NH2-MIL-101 (Fe), MIL-100 (Fe), and Cu-BDC. These MOFs have gained attention due to their excellent properties, such as high specific surface area, large pore volumes, and good stability in aqueous environments. Various research groups have concentrated on enhancing pristine MOFs by doping them with metals, metal oxides, mixed oxides, and even quantum dots. These materials modify pristine MOFs’ physical, chemical, optical, and textural properties. The primary focus has been changing their optical properties to reduce the bandgap and minimize the recombination of charge carriers, thereby enhancing their photocatalytic activity.
Recent advancements in the photocatalytic degradation of Rhodamine B (also known as “Basic Violet 10” according to color index) have been centered on identifying solid materials for its degradation and mineralization (Table 4). For example, Gayathri et al. investigated the band structure of Zr-based metal–organic frameworks (Zr-MOFs) with different ligands, such as terephthalic acid (UiO-66), 2-amino terephthalic acid (NH2-UiO-66), and 2-methylimidazole (UiO-66-2M), synthesized via the hydrothermal method. They applied them in the photocatalytic degradation of Rhodamine B [53] (Figure 11). They demonstrated that the optical properties of MOFs are influenced by the type of ligand, which can act as an antenna to capture photons and direct them to metal sites for oxidation and reduction reactions. NH2-UiO-66 exhibited the highest efficiency, achieving 95% degradation at 120 min (under reaction conditions: pH = 7, 10 ppm RhB, and catalyst concentration = 0.4 mg/L). The presence of amino groups in the organic ligand (2-amino terephthalic) used during MOF synthesis enables photon capture electron generation to fill the LUMO energy level and, subsequently, reduction in the bandgap, thereby improving photocatalytic activity in Rhodamine B degradation.
The photocatalytic degradation of Methyl Violet (also known as “Basic Violet 3” according to the Color Index) has primarily focused on synthesizing new MOFs using novel or mixed ligands. The development of MOF-composite materials has been a primary strategy to enhance the optical properties of pristine MOFs and their photocatalytic performance. Sofi et al. [50] demonstrated that incorporating Ag2WO4 and silver nanoparticles into UiO-66 improves the efficiency of Rhodamine B degradation (Figure 12). They observed that this enhancement was attributed to the synergistic effect between Ag2WO4 and UiO-66, along with the surface plasmonic impact of silver nanoparticles. Furthermore, this composite material exhibits excellent stability even after five cycles.
Another interesting study was conducted by Liu et al. [54] via a hydrothermal method and employed them in the solar light-driven photocatalytic degradation of Rhodamine B. They degraded 30 mg/L of RhB within 10 min using a catalyst of 0.3 g/L under solar light. While using sacrificial agents in contaminant degradation is not typically recommended, their presence enhances photocatalytic activity, as demonstrated in a study by Mahmoodi et al. [47]. They showed that adding H2O2 (20 mL) as a sacrificial agent considerably enhances the photocatalytic activity of Ag/CoFe2O4/NH2-MIL-125 (Ti) composite materials. Their results revealed that nearly 98.8% of 20 mg/L RhB could be removed within 10 min under visible light irradiation. Additionally, they established the stability of this composite material.
Figure 13 shows the semiconductor’s activation mechanism without and with a sacrificial agent such as hydrogen peroxide. Figure 13a illustrates the formation of the active species when they do not have a sacrificial agent. Moreover, Figure 13b shows the interaction of a sacrificial agent (H2O2) with electrons photogenerated in the conduction band of the semiconductor, allowing the formation of hydroxyl radicals, which are generally the active species that promote the degradation process of dyes [100,101]. In both mechanisms, the formation of the free radical is possible; however, when the sacrificial agent is present, the formation is direct.
In recent years, the search for new MOF-based semiconductors for the photocatalytic degradation of dyes has focused on using complex ligands [94,102,103]. For instance, the synthesis of 3D supramolecular structures via L (H4L = 1, 1′:4′, 1″:4″, 10″-quterphenyl-2, 5, 20″, 50″-tetracarboxylic acid) ligand linkage has been explored. Wu et al. [17] synthesized two isomorphic MOFs of manganese and cadmium with H4L ligand. These materials were used in the photocatalytic reaction of Rhodamine B under UV irradiation (Table 2). However, MOF-based manganese shows a higher catalytic performance compared with MOF-based cadmium. Also, the authors proposed a possible reaction mechanism for the photocatalytic degradation of dye using LC-MS. They found that the chains/rings in the main structure of RhB happened redox reactions to generate amine groups, which are oxidized by strong oxidizing groups, mainly OH radicals.
Other authors have studied MOFs using a mixed-ligand system [89,94,104,105], where these materials have shown good photocatalytic activity in the degradation of basic dyes such as MB, RhB, and MV. Chen et al. [94] show that MOF synthesis based on zinc clusters, benzene-p-dicarboxylic acid (H2BDC), and tris(4-(4H-1,2,4-triazol-4-yl) phenyl)amine (L) as organic ligands (Figure 14) offers excellent water stability. They also found that both ligands enable inter-ligand-based charge transfer and reduce the energy gap compared to MOF with the BDC ligand. This new material is promising in the degradation of basic dyes. Also, this can be activated with visible irradiation (Table 4).
For the photocatalytic degradation of cation dye azure I, Ren et al. [92] synthesized Mg-MOF using MgCl2, H4tptc, and CH3CN via solvothermal method at 150 °C. The degradation rate of Az I reached 100% within 50 min. After five cycles, the catalyst showed that its original structural framework remained intact during the photocatalytic degradation.

2.3. Reactive Dyes

In the photocatalytic degradation of the synthetic textile dye Reactive Yellow 145 in an aqueous solution, different MOF photocatalysts were used under UV and visible irradiation. Metallic (Co-BTC/CN and Fe- BTC/CN) and bimetallic (Co-Fe- BTC/CN) nanocomposites were examined in the degradation of Reactive Yellow 145 by Nguyen et al. [106] They introduce nanocomposites based on metallic and bimetallic MOF and CN sheet using the microwave-assisted hydrothermal method and investigated the effect of Co: Fe molar ratio on the decomposition of Reactive Yellow 145 in water. The Co-Fe-BTC/CN sample exhibited a higher efficiency than CN, Fe-BTC, and Fe-BTC/CN, with almost 99.38% removal of RY-145 after 40 min. The stability of the Co2-Fe8-BTC/CN catalyst was determined, and no change in RY-145 dye removal efficiency was observed after six reaction cycles, indicating that the catalyst is stable and reusable. On the other hand, α -Fe2O3@C@SiO2/TiO2 magnetic nanocomposites were successfully synthesized from MOFs and then coated with silica and titanium particles via sol-gel and solvothermal methods by Mousavi et al. [107]. The removal efficiency of Reactive Yellow 145 was about 97.5% at 150 mgL−1 of dye concentration, pH = 2, 15 mg100 mL−1 of photocatalyst, 25 °C reaction temperature, and 90 min under UV irradiation. The particles exhibited excellent stability and reusability because they can maintain a relatively high removal efficiency of about 97% of their initial performance after five reuse cycles. It can be concluded that TiO2 coating on magnetic particles increases their photocatalytic degradation efficiency and reusability.
For Reactive Red 198 degradation, AC, MIL-88B (Fe), and AC/MIL-88B (Fe) photocatalysts were synthesized by Mahmoodi et al. [108]. This study used a newly prepared activated carbon from Kiwi peel to synthesize AC/MOF nanocomposite to improve the MOF catalytic activity. The photocatalytic degradation of Reactive Red 198 (RR198) by the AC/MIL-88B (Fe) composite achieved 99% removal at 0.08 g/L of catalyst dosage, pH of 3, and a dye concentration of 50 mgL−1.

2.4. Other Dyes

To study the degradation of Direct Red 28 under sunlight at room temperature, Z. Saedi et al. [109] used TMU-16. The experimental results showed a dye degradation of 96% after 1 h with a 20 mgL−1 dye concentration. After three cycles, the degradation of the dye remained at about 95%. When TMU-16 is exposed to light photons, the Zn(O)4 metal–oxygen clusters act as semiconductors, forming electron–hole pairs. So, in this work, the ability of TMU-16 for significant photocatalytic degradation of Congo red (CR). Furthermore, researchers [110,111,112] reported strong adsorption of CR, which influences the degradation process and results in a negligible or prolonged degradation rate. The amino groups in CR achieved the adsorption of CR onto the catalyst as both donors and acceptors of electrons, enabling H-bonds to form between the amino organic dye and the uncoordinated O atoms of carboxyl groups in MOFs [112]. In the same research, Z. Saedi et al. [109] used TMU-16 to study the degradation of Eriochrome Black T under sunlight at room temperature. Eriochrome Black T showed 96% degradation after 180 min. Alternatively, a new metal–organic framework {[Co3 (BTC)2(Bimb)2.5]·2H2O}n (1) (Bimb = 1,4-bis[(1H-imidazol-1-yl)methyl]benzene and H3BTC = 1,3,5-Benzenetricarboxylic acid) was synthesized by Somnath et al. [113]. The mixed ligand 3D Cobalt was examined in the degradation of Eriochrome Black T in an aqueous solution with almost 77.5% removal of EBT after 135 min under UV light. The complex displayed high reusability, efficiency, and low cost of adsorbency with great potential in handling dye-contaminated wastewater.
Table 4. Review the most recent studies using MOFs for the photocatalytic degradation of dye.
Table 4. Review the most recent studies using MOFs for the photocatalytic degradation of dye.
Dye Common Name/Color IndexSampleBand Gap
(eV)
Time
(min)
Degradation
Efficiency (%)
Reaction ConditionsStabilityRef.
Acid Red 1Cl-M88A (MIL-88A (Fe)) (1)
NIT-M88A (MIL-88A (Fe)) (2)
2.17
2.19
120
90
100
100
20 mg/L dye: pH = 7; conc. catalyst = 1 g/L; 0.03 mL of H2O2; Xe lamp stable[40]
PBS/Acid Red 66
RhB/Basic Violet 10
Ag/Ag2WO4/UiO-662.0625 (PBS)
60 (RhB)
95 (PBS)
90 (RhB)
60 mg/L dye, pH = 7; conc. catalyst = 0.2 g/L; Halogen Lampstable[50]
Acid Black 1MIL-53 (Cr)/polymer (1)
HKUST-1 (Cu) (2)
2.1–2.43096
90
UV lamp stable[114]
MB/Basic Blue 9
MV/Basic Violet 3
RhB/Basic Violet 10
[Co4(bbibp)5(HCOO)8(H2O)2]n (1)
[Co(bimmb)(NO3)2]n (2)
2.50
3.07
75(1): 100 (MB); 50.8 (MV); 61.0 (RhB)
(2) 94 (MB); 68.7 (MV); 64.4 (RhB)
6 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; H2O2 = 0.5 mL; UV high pressure lamp -[57]
RhB/Basic Violet 10
CB/Crome blue
MB/Basic Blue 9
[Mn4(NDC)2(bb)4] 2.5310094.33 (RhB)
34.91 (CB)
29.35 (MB)
10 mg/L dye; pH = 7; conc. catalyst = 0.8 g/L; Hg Lampstable[102]
AO7/AO7
MB/Basic Blue 9
CF/MoS2/NH2-MIL-125(Ti)2.512067.9 (AO7)
94.3 (MB)
10 mg/L dye; pH = 7; conc. catalyst = 1.5 g/L; visible lightstable[49]
MB/Basic Blue 9SrAl2O4:Eu2+,Dy3+/g-C3N4@NH2-UiO-66 2.77309520 mg/L dye; pH = 7; conc. catalyst = 0.4 g/L; Xe lampstable[51]
MV/Basic Violet 3[Co23–OH) (L)(H2O)3·2•75H2O]n (1)
[Co23 –OH)(L)(bib)(H2O)2]n (2)
3.02
1.91
4055.3 (1)
61.3 (2)
10 mg/L dye; pH = 7; conc. catalyst = 0.1 g/L; Hg lampstable[115]
MV/Basic Violet 3[Zn2(1,4- BDC)2(bmp)•DMF] (1)
[Cd3(H2O)2(bmp)2(DMF)(BTC)2•3•5H2O•1•5DMF](2)
-
-
4070.1 (1)
37.83 (2)
- stable[100]
MV/Methyl Violet
CV/Basic Red 5
NR/Basic Red 5
CR/Direct Red 28
{[Co3 (BTC)2 (Bimb)2.5]·2H2O}n3.9413568.7 (MV)
77.5 (CV)
19.5 (NR)
15.8 (CR)
10 mg/L dye; pH = 7; 5 mL of methanolic solution (0.01 mmol powder in 50 mL methanol); UV lamp-[113]
MB/Basic Blue 9
MV/Basic Violet 3
[CoL (bimb)0.5]n (1)
{[CoL (bimmb)] DMF}n (2)
[CoL (bimmb)0.5]n (3)
[CoL (bbibp)]n (4)
-
-
120MV: 85.5 (1); 78.4 (2); 84.2 (3); 96.2 (4)
MB: 45.8 (1); 65.0 (2); 82.9 (3); 84.0 (4)
6 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; Hg lamp stable (4)[58]
RhB/Basic Violet 10{[Cu8Cl5(CPT)8(H2O)4](HSiW12) (H2O)20(CH3CN)4}n 2.69709710 mg/L dye; pH = 7; conc. catalyst = 0.375 g/L; H2O2 = 1 mL; visible lightstable[94]
CV/Basic Violet 3
AzI/52010
NR/Basic Red 5
Mgtptc based on MgO6 octahedral (Mgtptc)2.8455
50
55
90.94 (CV)
100 (Az1)
94.39 (NR)
-stable[92]
MB/Basic Blue 9
MV/Basic Violet 3
MG/Basic Green 4
{[Mn2(L)(tib)(H2O)] 3DMA 4H2O}n (1)
{[Co2(L)(bipd)2] 3H2O}n (2)
{[Co(L)0.5(bimb)] 3H2O 0.5O2}n (3)
3.35
3.37
3.22
120(1): 59.6 (MB); 63.6 (MV); 55.2 (MG);
(2) 85.8 (MB); 88.8 (MV);
82.3 (MG)
(3) 91.9 (MB); 93.8 (MV); 92.8 (MG)
6 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; UV lightstable (compound 2,3)[116]
MV/Basic Violet 3
MB/Basic Blue 9
Co(L)(tib)(-H2O)2]n (1)
[Co(L)(bip)0.5]n (2)
2.63
2.80
120(1): 92.1 (MV); 75.2 (MB)
(2): 89.7 (MV); 57.7 (MB)
10 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; Hg lampstable[59]
MB/Basic Blue 9
MV/Basic Violet 3
RhB/Basic Violet 10
Cu-tipe/GO composite 1.3530
45
60
90.4 (MB)
92.2 (MV)
91.3 (RhB)
10 mg/L dye; pH = 7; conc. catalyst = 0.4 g/L; H2O2 = 0.5 mL; visible light-[78]
MB/Basic Blue 9MIL-53(Al)@TiO2 3.052409520 mg/L dye; pH = 7; conc. catalyst = 0.1 g/L; visible light-[55]
MB/Basic Blue 9Cu(ttpa)-1@GO-3093.3 visible light-[79]
RhB/Basic Violet 10[EuCd(pbbp)2(SO4)2Cl2(H2O)3]-120921 × 10−5 M dye; pH = 7; conc. catalyst = 1 g/L; H2O2 = 1 mL; Xe lampstable[117]
CR/Direct Red 28
MB/Basic Blue 9
RhB/Basic Violet 10
Co-doped MIL-53(Al) 2.6316 (CR)
12 (MB)
16 (RhB)
99 (CR)
99 (MB)
99 (RhB)
30 mg/L dye; pH = 7; conc. catalyst = 0.3 g/L; Sunlightstable[54]
RhB/Basic Violet 10
MB/Basic Blue 9
MO/Acid Orange 52
{[Zn3(L)(BDC)3(H2O)3] 3.5H2O}n 2.947090 (RhB)
90 (MB)
90 (MO)
15 mg/L dye; pH = 7; conc. catalyst = 0.33 g/L; UV light-[94]
MB/Basic Blue 9[Cd(4- Hptz)2(H2O)2Cl2] (1)
[Cu(btx)2(ClO4)2]n (2)
[Cu(btx)(ClO4)]n (3)
4.12
3.58
2.59
10093.43 (1)
47.32 (2)
14.60 (3)
10 mg/L dye; pH = 7; conc. catalyst = 0.175 g/L; Hg lampstable[118]
MB/Basic Blue 9
MO/Acid Orange 52
RhB/Basic Violet 10
[Cu3(L)2(4,4′ -bipy)]2.8210023.19 (RhB)
9.26 (MO)
81.17 (MB)
10 mg/L dye; pH = 7; catalyst: 40 mgstable[89]
RhB/Basic Violet 10bimetallic MOF anchored corncob calcined derived activated carbon (CCAC)2.557510020 mg/L dye; pH = 7;conc. Catal yst = 0.05 g/L; Xe lampstable[119]
RhB/Basic Violet 10Ni-MOF (1)
NH2-MIL-101(Fe) (2)
Ti-MOF (3)
2.77
1.63
2.55
909 (1)
90 (2)
50 (3)
20 mg/L dye; pH = 7;conc. Catal yst = 1 g/L; Solar lightstable[56]
RhB/Basic Violet 10Zr6O4(OH)4 (ABDC)6/natural sepiolite (Sp) clay2.851206420 mg/L dye; pH = 5.9; conc. catal yst = 0.5 g/L visible light-[120]
NR/Basic Red 5{[Co(bcpt)(bib)3/2(H2O)] H2O}n (1)
{[Ni(bcpt)(bib)3/2(H2O)] H2O}n (2)
3.6
2.9
100 See information supplementary [121]
MV/Basic Violet 3[Co43-OH)2(H2O)2(L)2 (bib)1.5(CH3CN) 3H2O CH3CN] (1)
[Co83-OH)4(L)4 (bib)4 8CH3CN] (2)
3.64
2.88
10067.99
78.22
UV light-[91]
MV/Basic Violet 3[NH2(CH3)2][Zn2O(bmp)(BTC)] (1)
[NH2(CH3)2][Zn2O(bmp)(BTC)]·1·5DMF (2)
3.63
3.96
4038.25
55.95
10 mg/L dye; pH = 7; conc. catal yst = 0.8 g/L; UVlampStable[122]
MB/Basic Blue 9CdS/Cd-MOF 2.36100 20 mg/L dye; pH = 7; conc. catal yst = 0.1 g/L; Xe lampStable[123]
MB/Basic Blue 9
CR/Direct Red 28
MO/Acid Orange 52
RhB/Basic Violet 10
CuCd-BMOF/GO-x 2.74090.1 (MB)
85.4 (CR)
79.6 (MO)
87.6 (RhB)
20 mg/L dye; pH = 7; conc. catal yst = 0.6 g/L; Xe lampStable[41]
MV/Basic Violet 3
MG/
RhB/Basic Violet 10
NH2-UiO-66/FeOOH quantum dots2.5318082.2 (MV)
94.3 (MG)
81.5 (RhB)
20 mg/L dye; pH = 7; conc. catal yst = 0.6 g/L; 30 mM H2O2; LED lampstable[80]
MB/Basic Blue 9NH2-MIL-125(Ti)/polysulfone membrane -3008810 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; visible light -[48]
RhB/Basic Violet 10Bi-TATB 3.2118099.13 × 10−5 M dye; pH = 7; conc. catalyst = 0.1 g/L; LED lightstable[124]
MB/Basic Blue 9Ce-MOF/g-C3N4–TS 2.4512010010 mg/L dye; pH = 7; conc. catalyst = 0.25 g/L; UVlightstable[125]
MB/Basic Blue 9
MV/Basic Violet 3
RhB/Basic Violet 10
{[Ni(L)(bpt)(H2O)] 2H2O}n -9089.5 (MB)
84.7 (MV)
76.4 (RhB)
6 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; UV light-[104]
MO/Acid Orange 52STA-12Fe2.6401210 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; Sun lightstable (Fe)[126]
MB/Basic Blue 9
CR/Direct Red 28
MO/Acid Orange 52
RhB/Basic Violet 10
CuCd-BMOF/GO-x 2.74092 (Mb)
90.1 (CR)
85.4 (MO)
79.6 (RhB)
10 mg/L dye; pH = 7; conc. catalyst = 0.6 g/L; Xe lampstable[41]
MB/Basic Blue 9
MV/Basic Violet 3
RhB/Basic Violet 10
{[Cd(HL)(tib)(H2O)] 1.5H2O}}n (1)
{[Co3(L)2(tib)2] 4H2O}n (2)
[Cu(L)(3,5-bibp)]n (3)
3.13
3.32
3.37
90(1): 87.4 (MB); 67.3 (MV); 66.5 (RhB)
(2): 83.7 (MB); 48.4 (MV); 33.5 (RhB)
(3) 86.9 (MB); 55 (MV); 53.4 (RhB)
6 mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; UV lightstable[112]
RhB/Basic Violet 10[Mn(L)0.5(H2O)] (1)
[Cd(L)0.5(H2O)] (2)
2.84
3.05
15090.5 (1)
70.5 (2)
10 mg/L dye; pH = 7; conc. catalyst = 1 g/L; UV lightstable[105]
RhB/Basic Violet 10Ag/AgCl@CoFe2O4/NH2-MIL-125(Ti)2.73109720 mg/L dye; pH = 9; conc. catalyst = 0.8 g/L; H2O2 = 20 μL; Visible lightstable[47]
MB/Basic Blue 9
RhB/Basic Violet 10
MoS2@MIL-88(Fe)2.756098.5 (MB)
97.4 (RhB)
50 mg/L RhB; pH = 7; conc. catalyst = 1 g/L; Xe lampstable[127]
MB/Basic Blue 9Ag3PO4/UiO-66(Zr)/g-C3N42.924061.710 mg/L dye; pH = 7; conc. catalyst = 0.7 g/L; Visible light [128]
MB/Basic Blue 9
MV/Basic Violet 3
[Cd(bimd)(Hbipa)(H2O)]n (1)
{[Zn(bipd)(Hbipa)] H2O}n (2)
[Zn(bimb)(Hbipa)]n (3)
2.24
2.55
2.36
120(1): 90.3 (MB); 85.7 (MV)
(2): 89.8 (MB); 89 (MV)
(3): 92.9 (MB);95.3 (MV)
6mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; UV light-[111]
AO7/Acid Orange 7
MB/Basic Blue 9
CR/Direct Red 28
Ag2CrO4/MIL-53(Fe) 2.2214095.5 (AO7)
~60 (MB)
~80 (CR)
20mg/L dye; pH = 6; conc. catalyst = 1 g/L; Halogen lampstable[98]
MB/Basic Blue 9
MV/Basic Violet 3
NH2-MIL-101(Fe)@CuCoNi2.5612099.8 (MB)
93.8 (CV)
25mg/L dye; pH = 6; conc. catalyst = 1 g/L; 20 μL H2O2; Visible lightstable[129]
MG/Basic Green 4
MO/Acid Orange 52
{[Zn2(DIPA)2(bimp)5]·DMF 2H2O}n (1)
{[Zn2(HBPPA)2(bibp)2]·2H2O}n (2)
2.08
2.26
120(1): 96.6 (MB); 78.6 (MV)
(2): 98.3 (MB); 96.9 (MV)
6mg/L dye; pH = 7; conc. catalyst = 0.2 g/L; Hg lightstable[130]
RhB/Basic Violet 10g-C3N4/ZnFe2O4/UiO-662.86098visible lightstable[83]
RhB/Basic Violet 10NH2-MIL-101(Fe)/copper oxide 5010050 mg/L dye; pH = 6.5; conc. catalyst = 0.5 g/L; LED-[131]
MB/Basic Blue 9
MO/Acid Orange 52
Al-PcCl@NH2-MIL-125 (1)
CoPc@NH2-MIL-125 (2)
CoPcPA@NH2-MIL-125 (3)
2.47
2.42
2.54
120(1): 85(MB); 19 (MO)
(2): 79 (MB); 25 (MO)
(3): 99 (MB); 18 (MO)
30 mg/L dye; pH = 7; conc. catalyst = 0.5 g/L; Visible lightStable[46]
RhB/Basic Violet 10NH2-UiO-66/BiOBr/Bi2S3 609220 mg/L dye; pH = 7; conc. catalyst = 0.5 g/Lstable[52]
RhB/Basic Violet 10
CR/Direct Red 28
UiO-66 (1)
UiO-66-NH2 (2)
UiO-66-2M (3)
3.85
2.92
2.77
120(1): 48 (RhB); 52 (CR)
(2): 95 (RhB); 64 (CR)
(3): 69 (RhB); 48 (CR)
10 mg/L dye; pH = 7; conc. catalyst = 0.4 g/Lstable (3)[53]
BG/Green Basic 4Cu4(O) (OH)3(bb)4(NO3)30.3 H2O 3CH3CN]n 2.3212096.840 mg/L dye; pH = 5; mass catalyst = 30 mg-[132]
MB/Basic Blue 9
RhB/Basic Violet 10
Ag3VO4@MIL-125-NH2 (1)
Ag2WO4/MIL-125-NH2 (2)
2.27
2.57
60(1): 75 (MB); 75.3 (RhB)
(2): 65 (MB); 50.3 (RhB)
5 mg/L dye; pH = 7; conc. catalyst = 1 g/L; visible lightstable[45]
RhB/Basic Violet 10FCAU-17 (flakes) (Bi) (1)
CAU-17 (rods) (Bi) (2)
3.75
3.83
8078.6 (1)
96.1 (2)
5 mg/L dye; pH = 7; conc. catalyst = 0.1 g/L [133]
MB/Basic Blue 9amino acid-based carbon quantum dot/MIL-53 (Fe) binary composite2.631209835 mg/L dye; pH = 9; conc. catalyst = 0.1 g/L; Xe lampstable[81]
RhB/Basic Violet 10NH2-MIL-125 (Ti)/Ag/NiFe layered double hydroxide (Ti-MOF/Ag/NiFeLDH) 509515 mg/L dye; pH = 7; conc. catalyst = 0.5 g/L; Xe lampstable[134]
RhB/Basic Violet 10MIL-125/g-C3N4/sodium alginate-12099.41310 mg/L dye; pH = 5; Xe lampstable[82]
MB/Basic Blue 9Ag@{[Zn2(bta)(bpy)(H2O)2] 2H2O}n)2.8616086 10 mg/L dye; pH = 6; conc. catalyst = 1 g/L; Xe lamp-[77]
MB/Basic Blue 9[Zn4(tmlb)4(H2O)]n (1)
[Ni4(tmlb)4(H2O)]n (2)
[Cu(tmlb)]n (3)
[Co(tmlb)]n (4)
[Pb(tmlb) (H2O)]n (5)
3.92
3.89
3.9
3.75
3.8
12587.1 (1)
69.6 (2)
94.4 (3)
75.4 (4)
74.2 (5)
6 mg/L dye; pH = 6; conc. catalyst = 0.2 g/L; Hg lampstable[110]
MV/Basic Violet 3{[Co2(bdc)2(bmp)1.5]·DMF·1.5H2O}n 406810 mg/L dye; pH = 6; conc. catalyst = 0.8 g/L; Hg lampstable[101]
MV/Basic Violet 3[Co22- H2O)(L)(2,2′-bipy)2]n (1)
[Co22-H2O)(L)(phen)2]n (2)
-
-
4069 (1)
61 (2)
10 mg/L dye; pH = 6; conc. catalyst = 0.8 g/L; UV lamp-[135]
AO7/Acid Orange 7
MB/Basic Blue 9
RhB/Basic Violet 10
Ag2CrO4/Cu (BDC) 2.9210098 (AO7)
68 (MB)
46 (RhB)

20 mg/L RhB; pH = 6; mass catalyst = 15 mg; Halogen lampstable[99]
MB/Basic Blue 9AmCoPc/UiO-66-NH2
amino cobalt phthalocyanine (AmCoPc)
2.2118080.1710 mg/L RhB; pH = 6; conc. catalyst = 1 g/L; Xe lampstable[136]
MB/Basic Blue 9Ce/Co bimetallic MOFs with the mixed ligands 3.212082.310 mg/L RhB; pH = 6; conc. catalyst = 1 g/L; Xe lampstable[137]
Ponceau BS (PBS); acid orange 7 (AO7); Azure 1 (AzI); malachite green (MG).

3. Challenges, Perspectives, and Conclusions

Photocatalysis as a treatment method for the degradation/elimination of recalcitrant compounds of wastewater provides advantages compared to other methods, such as being a non-selective method, not generating sludges, and sometimes achieving complete dye mineralization. It is relatively inexpensive and has a simple configuration. Heterogeneous photocatalysis is understood to be the process that involves the joint use of light (light being understood as the entire electromagnetic spectrum) and a solid inorganic material called a semiconductor. Thus, employing MOFs (new MOFs, combination of ligands, and modification with other semiconductors) for the degradation of dyes offers advantages such as faster degradation, separation of the photo-generated species (electrons and holes) to make the degradation process more efficient, and providing stability to the MOF, thus increasing its useful life in the photocatalytic process. The dyes are relatively straightforward to characterize in the MOF reactions, as they can break the chromophore group in the dyes. The primary degradation mechanism for these dyes involves the generation of hydroxyl radicals. Those modifications to the MOFs have been explored and have shown promising results in treating recalcitrant compounds, some of which are persistent and considered carcinogenic in some instances.
In this study, we observed that the research on dye degradation using MOFs focuses on two key aspects. Firstly, the quest for new MOFs that allow working in the spectrum’s visible region and, perhaps most importantly, their stability in aqueous environments. Despite the hundreds or even thousands of studies conducted in recent years, many MOFs are unstable in aqueous media, as exemplified by HKUST (CuBTC) or MOF-5 (Zn). This limitation hinders their use in the photocatalytic degradation of dyes in wastewater. Hence, recent advancements are concentrated on synthesizing MOFs with favorable physicochemical and optical properties and good stability. As shown in Table 4, recent research primarily focuses on two axes: the synthesis of new MOFs with complex ligands and the synthesis of composite materials. Both approaches aim to enhance MOF stability and improve its optical properties, such as the band gap and the recombination of charge carriers. While the research results over the past five years have been significant, we believe that using complex ligands for MOF synthesis may significantly limit their practical application as photocatalysts for dye removal. This limitation arises from the challenges in obtaining these ligands due to specific synthesis conditions or their high acquisition costs. Therefore, we suggest that research should be directed toward developing new MOFs composed of simpler ligands, such as terephthalic or trimesic acid.

Author Contributions

Conceptualization, E.R.-G. and J.M.B.-A.; methodology, E.R.-G., A.M.M.-F., and R.L.-M.; investigation, J.M.B.-A., E.R.-G., A.M.M.-F., R.L.-M. and N.d.l.F.-M.; writing—original draft preparation, E.R.-G. and J.M.B.-A.; writing—review and editing, J.M.B.-A., E.R.-G., A.M.M.-F., R.L.-M. and N.d.l.F.-M.; supervision, E.R.-G.; project administration, E.R.-G.; funding acquisition, E.R.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Acknowledgments

All authors thank CONAHCYT for the granted incentive of SNI. J.M. Barrera-Andrade, N. de la Fuente-Maldonado, and E. Rojas-García thank to CONAHCYT for a postdoctoral fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pregi, E.; Kun, D.; Faludi, G.; Móczó, J.; Pukánszky, B. Modeling the Mechanical Properties of Polypropylene/Lignin/Flax Hybrid Composites. Mater. Des. 2022, 220, 110833. [Google Scholar] [CrossRef]
  2. Varzakas, T.; Tzia, C. Handbook of Food Processing: Food Safety, Quality, and Manufacturing Processes; CRC Press: Boca Raton, FL, USA, 2015; ISBN 978-1-4987-2178-3. [Google Scholar]
  3. Rodriguez-Amaya, D.B. Natural Food Pigments and Colorants. Curr. Opin. Food Sci. 2016, 7, 20–26. [Google Scholar] [CrossRef]
  4. Gürses, A.; Karaca, S.; Doğar, Ç.; Bayrak, R.; Açıkyıldız, M.; Yalçın, M. Determination of Adsorptive Properties of Clay/Water System: Methylene Blue Sorption. J. Colloid Interface Sci. 2004, 269, 310–314. [Google Scholar] [CrossRef]
  5. Lipskikh, O.I.; Korotkova, E.I.; Khristunova, Y.P.; Barek, J.; Kratochvil, B. Sensors for Voltammetric Determination of Food Azo Dyes—A Critical Review. Electrochim. Acta 2018, 260, 974–985. [Google Scholar] [CrossRef]
  6. Berradi, M.; Hsissou, R.; Khudhair, M.; Assouag, M.; Cherkaoui, O.; El Bachiri, A.; El Harfi, A. Textile Finishing Dyes and Their Impact on Aquatic Environs. Heliyon 2019, 5, e02711. [Google Scholar] [CrossRef] [PubMed]
  7. Sekar, N. 15—Acid Dyes. In Handbook of Textile and Industrial Dyeing; Clark, M., Ed.; Woodhead Publishing Series in Textiles; Woodhead Publishing: Sawston, UK, 2011; Volume 1, pp. 486–514. ISBN 978-1-84569-695-5. [Google Scholar]
  8. El Qada, E.N.; Allen, S.J.; Walker, G.M. Adsorption of Basic Dyes from Aqueous Solution onto Activated Carbons. Chem. Eng. J. 2008, 135, 174–184. [Google Scholar] [CrossRef]
  9. Hunger, K.; Mischke, P.; Rieper, W.; Raue, R.; Kunde, K.; Engel, A. Azo Dyes. In Ullmann’s Encyclopedia of Industrial Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2000; ISBN 978-3-527-30673-2. [Google Scholar]
  10. Lellis, B.; Fávaro-Polonio, C.Z.; Pamphile, J.A.; Polonio, J.C. Effects of Textile Dyes on Health and the Environment and Bioremediation Potential of Living Organisms. Biotechnol. Res. Innov. 2019, 3, 275–290. [Google Scholar] [CrossRef]
  11. Atchudan, R.; Jebakumar Immanuel Edison, T.N.; Perumal, S.; Karthikeyan, D.; Lee, Y.R. Effective Photocatalytic Degradation of Anthropogenic Dyes Using Graphene Oxide Grafting Titanium Dioxide Nanoparticles under UV-Light Irradiation. J. Photochem. Photobiol. Chem. 2017, 333, 92–104. [Google Scholar] [CrossRef]
  12. Natarajan, S.; Bajaj, H.C.; Tayade, R.J. Recent Advances Based on the Synergetic Effect of Adsorption for Removal of Dyes from Waste Water Using Photocatalytic Process. J. Environ. Sci. 2018, 65, 201–222. [Google Scholar] [CrossRef]
  13. Yagub, M.T.; Sen, T.K.; Afroze, S.; Ang, H.M. Dye and Its Removal from Aqueous Solution by Adsorption: A Review. Adv. Colloid Interface Sci. 2014, 209, 172–184. [Google Scholar] [CrossRef]
  14. Ahn, D.-H.; Chang, W.-S.; Yoon, T.-I. Dyestuff Wastewater Treatment Using Chemical Oxidation, Physical Adsorption and Fixed Bed Biofilm Process. Process Biochem. 1999, 34, 429–439. [Google Scholar] [CrossRef]
  15. Soloman, P.A.; Basha, C.A.; Velan, M.; Ramamurthi, V.; Koteeswaran, K.; Balasubramanian, N. Electrochemical Degradation of Remazol Black B Dye Effluent. CLEAN-Soil Air Water 2009, 37, 889–900. [Google Scholar] [CrossRef]
  16. Kornaros, M.; Lyberatos, G. Biological Treatment of Wastewaters from a Dye Manufacturing Company Using a Trickling Filter. J. Hazard. Mater. 2006, 136, 95–102. [Google Scholar] [CrossRef] [PubMed]
  17. Banat, F.; Al-Asheh, S.; Al-Rawashdeh, M.; Nusair, M. Photodegradation of Methylene Blue Dye by the UV/H2O2 and UV/Acetone Oxidation Processes. Desalination 2005, 181, 225–232. [Google Scholar] [CrossRef]
  18. Saeed, M.; Muneer, M.; ul Haq, A.; Akram, N. Photocatalysis: An Effective Tool for Photodegradation of Dyes—A Review. Environ. Sci. Pollut. Res. 2022, 29, 293–311. [Google Scholar] [CrossRef]
  19. Chen, J.; Hu, H.; Yang, J.; Xue, H.; Tian, Y.; Fan, K.; Zeng, Z.; Yang, J.; Wang, R.; Liu, Y. Removal Behaviors and Mechanisms for Series of Azo Dye Wastewater by Novel Nano Constructed Macro-Architectures Material. Bioresour. Technol. 2021, 322, 124556. [Google Scholar] [CrossRef]
  20. García, E.R.; Medina, R.L.; Lozano, M.M.; Hernández Pérez, I.; Valero, M.J.; Franco, A.M.M. Adsorption of Azo-Dye Orange II from Aqueous Solutions Using a Metal-Organic Framework Material: Iron- Benzenetricarboxylate. Materials 2014, 7, 8037–8057. [Google Scholar] [CrossRef]
  21. Ayati, A.; Shahrak, M.N.; Tanhaei, B.; Sillanpää, M. Emerging Adsorptive Removal of Azo Dye by Metal–Organic Frameworks. Chemosphere 2016, 160, 30–44. [Google Scholar] [CrossRef]
  22. Beydaghdari, M.; Hooriabad Saboor, F.; Babapoor, A.; Karve, V.V.; Asgari, M. Recent Advances in MOF-Based Adsorbents for Dye Removal from the Aquatic Environment. Energies 2022, 15, 2023. [Google Scholar] [CrossRef]
  23. Liu, W.; Fan, C.; Zong, Z.; Li, N.; Ma, K.; Zhu, B.; Zhang, X.; Fan, Y. Two Co(Ⅱ)-Based Metal Organic Frameworks for Highly Efficient Removal of Azo Dyes from Aqueous Environment: Synthesis, Selective Adsorption and Adsorption Mechanism. Colloids Surf. Physicochem. Eng. Asp. 2020, 603, 125236. [Google Scholar] [CrossRef]
  24. Timpua, T.; Jasman, J.; Pianaung, R.; Pongoh, D.; Pongoh, E.; Syafaruddin, S. Simple design of dye water filtering system for clean and healthy water quality. MATTER Int. J. Sci. Technol. 2017, 3, 546–559. [Google Scholar] [CrossRef]
  25. Soliman, E.S.; Hassan, R.A.; Farid, D.S. The Efficiency of Natural-Ecofriendly Clay Filters on Water Purification for Improving Performance and Immunity in Broiler Chickens. Open Vet. J. 2021, 11, 483–499. [Google Scholar] [CrossRef]
  26. Tan, K.B.; Vakili, M.; Horri, B.A.; Poh, P.E.; Abdullah, A.Z.; Salamatinia, B. Adsorption of Dyes by Nanomaterials: Recent Developments and Adsorption Mechanisms. Sep. Purif. Technol. 2015, 150, 229–242. [Google Scholar] [CrossRef]
  27. Abd-Elhamid, A.I.; Kamoun, E.A.; El-Shanshory, A.A.; Soliman, H.M.A.; Aly, H.F. Evaluation of Graphene Oxide-Activated Carbon as Effective Composite Adsorbent toward the Removal of Cationic Dyes: Composite Preparation, Characterization and Adsorption Parameters. J. Mol. Liq. 2019, 279, 530–539. [Google Scholar] [CrossRef]
  28. Cuenca Vargas, J.; David, J.; Werneck, M.; Franceschini, D.; Camargo, S., Jr. Síntese e caracterização de filmes de carbono para inibição de incrustações de carbonato de cálcio. In Proceedings of the Congresso Brasileiro de Carbono, São Paulo, Brazil, 10–12 December 2021. [Google Scholar]
  29. Wang, B.; Jin, C.; Shao, S.; Yue, Y.; Zhang, Y.; Wang, S.; Chang, R.; Zhang, H.; Zhao, J.; Li, X. Electron-Deficient Cu Site Catalyzed Acetylene Hydrochlorination. Green Energy Environ. 2023, 8, 1128–1140. [Google Scholar] [CrossRef]
  30. Ganjoo, R.; Sharma, S.; Kumar, A.; Daouda, M.M. Activated Carbon: Fundamentals, Classification, and Properties. In Activated Carbon: Progress and Applications; Royal Society of Chemistry: London, UK, 2023; pp. 1–22. ISBN 978-1-83916-780-5. [Google Scholar]
  31. Çeçen, F.; Aktaş, Ö. Activated Carbon for Water and Wastewater Treatment; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011; pp. i–xxvii. ISBN 978-3-527-63944-1. [Google Scholar]
  32. Faria, J.L.; Wang, W. Carbon Materials in Photocatalysis. In Carbon Materials for Catalysis; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2008; pp. 481–506. ISBN 978-0-470-40370-9. [Google Scholar]
  33. Robinson, T.; Chandran, B.; Nigam, P. Effect of Pretreatments of Three Waste Residues, Wheat Straw, Corncobs and Barley Husks on Dye Adsorption. Bioresour. Technol. 2002, 85, 119–124. [Google Scholar] [CrossRef] [PubMed]
  34. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.-G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A Critical Review on the Treatment of Dye-Containing Wastewater: Ecotoxicological and Health Concerns of Textile Dyes and Possible Remediation Approaches for Environmental Safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef] [PubMed]
  35. Sanromán, M.A.; Pazos, M.; Ricart, M.T.; Cameselle, C. Electrochemical Decolourisation of Structurally Different Dyes. Chemosphere 2004, 57, 233–239. [Google Scholar] [CrossRef]
  36. Dutta, S.; Bhattacharjee, J. Chapter 1—A Comparative Study between Physicochemical and Biological Methods for Effective Removal of Textile Dye from Wastewater. In Development in Wastewater Treatment Research and Processes; Shah, M.P., Rodriguez-Couto, S., Kapoor, R.T., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 1–21. ISBN 978-0-323-85657-7. [Google Scholar]
  37. Meghwal, K.; Agrawal, R.; Kumawat, S.; Jangid, N.K.; Ameta, C. Chemical and Biological Treatment of Dyes. In Impact of Textile Dyes on Public Health and the Environment; IGI Global: Hershey, PA, USA, 2020; pp. 170–204. ISBN 978-1-79980-311-9. [Google Scholar]
  38. Chen, Y.; Duan, X.; Zhou, X.; Wang, R.; Wang, S.; Ren, N.; Ho, S.-H. Advanced Oxidation Processes for Water Disinfection: Features, Mechanisms and Prospects. Chem. Eng. J. 2021, 409, 128207. [Google Scholar] [CrossRef]
  39. Konstantinou, I.K.; Albanis, T.A. TiO2-Assisted Photocatalytic Degradation of Azo Dyes in Aqueous Solution: Kinetic and Mechanistic Investigations: A Review. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  40. Tan, C.-E.; Su, E.-C.; Wey, M.-Y. Mixed Imidazole Ligand MIL-88A for Enhanced Photo-Fenton Decomposition of Azo Dye. Sol. Energy 2022, 246, 89–103. [Google Scholar] [CrossRef]
  41. Chen, J.; Wei, J.; Zhang, H.; Wang, X.; Fu, L.; Yang, T.-H. Construction of CuCd-BMOF/GO Composites Based on Phosphonate and Their Boosted Visible-Light Photocatalytic Degradation. Appl. Surf. Sci. 2022, 594, 153493. [Google Scholar] [CrossRef]
  42. 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. Physicochem. Eng. Asp. 2020, 589, 124484. [Google Scholar] [CrossRef]
  43. Ioannidou, T.; Anagnostopoulou, M.; Papoulis, D.; Christoforidis, K.C.; Vasiliadou, I.A. UiO-66/Palygorskite/TiO2 Ternary Composites as Adsorbents and Photocatalysts for Methyl Orange Removal. Appl. Sci. 2022, 12, 8223. [Google Scholar] [CrossRef]
  44. Andrade, P.H.M.; Palhares, H.; Volkringer, C.; Loiseau, T.; Hureau, M.; Nunes, E.; Moissette, A. State of the Art in Visible-Light Photocatalysis of Aqueous Pollutants Using Metal-Organic Frameworks. J. Photochem. Photobiol. C Photochem. Rev. 2023, 57, 100635. [Google Scholar] [CrossRef]
  45. Emam, H.E.; Ahmed, H.B.; Gomaa, E.; Helal, M.H.; Abdelhameed, R.M. Doping of Silver Vanadate and Silver Tungstate Nanoparticles for Enhancement the Photocatalytic Activity of MIL-125-NH2 in Dye Degradation. J. Photochem. Photobiol. Chem. 2019, 383, 111986. [Google Scholar] [CrossRef]
  46. Abdelhameed, R.M.; El-Shahat, M.; Abd El-Ghaffar, M.A. Boosting the Photocatalytic Activity of Ti-MOF via Emerging with Metal Phthalocyanine to Degrade Hazard Textile Pigments. J. Alloys Compd. 2022, 896, 162992. [Google Scholar] [CrossRef]
  47. Mahmoodi, N.M.; Taghizadeh, A.; Taghizadeh, M.; Abdi, J. In Situ Deposition of Ag/AgCl on the Surface of Magnetic Metal-Organic Framework Nanocomposite and Its Application for the Visible-Light Photocatalytic Degradation of Rhodamine Dye. J. Hazard. Mater. 2019, 378, 120741. [Google Scholar] [CrossRef]
  48. Ahmadi, A.; Sarrafzadeh, M.-H.; Hosseinian, A.; Ghaffari, S.-B. Foulant Layer Degradation of Dye in Photocatalytic Membrane Reactor (PMR) Containing Immobilized and Suspended NH2-MIL125(Ti) MOF Led to Water Flux Recovery. J. Environ. Chem. Eng. 2022, 10, 106999. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Cao, W.; Zhu, B.; Cai, J.; Li, X.; Liu, J.; Chen, Z.; Li, M.; Zhang, L. Fabrication of NH2-MIL-125(Ti) Nanodots on Carbon Fiber/MoS2-Based Weavable Photocatalysts for Boosting the Adsorption and Photocatalytic Performance. J. Colloid Interface Sci. 2022, 611, 706–717. [Google Scholar] [CrossRef]
  50. Sofi, F.A.; Majid, K. Plasmon Induced Interfacial Charge Transfer across Zr-Based Metal-Organic Framework Coupled Ag2WO4 Heterojunction Functionalized by Ag NPs: Efficient Visible Light Photocatalyst. Chem. Phys. Lett. 2019, 720, 7–14. [Google Scholar] [CrossRef]
  51. Chen, M.-L.; Li, S.-S.; Wen, L.; Xu, Z.; Li, H.-H.; Ding, L.; Cheng, Y.-H. Exploration of Double Z-Type Ternary Composite Long-Afterglow/Graphitic Carbon Nitride@metal–Organic Framework for Photocatalytic Degradation of Methylene Blue. J. Colloid Interface Sci. 2023, 629, 409–421. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, R.; Chen, Z.; Li, Y.; Zhao, C.; Cai, L.; Yu, J.; Yang, Z.; Jiang, J. A Novel Visible-Light-Induced Double Z-Scheme Photocatalytic System: NH2-UiO-66/BiOBr/Bi2S3 for Degradation of Tetracycline Hydrochloride and Rhodamine B. Colloids Surf. Physicochem. Eng. Asp. 2022, 649, 129350. [Google Scholar] [CrossRef]
  53. Gayathri, K.; Vinothkumar, K.; Teja, Y.N.; Al-Shehri, B.M.; Selvaraj, M.; Sakar, M.; Balakrishna, R.G. Ligand-Mediated Band Structure Engineering and Physiochemical Properties of UiO-66 (Zr) Metal-Organic Frameworks (MOFs) for Solar-Driven Degradation of Dye Molecules. Colloids Surf. Physicochem. Eng. Asp. 2022, 653, 129992. [Google Scholar] [CrossRef]
  54. Liu, C.; Zhu, C.; Wang, H.; Xie, S.; Zhou, J.; Fang, H. Synergistic Removal of Organic Pollutants by Co-Doped MIL-53(Al) Composite through the Integrated Adsorption/Photocatalysis. J. Solid State Chem. 2022, 316, 123582. [Google Scholar] [CrossRef]
  55. Zokaee, Z.; Mahmoodi, N.M.; Rahimpour, M.R.; Shariati, A. Synthesis of Visible Light Activated Metal-Organic Framework Coated on Titania Nanocomposite (MIL-53(Al)@TiO2) and Dye Photodegradation. J. Solid State Chem. 2022, 307, 122747. [Google Scholar] [CrossRef]
  56. Pattappan, D.; Vargheese, S.; Kavya, K.V.; Kumar, R.T.R.; Haldorai, Y. Metal-Organic Frameworks with Different Oxidation States of Metal Nodes and Aminoterephthalic Acid Ligand for Degradation of Rhodamine B under Solar Light. Chemosphere 2022, 286, 131726. [Google Scholar] [CrossRef]
  57. Zhang, X.; Meng, X.; Wang, J.; Ji, Z.; Lu, P.; Wang, X.; Chen, F. Rational Design of Two Novel Metal–Organic Frameworks as Photocatalysts for Degradation of Organic Dyes and Their Derived Materials as Electrocatalysts for OER. Inorganica Chim. Acta 2021, 523, 120416. [Google Scholar] [CrossRef]
  58. Ma, K.; Bi, C.; Zhang, X.; Zong, Z.; Xu, C.; Li, N.; Zhang, X.; Zhang, D.; Fan, Y. Four Co(II)—MOFs Based on 3,3′-Azodibenzoic Acid Ligand: Syntheses, Crystal Structure, Photodegradation Properties. J. Solid State Chem. 2019, 277, 115–122. [Google Scholar] [CrossRef]
  59. Cui, G.; Liu, W.; Wang, L.; Wu, R.; Bi, C.; Zhang, D.; Fan, Y. Two Novel Co (II) Bifunctional MOFs: Syntheses and Applications in Photocatalytic Degradation of Dyes and Electrocatalytic Water Oxidation. J. Solid State Chem. 2021, 304, 122562. [Google Scholar] [CrossRef]
  60. Laurier, K.G.M.; Vermoortele, F.; Ameloot, R.; De Vos, D.E.; Hofkens, J.; Roeffaers, M.B.J. Iron(III)-Based Metal–Organic Frameworks As Visible Light Photocatalysts. J. Am. Chem. Soc. 2013, 135, 14488–14491. [Google Scholar] [CrossRef]
  61. Nasalevich, A.; Becker, R.; Ramos-Fernandez, V.; Castellanos, S.; Veber, L.V.; Fedin, M.; Kapteijn, F.; Reek, H.J.N.; van der Vlugt, J.I.; Gascon, J. Co@NH 2 -MIL-125(Ti): Cobaloxime-Derived Metal–Organic Framework-Based Composite for Light-Driven H 2 Production. Energy Environ. Sci. 2015, 8, 364–375. [Google Scholar] [CrossRef]
  62. Baldovi, H.G.; Krüger, M.; Reinsch, H.; Alvaro, M.; Stock, N.; Garcia, H. Transient Absorption Spectroscopy and Photochemical Reactivity of CAU-8. J. Mater. Chem. C 2015, 3, 3607–3613. [Google Scholar] [CrossRef]
  63. Kent, C.A.; Liu, D.; Ma, L.; Papanikolas, J.M.; Meyer, T.J.; Lin, W. Light Harvesting in Microscale Metal–Organic Frameworks by Energy Migration and Interfacial Electron Transfer Quenching. J. Am. Chem. Soc. 2011, 133, 12940–12943. [Google Scholar] [CrossRef] [PubMed]
  64. Song, X.; Zhao, L.; Zhang, N.; Liu, L.; Ren, X.; Ma, H.; Luo, C.; Li, Y.; Wei, Q. Zinc-Based Metal–Organic Framework with MLCT Properties as an Efficient Electrochemiluminescence Probe for Trace Detection of Trenbolone. Anal. Chem. 2022, 94, 14054–14060. [Google Scholar] [CrossRef]
  65. Sinha, N.; Wenger, O.S. Photoactive Metal-to-Ligand Charge Transfer Excited States in 3d6 Complexes with Cr0, MnI, FeII, and CoIII. J. Am. Chem. Soc. 2023, 145, 4903–4920. [Google Scholar] [CrossRef] [PubMed]
  66. Sun, D.; Liu, W.; Qiu, M.; Zhang, Y.; Li, Z. Introduction of a Mediator for Enhancing Photocatalytic Performance via Post-Synthetic Metal Exchange in Metal–Organic Frameworks (MOFs). Chem. Commun. 2015, 51, 2056–2059. [Google Scholar] [CrossRef]
  67. Subalakshmi, K.; Senthilselvan, J.; Kumar, K.A.; Kumar, S.A.; Pandurangan, A. Solvothermal Synthesis of Hexagonal Pyramidal and Bifrustum Shaped ZnO Nanocrystals: Natural Betacyanin Dye and Organic Eosin Y Dye Sensitized DSSC Efficiency, Electron Transport, Recombination Dynamics and Solar Photodegradation Investigations. J. Mater. Sci. Mater. Electron. 2017, 28, 15565–15595. [Google Scholar] [CrossRef]
  68. Alvaro, M.; Carbonell, E.; Ferrer, B.; Llabrés i Xamena, F.X.; Garcia, H. Semiconductor Behavior of a Metal-Organic Framework (MOF). Chem.-Eur. J. 2007, 13, 5106–5112. [Google Scholar] [CrossRef]
  69. Huang, W.; Wang, X.; Zhang, W.; Zhang, S.; Tian, Y.; Chen, Z.; Fang, W.; Ma, J. Intraligand Charge Transfer Boosts Visible-Light-Driven Generation of Singlet Oxygen by Metal-Organic Frameworks. Appl. Catal. B Environ. 2020, 273, 119087. [Google Scholar] [CrossRef]
  70. Huo, P.; Chen, T.; Hou, J.-L.; Yu, L.; Zhu, Q.-Y.; Dai, J. Ligand-to-Ligand Charge Transfer within Metal–Organic Frameworks Based on Manganese Coordination Polymers with Tetrathiafulvalene-Bicarboxylate and Bipyridine Ligands. Inorg. Chem. 2016, 55, 6496–6503. [Google Scholar] [CrossRef]
  71. Chen, Y.; Sakata, O.; Nanba, Y.; Kumara, L.S.R.; Yang, A.; Song, C.; Koyama, M.; Li, G.; Kobayashi, H.; Kitagawa, H. Electronic Origin of Hydrogen Storage in MOF-Covered Palladium Nanocubes Investigated by Synchrotron X-rays. Commun. Chem. 2018, 1, 61. [Google Scholar] [CrossRef]
  72. Yang, L.-M.; Ganz, E.; Svelle, S.; Tilset, M. Computational Exploration of Newly Synthesized Zirconium Metal–Organic Frameworks UiO-66, -67, -68 and Analogues. J. Mater. Chem. C 2014, 2, 7111–7125. [Google Scholar] [CrossRef]
  73. Andrade, P.H.M.; Henry, N.; Volkringer, C.; Loiseau, T.; Vezin, H.; Hureau, M.; Moissette, A. Iodine Uptake by Zr-/Hf-Based UiO-66 Materials: The Influence of Metal Substitution on Iodine Evolution. ACS Appl. Mater. Interfaces 2022, 14, 29916–29933. [Google Scholar] [CrossRef] [PubMed]
  74. Li, Y.; Zhang, S.; Song, D. A Luminescent Metal–Organic Framework as a Turn-On Sensor for DMF Vapor. Angew. Chem. Int. Ed. 2013, 52, 710–713. [Google Scholar] [CrossRef]
  75. Zhang, X.; da Silva, I.; Fazzi, R.; Sheveleva, A.M.; Han, X.; Spencer, B.F.; Sapchenko, S.A.; Tuna, F.; McInnes, E.J.L.; Li, M.; et al. Iodine Adsorption in a Redox-Active Metal–Organic Framework: Electrical Conductivity Induced by Host−Guest Charge-Transfer. Inorg. Chem. 2019, 58, 14145–14150. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-Based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, L.; Chen, X.-L.; Cai, M.; Yan, R.-K.; Cui, H.-L.; Yang, H.; Wang, J.-J. Zn-MOFs Composites Loaded with Silver Nanoparticles Are Used for Fluorescence Sensing Pesticides, Trp, EDA and Photocatalytic Degradation of Organic Dyes. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 289, 122228. [Google Scholar] [CrossRef] [PubMed]
  78. Tian, H.-X.; Zha, M.; Ding, J.-G.; Zhu, L.-M.; Li, B.-L.; Wu, B. Visible-Light-Driven and Ultrasonic-Assisted Copper Metal-Organic Frameworks and Graphene Oxide Nanocomposite for Decolorization of Dyes. J. Solid State Chem. 2021, 304, 122627. [Google Scholar] [CrossRef]
  79. Zha, M.; Zhou, W.-J.; Ma, L.-X.; Li, L.-Y.; Li, B.-L.; Li, H.-Y.; Hu, C.-J. Syntheses of Two Copper Metal-Organic Frameworks with Tri(1,2,4-Triazole) and Biscarboxylate and Graphene Oxide Composites for Decomposition of Dye by Visible-Light Driven and Ultrasonic Assisted. J. Solid State Chem. 2022, 307, 122864. [Google Scholar] [CrossRef]
  80. Hosseini, M.-S.; Abbasi, A.; Masteri-Farahani, M. Decoration of NH2-UiO-66 with FeOOH Quantum Dots for Improving Photo-Degradation of Organic Dyes upon Visible Light Irradiation. Appl. Surf. Sci. 2022, 604, 154514. [Google Scholar] [CrossRef]
  81. Rezaei, H.; Zinatizadeh, A.A.; Joshaghani, M.; Zinadini, S. Amino Acid-Based Carbon Quantum Dot Modified MIL-53 (Fe): Investigation of Its Visible-Driven Photocatalytic Activity Provided by a Highly Efficient Photoreactor. Mater. Sci. Semicond. Process. 2023, 154, 107190. [Google Scholar] [CrossRef]
  82. Zhao, J.; Li, B.; Liu, Z.; Dai, D.; Li, Y.; Shi, R.; Zhang, H. A Novel Solar-Triggered MIL-125(Ti)/g-C3N4/SA Composite Aerogel with High Catalytic Activity for Degradation of Organic Contaminants. Sep. Purif. Technol. 2021, 279, 119696. [Google Scholar] [CrossRef]
  83. Sharafinia, S.; Farrokhnia, A.; Lemraski, E.G.; Rashidi, A. Decoration of ZnFe2O4 and UiO-66 over g-C3N4 as Magnetically Novel Reusable Visible Light Photocatalyst for Degradation of Rh–B. Opt. Mater. 2022, 132, 112838. [Google Scholar] [CrossRef]
  84. 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]
  85. 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]
  86. 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]
  87. 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]
  88. Healy, C.; Patil, K.M.; Wilson, B.H.; Hermanspahn, L.; Harvey-Reid, N.C.; Howard, B.I.; Kleinjan, C.; Kolien, J.; Payet, F.; Telfer, S.G.; et al. The Thermal Stability of Metal-Organic Frameworks. Coord. Chem. Rev. 2020, 419, 213388. [Google Scholar] [CrossRef]
  89. Shi, C.; Zhou, X.; Liu, D.; Li, L.; Xu, M.; Sakiyama, H.; Muddassir, M.; Wang, J. A New 3D High Connection Cu-Based MOF Introducing a Flexible Tetracarboxylic Acid Linker: Photocatalytic Dye Degradation. Polyhedron 2021, 208, 115441. [Google Scholar] [CrossRef]
  90. Li, Y.; Feng, J.; Zhang, Y.; Wang, C.; Hao, J.; Wang, Y.; Xu, Y.; Cheng, X. Covalent Organic frameworks@ZIF-67 Derived Novel Nanocomposite Catalyst Effectively Activated Peroxymonosulfate to Degrade Organic Pollu-tants. Chemosphere 2023, 311, 137038. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, J.; Zhou, S.; Chen, C.; Lu, L.; Li, B.; Hu, W.; Kumar, A.; Muddassir, M. Two New Uncommon 3D Cobalt-Based Metal Organic Frameworks: Temperature Induced Syntheses and Enhanced Photocatalytic Properties against Aromatic Dyes. Dye. Pigment. 2021, 187, 109068. [Google Scholar] [CrossRef]
  92. Ren, H.-Y.; Du, J.-M. An Efficiently Heterogeneous Photocatalyst for Degradation of Cation and Neutral Dyes under UV Light Based on Size-Dependent Effects of Tetracarboxyate Complex. J. Solid State Chem. 2020, 292, 121681. [Google Scholar] [CrossRef]
  93. Zulys, A.; Adawiah, A.; Gunlazuardi, J.; Yudhi, M.D.L. Light-Harvesting Metal-Organic Frameworks (MOFs) La-PTC for Photocatalytic Dyes Degradation. Bull. Chem. React. Eng. Catal. 2021, 16, 170–178. [Google Scholar] [CrossRef]
  94. Chen, D.-M.; Zhang, X.-J. A 8-Fold Interpenetrated Metal-Organic Framework: Luminescent Property and Photocatalytic Dye Degradation Performance. J. Solid State Chem. 2023, 321, 123919. [Google Scholar] [CrossRef]
  95. Jin, X.; Tang, T.; Tao, X.; Huang, L.; Xu, D. A Novel Dual-Ligand Fe-Based MOFs Synthesized with Dielectric Barrier Discharge (DBD) Plasma as Efficient Photocatalysts. J. Mol. Liq. 2021, 340, 117290. [Google Scholar] [CrossRef]
  96. Jin, X.; Tang, T.; Tao, X.; Huang, L.; Shang, S. Regulating N Content to Anchor Fe in Fe-MOFs: Obtaining Multiple Active Sites as Efficient Photocatalysts. J. Taiwan Inst. Chem. Eng. 2022, 132, 104133. [Google Scholar] [CrossRef]
  97. Murtaza, S.Z.M.; Shomal, R.; Sabouni, R.; Ghommem, M. Facile Metal Organic Framework Composites as Photocatalysts for Lone/Simultaneous Photodegradation of Naproxen, Ibuprofen and Methyl Orange. Environ. Technol. Innov. 2022, 27, 102751. [Google Scholar] [CrossRef]
  98. Najafidoust, A.; Abdollahi, B.; Sarani, M.; Darroudi, M.; Vala, A.M. MIL-(53)Fe Metal-Organic Framework (MOF)-Based Ag2CrO4 Hetrostructure with Enhanced Solar-Light Degradation of Organic Dyes. Opt. Mater. 2022, 125, 112108. [Google Scholar] [CrossRef]
  99. Abdollahi, B.; Farshnama, S.; Abbasi Asl, E.; Najafidoust, A.; Sarani, M. Cu(BDC) Metal–Organic Framework (MOF)-Based Ag2CrO4 Heterostructure with Enhanced Solar-Light Degradation of Organic Dyes. Inorg. Chem. Commun. 2022, 138, 109236. [Google Scholar] [CrossRef]
  100. Zhang, Y.; Hu, W.; Rao, C.; Zhou, D.; Hu, Y.; Jin, J.; Muddassir, M. Fast Photocatalytic Organic Dye by Two Metal-Organic Frameworks with 3D Two-Fold Interpenetrated Feature. J. Mol. Struct. 2021, 1227, 129538. [Google Scholar] [CrossRef]
  101. Zhou, E.-H.; Hu, W.; Ding, Q.; Li, L.; Liao, Z.; Jin, J.-C. Photocatalytic Decontamination of Methyl Violet by a 3D Twofold Interpenetrating Metal–Organic Framework. Inorg. Chem. Commun. 2020, 120, 108155. [Google Scholar] [CrossRef]
  102. Zhong, X.-L.; Wang, J.; Shi, C.; Lu, L.; Srivastava, D.; Kumar, A.; Afzal, M.; Alarifi, A. Photocatalytic Applications of a New 3D Mn(II)-Based MOF with Mab Topology. Inorg. Chim. Acta 2022, 540, 121063. [Google Scholar] [CrossRef]
  103. Chen, D.-M.; Zheng, Y.-P.; Shi, D.-Y.; Fang, S.-M. An Acid-Base Resistant Polyoxometalate-Based Metal–Organic Framework Constructed from {Cu4Cl}7+ and {Cu2(CO2)4} Clusters for Photocatalytic Degradation of Organic Dye. J. Solid State Chem. 2020, 287, 121384. [Google Scholar] [CrossRef]
  104. Zhang, X.; Zhang, X.; Fan, C.; Zong, Z.; Zhang, D.; Luo, Q.; Bi, C.; Fan, Y. A Novel Metal–Organic Frameworks Assembled by One Angular Ligand and 5-Aminoisophthalic Acid: Synthesis, Structure, Electrochemical and Photocatalytic Properties. Polyhedron 2019, 168, 21–27. [Google Scholar] [CrossRef]
  105. Wu, Y.; Long, J.; Wang, L.; Chen, C.; Xie, B.; Wang, J.; Wang, X.; Shi, C. Two New Coordination Polymers with Tetracarboxylate as Photocatalysts for Dye Degradation. Polyhedron 2021, 203, 115216. [Google Scholar] [CrossRef]
  106. Nguyen, M.B.; Sy, D.T.; Thoa, V.T.K.; Hong, N.T.; Doan, H.V. Bimetallic Co-Fe-BTC/CN Nanocomposite Synthesised via a Microwave-Assisted Hydrothermal Method for Highly Efficient Reactive Yellow 145 Dye Photodegradation. J. Taiwan Inst. Chem. Eng. 2022, 140, 104543. [Google Scholar] [CrossRef]
  107. Mousavi, S.E.; Younesi, H.; Bahramifar, N.; Tamunaidu, P.; Karimi-Maleh, H. A Novel Route to the Synthesis of α-Fe2O3@C@SiO2/TiO2 Nanocomposite from the Metal-Organic Framework as a Photocatalyst for Water Treatment. Chemosphere 2022, 297, 133992. [Google Scholar] [CrossRef]
  108. Mahmoodi, N.M.; Abdi, J.; Taghizadeh, M.; Taghizadeh, A.; Hayati, B.; Shekarchi, A.A.; Vossoughi, M. Activated Carbon/Metal-Organic Framework Nanocomposite: Preparation and Photocatalytic Dye Degradation Mathematical Modeling from Wastewater by Least Squares Support Vector Machine. J. Environ. Manag. 2019, 233, 660–672. [Google Scholar] [CrossRef]
  109. Saedi, Z.; Hajinia, N. Concurrent First- and Second-Order Photodegradation of Azo Dyes Using TMU-16 Pillared-Layer Microporous Metal Organic Framework under Visible Light. J. Solid State Chem. 2021, 300, 122210. [Google Scholar] [CrossRef]
  110. Zong, Z.; Fan, C.; Zhang, X.; Meng, X.; Jin, F.; Fan, Y. Synthesis, Crystal Structures and Dye Removal Properties of a Series of Metal-Organic Frameworks Based on N-Heterocyclic Carboxylic Acid Ligands. Microporous Mesoporous Mater. 2019, 282, 82–90. [Google Scholar] [CrossRef]
  111. Wu, R.; Bi, C.; Wang, L.; Wang, J.; Fan, C.; Li, N.; Wang, M.; Shao, F.; Chen, G.; Fan, Y. Five Functional Cd/Zn-Based MOFs Constructed from V-Shaped Tricarboxylate Ligand for Rapidly Selective Adsorption and Efficiently Photocatalytic Degradation of Hazardous Aromatic Dyes. Synth. Met. 2021, 277, 116786. [Google Scholar] [CrossRef]
  112. Zhang, X.; Zong, Z.; Zhang, X.; Zhang, D.; Luo, Q.; Bi, C.; Fan, Y. Rational Design of Three Bifunctional MOFs for Photocatalysis Degradation and Selective Adsorption of Wastewater Organic Dyes Removal. Polyhedron 2020, 191, 114816. [Google Scholar] [CrossRef]
  113. 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]
  114. Brahmi, C.; Benltifa, M.; Vaulot, C.; Michelin, L.; Dumur, F.; Millange, F.; Frigoli, M.; Airoudj, A.; Morlet-Savary, F.; Bousselmi, L.; et al. New Hybrid MOF/Polymer Composites for the Photodegradation of Organic Dyes. Eur. Polym. J. 2021, 154, 110560. [Google Scholar] [CrossRef]
  115. Wang, J.; Zhou, S.; Wang, G.; Hu, W.; Kumar, A.; Muddassir, M. Modular Construction and Photocatalytic Properties of Two Co(II) Metal-Organic Frameworks. J. Mol. Struct. 2021, 1223, 129218. [Google Scholar] [CrossRef]
  116. Xu, C.; Luo, R.; Zhang, D.; Zhang, X.; Zong, Z.; Fan, C.; Zhu, B.; Fan, Y. Mn(II)/Co(II)-Based Metal-Organic Frameworks Assembled by 5,5′-(1,4-Xylylenediamino) Diisophthalic Acid and Various Nitrogen-Containing Ligands for Photocatalytic and Magnetic Properties. J. Solid State Chem. 2021, 303, 122535. [Google Scholar] [CrossRef]
  117. Liu, J.-J.; Fu, J.-J.; Liu, T.; Cheng, F.-X. Europium-Cadmium Organic Framework with Zwitterionic Ligand Exhibiting Tunable Luminescence, CO2 Adsorption and Dye Degradation. J. Solid State Chem. 2022, 313, 123346. [Google Scholar] [CrossRef]
  118. Wang, L.; Zeng, T.; Liao, G.; Cheng, Q.; Pan, Z. Syntheses, Structures and Catalytic Mechanisms of Three New MOFs for Aqueous Cr(VI) Reduction and Dye Degradation under UV Light. Polyhedron 2019, 157, 152–162. [Google Scholar] [CrossRef]
  119. Sun, Z.; Wu, X.; Qu, K.; Huang, Z.; Liu, S.; Dong, M.; Guo, Z. Bimetallic Metal-Organic Frameworks Anchored Corncob-Derived Porous Carbon Photocatalysts for Synergistic Degradation of Organic Pollutants. Chemosphere 2020, 259, 127389. [Google Scholar] [CrossRef]
  120. Pawar, R.R.; Chuaicham, C.; Sekar, K.; Rajendran, S.; Sasaki, K. Synthesis, Characterization, and Application of MOF@clay Composite as a Visible Light-Driven Photocatalyst for Rhodamine B Degradation. Chemosphere 2022, 291, 132922. [Google Scholar] [CrossRef]
  121. Qin, L.; Hu, Q.; Zou, Y.-P.; Fu, W.-P.; Ye, T.-Q.; Zhang, M.-D. Two Co(II)/Ni(II) Isostructural Metal-Organic Frameworks with Bnn Topology for Photocatalysis and Electrocatalysis. Microporous Mesoporous Mater. 2021, 312, 110813. [Google Scholar] [CrossRef]
  122. Wang, J.; Zhou, L.; Rao, C.; Wang, G.-L.; Jiang, F.; Singh, A.; Kumar, A.; Liu, J. Two 3D Supramolecular Isomeric Zn(II)-MOFs as Photocatalysts for Photodegradation of Methyl Violet Dye. Dye. Pigment. 2021, 190, 109285. [Google Scholar] [CrossRef]
  123. Jing, C.; Zhang, Y.; Zheng, J.; Ge, S.; Lin, J.; Pan, D.; Naik, N.; Guo, Z. In-Situ Constructing Visible Light CdS/Cd-MOF Photocatalyst with Enhanced Photodegradation of Methylene Blue. Particuology 2022, 69, 111–122. [Google Scholar] [CrossRef]
  124. Nguyen, V.H.; Pham, A.L.H.; Nguyen, V.-H.; Lee, T.; Nguyen, T.D. Facile Synthesis of Bismuth(III) Based Metal-Organic Framework with Difference Ligands Using Microwave Irradiation Method. Chem. Eng. Res. Des. 2022, 177, 321–330. [Google Scholar] [CrossRef]
  125. Durmus, Z.; Köferstein, R.; Lindenberg, T.; Lehmann, F.; Hinderberger, D.; Maijenburg, A.W. Preparation and Characterization of Ce-MOF/g-C3N4 Composites and Evaluation of Their Photocatalytic Performance. Ceram. Int. 2023, 49, 24428–24441. [Google Scholar] [CrossRef]
  126. Farrokhi, A.; Jafarpour, M.; Alipour, M. Solar-Driven Advanced Oxidation Process Catalyzed by Metal–Organic Frameworks for Water Depollution. Polyhedron 2019, 170, 325–333. [Google Scholar] [CrossRef]
  127. Govarthanan, M.; Mythili, R.; Kim, W.; Alfarraj, S.; Alharbi, S.A. Facile Fabrication of (2D/2D) MoS2@MIL-88(Fe) Interface-Driven Catalyst for Efficient Degradation of Organic Pollutants under Visible Light Irradiation. J. Hazard. Mater. 2021, 414, 125522. [Google Scholar] [CrossRef]
  128. Yassin, J.M.; Taddesse, A.M.; Sánchez-Sánchez, M. Sustainable Synthesis of Semicrystalline Zr-BDC MOF and Heterostructural Ag3PO4/Zr-BDC/g-C3N4 Composite for Photocatalytic Dye Degradation. Catal. Today 2022, 390–391, 162–175. [Google Scholar] [CrossRef]
  129. Chen, J.; Xing, Z.; Han, J.; Su, M.; Li, Y.; Lu, A. Enhanced Degradation of Dyes by Cu-Co-Ni Nanoparticles Loaded on Amino-Modified Octahedral Metal–Organic Framework. J. Alloys Compd. 2020, 834, 155106. [Google Scholar] [CrossRef]
  130. Wu, R.; Zhang, X.; Wang, J.; Wang, L.; Zhu, B.; Xu, C.; Cui, G.; Zhang, D.; Fan, Y. Two Novel Zn (II)-Based Metal–Organic Frameworks for Rapidly Selective Adsorption and Efficient Photocatalytic Degradation of Hazardous Aromatic Dyes in Aqueous Phase. Adv. Powder Technol. 2022, 33, 103665. [Google Scholar] [CrossRef]
  131. Hoseinzadeh, H.; Bakhtiari, M.; Seifpanahi-Shabani, K.; Oveisi, M.; Hayati, B.; Rabeie, B.; Shahmoradi Ghaheh, F.; Salmani, R.; Ullah, H.; Mahmoodi, N.M. Synthesis of the Metal-Organic Framework—Copper Oxide Nanocomposite and LED Visible Light Organic Contaminants (Dye and Pharmaceutical) Destruction Ability in the Water. Mater. Sci. Eng. B 2021, 274, 115495. [Google Scholar] [CrossRef]
  132. Jin, J.-C.; Wang, J.; Guo, J.; Yan, M.-H.; Wang, J.; Srivastava, D.; Kumar, A.; Sakiyama, H.; Muddassir, M.; Pan, Y. A 3D Rare Cubane-like Tetramer Cu(II)-Based MOF with 4-Fold Dia Topology as an Efficient Photocatalyst for Dye Degradation. Colloids Surf. Physicochem. Eng. Asp. 2023, 656, 130475. [Google Scholar] [CrossRef]
  133. Dong, S.; Wang, L.; Lou, W.; Shi, Y.; Cao, Z.; Zhang, Y.; Sun, J. Bi-MOFs with Two Different Morphologies Promoting Degradation of Organic Dye under Simultaneous Photo-Irradiation and Ultrasound Vibration Treatment. Ultrason. Sonochem. 2022, 91, 106223. [Google Scholar] [CrossRef] [PubMed]
  134. 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. 2021, 278, 119525. [Google Scholar] [CrossRef]
  135. Zhang, S.; Lu, L.; Wang, J.; Tan, X.; An, B.; Singh, A.; Kumar, A.; Sakiyama, H.; Wang, J. Photocatalytic and Magnetic Properties of Two New Co(II) Cluster-Based Metal-Organic Frameworks. Inorg. Chem. Commun. 2020, 111, 107563. [Google Scholar] [CrossRef]
  136. Yin, Y.; Liu, Y.; Xin, Z.; Xu, G.; Feng, Y.; Jiang, B.; He, X.; Zhang, H.; Ma, J. A New Type of Composite Catalyst AmCoPc/UiO-66-NH2 Synergistic Photocatalytic Degradation of Dyes. Inorg. Chem. Commun. 2022, 145, 109914. [Google Scholar] [CrossRef]
  137. Ma, S.; Shi, Y.; Xia, X.; Song, Q.; Yang, J. Cerium-Cobalt Bimetallic Metal–Organic Frameworks with the Mixed Ligands for Photocatalytic Degradation of Methylene Blue. Inorg. Chem. Commun. 2023, 152, 110664. [Google Scholar] [CrossRef]
Figure 1. Structures of organic dyes that will be studied in this work.
Figure 1. Structures of organic dyes that will be studied in this work.
Colorants 02 00035 g001
Figure 2. Main treatments of dyes.
Figure 2. Main treatments of dyes.
Colorants 02 00035 g002
Figure 3. Historic publications of the MOFs as semiconductors. (Source: Scopus; keywords: MOFs and photocatal * and dyes).
Figure 3. Historic publications of the MOFs as semiconductors. (Source: Scopus; keywords: MOFs and photocatal * and dyes).
Colorants 02 00035 g003
Figure 4. Scheme of charge transfer pathway in composite based MOFs; (a) MOF and semiconductor; (b) MOF, semiconductor and metal; and (c) MOF and semiconductor with Z-scheme.
Figure 4. Scheme of charge transfer pathway in composite based MOFs; (a) MOF and semiconductor; (b) MOF, semiconductor and metal; and (c) MOF and semiconductor with Z-scheme.
Colorants 02 00035 g004
Figure 5. Strategies to improve the degradation of dyes using MOFs as semiconductors.
Figure 5. Strategies to improve the degradation of dyes using MOFs as semiconductors.
Colorants 02 00035 g005
Figure 6. Step diagram of gradual degradation of MB dye. Taken with permission of Ref. [89].
Figure 6. Step diagram of gradual degradation of MB dye. Taken with permission of Ref. [89].
Colorants 02 00035 g006
Figure 7. A possible degradation pathway of RhB. Taken with permission of Ref. [90].
Figure 7. A possible degradation pathway of RhB. Taken with permission of Ref. [90].
Colorants 02 00035 g007
Figure 8. Acid dyes are most used in the characterization of MOFs in this review.
Figure 8. Acid dyes are most used in the characterization of MOFs in this review.
Colorants 02 00035 g008
Figure 9. (a) positions of valence and conduction band of La-PTC photocatalyst and dyes, and (b) charge transfer mechanism for MO degradation over La-PTC photocatalyst under visible-light irradiation. Taken with permission of Ref. [93].
Figure 9. (a) positions of valence and conduction band of La-PTC photocatalyst and dyes, and (b) charge transfer mechanism for MO degradation over La-PTC photocatalyst under visible-light irradiation. Taken with permission of Ref. [93].
Colorants 02 00035 g009
Figure 10. (a) The solid-state luminescent emission spectra of MOF and the ligands. (b) The calculated HOMO and LUMO for the periodic network of MOF. (c) UV–vis diffuse reflectance spectra for MOF and ligands. (d) Diffuse reflectance spectra of Kubelka–Munk function versus energy of MOF. Taken with permission of Ref. [94].
Figure 10. (a) The solid-state luminescent emission spectra of MOF and the ligands. (b) The calculated HOMO and LUMO for the periodic network of MOF. (c) UV–vis diffuse reflectance spectra for MOF and ligands. (d) Diffuse reflectance spectra of Kubelka–Munk function versus energy of MOF. Taken with permission of Ref. [94].
Colorants 02 00035 g010
Figure 11. (a) UV-visible absorption spectra (insert-Tauc’s plots for band gap estimation) and (b) photoluminescence spectra of the synthesized different liganded-UiO-66 MOFs. Taken with permission of Ref. [53].
Figure 11. (a) UV-visible absorption spectra (insert-Tauc’s plots for band gap estimation) and (b) photoluminescence spectra of the synthesized different liganded-UiO-66 MOFs. Taken with permission of Ref. [53].
Colorants 02 00035 g011
Figure 12. Schematic diagram showing the band structure and Z-scheme separation of photoinduced electron–hole pairs at interface of Ag/Ag2WO4/UIO-66 catalyst under visible light irradiation. Taken with permission of Ref. [50].
Figure 12. Schematic diagram showing the band structure and Z-scheme separation of photoinduced electron–hole pairs at interface of Ag/Ag2WO4/UIO-66 catalyst under visible light irradiation. Taken with permission of Ref. [50].
Colorants 02 00035 g012
Figure 13. Scheme of the generation of active species for the degradation of dyes, (a) without sacrificial agents (hydrogen peroxide) and (b) with sacrificial agents (hydrogen peroxide).
Figure 13. Scheme of the generation of active species for the degradation of dyes, (a) without sacrificial agents (hydrogen peroxide) and (b) with sacrificial agents (hydrogen peroxide).
Colorants 02 00035 g013
Figure 14. (a) View for the six coordinated Zn (II) ions of MOF; (b) the left-handed helical chains of MOF; (c) the 3D porous framework ligands and helical chains; (d) the 8-fold interpenetrated srs net of MOF. Taken with permission of Ref. [94].
Figure 14. (a) View for the six coordinated Zn (II) ions of MOF; (b) the left-handed helical chains of MOF; (c) the 3D porous framework ligands and helical chains; (d) the 8-fold interpenetrated srs net of MOF. Taken with permission of Ref. [94].
Colorants 02 00035 g014
Table 1. Structure of any azo-type organic dyes according to their classification.
Table 1. Structure of any azo-type organic dyes according to their classification.
Acid Dye
MO, Methyl Orange, C.I. Acid Orange 52
Colorants 02 00035 i001
Acid Orange, C.I. Acid Orange 7
Colorants 02 00035 i002
AB92, Acid Blue, C.I. Acid Blue 92
Colorants 02 00035 i003
Ponceau BS, C.I. Acid Red 66
Colorants 02 00035 i004
New Coccine, NC, C.I. Acid Red 18 = 1
Colorants 02 00035 i005
Tartrazine, C.I. Acid Yellow 23
Colorants 02 00035 i006
Acid Black, C.I. Acid Black 1
Colorants 02 00035 i007
Acid Red 1, C.I. Acid Red 1
Colorants 02 00035 i008
Basic dye
MB, Methylene Blue, C.I. Basic Blue 9
Colorants 02 00035 i009
Rhodamine B, RhB, C.I. Basic Violet 10
Colorants 02 00035 i010
Methyl violet, MV, CV, C.I. Basic Violet 3
Colorants 02 00035 i011
Malachite Green (MG), Green Basic, C.I. Basic Green 4
Colorants 02 00035 i012
Basic Violet (BV14), C.I. Basic Violet 14
Colorants 02 00035 i013
Neutral Red, NR, C.I. Basic Red 5
Colorants 02 00035 i014
Reactive dye
Reactive Red (RR)198, C.I. Reactive Red 198
Colorants 02 00035 i015
Reactive Yellow (RY), C.I. Reactive Yellow 145
Colorants 02 00035 i016
Direct dyeMordant dye
Congo Red, CR, C.I. Direct Red 28
Colorants 02 00035 i017
Eriochrome Black T, C.I. Mordant Black
Colorants 02 00035 i018
Blue = Nitrogen, Red = Oxygen, Yellow = Sulfur, White = Hydrogen, Gray = Carbon, Green = Chloro.
Table 2. Advantages and disadvantages of any of the methods used for the elimination/degradation of azo-type dyes in wastewater [13,14,15,16,17,18].
Table 2. Advantages and disadvantages of any of the methods used for the elimination/degradation of azo-type dyes in wastewater [13,14,15,16,17,18].
TreatmentAdvantageDisadvantage
Adsorption-Able to adsorb all the concentration of the dye.
-One step to recover the absorbent with the dye.
The dye only changes the medium; the environmental problem persisted throughout the time using this treatment.
Chemical oxidation-Oxidation can be achieved in most of the dyes.
-One or more oxidant substances can be used to oxidize the dye.
-In many cases, the sub-products are more toxic than the molecule to be degraded
-This treatment can be expensive because it uses reactive in huge quantities, and they will have a high concentration.
Electrochemical-The use of electric power can break down the dyes.
-This method can treat vast quantities of effluents with dyes.
After applying this treatment, the formation of a massive amount of sludge that has the sub-products of the degradation of the dye
Biological-Use more than one kind of bacteria.
-The system does not need the addition of more reactive because the dye can be used as food for the bacteria.
-This methodology can apply only to a few quantities of dyes.
-The biodegradation process is slow.
UV oxidation-Using UV-C light to break the links of dye molecules.
-The treatment can be applied for a short time.
-The high energy of the light can produce a health problem for the people around it.
-High consumption of electricity.
Photocatalysis-Require only a semiconductor and light.
-This methodology can eliminate a considerable volume with a low concentration.
-The system can be activated using solar radiation.
-Generally, the system does not generate more toxic compounds.
-The method can be achieved in many cases via the mineralization process of the dye.
-Some semiconductors must be modified to achieve a high percentage of dye degradation.
-In some cases, the separation process of the semiconductor from the water.
Table 3. Some types of charge transfer in MOFs and their characteristics.
Table 3. Some types of charge transfer in MOFs and their characteristics.
Charge TransferCharacteristics ApplicationsReference
Metal-to-Ligand Charge Transfer (MLCT)In MLCT, electrons transfer from the metal center to the ligand. This typically involves a transition metal cation donating electrons to ligand molecules.MLCT can enhance the photocatalytic activity of MOFs, making them useful for light-driven reactions.[63,64,65]
Ligand-to-Metal Charge Transfer (LMCT) or Ligand-to- Cluster Charge Transfer (LCCT)In LMCT, electrons transfer from the ligand to the metal center. This occurs when ligands act as donors to the metal ions.LMCT can affect MOFs’ electronic and optical properties, making them suitable for sensors and photoluminescent materials.[66,67,68,69]
Ligand-to-Ligand Charge Transfer (LLCT)LLCT involves electron transfer between two or more ligands within the MOF structure. This can lead to changes in the color and electronic properties of the MOF.LLCT is often associated with color changes and can be utilized in chemical sensing and colorimetric indicators.[70]
Guest-to-Host Charge Transfer (GHCT)GHCT occurs when guest molecules or ions transfer electrons to or from the MOF framework. This type of charge transfer is essential in gas adsorption and separation.GHCT is essential in gas storage and sensing applications, where guest molecules interact with the MOF structure.[71]
Linker-Based Transition (LBT) linker orbitals form the VB and CB. This is valuable in applications like water and pollutant degradation [72,73]
Metal-Based Transition (MBT)In MBT, electrons transfer between metal centers within the MOF structure. This is valuable in applications like pollutant degradation[74]
Host-to-Guest Charge
Transfer
(HGCT)
HGCT occurs when the MOF’s host structure transfers electrons to guest molecules or particles. HGTC is significant in applications like gas sensing and photocatalytic degradation of dyes. [73,75]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Barrera-Andrade, J.M.; de la Fuente-Maldonado, N.; Lopez-Medina, R.; Maubert-Franco, A.M.; Rojas-Garcia, E. Revolutionizing Wastewater Treatment: Harnessing Metal–Organic Frameworks for Exceptional Photocatalytic Degradation of Azo-Type Dyes. Colorants 2023, 2, 674-704. https://doi.org/10.3390/colorants2040035

AMA Style

Barrera-Andrade JM, de la Fuente-Maldonado N, Lopez-Medina R, Maubert-Franco AM, Rojas-Garcia E. Revolutionizing Wastewater Treatment: Harnessing Metal–Organic Frameworks for Exceptional Photocatalytic Degradation of Azo-Type Dyes. Colorants. 2023; 2(4):674-704. https://doi.org/10.3390/colorants2040035

Chicago/Turabian Style

Barrera-Andrade, Jose Manuel, Natali de la Fuente-Maldonado, Ricardo Lopez-Medina, Ana Marisela Maubert-Franco, and Elizabeth Rojas-Garcia. 2023. "Revolutionizing Wastewater Treatment: Harnessing Metal–Organic Frameworks for Exceptional Photocatalytic Degradation of Azo-Type Dyes" Colorants 2, no. 4: 674-704. https://doi.org/10.3390/colorants2040035

Article Metrics

Back to TopTop