Next Article in Journal
Interior of Amylopectin and Nano-Sized Amylopectin Fragments Probed by Viscometry, Dynamic Light Scattering, and Pyrene Excimer Fluorescence
Next Article in Special Issue
Influence of Multidimensional Graphene Oxide (GO) Sheets on Anti-Biofouling and Desalination Performance of Thin-Film Composite Membranes: Effects of GO Lateral Sizes and Oxidation Degree
Previous Article in Journal
Manufacture of Reduced Fat White-Brined Cheese with the Addition of β-Glucans Biobased Polysaccharides as Textural Properties Improvements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Critical Review on Metal-Organic Frameworks and Their Composites as Advanced Materials for Adsorption and Photocatalytic Degradation of Emerging Organic Pollutants from Wastewater

1
Fundamental and Applied Sciences Department, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Malaysia
2
Chemistry Department, Al-Qalam University Katsina, Katsina 2137, Nigeria
3
Chemical Engineering Department, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Malaysia
4
School of Chemical Sciences, Universiti Sains Malaysia, Gelugor 11800, Malaysia
5
Civil Engineering Department, Abubakar Tafawa Balewa University, Bauchi 740272, Nigeria
6
Physics Department, College of Science, Al-Imam Muhammad Ibn Saud Islamic University, Riyadh 11432, Saudi Arabia
7
Radiology and Medical Imaging Department, College of Applied Medical Sciences, Prince Sattam Bin Abduaziz University, Alkharj 11942, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Polymers 2020, 12(11), 2648; https://doi.org/10.3390/polym12112648
Submission received: 4 October 2020 / Revised: 1 November 2020 / Accepted: 6 November 2020 / Published: 10 November 2020

Abstract

:
Water-borne emerging pollutants are among the greatest concern of our modern society. Many of these pollutants are categorized as endocrine disruptors due to their environmental toxicities. They are harmful to humans, aquatic animals, and plants, to the larger extent, destroying the ecosystem. Thus, effective environmental remediations of these pollutants became necessary. Among the various remediation techniques, adsorption and photocatalytic degradation have been single out as the most promising. This review is devoted to the compilations and analysis of the role of metal-organic frameworks (MOFs) and their composites as potential materials for such applications. Emerging organic pollutants, like dyes, herbicides, pesticides, pharmaceutical products, phenols, polycyclic aromatic hydrocarbons, and perfluorinated alkyl substances, have been extensively studied. Important parameters that affect these processes, such as surface area, bandgap, percentage removal, equilibrium time, adsorption capacity, and recyclability, are documented. Finally, we paint the current scenario and challenges that need to be addressed for MOFs and their composites to be exploited for commercial applications.

1. Introduction

Emerging organic pollutants have received much concern due to their ubiquitous detection in various water spheres. They are toxic species produced from both natural and anthropogenic sources via; volcanoes, bush burning, petroleum exploration and refining, coal mining and processing, petrochemicals production, agrochemical application, textile, and leather dyeing, pharmaceutical production among others. They are widely discharged into the environment and conversely get deposited into the water bodies. Most of these pollutants are highly hydrophobic; thus, they bioaccumulate and magnify in the water and consequently get into the tissues of various aquatic organisms, as well as humans. Among the prominence includes dyes [1,2], pharmaceuticals and personal care products (PPCPs) [3,4], phenolics [5], herbicides and pesticides [6,7], polycyclic aromatic hydrocarbons (PAHs) [8,9], and perfluoroalkyl carboxylates and sulfonates [10,11]. These pollutants had been classified as endocrine disruptors (EDCs), due to their tendency to interfere with the function of the natural hormones [12,13]. They are highly resistant to naturally occurring processes of biodegradation and photolysis [14]. Toxicity studies have linked these compounds with many forms of ailments, such as genotoxicity, neurotoxicity, reproductive toxicity, development toxicity, cancerous tumors, etc. [15,16]. Thus, due to their frequent detection in the water and high toxicities, they are classified as emerging pollutants.
Environmental scientists, engineers, as well as environmental control and monitoring agencies, were challenged to provide effective remediations of these toxic pollutants. Thus, various methods have been put forward to achieve the tasks. Flocculation as an alternative have been practiced for decades [17]. The method is based on the formation of suspended solid particles (known as flocculants) using alumina, biopolymeric pectin, polyacrylamide, etc. [18]. Similarly, coagulation has also been considered [19]. However, the two suffered disadvantages of incomplete removal of the pollutants, as well as the formation of secondary pollution in form of sludge [20,21]. Other physical techniques, such as sedimentation, filtration, and reverse osmosis, have also been applied [22,23]. In most cases, they are not without drawbacks. Reverse osmosis, for example, requires periodic maintenance due to the clogging of the membranes [24,25]. The use of bioremediation using naturally occurring microorganisms, such as algae, bacteria, and fungi, to degrade the organic pollutants have been put forward [26,27]. However, some of these pollutants are resistant to biodegradations.
Due to the shortcomings of the aforementioned techniques, and driven by the need for a cheaper, sustainable, and effective treatment process, alternative approaches are necessary. Of these, adsorption and photocatalytic degradation are attractive as they could offer complete removal and mineralization of the toxic contaminants. This article is aimed at reviewing the application of metal-organic frameworks (MOFs) as versatile and highly efficient materials remediations of toxic organic pollutants from wastewater. Different classes of the pollutants have been discussed, and the literature reported on their removals by the MOFs has been detailed. Emphasis has been paid to adsorption and photocatalytic degradation using various pristine MOFs and their composites.

1.1. Adsorption

The application of adsorption techniques as an alternative wastewater remediation process has been well discovered. It has been proposed to solve the challenging task of incomplete removal of pollutants during wastewater processing. Organic pollutants are particularly more resistant to many forms of water remediation due to their hydrophobicity and lower molecular weight. For adsorption, process, pollutant molecules are attracted onto the surfaces of the adsorbent materials through diffusion process from the bulk of the solution to the active pores of the adsorbents [28,29]. Usually, the mechanism takes place through intermolecular forces of attraction, such as chemisorption (e.g., ionic interactions) and physisorption (e.g., van der Waals and π–π interactions) [30,31]. Adsorption has been emphasized by the unique properties of the adsorbent materials, such as high porosity, large specific Brunner Emmett Teller (BET) surface area, moisture and thermal stabilities, good selectivity for the target pollutants, availability, and low-cost, easy to handle and regenerated, etc. [32,33]. Among the desirable properties of ideal adsorbent materials is the physical state in form of either powder, cake or beads.
Among the most widely applied carbonaceous porous materials include biochar, activated carbon (AC), graphene, and carbon nanotubes [34,35]. They are usually obtained or synthesized from agricultural waste products. AC has been the most reported carbon-adsorbent. It has well-developed pore size distribution, with high surface functional groups that provide binding sites for adsorption of pollutants in water (surface area up to 1100 m2 g−1, and specific pore volumes up to 0.40 m3/g) [36]. Thus, it has found wide applications in water and gas purification, as well as separation processes [37]. Commercial AC is obtainable from non-renewable starting materials, such as lignite, coal, and petroleum coke. Although, there is a strong drive in using renewable materials, such as agricultural wastes (e.g., rice husks, fruit peels, sugarcane bagasse) [38,39]. AC, unfortunately, is not the ideal adsorbent material for treating emerging organic pollutants in water mainly due to the lack of complete removal at low concentrations. Furthermore, the time required for the adsorption is rather slow and the difficulty of regeneration of the used adsorbent. Progress in materials science has resulted in the introduction of new generation of adsorbents with abnormally high surface areas and porosity. These materials include mesoporous silica [40,41], halloysite nanotubes [42,43] graphene [44], molecularly imprinted polymers (MIPs) [45], and MOFs (e.g., MOF-5, HKUST-1, MIL-100, UiO-66, etc.) [46,47]. Significant selectivity can be achieved from the cavity size of the MOFs frameworks. Surface chemical modifications of these adsorbents usually brought about higher removal capacities and selectivity of the composites towards the organic pollutants.

1.2. Photocatalysis

Photocatalysis is a general term used to a defined catalytic reaction that is induced by light energy [48]. Of much interest is the potential of harnessing solar energy. It is an advanced oxidation process for the efficient degradation of toxic pollutants from wastewater using photocatalytic materials. In the process, the light energy is converted into chemical energy with the generation of free radicals, such as hydroxyl radicals, which attack the pollutants and subsequently degrade them into non-toxic by-products [49,50]. Thus, the field has attracted tremendous interest because of its advantageous features as summarized below:
(i)
Ability to degrade pollutants within a short time with the help of light or solar energy.
(ii)
Operates under ambient conditions.
(iii)
Mineralization of organic pollutants into carbon dioxide and water; thus, no secondary pollutants are produced.
An ideal photocatalyst should be stable in both aqueous and organic solvents under acidic or alkaline solutions and be able to tolerate strong light irradiation. Additionally, it must be of high porosity, low-cost, have simplicity in applications, and be easily regenerated. Thus, various porous materials have been discovered. Among them, those containing mesopores and microspores have received much attention due to their uniformity in their surface morphology, particle size, pore volume, and diameters [51]. Some of these materials, such as MOFs, zeolites, silicates, graphene and reduced graphene oxide (GO and RGO), metal-oxide nanoparticles (MNPs), carbon quantum dots (CQDs), and other nanoporous carbon materials, can be chemically modified for the intended application. Of these, MOFs have shown lots of promise.

1.3. Metal-Organic Frameworks

MOFs are advanced porous hybrid materials that are formed from coordination interactions of the metal node with organic linkers (Figure 1) forming two or three-dimensional structures of porous frameworks [52]. They are also referred to as a special group of Coordination polymers (CPs) involving strong metal-ligand interactions [53] and possessed metal-ligand coordinative bonds which are stronger than hydrogen bonds, and they have more directionality than other weak interactions, such as π-π stacking [54]. The development of porous materials can be traced back to 1990 from the work of Hoskin and Robson (1990) for the synthesis of scaffolding-like structural 3D frameworks by linking tetrahedral or octahedral arrays of metals centers with the organic moieties. A diamond-like framework, [N(CH3)4][CuZn(CN)4), having several cavities, was successfully synthesized and analyzed by single-crystal x-ray diffraction [55]. The group of Yaghi (1995) has been instrumental in the design new structures from the assembling of metal ions coordinated to the organic moieties as linkers. In 1999, the famous MOF-5 was successfully synthesized by the group [56], heralding the beginning of the exploration of novel structures of various dimensional frameworks.
Interest in MOFs is due to their peculiarities, uncommon to other synthetic materials, possessing ultra-high surface area, high crystallinity, uniformity of pore sizes, and tunability of volumes. Their microporous structures provide surface area of up to 9000 m2 g−1 and specific pore volumes of up to 2 cm3 g−1, together with a large variety of pore dimensions and topologies. The unique features of MOFs found numerous applications in gas storage, CO2 capture and conversions, chemical separations, drug delivery, nerve agents, sensing, energy conversion, pre-concentrators of explosive vapor, catalysis, wastewater remediations, etc. [58,59].
MOFs possessed open-framework structures that can allow for the inclusion of guest species, particularly solvents during synthesis. These guest species could be removed via desolvation that may result in an empty framework [60]. Therefore, the nature of the framework is determined by the extent to which the volatile solvents are sufficiently removed or exchanged to permit either the generation of a truly porous material or other molecules to occupy the pore structure [61,62]. The MOFs system allows access to open-framework structures with network topologies and connectivity that are not usually observed in classical porous materials [63]. Of much interest is the possibility of generating large-diameter channels and cavities. By controlling the size and functionalization of the organic linkers, well defined MOF structures with high surface areas and tunable pore sizes can be achieved [64,65].
Few reviews were found in the literature highlighting the applications of MOFs for wastewater remediation. Kumar et al. (2018) focused on inorganic contaminants removal using MOFs in the wastewater system [66]. A review by Dhaka et al. (2019) also discussed more on the performance of MOFs for the adsorptive removal of several emerging pollutants [67]. In addition, the performance of MOFs on heavy metals and other inorganic pollutants removal compared to other adsorbents. Joseph et al. (2019) also reviewed the removal of pharmaceuticals drugs in wastewater [68]. However, those reviews have not discussed details on adsorption of various classes of emerging organic and that the photocatalytic degradation of the pollutants was not considered. The present review is aimed at filling the gaps that were not provided by the earlier reports. Thus, a comprehensive update on the adsorptive removal of emerging organic pollutants, using MOFs and their composites are presented. Additionally, the photocatalytic degradation of these pollutants by the MOFs and composites will be discussed. Since the effectiveness of an adsorbent is normally evaluated based on adsorption capacity, selectivity for the specific compound, and regenerability, these relevant data and others are provided in our compilations.

2. MOFs for Remediation of Emerging Pollutants in Water

2.1. MOFs for Adsorption

The possibility to synthesize hundreds of frameworks from various clusters of metal ions with organic linkers gives rise to an unlimited number of crystalline MOFs with microporous or mesoporous structures. Additionally, different functional groups in the organic linkers and metal node serves as adsorption centers for various types of organic contaminants [69].
MOFs also offer selective adsorption of organic molecules due to the functionalities of the organic linkers, possibly forming inclusion complexes with the guest adsorbate molecules. The mode of adsorption interactions is usually through covalent bonding, hydrogen bonding, dative bonding, Van der Waals forces, and π-π interactions [70,71] (Figure 2). Molecular modeling has shown that when the pore sizes of the MOF is bigger than the pollutant molecule, the guest molecule to preferably resides in the pores of MOFs [72]. Alternatively, the guest molecule is adsorbed on the outside if it is bigger than the pores of the MOF. Thus, choosing the MOF for the adsorption of an analyte is important to optimize the adsorption [73,74]. MOFs with promising adsorption properties have been selectively used for the removal of contaminants in water. Their stabilities, adsorption capacities, and ease of reusability have been reported [75].
For the past 10 years, MOFs have received considerable attention as potential adsorbent materials for the removal of pollutants in water. The number of articles that were published from 2010–2020 on the adsorption and photocatalytic degradation by MOFs according to the category of pollutants is shown in Figure 3. It can be readily seen that publications were predominantly on adsorptions compared to photocatalytic degradation. Dyes were also popular topics of research both for adsorption and photocatalytic degradation. This is not surprising as studies on removal and degradation of dyes are easy to be executed using spectrophotometers, and the effects can be seen with the naked eye. On the other hand, studies on pollutants that are not chromogenic, such as the Perfluorooctane sulfonates (PFOS) and Perfluoroalkyl substances (PFAS), will require less readily available instruments, such as High performance liquid chromatography (HPLC)-conductivity or tandem HPLC-MS. Nevertheless, it can be expected that studies using MOFs for other categories of pollutants will grow significantly in the coming years.

2.2. MOFs for Photocatalysis

The idea of using MOFs as photocatalysts were first conceived by Alvaro et al., 2007 [77], when investigating the semiconducting properties of MOF-5. In their pioneering studies, terephthalate organic linker of the MOF, when in solution tends to generate some changes. This is suggested by the fact that electrons are ejected from the excited terephthalate molecule. This finding was the catalyst for investigations on the use of MOFs as photocatalyst for the degradation of different contaminants in water.
Generally, MOFs exhibit semiconductor-like behavior upon light irradiation. The organic linker can act as an antenna to harvest light from the either natural or artificial sources and subsequently activate the metal sites via ligand to metal cluster charge transition (LMCT) [78]. The mechanism can be viewed in terms of excitation of an electron from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) when light is irradiated on the MOF, thus leaving a hole in the HOMO. This hole can interact with OH, forming an OH radical which oxidizes the organic compounds [79]. Thus, the photocatalytic performance of the photoactive MOF involves the generation of electron-hole pairs in the conduction and valence bands of the MOF respectively. In an aqueous medium, the generated electrons (e) interact with oxygen to produce oxygen radicals which in turn transform to hydroxyl free radicals (OH). Similarly, the generated holes (h+) could undergo a reduction upon interactions with the hydroxyl molecules in the solution to form the hydroxyl free radical or act directly or the pollutant. In both cases, the OH and h+ active species could sufficiently attack the target pollutant and subsequently breaks all the bonds in the analyte to ultimately form non-toxic species (CO2 and H2O). Thus, an important criterion in the choice of MOFs for photocatalytic applications is the ability of the MOFs to harvest and channel the light energy.
The high porosity of MOFs contributes extensively to the photocatalytic process by trapping the pollutants. Some MOFs containing Fe, Cr, Zr, and Ti metal ions exhibit good stability in water and can harvest and channel solar energy [80]. They usually possess a small bandgap which enables visible light excitation; hence, they are considered as good candidates for photocatalytic degradations of organic pollutants [81].

2.3. MOF Composites for Adsorption and Photocatalytic Degradation

Even though the fact that MOFs have displayed good potential as adsorbents and photocatalysts for pollutant remediation, some MOFs are plagued by poor chemical and moisture stability and the inability to harness energy from sunlight. To overcome these shortcomings, MOFs have been incorporated with other functional materials, such as metal and metal-oxide nanoparticles (MIL-101(Cr/Al)) [82,83], carbon quantum dots (CQDs/NH2-MIL-125(Ti)) [84], graphene and graphene oxides, zeolite (ZIF-67@MIL-125-NH2) [85], (CNT@MIL-68(Al) [86], molecular imprinted materials e.g., polydopamine (PDA/Fe-MOF/RGO) [87], and ionic liquids, to form composites. These MOF composites were prepared using techniques, such as fabrication, impregnation, surface functionalization, immobilization, and deposition. Some of the methods were able to produce composite MOFs with remarkable properties than the precursor materials. Nevertheless, applications of composites of MOF as photocatalysts are still at the infancy stage. An important target of photocatalytic activities is low bandgaps (<3.0 eV) that allow visible light from the sun to be harnessed.
The MOFs composites usually possessed some synergistic effects, such as the reduction of bandgap, lower photoluminescence, and photocurrent response, to harness light energy and prevent electron-hole recombination. Thus, the composites are highly efficient in utilizing light energy from both visible and ultra-violet regions and higher stability in harsh environments as compared to the counterpart pristine MOFs [88,89].
Similarly, MOF composites with other active materials, such as metal-oxide nanoparticles, carbon quantum dots (CQDs), and graphene oxides (GO), have proven to be effective photocatalysts for the degradation of organic pollutants. This is because the incorporated semi-conductor materials help to facilitate electron transfer in the MOF, resulting into effective separation of the photogenerated electron-hole pairs. On this basis, Wang et al. (2019) [90] proposed on the mechanism of enhanced photocatalytic degradation of rhodamine blue dye using CQDs supported on NH2-MIL-125(Ti) as follows:
CQDs/NH2-MIL-125(Ti) + hv → CQDs/NH2-MIL-125 + e + h+
e + O2 → O2•−
O2 +H2O → HO2 + OH
HO2 + H2O → H2O2 + OH
H2O2 → 2OH
h+ +OH → OH
OH + RhB(dye) → CO2 + H2O
h+ + RhB(dye) → CO2 + H2O
Recently, Li et al. (2019) prepared an interesting heterojunction composite of the MOF (NH2-MIL-53(Fe)) with graphitic carbon nitride doped pyromellitic to form the composite (g-C3N4/PDI@NH2-MIL-53(Fe)) using the facile hydrothermal technique. The composite exhibited photoactive for the removal of tetracycline (90% in 1 h), carbamazepine (78% in 2.5 h), bisphenol A (100% in 10 min), and p-nitrophenol (100% in 30 min). Additionally, the composite MOF was more efficient in terms of reusability (5th cycles for each pollutant) than the pristine NH2-MIL-53(Fe) MOF [91]. Similarly, the synthesis of hybrid MOF/COF composites of NH2- MIL-53(Al), NH2-MIL-125(Ti), and NH2-UiO-66(Zr) with (1,3,5-triazine-2,4,6-triyl)tribenzaldehyde (TTB) and 4,4′,4″-(1,3,5-triazine-2,4,6-triyl)trianiline(TTA) to form N/TTB-TTA (N = NH2-MIL-53(Al), NH2-MIL-125(Ti), and NH2-UiO- 66(Zr)) was reported by the group of He et al. (2019). These hybrids MOFs have shown improved porosity and photocatalytic efficiency for the complete mineralization of methyl orange in aqueous medium [92]. Figure 4 depicted the mechanism of methyl orange and 4-nitrophenol degradation using MOF-199-NH2/BaWO4 composite synthesized from MOF-199-NH2 and BaWO4 by the immobilization technique [93]. It is interesting to note that the immobilization of BaWO4 into the MOF-199-NH2 has caused a red-shift in the absorption maximum of the composites with lower optical property than the pristine MOF. In addition, the calculated bandgap of the composite is lower (3.0 eV) compared to the MOF-199-NH2 (3.2 eV) (Figure 4). Thus, complete degradation within 50 and 80 min were achieved using the MOF-199-NH2/BaWO4 composite for methyl orange and 4-nitrophenol, respectively.

3. MOFs and Composites for Adsorption and Photocatalytic Degradation of Emerging Pollutants in Water

3.1. MOFs and Composites for Adsorption and Photocatalytic Degradation of Dyes

Globally, water contamination from dyes has been one of the biggest sources of environmental pollution. Despite various regulations on the use of dyes, the discharge of effluents containing dyes, particularly from small-scale textile, cosmetics, leather, and food industries, has been a major source of water pollution. These dyes, when discharge into the environmental water, usually cause significant ecological threats, such as destruction of aquatic life, impeding plant growth, and posing various forms of toxicity to humans, including genotoxicity, reproductive toxicity, neurotoxicity, and other forms of diseases [21]. Thus, concerted efforts are needed to address the problem at the source and to remediate the already polluted water to safe levels. Figure 5 depicted the trends in publications on adsorption and photocatalytic degradations of dyes for the last decade. Exponential growth in the number of publications has been observed each year for both adsorptions and photocatalytic degradations. For instance, in 2020 alone, 2131 and 834 the number of articles has been reported on the adsorption and photocatalytic degradations of dyes, respectively, according to the data obtained from science direct repository.
The significant porosity of MOFs due to the number of empty spaces within the frameworks rendered them a suitable candidate for dye adsorption [94]. The MOFs can provide larger adsorption sites for various kinds of dye molecules, including both cationic and anionic [95,96]. The simultaneous adsorption and photocatalytic degradation of methyl orange (Figure 6) using Co- and Zn-based MOFs, (M(tpbpc)(bdc)0.5·H2O) was reported by Liu et al. (2017), with complete mineralization of the dye achieved at 90 min of irradiations [97].
Table 1 summarizes some of the properties of MOFs as adsorbents for the removal of dyes from water. Some of the MOFs exhibited abnormally high surface area (up to 3500 m2 g−1). More so, they have shown higher adsorption capacities than other conventional adsorbents. For example, UiO-67(Zr) was able to achieve an equilibrium adsorption capacity of 799 mg g−1, for Congo red adsorption [98]. Adsorption capacity with (qe) value of 1045 mg g−1 was achieved for the adsorption of methylene blue by MIL-100(Fe) [99]. It is heartening to note that some of the MOFs were able to achieve almost or complete removal of the dyes within a relatively shorter time than the other adsorbents, which take several days to achieve complete removal. Many authors did not report the regeneration of their adsorbents; nevertheless, some of these MOFs could be reused a number of times without significant reduction in their efficiencies.
With the discovery of the photocatalytic properties of the MOF-5 in 2007, researchers continue exploring the photocatalytic efficiencies of other classes of MOFs for the degradations of contaminants from wastewater, of which dyes received considerable attention. The photocatalytic degradation offers an interesting option to completely breakdown the persistent dyes into neutral species. Some of the MOFs reported were able to degrade the contaminants under sunlight irradiations due to their lower band-gap, higher surface area, and pore volume, as well as good stability in aqueous medium. However, a major shortcoming encountered was the inability of some MOFs to be activated under visible light irradiations. Similarly, some of the MOFs were unstable in an aqueous medium. As such modifications using functionalized materials were considered [128]. Thus, various MOF composites, such as bi-metallic MOFs [129], NPs@MOFs [130], CQDs@MOFs, etc., with different active species were found to be more effective than the corresponding pristine MOFs, particularly in terms of harvesting visible light, preventing electron-hole recombination, and reusability.
Some MOFs and their composites reported for the photocatalytic degradation of dyes are summarized in Table 2. It can be seen that some of the pristine MOFs possessed high bandgaps (>3.0 eV); thus, they cannot utilize visible light effectively for photodegradation to occur, particularly under the sunlight irradiations. However, it is worthy to note that, the higher surface area of the MOFs might result in their higher absorption profile which can be extended to the visible region. Thus, they can absorb a few photons of visible light energy, capable to generate some holes on the surface of the MOFs to form free radicals that can act on the dyes. The functionalization of the organic linker in the MOFs was also responsible for the photocatalytic degradation. As an example, the presence of NH2 in NH2-MIL-88(Fe) has been claimed as the contributing factor to the adsorption capacity due to the shift in the absorption maximum of the MOF [50]. It is interesting to note that, modifications of the MOFs with light active species, such as metals, metal oxides, sulfides, etc., resulted in MOF composites with much lower bandgaps than the pristine MOFs or the active materials themselves [131]. Of all the MOFs reported in Table 2, only 16% were able to achieve the bandgap of less than 3.0 eV. This underscores the need for new materials with reduced bandgaps to tap sunlight irradiation for their degradation.

3.2. MOFs and Composites for Adsorptive Removal and Photocatalytic Degradation of Phenols and Other Miscellaneous Emerging Pollutants

Phenolic compounds are widely used by chemical and allied industries in making useful products, such as petrochemicals and plastics. Phenols and its derivatives are also used as a precursor in chemical industries in the production of pharmaceuticals, dyes, herbicides, pesticides, detergents, epoxies, among others. It has been estimated that more than 10 million tons of phenolic compounds are discharged annually into the environment, thus polluting the soil, surface water, and underground water [164]. The presence of these toxic endocrine-disrupting compounds, such as phenol, bisphenol A, 2,4-dinitrophenol, and 2,3,4,5-tetrachlorophenol, in the wastewater poses negative effects to living organisms, threatening the harmony of ecosystems [165]. The United States Environmental Protection Agency (USEPA) stipulates the threshold level of phenolic effluents to be discharged into public sewage systems should not exceed 5 ppm, and the maximum permissible limit in potable drinking water should not exceed 1 ppb [166].
Modern agricultural practice requires the use of agrochemicals, such as pesticides and herbicides, that help to protect farm products from pests, controlling unwanted weeds, as well as boosting the yield of crops. Herbicides are chemicals that are primarily produced to inhibit weeds that compete with the plant’s growth, while insecticides are aimed at repelling or mitigating insects and other pests from attacking the agricultural products, such as fruits, vegetables, cotton, etc. Commonly used agrochemicals are the neonicotinoids (e.g., thiamethoxam, imidacloprid, acetamiprid, nitenpyram, dinotefuran, clothianidin, and thiacloprid), organophosphates (e.g., diazinon, parathion, methyl parathion, paraoxon, and fenitrothion) and carbamates (e.g., aldicarb, carbaryl, and methomyl). When applied, these chemicals accumulate in the soil and subsequently washed into the environmental waters, such as lakes, lagoon river, and groundwater, posing potential hazards to the ecosystem [167]. Glyphosate, the most widely used herbicide in the USA, has been listed as a likely human carcinogenic agrochemical by the World Health Organization [168]. Similarly, atrazine also has been reported to show endocrine-disrupting property to aquatic animals even at low concentrations [169].
Other emerging pollutants of high toxic effects in water are the polycyclic aromatic hydrocarbons (PAHs). They are a group of hydrophobic compounds with two or more benzene rings. PAHs are known to originate extensively from anthropogenic sources, particularly from crude oil exploration, petrochemical effluents, oil spillage, etc. [170,171]. Due to their lipophilic nature, they are prone to be accumulated in the fatty tissues of living organisms. Long-term exposure to PAHs results in eye irritations, nausea, vomiting, and, in severe cases, may lead to liver and kidney failure and lung cancer [172,173]. Hence, they are categorized as emerging contaminants by the European Union, the USEPA, and other environmental regulatory bodies [174].
Another group of highly recalcitrant emerging pollutants that have currently gained world-wide attention are the poly and perfluorinated alkyl substances (PFAS). PFAS made headlines because they were found in the drinking water across many cities in the US and other countries of the world. Removing them is so difficult that scientists have nicknamed them “forever chemicals.”
PFAS are fluorinated chemicals that have been widely used for the production of industrial (e.g., surfactants) and consumer products (e.g., non-stick coatings). The most toxic of these groups are the perfluoroalkyl carboxylates (PFCAs) and perfluoroalkyl sulfonates (PFAS). Perfluorooctanoic acid (PFOA) tends to bioaccumulate in human tissues and possessed a half-life of 4 years [175]. PFOA and PFOS are highly water-soluble; thus, they are readily transported in the aquatic environment. These compounds are detected in surface water [176], groundwater [177], rainwater [178], wastewater [179], and drinking water [180]. They have also been detected in a number of food matrices [181], human serum, breast milk, and other biological samples [182]. The USEPA has recommended clean-up of underground water that is contaminated with 70 parts per trillion of PFOA and PFOS [183]. The recommendation, however, is applied to groundwater that is a current or potential source of drinking water. The structures of PFOA and PFOS are shown in Figure 7. Typical of perfluoro compounds, it is the high-energy C–F bonds that render them persistent in the environment.
The toxicological impacts of these emerging pollutants have motivated researchers to look for green and environmentally sustainable methods for their remediations. Some water-insoluble MOFs and their composites offer good removal and photo-active degradations of herbicides and pesticides from wastewater. As an example, rapid (20–60 min) and complete removal (99%) of glyphosate were achieved using the highly porous zirconium MOFs NU-100(Zr) and UiO-67(Zr) [184]. Similarly, high removal of bisphenol A (473 mg/g) was achieved in 30 min using MIL-53(Al)-F127 composite MOF [185].
Studies by Apkinar et al. [186] exemplify the synthetic tunability of MOFs on the role of chemical functionality in the adsorptive removal of pollutants from water. The team investigated the adsorption in several Zr-based MOFs with a variety of pore sizes and with increasingly large conjugated π- systems and framework topologies. The unusually fast equilibration adsorption of 1 min exhibited by NU-1000 is due to the rapid diffusion through the hierarchically porous MOF structure although its capacity is comparable to that of other adsorbents that have been used for atrazine adsorption. The studies further corroborated that the presence of linkers with extended π-systems, rather than large pores results in the exceptional atrazine uptake by NU-1000. The applications of some of the MOFs and their composites as adsorbents and photocatalysts for the remediations of these water pollutants are summarized in Table 3 and Table 4.
Recently, we reported the adsorptions of PAHs in aqueous medium using the highly porous Zr-based UiO (UiO-66(Zr), NH2-UiO-66(Zr)) [187], and MILs (MIL-88(Fe) and NH2-MIL-88(Fe) [188,189]. In most cases, rapid adsorption of the pollutants was achieved within a short time (30 min), which were attributed to the availability of the active adsorption sites in the MOFs. Molecular docking simulation was used to study the fundamental interactions between the MOFs with chrysene as a PAH model compound (Figure 8). The binding interaction studies show that the chrysene preferably resides in the inner and outer pores UiO-66(Zr) and NH2-UiO-66(Zr), respectively. The preference has resulted from the pore diameters of the MOFs concerning the molecular size of the pollutant [187].
Very limited reports can be found on the use of MOFs for the adsorption of the perfluoro compounds (Table 3). Jun et al. (2019) investigated the competitive adsorption of three adsorbates (i.e., bisphenol A, 17α-ethynyl estradiol, and PFOA) using Al-MOF. The effects of various water chemistry parameters, such as solution temperature, pH, background ions, and natural organic matter (i.e., humic acid), were also studied. The authors concluded that the synergetic effects of hydrophobic and electrostatic interactions were important factors in the adsorption process. Three MOFs, zeolitic imidazolate framework-7 (ZIF-7), ZIF-8, and ZIF-L were investigated for the adsorption of PFOA in an aqueous solution by Chen et al. (2016). The PFOA sorption performance of ZIF-7, ZIF-8, and ZIF-L was then compared with the performance of two commercialized adsorbents, zeolite 13X and activated carbon. ZIF-8 and ZIF-L were shown to outperform the two commercial sorbents. Their work demonstrates that the crystal structure and the surface functionality of MOFs influence, PFOA adsorption performance. To date, there is yet to be found reports on the photocatalytic degradation of perfluoro compounds using MOFs and composites.
Some articles published for photocatalytic degradations of phenols, pesticides, herbicides, and PAHs using MOFs and their composites are found in Table 4. According to a report by Mei et al., 2019, complete mineralization of thiamethoxam was achieved within 60 min of visible light irradiation in the presence of MIL-53(Fe) [190]. Before that, Ahmad et al. (2018) decorated MIL-100(Fe) with for ZnO nanosphere for the degradation of phenol, bisphenol A and atrazine. The introduction of the ZnO into the MOF has boosted its optical property; hence the composite was able to absorb visible light. More than 90% of the pollutants were degraded within 120 min [191]. Recently, photocatalytic degradation of bisphenol A was reported using MOF@COF hybrid composites of Fe-MIL-101-NH2@TPMA and Zr-UiO-66-NH2@TPMA. The synergetic effect of the persulfate (PS) added to the medium coupled with the optical properties of the composites was able to degrade 99% of the pollutant within 240 min under visible light irradiation [192]. To date, MOF has not been reported for the photocatalytic degradations of PAHs and PFASs. A difficulty in the detection of PFASs has been considered as a challenging factor, as it requires sophisticated tandem mass spectrometry.
Table 3. MOFs and composites used for the adsorptions of phenols, herbicides, pesticides, and other miscellaneous organic pollutants.
Table 3. MOFs and composites used for the adsorptions of phenols, herbicides, pesticides, and other miscellaneous organic pollutants.
Type of MOFSynthesis MethodSurface Area (m2 g−1)PollutantsConcentration (mg L−1)% RemovalQe (mg g−1)Equilibrium TimeReusedRef
Phenolics
MIL-53(Al)
MIL-53(Al)-F127
Hydrothermal931
1008
Bisphenol A250-329
473
90 min
30 min
3
3
[185]
MIL-68(Al)/PVDFCasting-P-nitrophenol1094126720 min6[125]
HKUST-1(Cu)Microwave-P-nitrophenol200 400 30 min-[193]
SiO2@MIL-68(Al)Solvothermal1156Aniline3000-53240 s5[194]
[Zn(ATA)(BPD)]
MOF-VII
Ultrasound170
675
2,4-dichlorophenol6068
91
-
-
90 min
90 min
5[195]
[Zn(TDC) MOFVapor-diffusion2352, 4-dichloropheno6095-180 min-[196]
MIL-68(Al)
CNT@MIL-68(Al)
Solvothermal1283
1407
Phenol
Phenol
1000-
-
118
257
120 min5[86]
NH2-UiO-66(Zr)Solvothermal 2,4,6-trinitrophenol
Styphnic acid
2,4-dinitrotoluene
100-23
24
0.5
2
36 h-[197]
MIL–68(Al)
MIL–68(Al)/GO
Solvothermal 550
762
p–nitrophenol300-
-
271
332
17 h
17 h
5[198]
NH2-MIL-88(Fe)Hydrothermal4142,4,6-trinitrophenol35-16440 min5[199]
MOF-199(Cu)Solvothermal2271Phenol
p-nitro phenol
50080
89
58
68
300 min
30 min
-
-
[200]
Al-MOF/SA-CSHydrothermal688Bisphenol A50-13718 h6[201]
Cu-BDC MOF
Cu-BDC@GrO
Cu-BDC@CNT
Solvothermal-
-
-
Bisphenol A
Bisphenol A
Bisphenol A
2009760
182
164
40 min5[202]
laccase@HKUST-1Immobilization-Bisphenol A20074-4 hNA[203]
Pesticides
M-MOFRoom temperature250Thiamethoxam
Acetamiprid
Nitenpyram
Dinotefuran
Clothianidin
Thiacloprid
100-3
3
3
3
2
3
60 min-[204]
MIL-101(Cr)Hydrothermal2612Diazinon5054158 45 min4[205]
Cr-MIL-101-BTPHydrothermal1113Acetochlor120100322200 min6[206]
MIL-101(Cr)
TS-MIL-101(Cr)
Hydrothermal-Atrazine3037
69
-60 min-[207]
Herbicides
HKUST-1(Cu)
ZrO2@HKUST-1
Room temperature 1484
1152
Cyhalothrin60-140
138
2 h-[208]
UiO-67(Zr)Hydrothermal2172Glyphosate Glufosinate20096
92
537
360
150 min
200 min
-[209]
NU-100(Zr)
UiO-67(Zr)
SolvothermalN/A
N/A
Glyphosate1117.5100
100
1340
1500
20 min
60 min
-
-
[184]
UiO-66(Zr)
UiO-67(Zr)
Solvothermal1640
2345
Atrazine2520
98
3
12
50 min
2 min
1
4
[210]
DUT-52(Zr)
NU-1008(Zr)
NU-901(Zr)
NU-1000(Zr)
Solvothermal1960
1400
2110
2110
Atrazine1082
69
85
93
-1 min3[186]
PAHs
Zn-BDC MOF
Cu-BDC MOF
Mechanical
Mechanical
-Naphthalene
Anthracene
Naphthalene
Anthracene
10088
50
84
52
87
52
84
52
210 min
120 min
210 min
120 min
3[211]
UiO-66(Zr)
NH2-UiO-66(Zr)
Solvothermal1420
985
Anthracene
Chrysene
Anthracene
Chrysene
499
96
98
96
24
22
24
19
25 min
25 min
30 min
30 min
5
5
[187]
MIL-88(Fe)
NH2-MIL-88(Fe)
Microwave1240
941
Pyrene
Pyrene
499
96
24
23
40 min5[212]
MIL-88(Fe)
NH2-MIL-88(Fe)
Microwave1240
941
Chrysene
Chrysene
499
95
24
22
25 min5[188]
MIL-88(Fe)
NH2-MIL-88(Fe)
Mixed-MIL-88(Fe)
Microwave1240
941
1025
Anthracene
Anthracene
Anthracene
498
92
96
24
21
23
25 min-[189]
PFCAs
ZIF-7
ZIF-8
ZIF-L
Room temperature14
1291
12
Perfluorooctanoic acid250 40
45
97
26
214
295
60 min-[213]
Basolite A-100Commercial630Perfluorooctanoic acid1100169 4
Table 4. MOFs and composites reported for the photocatalytic degradations of phenols, herbicides, pesticides, and other miscellaneous organic pollutants.
Table 4. MOFs and composites reported for the photocatalytic degradations of phenols, herbicides, pesticides, and other miscellaneous organic pollutants.
MOFSynthesis MethodSurface Area (m2 g−1)Bandgap (eV)PollutantsConcentration (mg L−1)Light Source(%) RemovalIrradiation TimeReusedRef
Phenolics
NH2-MIL-125 (Ti)@Bi2MSolvothermal881.89Dichlorophen10 Visible93180 min-[214]
[CoNi(m3-tp)2(m2-pyz)2]
MOF/CuWO4
Hydrothermal1054
801
2.5
2.4
4-nitrophenol10 Visible24
81
105 min6 [152]
MIL-88B(Fe)
CNT@MIL-88B(Fe)
Hydrothermal
Hydrothermal
118-Phenol25 55
100
30 min
10 min
3[215]
CdS@NH2-MIL-125(Ti)Solvothermal13752.36Phenol180 Visible-120 min5 [147]
HOQ@MOF-5(Zn)Room temperature-3.12Phenol1 Visible10070 min5[216]
MIL-100(Fe)@ZnOSolvothermal6542.63Phenol,
Bisphenol A
5 Visible95
84
120 min5[191]
MIL-101-NH2@TpMA
UiO-66-NH2@TpMA
Hydrothermal
Hydrothermal
129
531
2.12
2.01
Bisphenol A50 Visible99
82
240 min
240 min
5
5
[192]
MIL-88(Fe)/PS/UVMicrowave-1.78Bisphenol A10 Visible10030 min3 [217]
MIL-101(Fe)
Pd@MIL-100(Fe)
Hydrothermal2006
2102
-Bisphenol A20 Visible47
68
240 min4[218]
Cu-hemin-MOFs/BNRoom temperature--Bisphenol A40 Visible9930 min4 [219]
laccase@HKUST-1(Cu)Immobilization--Bisphenol A200 Visible1004 h10 [203]
AQS-NH-MIL-101(Fe)Solvothermal--Bisphenol A60 Visible98180 min3 [220]
Pesticides
UiO- 66@WGSolvothermal3802.3Malathion20 Visible8370 min4 [221]
AgIO3/MIL-53(Fe)Room temperature2082.43Malathion
Chlorpyrifos
20 Solar93
98
120 min-[222]
Fe3O4@MOF-2Room temperature--Diazinon30 Visible9960 min15 [223]
MIL-53(Fe)Solvothermal6682.89Thiamethoxam5 Visible9660 min-[190]
HKUST-1(Cu)
ZrO2@HKUST-1(Cu)
Room temperature
Solvothermal
1484
1152
3.87
2.27
Cyhalothrin60Visible34
100
6 h4[208]
Herbicides
MIL-100(Fe)@ZnOSolvothermal6542.63Atrazine5 Visible79120 min5[191]
TiO2@NH2-MIL-101(Cr)Solvothermal--Atrazine30 Visible4560 min-[84]
It has long been recognized that the catalytic activity of enzymes can be extended by immobilizing onto solid supports, such as polymers and inorganic materials. The superior performance of MOF HKUST-1 for the encapsulation of the enzyme laccase to enhance its catalytic activity, stability, and reusability compared with other conventional polymers or inorganic carriers was demonstrated by Zhang et al. (2020). The MOF not only acted as protective layer against high temperatures, continuous operation, and long-term storage but also could enhance the accessibility of active site of laccase due to its flower-like structure and high exposed surface area. The laccase@HKUST-1 still maintained 75.9% of its original degradation efficiency after 10 cycles, suggesting the effectiveness of the MOF to act as a protective layer to protect the laccase against the possible industrial environment. Unfortunately, the rapid breakdown of bisphenol using this composite material did not materialize (4 h).

3.3. MOFs and Composites for Adsorption and Photocatalytic Degradation of Pharmaceutical and Personal Care Products (PPCPs)

PPCPs are produced and used worldwide primarily for the remediation of ailments, as supplements, and as body care. These chemicals are usually discharged as wastewater from the manufacturing industries, hospitals, landfill leachates into the environment, either in their native form or as metabolites. The fundamental pathway for the release of these contaminants is through excretions. Thus, municipal wastewater is the major route bringing human pharmaceuticals into the environment. Of the various class of pharmaceuticals, antibiotics, such as penicillin, amoxicillin, tetracyclines, sulfonamides, etc., are found to be persistent in water due to their resistance to biological treatments from wastewater treatments plants. They usually remained untreated in the municipal wastewater for a long time; hence, they pose toxic effects even at low concentrations (ng L−1). Although the concentration of these pharmaceutical residues in the environment is low, its uninterrupted input to the environment may result in the long-term risk for terrestrial and aquatic organisms. In human beings, these pollutants may cause mutations in the genomic texture by disrupting the endocrine glands; hence, they are classified as endocrine disruptors.
The applications of MOFs as adsorbents, as well as photocatalysts, for the remediation of PPCPs have been reported (Table 5). Many MOFs were proven to be efficient for the adsorption of these pollutants within short time with high removal capacities. Similarly, the use of pristine MOFs and their corresponding functionalized derivatives and composites have been studied. MOFs composites have demonstrated better photocatalytic activities than the pristine MOFs. Some of these MOFs have also displayed good reusability which could be employed for industrial and large-scale applications. Figure 9 illustrates the versatility of MOFs, such as UiO-66(Zr), MOF-88(Fe), and MOF-808(Fe), for the removal of some common pharmaceuticals [224].
Photocatalysts of high porosity, ordered crystallinity, visible light harvesting capabilities and mechanical stability are desirable for the complete mineralization of the pharmaceutical drugs. The presence of the metallic node and organic linker can enhance the utilization of the solar energy through HOMO and LUMO interactions. The interactions generate the photon energy that are responsible for the excites the electrons from the contaminants to produce the active species of H+ and OH that mineralize the organic species. Figure 10 illustrates the mechanism for the photocatalytic degradation of ibuprofen using MIL-88(Fe) and corresponding composites, Ag/AgCl@MIL-88(Fe). The incorporation of AgCl into the framework or the MIL-88(Fe) MOF caused reduction in the bandgap (2.51 eV) of the MOF, which improved the photocatalytic capability of the MOF [251]. The applications of MOFs and their composites for the photocatalytic degradation of pharmaceutical drugs is highlighted in Table 6. In most cases, several hours are required for the complete mineralization of the pharmaceuticals.

4. Patent Search

The diversity in MOFs and their versatile functionalities has prompted researchers to explore their potentialities in synthesis and applications. Thus, number of literatures has been written and patented on the synthesis and applications of MOFs and their composites. The advancement in the synthesis and characterizations of MOFs and frontier applications in adsorption and photocatalytic degradation. The area of research remains active among community of scientists and engineers. Thus, the number of published articles for MOFs application in wastewater remediations have been well patented. A search using the website lens.org reveals that most patents were granted for the past 10 years on adsorption using MOFs were on dyes, followed by phenols, PPCPs, and then pesticides and herbicide. Similarly, with the photocatalytic degradation (Figure 11a). Patents granted for the adsorption and photocatalytic degradation of dyes using MOFs-based materials are shown in Figure 11b. The growth was exponential until 2016, with a gradual decrease from then on. The reason for the decreased in patenting could be due to the discovery of a large number of promising MOFs for various laboratory and pilot-scale wastewater applications.

5. Conclusions

The motivation for the development of improved technologies for the remediation of waters is driven by the frequent occurrence of emerging pollutants in drinking water. This is because the conventional wastewater treatment facilities are ill-equipped for the complete removal of these pollutants in water. Adsorption using conventional adsorbents, despite being the gold standard in water treatment technology, is not suited for the task. MOFs and/or their composites, on the other hand, have shown very encouraging results not only as super adsorbents but also as super photocatalysts. The extreme porosity and large interior surface area of MOFs offer unique prospects for adsorption and photocatalysis. Unlike conventional adsorbents which rely to a large extent on the unspecific van der Waals force, the simultaneous use of various interactions, such as cationic, π–π stacking, hydrogen-bonding, and Van der Waals interactions, has been associated with MOFs adsorption. MOFs can also offer more selectivity to the organic pollutants than other conventional adsorbents due to the orientation of their frameworks. They provide large number of pores with uniform sizes. The ‘breathing effects’ of MOFs cavity allow for the adsorption of larger molecules of pollutant from wastewater. For photocatalytic application, their visible light adsorption capacity and moderate bandgap has been commended. To a larger extent, composites of MOFs offer great advantage than their pristine forms due their multiple functionalities. Thus, MOFs have proven to be promising materials for adsorption and photocatalytic degradation of different classes of organic pollutants.
A few start-up companies which are predominantly spin-offs from university laboratories and the German chemical company BASF have started commercializing several kinds of MOFs, mainly for applications as gas storage and adsorption of toxic gases. MOFs, such as MOF-5, MIL-53, HKUST-1, ZIF-90, and UIO-66, can be obtained from the open market. It must be pointed out that most evaluations cited in this article were conducted under normal laboratory conditions. The actual performance of the MOFs in real water samples with complex matrices, such as wastewater and under industrial-scale operations, are virtually unknown. For commercial exploitation, it would perhaps be easier for these adsorbents materials to be applied as super filters in the household water purification system due to the smaller amounts of adsorbents/photocatalysts required. For large-scale productions, such as wastewater treatment facilities, the cost will be a primary factor on the commercial exploitation of these materials. However, if savings from mass production and reusability are factored, it might be cost-effective on the long run. The use of cheaper metals (e.g., potassium, sodium) and, at the same time, not compromising the qualities of the MOFs will be the way forward. Photocatalysts can able to harness direct sunlight and significantly reduce the degradation time are much welcome. Other major challenges that must be overcome are the often complicated and lengthy synthesis processes, poor long-term physicochemical stability of the MOFs, and the limited prospects for reuse. Typical of any new materials, long term safety issues, such as the liberation of chemicals and metals from the degradation of MOFs, as well as risks to exposure to trapped organic solvents (e.g., chloroform, acetone, dimethylformamide), are virtually unknown.
The application of MOFs for industrial wastewater treatments have been established. The major form for the adsorbents and photocatalysts desired includes pellets, spherical, mold, nanorods, beads, etc. Thus, the use of MOFs composites has demonstrated many advantages, particularly in photocatalysis, where low bandgap is required. The requirements include high surface area and small pore diameters with distinct pore structures to enable faster transport of the MOFs in the aqueous phase. Along with that, thermal stability, abrasion, and moisture resistance are prerequisites to the industrial application of the MOFs.
Thus, adsorption with simultaneous photocatalytic degradation under sunlight irradiation is certainly a novel idea as it offers a complete solution to the problem of removal of pollutants from wastewater and their safe remediation into environmentally benign species. MOFs and their composites seem destined to play these roles.

Author Contributions

Z.U.Z.; writing—review and editing, K.J.; validation, N.S.S.; Original data draft preparation, A.R.; Conceptualization, formal analysis, N.H.H.A.B.; resources, B.S.; Supervision, M.N.H.R.; Formal analysis, H.A.I.; Investigation, A.H.J.; project administration, O.A.; Funding acquisition, A.S.; Visualization. All authors have read and agreed to the published version of the manuscript.

Funding

We would like to acknowledge grants received from Universiti Teknologi PETRONAS under YUTP and UTP-UIR scheme with cost center 015LCO-211 and 015MEO-166, respectively.

Acknowledgments

The authors also wish to acknowledge King Khalid University, Abha, Kingdom of Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest

Abbreviations

BaWO4Barium tungstate
BETBrunner Emmett Teller
COFCovalent organic framework
CNTsCarbon nanotubes
CPsCoordination polymers
CQDsCarbon quantum dots
EDCsEndocrine disrupting compounds
GOGraphene oxide
HKUSTHongkong University of Science and Technology
HOMOHighest occupied molecular orbital
HPLCHigh performance liquid chromatography
LMCTligand to metal cluster charge transition
LUMOLowest occupied molecular orbital
MILMaterial institute Lavoisier
MIPsMolecularly impregnated polymers
MNPsMetal-oxide nanoparticles
MOFsMetal-organic frameworks
PAHsPolycyclic aromatic hydrocarbons
PDIPyromellitic diimide
PFASPerfluoroalkyl substances
PFCsPerfluorinated compounds
PFCAsPerfluoro carboxylic acids
PFOAPerfluorooctanoic acid
PFOSPerfluorooctane sulfonates
PPCPsPharmaceutical and Personal Care Products
RGOReduced graphene oxide
UiOUniversiti i Oslo
USEPAUnited-states environmental protection agency
ZIFsZeolite imidazole framework

References

  1. Chen, Y.; Zhai, B.; Liang, Y. Enhanced degradation performance of organic dyes removal by semiconductor/MOF/graphene oxide composites under visible light irradiation. Diam. Relat. Mater. 2019, 98. [Google Scholar] [CrossRef]
  2. Zango, Z.U.; Shehu Imam, S. Evaluation of Microcrystalline Cellulose from Groundnut Shell for the Removal of Crystal Violet and Methylene Blue. Nanosci. Nanotechnol. 2018, 8, 1–6. [Google Scholar] [CrossRef]
  3. Peng, Y.; Zhang, Y.; Huang, H.; Zhong, C. Flexibility induced high-performance MOF-based adsorbent for nitroimidazole antibiotics capture. Chem. Eng. J. 2018, 333, 678–685. [Google Scholar] [CrossRef]
  4. Seo, P.W.; Bhadra, B.N.; Ahmed, I.; Khan, N.A.; Jhung, S.H. Adsorptive Removal of Pharmaceuticals and Personal Care Products from Water with Functionalized Metal-organic Frameworks: Remarkable Adsorbents with Hydrogen-bonding Abilities. Sci. Rep. 2016, 6, 34462. [Google Scholar] [CrossRef] [Green Version]
  5. Zhao, L.; Wu, Q.; Ma, A. Biodegradation of Phenolic Contaminants: Current Status and Perspectives. IOP Conf. Ser. Earth Environ. Sci. 2018, 111. [Google Scholar] [CrossRef]
  6. Derylo-Marczewska, A.; Blachnio, M.; Marczewski, A.W.; Seczkowska, M.; Tarasiuk, B. Phenoxyacid pesticide adsorption on activated carbon – Equilibrium and kinetics. Chemosphere 2019, 214, 349–360. [Google Scholar] [CrossRef]
  7. Mojiri, A.; Zhou, J.L.; Robinson, B.; Ohashi, A.; Ozaki, N.; Kindaichi, T.; Farraji, H.; Vakili, M. Pesticides in aquatic environments and their removal by adsorption methods. Chemosphere 2020, 253, 126646. [Google Scholar] [CrossRef]
  8. Zango, Z.U.; Sambudi, N.S.; Jumbri, K.; Ramli, A.; Hana, N.; Abu, H.; Saad, B.; Nur, M.; Rozaini, H.; Isiyaka, H.A.; et al. An Overview and Evaluation of Highly Porous Adsorbent Materials for Polycyclic Aromatic Hydrocarbons and Phenols Removal from Wastewater. Water 2020, 12, 2921. [Google Scholar] [CrossRef]
  9. Alegbeleye, O.O.; Opeolu, B.O.; Jackson, V.A. Polycyclic Aromatic Hydrocarbons: A Critical Review of Environmental Occurrence and Bioremediation. Environ. Manag. 2017, 60, 758–783. [Google Scholar] [CrossRef]
  10. Arvaniti, O.S.; Hwang, Y.; Andersen, H.R.; Stasinakis, A.S.; Thomaidis, N.S.; Aloupi, M. Reductive degradation of perfluorinated compounds in water using Mg-aminoclay coated nanoscale zero valent iron. Chem. Eng. J. 2015, 262, 133–139. [Google Scholar] [CrossRef] [Green Version]
  11. Lath, S.; Navarro, D.A.; Losic, D.; Kumar, A.; Mclaughlin, M.J. Sorptive remediation of perfluorooctanoic acid (PFOA) using mixed mineral and graphene/carbon-based materials. Environ. Chem. 2018, 15, 472–480. [Google Scholar] [CrossRef] [Green Version]
  12. Jun, B.M.; Hwang, H.S.; Heo, J.; Han, J.; Jang, M.; Sohn, J.; Park, C.M.; Yoon, Y. Removal of selected endocrine-disrupting compounds using Al-based metal organic framework: Performance and mechanism of competitive adsorption. J. Ind. Eng. Chem. 2019, 79, 345–352. [Google Scholar] [CrossRef]
  13. Canle, M.; Fernández Pérez, M.I.; Santaballa, J.A. Photocatalyzed degradation/abatement of endocrine disruptors. Curr. Opin. Green Sustain. Chem. 2017, 6, 101–138. [Google Scholar] [CrossRef]
  14. Imam, S.S.; Zango, Z.U. Magnetic Nanoparticle (Fe3O4) Impregnated onto Coconut Shell Activated Carbon for the Removal of Ni (II) from Aqueous Solution. Int. J. Res. Chem. Environ. 2018, 8, 9–15. [Google Scholar]
  15. Hsieh, H.Y.; Huang, K.C.; Cheng, J.O.; Lo, W.T.; Meng, P.J.; Ko, F.C. Environmental effects on the bioaccumulation of PAHs in marine zooplankton in Gaoping coastal waters, Taiwan: Concentration, distribution, profile, and sources. Mar. Pollut. Bull. 2019, 144, 68–78. [Google Scholar] [CrossRef] [PubMed]
  16. Yali, Z.P.; Jadid, A.P.; Samin, L.A. Modeling of retention time for polychlorinated biphenyl congeners in human adipose tissue using quantitative structure–retention relationship methodology. Int. J. Environ. Sci. Technol. 2017, 14, 2357–2366. [Google Scholar] [CrossRef]
  17. Ho, Y.C.; Norli, I.; Alkarkhi, A.F.M.; Morad, N. Characterization of biopolymeric flocculant (pectin) and organic synthetic flocculant (PAM): A comparative study on treatment and optimization in kaolin suspension. Bioresour. Technol. 2010, 101, 1166–1174. [Google Scholar] [CrossRef]
  18. Ho, Y.C. New Vegetal Biopolymeric Flocculant: A Degradation and Flocculation Study. Iran. J. Energy Environ. 2014, 5, 2–3. [Google Scholar] [CrossRef] [Green Version]
  19. Rosińska, A.; Dabrowska, L. Selection of coagulants for the removal of chosen PAH from drinking water. Water 2018, 10, 886. [Google Scholar] [CrossRef] [Green Version]
  20. Hussaini Jagaba, A. Wastewater Treatment Using Alum, the Combinations of Alum-Ferric Chloride, Alum-Chitosan, Alum-Zeolite and Alum-Moringa Oleifera as Adsorbent and Coagulant. Int. J. Eng. Manag. 2018, 2, 67. [Google Scholar] [CrossRef] [Green Version]
  21. Pavithra, K.G.; Kumar, P.S.; Jaikumar, V.; Rajan, P.S. Removal of colorants from wastewater: A review on sources and treatment strategies. J. Ind. Eng. Chem. 2019, 75, 1–19. [Google Scholar] [CrossRef]
  22. Guo, D.; Wang, H.; Fu, P.; Huang, Y.; Liu, Y.; Lv, W.; Wang, F. Diatomite precoat filtration for wastewater treatment: Filtration performance and pollution mechanisms. Chem. Eng. Res. Des. 2018, 137, 403–411. [Google Scholar] [CrossRef]
  23. Pronk, W.; Ding, A.; Morgenroth, E.; Derlon, N.; Desmond, P.; Burkhardt, M.; Wu, B.; Fane, A.G. Gravity-driven membrane filtration for water and wastewater treatment: A review. Water Res. 2019, 149, 553–565. [Google Scholar] [CrossRef] [PubMed]
  24. Pervov, A.; Tikhonov, K.; Makisha, N. Application of reverse osmosis techniques to treat and reuse biologically treated wastewater. IOP Conf. Ser. Mater. Sci. Eng. 2018, 365. [Google Scholar] [CrossRef]
  25. Jafarinejad, S. A Comprehensive Study on the Application of Reverse Osmosis (RO) Technology for the Petroleum Industry Wastewater Treatment. J. Water Environ. Nanotechnol. 2017, 2, 243–264. [Google Scholar] [CrossRef]
  26. Zeneli, A.; Kastanaki, E.; Simantiraki, F.; Gidarakos, E. Monitoring the biodegradation of TPH and PAHs in refinery solid waste by biostimulation and bioaugmentation. J. Environ. Chem. Eng. 2019, 7. [Google Scholar] [CrossRef]
  27. Gaur, N.; Narasimhulu, K.; PydiSetty, Y. Recent advances in the bio-remediation of persistent organic pollutants and its effect on environment. J. Clean. Prod. 2018, 198, 1602–1631. [Google Scholar] [CrossRef]
  28. Siipola, V.; Pflugmacher, S.; Romar, H.; Wendling, L.; Koukkari, P. Low-Cost Biochar Adsorbents for Water Purification Including Microplastics Removal. Appl. Sci. 2020, 10, 788. [Google Scholar] [CrossRef] [Green Version]
  29. Tsang, D.C.W.; Kumar, S.; Lee, S.-S.; Kim, K.-H.; Kumar, V. Metal organic frameworks as potent treatment media for odorants and volatiles in air. Environ. Res. 2018, 168, 336–356. [Google Scholar] [CrossRef]
  30. Lv, S.-W.; Liu, J.-M.; Ma, H.; Wang, Z.-H.; Li, C.-Y.; Zhao, N.; Wang, S. Simultaneous adsorption of methyl orange and methylene blue from aqueous solution using amino functionalized Zr-based MOFs. Microporous Mesoporous Mater. 2019, 282, 179–187. [Google Scholar] [CrossRef]
  31. Fu, L.; Wang, S.; Lin, G.; Zhang, L.; Liu, Q.; Fang, J.; Wei, C.; Liu, G. Post-functionalization of UiO-66-NH 2 by 2,5-Dimercapto-1,3,4-thiadiazole for the high efficient removal of Hg(II) in water. J. Hazard. Mater. 2019, 368, 42–51. [Google Scholar] [CrossRef] [PubMed]
  32. Zhan, X.; Zhang, Y.; Xie, L.; Liu, H.; Zhang, X.; Ruan, B.; Ding, H.; Wu, J.; Shi, D.; Jiang, T.; et al. Magnetically treated Zr-based UiO-type porous coordination polymers study on adsorption of azo dye. Microporous Mesoporous Mater. 2020, 110291. [Google Scholar] [CrossRef]
  33. Xin, S.; Yang, N.; Gao, F.; Zhao, J.; Li, L.; Teng, C. Three-dimensional polypyrrole-derived carbon nanotube framework for dye adsorption and electrochemical supercapacitor. Appl. Surf. Sci. 2017, 414, 218–223. [Google Scholar] [CrossRef]
  34. Jung, C.; Son, A.; Her, N.; Zoh, K.D.; Cho, J.; Yoon, Y. Removal of endocrine disrupting compounds, pharmaceuticals, and personal care products in water using carbon nanotubes: A review. J. Ind. Eng. Chem. 2015, 27, 1–11. [Google Scholar] [CrossRef]
  35. Bedia, J.; Peñas-Garzón, M.; Gómez-Avilés, A.; Rodriguez, J.; Belver, C. A Review on the Synthesis and Characterization of Biomass-Derived Carbons for Adsorption of Emerging Contaminants from Water. C J. Carbon Res. 2018, 4, 63. [Google Scholar] [CrossRef] [Green Version]
  36. Bernal, V.; Giraldo, L.; Moreno-Piraján, J. Physicochemical Properties of Activated Carbon: Their Effect on the Adsorption of Pharmaceutical Compounds and Adsorbate–Adsorbent Interactions. J. Carbon Res. 2018, 4, 62. [Google Scholar] [CrossRef] [Green Version]
  37. Garba, Z.N.; Zango, Z.U.; Babando, A.A.; Galadima, A. Competitive adsorption of dyes onto granular activated carbon. J. Chem. Pharm. Res. 2015, 7, 710–717. [Google Scholar]
  38. Sophia, A.C.; Lima, E.C. Removal of emerging contaminants from the environment by adsorption. Ecotoxicol. Environ. Saf. 2018, 150, 1–17. [Google Scholar] [CrossRef]
  39. Garba, Z.N.; Tanimu, A.; Zango, Z.U. Borassus aethiopum shell-based activated carbon as efficient adsorbent for carbofuran. Bull. Chem. Soc. Ethiop. 2019, 33, 425–436. [Google Scholar] [CrossRef]
  40. Kaur, S.; Rani, S.; Mahajan, R.K.; Asif, M.; Gupta, V.K. Synthesis and adsorption properties of mesoporous material for the removal of dye safranin: Kinetics, equilibrium, and thermodynamics. J. Ind. Eng. Chem. 2015, 22, 19–27. [Google Scholar] [CrossRef]
  41. Peres, E.C.; Slaviero, J.C.; Cunha, A.M.; Hosseini-Bandegharaei, A.; Dotto, G.L. Microwave synthesis of silica nanoparticles and its application for methylene blue adsorption. J. Environ. Chem. Eng. 2018, 6, 649–659. [Google Scholar] [CrossRef]
  42. Zango, Z.U.; Abu Bakar, N.H.H.; Tan, W.L.; Bakar, M.A. Enhanced removal efficiency of methyl red via the modification of halloysite nanotubes by copper oxide. J. Dispers. Sci. Technol. 2017. [Google Scholar] [CrossRef]
  43. Zango, Z.U.; Garba, Z.N.; Abu Bakar, N.H.H.; Tan, W.L.; Abu Bakar, M. Adsorption studies of Cu2+–Hal nanocomposites for the removal of 2,4,6-trichlorophenol. Appl. Clay Sci. 2016, 132–133, 68–78. [Google Scholar] [CrossRef]
  44. Park, C.M.; Wang, D.; Han, J.; Heo, J.; Su, C. Evaluation of the colloidal stability and adsorption performance of reduced graphene oxide–elemental silver/magnetite nanohybrids for selected toxic heavy metals in aqueous solutions. Appl. Surf. Sci. 2019, 471, 8–17. [Google Scholar] [CrossRef] [PubMed]
  45. Krupadam, R.J. Nanoporous Polymeric Material for Remediation of PAHs Polluted Water. Polycycl. Aromat. Compd. 2012, 32, 313–333. [Google Scholar] [CrossRef]
  46. Wu, G.; Ma, J.; Li, S.; Guan, J.; Jiang, B.; Wang, L.; Li, J.; Wang, X.; Chen, L. Magnetic copper-based metal organic framework as an effective and recyclable adsorbent for removal of two fluoroquinolone antibiotics from aqueous solutions. J. Colloid Interface Sci. 2018, 528, 360–371. [Google Scholar] [CrossRef]
  47. Zango, Z.U.; Ramli, A.; Jumbri, K.; Soraya, N.; Ahmad, H.I.; Hana, N.; Abu, H.; Saad, B. Optimization studies and artificial neural network modeling for pyrene adsorption onto UiO-66(Zr) and NH2-UiO-66(Zr) metal organic frameworks. Polyhedron 2020, 192, 114857. [Google Scholar] [CrossRef]
  48. Ravelli, D.; Dondi, D.; Fagnoni, M.; Albini, A. Photocatalysis. A multi-faceted concept for green chemistry. Chem. Soc. Rev. 2009, 38, 1999–2011. [Google Scholar] [CrossRef]
  49. Sharma, K.; Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, P.; Singh, P. Recent advances in enhanced photocatalytic activity of bismuth oxyhalides for efficient photocatalysis of organic pollutants in water: A review. J. Ind. Eng. Chem. 2019, 78, 1–20. [Google Scholar] [CrossRef]
  50. He, J.; Zhang, Y.; Zhang, X.; Huang, Y. Highly efficient Fenton and enzyme-mimetic activities of NH2-MIL-88B(Fe) metal organic framework for methylene blue degradation. Sci. Rep. 2018, 8, 5159. [Google Scholar] [CrossRef] [Green Version]
  51. Debnath, D.; Gupta, A.K.; Ghosal, P.S. Recent advances in the development of tailored functional materials for the treatment of pesticides in aqueous media: A review. J. Ind. Eng. Chem. 2019, 70, 51–69. [Google Scholar] [CrossRef]
  52. García, E.; Medina, R.; Lozano, M.; Hernández Pérez, I.; Valero, M.; Franco, A. 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] [PubMed] [Green Version]
  53. Hu, J.; Liu, Y.; Liu, J.; Gu, C.; Wu, D. High CO2 adsorption capacities in UiO type MOFs comprising heterocyclic ligand. Microporous Mesoporous Mater. 2018, 256, 25–31. [Google Scholar] [CrossRef]
  54. Fan, Y.H.; Zhang, S.W.; Qin, S.-B.; Li, X.S.; Qi, S.H. An enhanced adsorption of organic dyes onto NH2 functionalization titanium-based metal-organic frameworks and the mechanism investigation. Microporous Mesoporous Mater. 2018, 263, 120–127. [Google Scholar] [CrossRef]
  55. Hoskins, B.F.; Robson, R. Design and Construction of a New Class of Scaffolding-like Materials Comprising Infinite Polymeric Frameworks of 3D-Linked Molecular Rods. A Reappraisal of the Zn(CN)2 and Cd(CN)2 Structures and the Synthesis and Structure of the Diamond-Related Framework. J. Am. Chem. Soc. 1990, 112, 1546–1554. [Google Scholar] [CrossRef]
  56. Tranchemontagne, D.J.; Hunt, J.R.; Yaghi, O.M. Room temperature synthesis of metal-organic frameworks: MOF-5, MOF-74, MOF-177, MOF-199, and IRMOF-0. Tetrahedron 2008, 64, 8553–8557. [Google Scholar] [CrossRef]
  57. Ghanbari, T.; Abnisa, F.; Wan Daud, W.M.A. A review on production of metal organic frameworks (MOF) for CO2 adsorption. Sci. Total Environ. 2020, 707, 135090. [Google Scholar] [CrossRef]
  58. Neshastehgar, M.; Rahmani, P.; Shojaei, A.; Molavi, H. Enhanced adsorption removal performance of UiO-66 by rational hybridization with nanodiamond. Microporous Mesoporous Mater. 2020, 296. [Google Scholar] [CrossRef]
  59. Zhao, R.; Ma, T.; Zhao, S.; Rong, H.; Tian, Y.; Zhu, G. Uniform and stable immobilization of metal-organic frameworks into chitosan matrix for enhanced tetracycline removal from water. Chem. Eng. J. 2020, 382, 122893. [Google Scholar] [CrossRef]
  60. Hu, M.L.; Masoomi, M.Y.; Morsali, A. Template strategies with MOFs. Coord. Chem. Rev. 2019, 387, 415–435. [Google Scholar] [CrossRef]
  61. Biserčić, M.S.; Marjanović, B.; Vasiljević, B.N.; Mentus, S.; Zasońska, B.A.; Ćirić-Marjanović, G. The quest for optimal water quantity in the synthesis of metal-organic framework MOF-5. Microporous Mesoporous Mater. 2019, 278, 23–29. [Google Scholar] [CrossRef]
  62. Kumar, P.; Kim, K.H.; Lee, J.; Shang, J.; Khazi, M.I.; Kumar, N.; Lisak, G. Metal-organic framework for sorptive/catalytic removal and sensing applications against nitroaromatic compounds. J. Ind. Eng. Chem. 2020, 84, 87–95. [Google Scholar] [CrossRef]
  63. Li, X.; Wang, B.; Cao, Y.; Zhao, S.; Wang, H.; Feng, X.; Zhou, J.; Ma, X. Water Contaminant Elimination Based on Metal-Organic Frameworks and Perspective on Their Industrial Applications. ACS Sustain. Chem. Eng. 2019, 7, 4548–4563. [Google Scholar] [CrossRef]
  64. Mu, X.; Chen, Y.; Lester, E.; Wu, T. Optimized synthesis of nano-scale high quality HKUST-1 under mild conditions and its application in CO2 capture. Microporous Mesoporous Mater. 2018, 270, 249–257. [Google Scholar] [CrossRef] [Green Version]
  65. Cohen, S.M. Postsynthetic methods for the functionalization of metal–organic frameworks. Chem. Rev. 2012, 112, 970–1000. [Google Scholar] [CrossRef] [PubMed]
  66. Kumar, P.; Bansal, V.; Kim, K.H.; Kwon, E.E. Metal-organic frameworks (MOFs) as futuristic options for wastewater treatment. J. Ind. Eng. Chem. 2018, 62, 130–145. [Google Scholar] [CrossRef]
  67. Dhaka, S.; Kumar, R.; Deep, A.; Kurade, M.B.; Ji, S.W.; Jeon, B.H. Metal–organic frameworks (MOFs) for the removal of emerging contaminants from aquatic environments. Coord. Chem. Rev. 2019, 380, 330–352. [Google Scholar] [CrossRef]
  68. Joseph, L.; Jun, B.M.; Jang, M.; Park, C.M.; Muñoz-Senmache, J.C.; Hernández-Maldonado, A.J.; Heyden, A.; Yu, M.; Yoon, Y. Removal of contaminants of emerging concern by metal-organic framework nanoadsorbents: A review. Chem. Eng. J. 2019, 369, 928–946. [Google Scholar] [CrossRef]
  69. Hasan, Z.; Jhung, S.H. Removal of hazardous organics from water using metal-organic frameworks (MOFs): Plausible mechanisms for selective adsorptions. J. Hazard. Mater. 2015, 283, 329–339. [Google Scholar] [CrossRef]
  70. Ghosh, A.; Das, G. Green synthesis of Sn(II)-BDC MOF: Preferential and efficient adsorption of anionic dyes. Microporous Mesoporous Mater. 2020, 297. [Google Scholar] [CrossRef]
  71. Xu, F.; Yu, Y.; Yan, J.; Xia, Q.; Wang, H.; Li, J.; Li, Z. Ultrafast room temperature synthesis of GrO@HKUST-1 composites with high CO2 adsorption capacity and CO2/N2 adsorption selectivity. Chem. Eng. J. 2016, 303, 231–237. [Google Scholar] [CrossRef]
  72. Gaikwad, S.; Kim, S.J.; Han, S. Novel metal–organic framework of UTSA-16(Zn) synthesized by a microwave method: Outstanding performance for CO2 capture with improved stability to acid gases. J. Ind. Eng. Chem. 2020, 87, 250–263. [Google Scholar] [CrossRef]
  73. Jang, S.; Song, S.; Lim, J.H.; Kim, H.S.; Phan, B.T.; Ha, K.T.; Park, S.; Park, K.H. Application of various metal-organic frameworks (MOFs) as catalysts for air and water pollution environmental remediation. Catalysts 2020, 10, 195. [Google Scholar] [CrossRef] [Green Version]
  74. Petit, C. Present and future of MOF research in the field of adsorption and molecular separation. Curr. Opin. Chem. Eng. 2018, 20, 132–142. [Google Scholar] [CrossRef]
  75. Zhao, H.; Li, Q.; Wang, Z.; Wu, T.; Zhang, M. Synthesis of MIL-101(Cr) and its water adsorption performance. Microporous Mesoporous Mater. 2020, 297. [Google Scholar] [CrossRef]
  76. Yoon, S.; Calvo, J.J.; So, M.C. Removal of acid orange 7 from aqueous solution by metal-organic frameworks. Crystals 2019, 9, 17. [Google Scholar] [CrossRef] [Green Version]
  77. Alvaro, M.; Carbonell, E.; Ferrer, B.Ø.; Llabr, F.X. Semiconductor Behavior of a Metal-Organic Framework (MOF). Chem. Eur. J. 2007, 13, 5106–5112. [Google Scholar] [CrossRef]
  78. Quan, X.; Sun, Z.; Meng, H.; Han, Y.; Wu, J.; Xu, J.; Xu, Y.; Zhang, X. Polyethyleneimine (PEI) incorporated Cu-BTC composites: Extended applications in ultra-high efficient removal of congo red. J. Solid State Chem. 2019, 270, 231–241. [Google Scholar] [CrossRef]
  79. Qin, J.-S.; Yuan, S.; Lollar, C.; Pang, J.; Alsalme, A.; Zhou, H.C. Stable metal–organic frameworks as a host platform for catalysis and biomimetics. Chem. Commun. 2018, 54, 4231–4249. [Google Scholar] [CrossRef]
  80. Zahn, G.; Schulze, H.A.; Lippke, J.; König, S.; Sazama, U.; Fröba, M.; Behrens, P. A water-born Zr-based porous coordination polymer: Modulated synthesis of Zr-fumarate MOF. Microporous Mesoporous Mater. 2015, 203, 186–194. [Google Scholar] [CrossRef]
  81. Liang, R.; Jing, F.; Shen, L.; Qin, N.; Wu, L. MIL-53(Fe) as a highly efficient bifunctional photocatalyst for the simultaneous reduction of Cr(VI) and oxidation of dyes. J. Hazard. Mater. 2015, 287, 364–372. [Google Scholar] [CrossRef] [PubMed]
  82. Xia, X.; Xu, Y.; Chen, Y.; Liu, Y.; Lu, Y.; Shao, L. Fabrication of MIL-101(Cr/Al) with flower-like morphology and its catalytic performance. Appl. Catal. A Gen. 2018, 559, 138–145. [Google Scholar] [CrossRef]
  83. Liu, Y.; Liu, Z.; Huang, D.; Cheng, M.; Zeng, G.; Lai, C.; Zhang, C.; Zhou, C.; Wang, W.; Jiang, D.; et al. Metal or metal-containing nanoparticle@MOF nanocomposites as a promising type of photocatalyst. Coord. Chem. Rev. 2019, 388, 63–78. [Google Scholar] [CrossRef]
  84. Wu, T.; Liu, X.; Liu, Y.; Cheng, M.; Liu, Z.; Zeng, G.; Shao, B.; Liang, Q.; Zhang, W.; He, Q. Application of QD-MOF composites for photocatalysis: Energy production and environmental remediation. Coord. Chem. Rev. 2020, 403, 213097. [Google Scholar] [CrossRef]
  85. Abdelhameed, R.M.; El-Shahat, M. Fabrication of ZIF-67@MIL-125-NH2 nanocomposite with enhanced visible light photoreduction activity. J. Environ. Chem. Eng. 2019, 7. [Google Scholar] [CrossRef]
  86. Han, T.; Xiao, Y.; Tong, M.; Huang, H.; Liu, D.; Wang, L.; Zhong, C. Synthesis of CNT@MIL-68(Al) composites with improved adsorption capacity for phenol in aqueous solution. Chem. Eng. J. 2015, 275, 134–141. [Google Scholar] [CrossRef]
  87. Meng, Z.; Liu, B.; Li, M.; Liu, X.; Li, S.; Su, B. Molecular imprinted materials PDA/Fe-MOFs/RGO for the selective and high removal of phenol. Desalin. Water Treat. 2019, 169, 279–286. [Google Scholar] [CrossRef]
  88. Yang, Z.; Xu, X.; Liang, X.; Lei, C.; Gao, L.; Hao, R.; Lu, D.; Lei, Z. Fabrication of Ce doped UiO-66/graphene nanocomposites with enhanced visible light driven photoactivity for reduction of nitroaromatic compounds. Appl. Surf. Sci. 2017, 420, 276–285. [Google Scholar] [CrossRef]
  89. Alfonso-Herrera, L.A.; Huerta-Flores, A.M.; Torres-Martínez, L.M.; Rivera-Villanueva, J.M.; Ramírez-Herrera, D.J. Hybrid SrZrO3-MOF heterostructure: Surface assembly and photocatalytic performance for hydrogen evolution and degradation of indigo carmine dye. J. Mater. Sci. Mater. Electron. 2018, 29, 10395–10410. [Google Scholar] [CrossRef]
  90. Wang, Q.; Wang, G.; Liang, X.; Dong, X.; Zhang, X. Supporting carbon quantum dots on NH2-MIL-125 for enhanced photocatalytic degradation of organic pollutants under a broad spectrum irradiation. Appl. Surf. Sci. 2019, 467–468, 320–327. [Google Scholar] [CrossRef]
  91. Li, Y.; Fang, Y.; Cao, Z.; Li, N.; Chen, D.; Xu, Q.; Lu, J. Construction of g-C3N4/PDI@MOF heterojunctions for the highly efficient visible light-driven degradation of pharmaceutical and phenolic micropollutants. Appl. Catal. B Environ. 2019, 250, 150–162. [Google Scholar] [CrossRef]
  92. He, S.; Rong, Q.; Niu, H.; Cai, Y. Platform for molecular-material dual regulation: A direct Z-scheme MOF/COF heterojunction with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2019, 247, 49–56. [Google Scholar] [CrossRef]
  93. Ramezanalizadeh, H.; Zakeri, F.; Manteghi, F. Immobilization of BaWO4 nanostructures on a MOF-199-NH2: An efficient separable photocatalyst for the degradation of organic dyes. Optik 2018, 174, 776–786. [Google Scholar] [CrossRef]
  94. 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]
  95. Jiang, D.; Chen, M.; Wang, H.; Zeng, G.; Huang, D.; Cheng, M.; Liu, Y.; Xue, W.; Wang, Z.W. The application of different typological and structural MOFs-based materials for the dyes adsorption. Coord. Chem. Rev. 2019, 380, 471–483. [Google Scholar] [CrossRef]
  96. Nandasiri, M.I.; Jambovane, S.R.; McGrail, B.P.; Schaef, H.T.; Nune, S.K. Adsorption, separation, and catalytic properties of densified metal-organic frameworks. Coord. Chem. Rev. 2016, 311, 38–52. [Google Scholar] [CrossRef] [Green Version]
  97. Liu, J.; Xiao, J.; Wang, D.; Sun, W.; Gao, X.; Yu, H.; Liu, H.; Liu, Z. Construction and Photocatalytic Activities of a Series of Isostructural Co2+/Zn2+ Metal-Doped Metal-Organic Frameworks. Cryst. Growth Des. 2017, 17, 1096–1102. [Google Scholar] [CrossRef]
  98. Yang, J.M.; Yang, B.C.; Zhang, Y.; Yang, R.N.; Ji, S.S.; Wang, Q.; Quan, S.; Zhang, R.Z. Rapid adsorptive removal of cationic and anionic dyes from aqueous solution by a Ce(III)-doped Zr-based metal–organic framework. Microporous Mesoporous Mater. 2020, 292. [Google Scholar] [CrossRef]
  99. Tong, M.; Liu, D.; Yang, Q.; Devautour-Vinot, S.; Maurin, G.; Zhong, C. Influence of framework metal ions on the dye capture behavior of MIL-100 (Fe,Cr) MOF type solids. J. Mater. Chem. A 2013, 1, 8534–8537. [Google Scholar] [CrossRef]
  100. Yilmaz, E.; Sert, E.; Atalay, F.S. Synthesis, characterization of a metal organic framework: MIL-53(Fe) and adsorption mechanisms of methyl red onto MIL-53(Fe). J. Taiwan Inst. Chem. Eng. 2016, 65, 323–330. [Google Scholar] [CrossRef]
  101. Haque, E.; Jun, J.W.; Jhung, S.H. Adsorptive removal of methyl orange and methylene blue from aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard. Mater. 2011, 185, 507–511. [Google Scholar] [CrossRef] [PubMed]
  102. Wen, G.; Guo, Z.G. Facile modification of NH2-MIL-125(Ti) to enhance water stability for efficient adsorptive removal of crystal violet from aqueous solution. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 541, 58–67. [Google Scholar] [CrossRef]
  103. Shen, T.; Luo, J.; Zhang, S.; Luo, X. Hierarchically mesostructured MIL-101 metal-organic frameworks with different mineralizing agents for adsorptive removal of methyl orange and methylene blue from aqueous solution. J. Environ. Chem. Eng. 2015, 3, 1372–1383. [Google Scholar] [CrossRef]
  104. Karmakar, S.; Roy, D.; Janiak, C.; De, S. Insights into multi-component adsorption of reactive dyes on MIL-101-Cr metal organic framework: Experimental and modeling approach. Sep. Purif. Technol. 2019, 215, 259–275. [Google Scholar] [CrossRef]
  105. Zhao, X.; Wang, K.; Gao, Z.; Gao, H.; Xie, Z.; Du, X.; Huang, H. Reversing the Dye Adsorption and Separation Performance of Metal-Organic Frameworks via Introduction of -SO3H Groups. Ind. Eng. Chem. Res. 2017, 56, 4496–4501. [Google Scholar] [CrossRef]
  106. Kaur, R.; Kaur, A.; Umar, A.; Anderson, W.A.; Kansal, S.K. Metal organic framework (MOF) porous octahedral nanocrystals of Cu-BTC: Synthesis, properties and enhanced absorption properties. Mater. Res. Bull. 2019, 109, 124–133. [Google Scholar] [CrossRef]
  107. Jabbari, V.; Veleta, J.M.; Zarei-Chaleshtori, M.; Gardea-Torresdey, J.; Villagrán, D. Green synthesis of magnetic MOF@GO and MOF@CNT hybrid nanocomposites with high adsorption capacity towards organic pollutants. Chem. Eng. J. 2016, 304, 774–783. [Google Scholar] [CrossRef] [Green Version]
  108. Azhdari, R.; Mojtaba, S.; Alireza, S.; Bahrani, S. Decorated graphene with aluminum fumarate metal organic framework as a superior non-toxic agent for e ffi cient removal of Congo Red dye from wastewater. J. Environ. Chem. Eng. 2019, 7, 103437. [Google Scholar] [CrossRef]
  109. Oveisi, M.; Asli, M.A.; Mahmoodi, N.M. MIL-Ti metal-organic frameworks (MOFs) nanomaterials as superior adsorbents: Synthesis and ultrasound-aided dye adsorption from multicomponent wastewater systems. J. Hazard. Mater. 2018, 347, 123–140. [Google Scholar] [CrossRef]
  110. Niu, P.; Lu, N.; Liu, J.; Jia, H.; Zhou, F.; Fan, B.; Li, R. Water-induced synthesis of hierarchical Zr-based MOFs with enhanced adsorption capacity and catalytic activity. Microporous Mesoporous Mater. 2019, 281, 92–100. [Google Scholar] [CrossRef]
  111. Zhang, J.; Li, F.; Sun, Q. Rapid and selective adsorption of cationic dyes by a unique metal-organic framework with decorated pore surface. Appl. Surf. Sci. 2018, 440, 1219–1226. [Google Scholar] [CrossRef]
  112. Tian, S.; Xu, S.; Liu, J.; He, C.; Xiong, Y.; Feng, P. Highly efficient removal of both cationic and anionic dyes from wastewater with a water-stable and eco-friendly Fe-MOF via host-guest encapsulation. J. Clean. Prod. 2019, 239. [Google Scholar] [CrossRef]
  113. Chen, C.; Zhang, M.; Guan, Q.; Li, W. Kinetic and thermodynamic studies on the adsorption of xylenol orange onto MIL-101(Cr). Chem. Eng. J. 2012, 183, 60–67. [Google Scholar] [CrossRef]
  114. He, J.; Li, J.; Du, W.; Han, Q.; Wang, Z.; Li, M. A mesoporous metal-organic framework: Potential advances in selective dye adsorption. J. Alloys Compd. 2018, 750, 360–367. [Google Scholar] [CrossRef]
  115. Qi, Z.P.; Kang, Y.S.; Guo, F.; Sun, W.Y. Controlled synthesis of NbO-type metal-organic framework nano/microcrystals with superior capacity and selectivity for dye adsorption from aqueous solution. Microporous Mesoporous Mater. 2019, 273, 60–66. [Google Scholar] [CrossRef]
  116. Yang, M.; Bai, Q. Flower-like hierarchical Ni-Zn MOF microspheres: Efficient adsorbents for dye removal. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 582. [Google Scholar] [CrossRef]
  117. Shi, Z.; Xu, C.; Guan, H.; Li, L.; Fan, L.; Wang, Y.; Liu, L.; Meng, Q.; Zhang, R. Magnetic metal organic frameworks (MOFs) composite for removal of lead and malachite green in wastewater. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 539, 382–390. [Google Scholar] [CrossRef] [Green Version]
  118. Zhao, S.; Chen, D.; Wei, F.; Chen, N.; Liang, Z.; Luo, Y. Removal of Congo red dye from aqueous solution with nickel-based metal-organic framework/graphene oxide composites prepared by ultrasonic wave-assisted ball milling. Ultrason. Sonochem. 2017, 39, 845–852. [Google Scholar] [CrossRef]
  119. Haque, E.; Lee, J.E.; Jang, I.T.; Hwang, Y.K.; Chang, J.S.; Jegal, J.; Jhung, S.H. Adsorptive removal of methyl orange from aqueous solution with metal-organic frameworks, porous chromium-benzenedicarboxylates. J. Hazard. Mater. 2010, 181, 535–542. [Google Scholar] [CrossRef]
  120. Azad, F.N.; Ghaedi, M.; Dashtian, K.; Hajati, S.; Pezeshkpour, V. Ultrasonically assisted hydrothermal synthesis of activated carbon-HKUST-1-MOF hybrid for efficient simultaneous ultrasound-assisted removal of ternary organic dyes and antibacterial investigation: Taguchi optimization. Ultrason. Sonochem. 2016, 31, 383–393. [Google Scholar] [CrossRef]
  121. Hamedi, A.; Zarandi, M.B.; Nateghi, M.R. Highly efficient removal of dye pollutants by MIL-101(Fe) metal-organic framework loaded magnetic particles mediated by Poly L-Dopa. J. Environ. Chem. Eng. 2019, 7. [Google Scholar] [CrossRef]
  122. Liu, X.; Gong, W.; Luo, J.; Zou, C.; Yang, Y.; Yang, S. Selective adsorption of cationic dyes from aqueous solution by polyoxometalate-based metal-organic framework composite. Appl. Surf. Sci. 2016, 362, 517–524. [Google Scholar] [CrossRef]
  123. Aslam, S.; Zeng, J.; Subhan, F.; Li, M.; Lyu, F.; Li, Y.; Yan, Z. In situ one-step synthesis of Fe3O4@MIL-100(Fe) core-shells for adsorption of methylene blue from water. J. Colloid Interface Sci. 2017, 505, 186–195. [Google Scholar] [CrossRef] [PubMed]
  124. Mahmoodi, N.M.; Oveisi, M.; Asadi, E. Synthesis of NENU metal-organic framework-graphene oxide nanocomposites and their pollutant removal ability from water using ultrasound. J. Clean. Prod. 2019, 211, 198–212. [Google Scholar] [CrossRef]
  125. Tan, Y.; Sun, Z.; Meng, H.; Han, Y.; Wu, J.; Xu, J.; Xu, Y.; Zhang, X. A new MOFs/polymer hybrid membrane: MIL-68 (Al)/PVDF, fabrication and application in high-efficient removal of p-nitrophenol and methylene blue. Sep. Purif. Technol. 2019, 68, 217–226. [Google Scholar] [CrossRef]
  126. Tambat, S.N.; Sane, P.K.; Suresh, S.; Varadan, O.N.; Pandit, A.B.; Sontakke, S.M. Hydrothermal synthesis of NH2-UiO-66 and its application for adsorptive removal of dye. Adv. Powder Technol. 2018, 29, 2626–2632. [Google Scholar] [CrossRef]
  127. Chang, N.; Zhang, H.; Shi, M.S.; Li, J.; Yin, C.J.; Wang, H.T.; Wang, L. Regulation of the adsorption affinity of metal-organic framework MIL-101 via a TiO2 coating strategy for high capacity adsorption and efficient photocatalysis. Microporous Mesoporous Mater. 2018, 266, 47–55. [Google Scholar] [CrossRef]
  128. Li, T.T.; Liu, Y.M.; Wang, T.; Wu, Y.L.; He, Y.L.; Yang, R.; Zheng, S.R. Regulation of the surface area and surface charge property of MOFs by multivariate strategy: Synthesis, characterization, selective dye adsorption and separation. Microporous Mesoporous Mater. 2018, 272, 101–108. [Google Scholar] [CrossRef]
  129. Jalali, S.; Rahimi, M.R.; Dashtian, K.; Ghaedi, M.; Mosleh, S. One step integration of plasmonic Ag2CrO4/Ag/AgCl into HKUST-1-MOF as novel visible-light driven photocatalyst for highly efficient degradation of mixture dyes pollutants: Its photocatalytic mechanism and modeling. Polyhedron 2019, 166, 217–225. [Google Scholar] [CrossRef]
  130. Xiang, W.; Zhang, Y.; Lin, H.; Liu, C.J. Nanoparticle/metal-organic framework composites for catalytic applications: Current status and perspective. Molecules 2017, 22, 2103. [Google Scholar] [CrossRef] [Green Version]
  131. Huang, J.; Zhang, X.; Song, H.; Chen, C.; Han, F.; Wen, C. Protonated graphitic carbon nitride coated metal-organic frameworks with enhanced visible-light photocatalytic activity for contaminants degradation. Appl. Surf. Sci. 2018, 441, 85–98. [Google Scholar] [CrossRef]
  132. Xu, W.T.; Ma, L.; Ke, F.; Peng, F.M.; Xu, G.S.; Shen, Y.H.; Zhu, J.F.; Qiu, L.G.; Yuan, Y.P. Metal-organic frameworks MIL-88A hexagonal microrods as a new photocatalyst for efficient decolorization of methylene blue dye. Dalt. Trans. 2014, 43, 3792–3798. [Google Scholar] [CrossRef] [PubMed]
  133. Mahmoodi, N.M.; Abdi, J.; Oveisi, M.; Alinia Asli, M.; Vossoughi, M. Metal-organic framework (MIL-100(Fe)): Synthesis, detailed photocatalytic dye degradation ability in colored textile wastewater and recycling. Mater. Res. Bull. 2018, 100, 357–366. [Google Scholar] [CrossRef]
  134. Prince, G.; Nikhil, R.; Dhabarde, P.C. Rapid synthesis of Titanium based Metal Organic framework (MIL-125) via crossmark microwave route and its performance evaluation in photocatalysis. Mater. Lett. 2017, 186, 151–154. [Google Scholar] [CrossRef]
  135. Li, X.; Guo, W.; Liu, Z.; Wang, R.; Liu, H. Fe-based MOFs for efficient adsorption and degradation of acid orange 7 in aqueous solution via persulfate activation. Appl. Surf. Sci. 2016, 369, 130–136. [Google Scholar] [CrossRef]
  136. Wan, Y.; Wan, J.; Ma, Y.; Wang, Y.; Luo, T. Sustainable synthesis of modulated Fe-MOFs with enhanced catalyst performance for persulfate to degrade organic pollutants. Sci. Total Environ. 2020, 701. [Google Scholar] [CrossRef]
  137. Huang, J.; Song, H.; Chen, C.; Yang, Y.; Xu, N.; Ji, X.; Li, C.; You, J.A. Facile synthesis of N-doped TiO2 nanoparticles caged in MIL-100(Fe) for photocatalytic degradation of organic dyes under visible light irradiation. J. Environ. Chem. Eng. 2017, 5, 2579–2585. [Google Scholar] [CrossRef]
  138. Yang, Y.; Wang, W.; Li, H.; Jin, X.; Wang, H.; Zhang, L.; Zhang, Y. NH2-MIL-53(Al) nanocrystals anchored on the surface of RGO hollow spheres and its visible light degradation of methylene blue. Mater. Lett. 2017, 197, 17–20. [Google Scholar] [CrossRef]
  139. Abdpour, S.; Kowsari, E.; Reza, M.; Moghaddam, A.; Schmolke, L.; Janiak, C. Mil-100(Fe) nanoparticles supported on urchin like Bi2S3 structure for improving photocatalytic degradation of rhodamine-B dye under visible light irradiation. J. Solid State Chem. 2018, 266, 54–62. [Google Scholar] [CrossRef]
  140. Mahmoodi, N.M.; Abdi, J. Nanoporous metal-organic framework (MOF-199): Synthesis, characterization and photocatalytic degradation of Basic Blue 41. Microchem. J. 2019, 144, 436–442. [Google Scholar] [CrossRef]
  141. Chang, N.; Zhang, H.; Shi, M.S.; Li, J.; Shao, W.; Wang, H.T. Metal-organic framework templated synthesis of TiO2@MIL-101 core-shell architectures for high-efficiency adsorption and photocatalysis. Mater. Lett. 2017, 200, 55–58. [Google Scholar] [CrossRef]
  142. Wu, W.; Li, B.; Gu, C.; Wang, J.; Singh, A.; Kumar, A. Luminescent sensing of Cu2+, CrO24 and photocatalytic degradation of methyl violet by Zn (II) metal-organic framework (MOF) having 5,5-(1H-2,3,5-triazole-1,4-diyl) diisophthalic acid ligand. J. Mol. Struct. 2017, 1148, 531–536. [Google Scholar] [CrossRef]
  143. Du, X.; He, H.; Du, L.; Li, W.; Wang, Y.; Jiang, Q.; Yang, L.; Zhang, J.; Guo, S. Porous Pr(III)-based organic framework for dye-adsorption and photo degradation with (4,5)-c net. Polyhedron 2019, 171, 221–227. [Google Scholar] [CrossRef]
  144. Zhang, Y.; Zhou, J.; Feng, Q.; Chen, X.; Hu, Z. Visible light photocatalytic degradation of MB using UiO-66/g-C3N4 heterojunction nanocatalyst. Chemosphere 2018, 212, 523–532. [Google Scholar] [CrossRef]
  145. Ding, J.; Yang, Z.; He, C.; Tong, X.; Li, Y.; Niu, X.; Zhang, H. UiO-66(Zr) coupled with Bi2MoO6 as photocatalyst for visible-light promoted dye degradation. J. Colloid Interface Sci. 2017, 497, 126–133. [Google Scholar] [CrossRef]
  146. Zhang, X.; Zhang, N.; Gan, C.; Liu, Y.; Chen, L.; Zhang, C.; Fang, Y. Synthesis of In2S3/UiO-66 hybrid with enhanced photocatalytic activity towards methyl orange and tetracycline hydrochloride degradation under visible-light irradiation. Mater. Sci. Semicond. Process. 2019, 91, 212–221. [Google Scholar] [CrossRef]
  147. Wang, H.; Cui, P.H.; Shi, J.X.; Tan, J.Y.; Zhang, J.Y.; Zhang, N.; Zhang, C. Controllable self-assembly of CdS@NH2-MIL-125(Ti) heterostructure with enhanced photodegradation efficiency for organic pollutants through synergistic effect. Mater. Sci. Semicond. Process. 2019, 97, 91–100. [Google Scholar] [CrossRef]
  148. Akbarzadeh, E.; Soheili, H.Z.; Hosseinifard, M.; Gholami, M.R. Preparation and characterization of novel Ag3VO4/Cu-MOF/rGO heterojunction for photocatalytic degradation of organic pollutants. Mater. Res. Bull. 2020, 121. [Google Scholar] [CrossRef]
  149. Li, H.; Li, Q.; He, X.; Xu, Z.; Wang, Y.; Jia, L. Synthesis of AgBr@MOFs nanocomposite and its photocatalytic activity for dye degradation. Polyhedron 2019, 165, 31–37. [Google Scholar] [CrossRef]
  150. Mosleh, S.; Rahimi, M.R.; Ghaedi, M.; Dashtian, K.; Hajati, S.; Wang, S. Ag3PO4/AgBr/Ag-HKUST-1-MOF composites as novel blue LED light active photocatalyst for enhanced degradation of ternary mixture of dyes in a rotating packed bed reactor. Chem. Eng. Process. Process. Intensif. 2017, 114, 24–38. [Google Scholar] [CrossRef]
  151. Mosleh, S.; Rahimi, M.R.; Ghaedi, M.; Dashtian, K. Sonophotocatalytic degradation of trypan blue and vesuvine dyes in the presence of blue light active photocatalyst of Ag3PO4/Bi2S3-HKUST-1-MOF: Central composite optimization and synergistic effect study. Ultrason. Sonochem. 2016, 32, 387–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Ramezanalizadeh, H.; Manteghi, F. Synthesis of a novel MOF/CuWO4 heterostructure for efficient photocatalytic degradation and removal of water pollutants. J. Clean. Prod. 2016, 172, 2655–2666. [Google Scholar] [CrossRef]
  153. Kaur, R.; Vellingiri, K.; Kim, K.H.; Paul, A.K.; Deep, A. Efficient photocatalytic degradation of rhodamine 6G with a quantum dot-metal organic framework nanocomposite. Chemosphere 2016, 154, 620–627. [Google Scholar] [CrossRef] [PubMed]
  154. Araya, T.; Chen, C.C.; Jia, M.K.; Johnson, D.; Li, R.; Huang, Y. ping Selective degradation of organic dyes by a resin modified Fe-based metal-organic framework under visible light irradiation. Opt. Mater. 2017, 64, 512–523. [Google Scholar] [CrossRef]
  155. Araya, T.; Jia, M.; Yang, J.; Zhao, P.; Cai, K.; Ma, W.; Huang, Y. Resin modified MIL-53(Fe) MOF for improvement of photocatalytic performance. Appl. Catal. B Environ. 2017, 203, 768–777. [Google Scholar] [CrossRef]
  156. Du, J.; Yuan, Y.; Sun, J.; Peng, F.; Jiang, X.; Qiu, L. New photocatalysts based on MIL-53 metal—Organic frameworks for the decolorization of methylene blue dye. J. Hazard. Mater. 2011, 190, 945–951. [Google Scholar] [CrossRef]
  157. Liu, N.; Jing, C.; Li, Z.; Huang, W.; Gao, B.; You, F.; Zhang, X. Effect of synthesis conditions on the photocatalytic degradation of Rhodamine B of MIL-53(Fe). Mater. Lett. 2019, 237, 92–95. [Google Scholar] [CrossRef]
  158. Pu, M.; Guan, Z.; Ma, Y.; Wan, J.; Wang, Y.; Brusseau, M.L. General Synthesis of iron-based metal-organic framework MIL-53 as an efficient catalyst to activate persulfate for the degradation of Orange G in aqueous solution. Appl. Catal. A 2017, 549, 82–92. [Google Scholar] [CrossRef]
  159. Abdpour, S.; Kowsari, E.; Reza, M.; Moghaddam, A. Synthesis of MIL-100(Fe)@ MIL-53(Fe) as a novel hybrid photocatalyst and evaluation photocatalytic and photoelectrochemical performance under visible light irradiation. J. Solid State Chem. 2018, 262, 172–180. [Google Scholar] [CrossRef]
  160. Zhang, R.; Du, B.; Li, Q.; Cao, Z.; Feng, G.; Wang, X. α-Fe2O3 nanoclusters confined into UiO-66 for efficient visible-light photodegradation performance. Appl. Surf. Sci. 2019, 466, 956–963. [Google Scholar] [CrossRef]
  161. Chen, J.; Chao, F.; Ma, X.; Zhu, Q.; Jiang, J.; Ren, J.; Guo, Y.; Lou, Y. Synthesis of flower-like CuS/UiO-66 composites with enhanced visible-light photocatalytic performance. Inorg. Chem. Commun. 2019, 104, 223–228. [Google Scholar] [CrossRef]
  162. Huu, V.; Giang, L.; Thi, Q.; Bui, P.; Duy, T. Composite photocatalysts containing MIL-53(Fe) as a heterogeneous photo-Fenton catalyst for the decolorization of rhodamine B under visible light irradiation. J. Environ. Chem. Eng. 2018, 53, 2–9. [Google Scholar]
  163. Liu, X.; Dang, R.; Dong, W.; Huang, X.; Tang, J.; Gao, H.; Wang, G. A sandwich-like heterostructure of TiO2 nanosheets with MIL-100(Fe): A platform for efficient visible-light-driven photocatalysis. Appl. Catal. B Environ. 2017, 209, 506–513. [Google Scholar] [CrossRef]
  164. Michałowicz, J.; Włuka, A.; Cyrkler, M.; Maćczak, A.; Sicińska, P.; Mokra, K. Phenol and chlorinated phenols exhibit different apoptotic potential in human red blood cells (in vitro study). Environ. Toxicol. Pharmacol. 2018, 61, 95–101. [Google Scholar] [CrossRef]
  165. Maćczak, A.; Cyrkler, M.; Bukowska, B.; Michałowicz, J. Eryptosis-inducing activity of bisphenol A and its analogs in human red blood cells (in vitro study). J. Hazard. Mater. 2016, 307, 328–335. [Google Scholar] [CrossRef]
  166. Alshabib, M.; Onaizi, S.A. A review on phenolic wastewater remediation using homogeneous and heterogeneous enzymatic processes: Current status and potential challenges. Sep. Purif. Technol. 2019, 219, 186–207. [Google Scholar] [CrossRef]
  167. Ahmed, S.; Rasul, M.G.; Brown, R.; Hashib, M.A. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef] [Green Version]
  168. De Roos, A.J.; Blair, A.; Rusiecki, J.A.; Hoppin, J.A.; Svec, M.; Dosemeci, M.; Sandler, D.P.; Alavanja, M.C. Cancer incidence among glyphosate-exposed pesticide applicators in the Agricultural Health Study. Environ. Health Perspect. 2005, 113, 49–54. [Google Scholar] [CrossRef]
  169. Drout, R.J.; Robison, L.; Chen, Z.; Islamoglu, T.; Farha, O.K. Zirconium Metal–Organic Frameworks for Organic Pollutant Adsorption. Trends Chem. 2019, 1, 304–317. [Google Scholar] [CrossRef]
  170. Xiao, Y.; Tong, F.; Kuang, Y.; Chen, B. Distribution and source apportionment of polycyclic aromatic hydrocarbons (PAHs) in forest soils from urban to rural areas in the Pearl River Delta of southern China. Int. J. Environ. Res. Public Health 2014, 11, 2642–2656. [Google Scholar] [CrossRef]
  171. Yali, Z.P.; Fatemi, M.H. Prediction of the sorption coefficient for the adsorption of PAHs on MWCNT based on hybrid QSPR-molecular docking approach. Adsorption 2019, 25, 737–743. [Google Scholar] [CrossRef]
  172. Zhang, Y.; Tao, S.; Shen, H.; Ma, J. Inhalation exposure to ambient polycyclic aromatic hydrocarbons and lung cancer risk of Chinese population. Proc. Natl. Acad. Sci. USA 2009, 106, 21063–21067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Lawal, A.T. Polycyclic aromatic hydrocarbons. A review. Cogent Environ. Sci. 2017, 3, 1–89. [Google Scholar] [CrossRef]
  174. Mezzanotte, V.; Anzano, M.; Collina, E.; Marazzi, A.; Lasagni, M. Distribution and Removal of Polycyclic Aromatic Hydrocarbons in Two Italian Municipal Wastewater Treatment Plants in 2011–2013. Polycycl. Aromat. Compd. 2015. [Google Scholar] [CrossRef]
  175. Liu, L.; Li, D.; Li, C.; Ji, R.; Tian, X. Metal nanoparticles by doping carbon nanotubes improved the sorption of perfluorooctanoic acid. J. Hazard. Mater. 2018, 351, 206–214. [Google Scholar] [CrossRef]
  176. Moody, C.A.; Kwan, W.C.; Martin, J.W.; Muir, D.C.G.; Mabury, S.A. Determination of perfluorinated surfactants in surface water samples by two independent analytical techniques: Liquid chromatography/tandem mass spectrometry and 19F NMR. Anal. Chem. 2002, 73, 2200–2206. [Google Scholar] [CrossRef]
  177. Enevoldsen, R.; Juhler, R.K. Perfluorinated compounds (PFCs) in groundwater and aqueous soil extracts: Using inline SPE-LC-MS/MS for screening and sorption characterisation of perfluorooctanesulphonate and related compounds. Anal. Bioanal. Chem. 2010, 398, 1161–1172. [Google Scholar] [CrossRef]
  178. Loewen, M.; Halldorson, T.; Wang, F.; Tomy, G. Fluorotelomer carboxylic acids and PFOS in rainwater from an urban center in Canada. Environ. Sci. Technol. 2005, 39, 2944–2951. [Google Scholar] [CrossRef]
  179. Zhang, W.; Zhang, Y.; Taniyasu, S.; Yeung, L.W.Y.; Lam, P.K.S.; Wang, J.; Li, X.; Yamashita, N.; Dai, J. Distribution and fate of perfluoroalkyl substances in municipal wastewater treatment plants in economically developed areas of China. Environ. Pollut. 2013, 176, 10–17. [Google Scholar] [CrossRef]
  180. Mak, Y.L.; Taniyasu, S.; Yeung, L.W.; Lu, G.; Jin, L.; Yang, Y.; Lam, P.K.; Kannan, K.; Yamashita, N. Perfluorinated compounds in tap water from China and several other countries. Environ. Sci. Technol. 2009, 43, 4824–4829. [Google Scholar] [CrossRef]
  181. Zabaleta, I.; Bizkarguenaga, E.; Iparragirre, A.; Navarro, P.; Prieto, A.; Fernandez, L.A.; Zuloaga, O. Focused ultrasound solid-liquid extraction for the determination of perfluorinated compounds in fish, vegetables and amended soil. J. Chromatogr. A 2014, 1331, 27–37. [Google Scholar] [CrossRef] [PubMed]
  182. Huber, S.; Brox, J. An automated high-throughput SPE micro-elution method for perfluoroalkyl substances in human serum. Anal. Bioanal. Chem. 2015, 407, 3751–3761. [Google Scholar] [CrossRef] [PubMed]
  183. Schaefer, C.E.; Andaya, C.; Burant, A.; Condee, C.W.; Urtiaga, A.; Strathmann, T.J.; Higgins, C.P. Electrochemical treatment of perfluorooctanoic acid and perfluorooctane sulfonate: Insights into mechanisms and application to groundwater treatment. Chem. Eng. J. 2017, 317, 424–432. [Google Scholar] [CrossRef]
  184. Pankajakshan, A.; Sinha, M.; Ojha, A.A.; Mandal, S. Water-Stable Nanoscale Zirconium-Based Metal-Organic Frameworks for the Effective Removal of Glyphosate from Aqueous Media. ACS Omega 2018, 3, 7832–7839. [Google Scholar] [CrossRef] [PubMed]
  185. Zhou, M.; Wu, Y.-N.; Qiao, J.; Zhang, J.; McDonald, A.; Li, G.; Li, F. The removal of bisphenol A from aqueous solutions by MIL-53(Al) and mesostructured MIL-53(Al). J. Colloid Interface Sci. 2013, 405, 157–163. [Google Scholar] [CrossRef] [PubMed]
  186. Akpinar, I.; Drout, R.J.; Islamoglu, T.; Kato, S.; Lyu, J.; Farha, O.K. Exploiting π-π Interactions to Design an Efficient Sorbent for Atrazine Removal from Water. ACS Appl. Mater. Interfaces 2019, 11, 6097–6103. [Google Scholar] [CrossRef] [PubMed]
  187. Zango, Z.U.; Sambudi, N.S.; Jumbri, K.; Abu Bakar, N.H.H.; Abdullah, N.A.F.; Negim, E.S.M.; Saad, B. Experimental and molecular docking model studies for the adsorption of polycyclic aromatic hydrocarbons onto UiO-66(Zr) and NH2-UiO-66(Zr) metal-organic frameworks. Chem. Eng. Sci. 2020, 220, 115608. [Google Scholar] [CrossRef]
  188. Zango, Z.U.; Abu Bakar, N.H.H.; Sambudi, N.S.; Jumbri, K.; Abdullah, N.A.F.; Abdul Kadir, E.; Saad, B. Adsorption of chrysene in aqueous solution onto MIL-88(Fe) and NH2-MIL-88(Fe) metal-organic frameworks: Kinetics, isotherms, thermodynamics and docking simulation studies. J. Environ. Chem. Eng. 2019. [Google Scholar] [CrossRef]
  189. Zango, Z.U.; Jumbri, K.; Sambudi, N.S.; Abu Bakar, N.H.H.; Abdullah, N.A.F.; Basheer, C.; Saad, B. Removal of anthracene in water by MIL-88(Fe), NH2-MIL-88(Fe), and mixed-MIL-88(Fe) metal–organic frameworks. RCS Adv. 2019, 9, 41490–41501. [Google Scholar] [CrossRef] [Green Version]
  190. Mei, W.; Song, H.; Tian, Z.; Zuo, S.; Li, D.; Xu, H.; Xia, D. Efficient photo-Fenton like activity in modified MIL-53(Fe) for removal of pesticides: Regulation of photogenerated electron migration. Mater. Res. Bull. 2019, 119. [Google Scholar] [CrossRef]
  191. Ahmad, M.; Chen, S.; Ye, F.; Quan, X.; Afzal, S.; Yu, H.; Zhao, X. Efficient photo-Fenton activity in mesoporous MIL-100(Fe) decorated with ZnO nanosphere for pollutants degradation. Appl. Catal. B Environ. 2018, 245, 428–438. [Google Scholar] [CrossRef]
  192. Lv, S.W.; Liu, J.M.; Li, C.Y.; Zhao, N.; Wang, Z.H.; Wang, S. Two novel MOFs@COFs hybrid-based photocatalytic platforms coupling with sulfate radical-involved advanced oxidation processes for enhanced degradation of bisphenol A. Chemosphere 2020, 243. [Google Scholar] [CrossRef] [PubMed]
  193. Lin, K.A.; Hsieh, Y. Copper-based metal organic framework (MOF), HKUST-1, as an efficient adsorbent to remove p-nitrophenol from water. J. Taiwan Inst. Chem. Eng. 2015, 50, 223–228. [Google Scholar]
  194. Han, T.; Li, C.; Guo, X.; Huang, H.; Liu, D.; Zhong, C. In-situ synthesis of SiO2@MOF composites for high-efficiency removal of aniline from aqueous solution. Appl. Surf. Sci. 2016, 390, 506–512. [Google Scholar] [CrossRef]
  195. Abazari, R.; Salehi, G.; Mahjoub, A.R. Ultrasound-assisted preparation of a nanostructured zinc(II) amine pillar metal-organic framework as a potential sorbent for 2,4-dichlorophenol adsorption from aqueous solution. Ultrason. Sonochem. 2018, 46, 59–67. [Google Scholar] [CrossRef] [PubMed]
  196. Abazari, R.; Mahjoub, A.R. Ultrasound-assisted synthesis of Zinc(II)-based metal organic framework nanoparticles in the presence of modulator for adsorption enhancement of 2,4-dichlorophenol and amoxicillin. Ultrason. Sonochem. 2018, 42, 577–584. [Google Scholar] [CrossRef]
  197. Xu, Z.; Wen, Y.; Tian, L.; Li, G. Efficient and selective adsorption of nitroaromatic explosives by Zr-MOF. Inorg. Chem. Commun. 2017, 77, 11–13. [Google Scholar] [CrossRef]
  198. Wu, Z.; Yuan, X.; Zhong, H.; Wang, H.; Zeng, G.; Chen, X.; Wang, H.; Zhang, L.; Shao, J. Enhanced adsorptive removal of p-nitrophenol from water by aluminum metal-organic framework/reduced graphene oxide composite. Sci. Rep. 2016, 6, 25638. [Google Scholar] [CrossRef] [Green Version]
  199. Guo, H.; Niu, B.; Wu, X.; Zhang, Y.; Ying, S. Effective removal of 2, 4, 6-Trinitrophenol over hexagonal metal—Organic framework NH2--MIL--88B(Fe). Appl. Organomet. Chem. 2018, 33, e4580. [Google Scholar] [CrossRef] [Green Version]
  200. Giraldo, L.; Bastidas-Barranco, M.; Húmpola, P.; Moreno-Piraján, J.C. Design, synthesis and characterization of MOF-199 and ZIF-8: Applications in the adsorption of phenols derivatives in aqueous solution. Eur. J. Chem. 2017, 8, 293–304. [Google Scholar] [CrossRef] [Green Version]
  201. Luo, Z.; Chen, H.; Wu, S.; Yang, C.; Cheng, J. Enhanced removal of bisphenol A from aqueous solution by aluminum-based MOF/sodium alginate-chitosan composite beads. Chemosphere 2019, 237. [Google Scholar] [CrossRef] [PubMed]
  202. Ahsan, M.A.; Jabbari, V.; Islam, M.T.; Turley, R.S.; Dominguez, N.; Kim, H.; Castro, E.; Hernandez-Viezcas, J.A.; Curry, M.L.; Lopez, J.; et al. Sustainable synthesis and remarkable adsorption capacity of MOF/graphene oxide and MOF/CNT based hybrid nanocomposites for the removal of Bisphenol A from water. Sci. Total Environ. 2019, 673, 306–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Zhang, R.; Wang, L.; Han, J.; Wu, J.; Li, C.; Ni, L.; Wang, Y. Improving laccase activity and stability by HKUST-1 with cofactor via one-pot encapsulation and its application for degradation of bisphenol A. J. Hazard. Mater. 2020, 383. [Google Scholar] [CrossRef] [PubMed]
  204. Liu, G.; Li, L.; Xu, D.; Huang, X.; Xu, X.; Zheng, S.; Zhang, Y.; Lin, H. Metal–organic framework preparation using magnetic graphene oxide–β-cyclodextrin for neonicotinoid pesticide adsorption and removal. Carbohydr. Polym. 2017, 175, 584–591. [Google Scholar] [CrossRef] [PubMed]
  205. Mirsoleimani-azizi, S.M.; Setoodeh, P.; Samimi, F.; Shadmehr, J. Diazinon removal from aqueous media by mesoporous MIL-101(Cr) in a continuous fi xed-bed system. J. Environ. Chem. Eng. 2018, 6, 4653–4664. [Google Scholar] [CrossRef]
  206. Wang, B.; Yang, Y.; Lu, Y.; Wang, W.; Wang, Q.; Dong, X.; Zhao, J. Rapid and ef ficient removal of acetochlor from environmental water using Cr-MIL-101 sorbent modi fi ed with 3, 5-Bis (trifluoromethyl) phenyl isocyanate. Sci. Total Environ. 2019, 710, 135512. [Google Scholar]
  207. Moeini, Z.; Azhdarpoor, A.; Yousefinejad, S.; Hashemi, H. Removal of atrazine from water using titanium dioxide encapsulated in salicylaldehyde–NH 2 –MIL-101(Cr): Adsorption or oxidation mechanism. J. Clean. Prod. 2019, 224, 238–245. [Google Scholar] [CrossRef]
  208. Fan, C.; Dong, H.; Liang, Y.; Yang, J.; Tang, G.; Zhang, W.; Cao, Y. Sustainable synthesis of HKUST-1 and its composite by biocompatible ionic liquid for enhancing visible-light photocatalytic performance. J. Clean. Prod. 2019, 208, 353–362. [Google Scholar] [CrossRef]
  209. Zhu, X.; Li, B.; Yang, J.; Li, Y.; Zhao, W.; Shi, J.; Gu, J. Effective adsorption and enhanced removal of organophosphorus pesticides from aqueous solution by Zr-Based MOFs of UiO-67. ACS Appl. Mater. Interfaces 2015, 7, 223–231. [Google Scholar] [CrossRef]
  210. Akpinar, I.; Yazaydin, A.O. Adsorption of Atrazine from Water in Metal-Organic Framework Materials. J. Chem. Eng. Data 2018, 63, 2368–2375. [Google Scholar] [CrossRef]
  211. Okoro, H.K.; Tella, A.C.; Ajibola, O.A.; Zvinowanda, C.; Ngila, J.C. Adsorptive removal of naphthalene and anthracene from aqueous solution with zinc and copper-terephthalate metal-organic frameworks. Bull. Chem. Soc. Ethiop. 2019, 33, 229–241. [Google Scholar] [CrossRef] [Green Version]
  212. Zango, Z.U.; Sambudi, N.S.; Jumbri, K.; Abu Bakar, N.H.H.; Saad, B. Removal of Pyrene from Aqueous Solution Using Fe-based Metal-organic Frameworks. IOP Conf. Ser. Earth Environ. Sci. 2020, 549, 012061. [Google Scholar] [CrossRef]
  213. Chen, M.J.; Yang, A.C.; Wang, N.H.; Chiu, H.C.; Li, Y.L.; Kang, D.Y.; Lo, S.L. Influence of crystal topology and interior surface functionality of metal-organic frameworks on PFOA sorption performance. Microporous Mesoporous Mater. 2016, 236, 202–210. [Google Scholar] [CrossRef]
  214. Zhang, S.; Du, M.; Kuang, J.; Xing, Z.; Li, Z.; Pan, K.; Zhu, Q.; Zhou, W. Surface-defect-rich mesoporous NH2-MIL-125(Ti)@Bi2MoO6 core-shell heterojunction with improved charge separation and enhanced visible-light-driven photocatalytic performance. J. Colloid Interface Sci. 2019, 554, 324–334. [Google Scholar] [CrossRef] [PubMed]
  215. Zhang, H.; Chen, S.; Zhang, H.; Fan, X.; Gao, C.; Yu, H.; Quan, X. Carbon nanotubes-incorporated MIL-88B-Fe as highly ef fi cient Fenton-like catalyst for degradation of organic pollutants. Front. Environ. Sci. Eng. 2019, 13, 18. [Google Scholar] [CrossRef]
  216. Thakare, S.R.; Ramteke, S.M. Postmodification of MOF-5 using secondary complex formation using 8-hydroxyquinoline (HOQ) for the development of visible light active photocatalysts. J. Phys. Chem. Solids 2018, 116, 264–272. [Google Scholar] [CrossRef]
  217. Lin, J.; Hu, Y.; Wang, L.; Liang, D.; Ruan, X.; Shao, S. M88/PS/Vis system for degradation of bisphenol A: Environmental factors, degradation pathways, and toxicity evaluation. Chem. Eng. J. 2019. [Google Scholar] [CrossRef]
  218. Liang, R.; Luo, S.; Jing, F.; Shen, L.; Qin, N.; Wu, L. A simple strategy for fabrication of Pd@MIL-100(Fe) nanocomposite as a visible-light-driven photocatalyst for the treatment of pharmaceuticals and personal care products (PPCPs). Appl. Catal. B Environ. 2015, 176–177, 240–248. [Google Scholar] [CrossRef]
  219. Ke, Q.; Shi, Y.; Liu, Y.; Chen, F.; Wang, H.; Wu, X.L.; Lin, H.; Chen, J. Enhanced catalytic degradation of bisphenol A by hemin-MOFs supported on boron nitride via the photo-assisted heterogeneous activation of persulfate. Sep. Purif. Technol. 2019, 229. [Google Scholar] [CrossRef]
  220. Li, X.; Guo, W.; Liu, Z.; Wang, R.; Liu, H. Quinone-modified NH2-MIL-101(Fe) composite as a redox mediator for improved degradation of bisphenol A. J. Hazard. Mater. 2017, 324, 665–672. [Google Scholar] [CrossRef]
  221. Fakhri, H.; Bagheri, H. Highly efficient Zr-MOF@WO3/graphene oxide photocatalyst: Synthesis, characterization and photodegradation of tetracycline and malathion. Mater. Sci. Semicond. Process. 2020, 107. [Google Scholar] [CrossRef]
  222. Oladipo, A.A.; Vaziri, R.; Abureesh, M.A. Highly robust AgIO3/MIL-53(Fe) nanohybrid composites for degradation of organophosphorus pesticides in single and binary systems: Application of artificial neural networks modelling. J. Taiwan Inst. Chem. Eng. 2018, 83, 133–142. [Google Scholar] [CrossRef]
  223. Sajjadi, S.; Khataee, A.; Bagheri, N.; Kobya, M.; Şenocak, A.; Demirbas, E.; Karaoğlu, A.G. Degradation of diazinon pesticide using catalyzed persulfate with Fe3O4@MOF-2 nanocomposite under ultrasound irradiation. J. Ind. Eng. Chem. 2019, 77, 280–290. [Google Scholar] [CrossRef]
  224. Lin, S.; Zhao, Y.; Yun, Y.S. Highly Effective Removal of Nonsteroidal Anti-inflammatory Pharmaceuticals from Water by Zr(IV)-Based Metal-Organic Framework: Adsorption Performance and Mechanisms. ACS Appl. Mater. Interfaces 2018, 10, 28076–28085. [Google Scholar] [CrossRef]
  225. Jun, B.M.; Heo, J.; Park, C.M.; Yoon, Y. Comprehensive evaluation of the removal mechanism of carbamazepine and ibuprofen by metal organic framework. Chemosphere 2019, 235, 527–537. [Google Scholar] [CrossRef]
  226. Lv, Y.; Zhang, R.; Zeng, S.; Liu, K.; Huang, S.; Liu, Y.; Xu, P.; Lin, C.; Cheng, Y.; Liu, M. Removal of p-arsanilic acid by an amino-functionalized indium-based metal–organic framework: Adsorption behavior and synergetic mechanism. Chem. Eng. J. 2018, 339, 359–368. [Google Scholar] [CrossRef]
  227. Li, S.; Cui, J.; Wu, X.; Zhang, X.; Hu, Q.; Hou, X. Rapid in situ microwave synthesis of Fe3O4 @MIL-100(Fe) for aqueous diclofenac sodium removal through integrated adsorption and photodegradation. J. Hazard. Mater. 2019, 373, 408–416. [Google Scholar] [CrossRef]
  228. Hasan, Z.; Choi, E.J.; Jhung, S.H. Adsorption of naproxen and clofibric acid over a metal–organic framework MIL-101 functionalized with acidic and basic groups. Chem. Eng. J. 2013, 219, 537–544. [Google Scholar] [CrossRef]
  229. Gao, Y.; Xia, J.; Liu, D.; Kang, R.; Yu, G.; Deng, S. Synthesis of mixed-linker Zr-MOFs for emerging contaminant adsorption and photodegradation under visible light. Chem. Eng. J. 2019, 378, 122118. [Google Scholar] [CrossRef]
  230. Liu, W.; Shen, X.; Han, Y.; Liu, Z.; Dai, W.; Dutta, A.; Kumar, A.; Liu, J. Selective adsorption and removal of drug contaminants by using an extremely stable Cu(II)-based 3D metal-organic framework. Chemosphere 2019, 215, 524–531. [Google Scholar] [CrossRef]
  231. Tella, A.C.; Owalude, S.O.; Olatunji, S.J.; Adimula, V.O.; Elaigwu, S.E.; Alimi, L.O.; Ajibade, P.A.; Oluwafemi, O.S. Synthesis of zinc-carboxylate metal-organic frameworks for the removal of emerging drug contaminant (amodiaquine) from aqueous solution. J. Environ. Sci. 2018, 64, 264–275. [Google Scholar] [CrossRef] [PubMed]
  232. Abazari, R.; Mahjoub, A.R.; Shariati, J. Synthesis of a nanostructured pillar MOF with high adsorption capacity towards antibiotics pollutants from aqueous solution. J. Hazard. Mater. 2019, 366, 439–451. [Google Scholar] [CrossRef] [PubMed]
  233. Zhao, X.; Wei, Y.; Zhao, H.; Gao, Z.; Zhang, Y. Functionalized metal-organic frameworks for effective removal of rocephin in aqueous solutions. J. Colloid Interface Sci. 2018, 514, 234–239. [Google Scholar] [CrossRef] [PubMed]
  234. Zhou, Y.; Yang, Q.; Zhang, D.; Gan, N.; Li, Q.; Cuan, J. Detection and removal of antibiotic tetracycline in water with a highly stable luminescent MOF. Sens. Actuators B Chem. 2018, 262, 137–143. [Google Scholar] [CrossRef]
  235. Mirsoleimani-Azizi, S.M.; Setoodeh, P.; Zeinali, S.; Rahimpour, M.R. Tetracycline antibiotic removal from aqueous solutions by MOF-5: Adsorption isotherm, kinetic and thermodynamic studies. J. Environ. Chem. Eng. 2018, 6, 6118–6130. [Google Scholar] [CrossRef]
  236. Gao, Y.; Kang, R.; Xia, J.; Yu, G.; Deng, S. Understanding the adsorption of sulfonamide antibiotics on MIL-53s: Metal dependence of breathing effect and adsorptive performance in aqueous solution. J. Colloid Interface Sci. 2019, 535, 159–168. [Google Scholar] [CrossRef]
  237. Naeimi, S.; Faghihian, H. Application of novel metal organic framework, MIL-53(Fe) and its magnetic hybrid: For removal of pharmaceutical pollutant, doxycycline from aqueous solutions. Environ. Toxicol. Pharmacol. 2017, 53, 121–132. [Google Scholar] [CrossRef]
  238. Gao, Y.; Liu, K.; Kang, R.; Xia, J.; Yu, G.; Deng, S. A comparative study of rigid and fl exible MOFs for the adsorption of pharmaceuticals: Kinetics, isotherms and mechanisms. J. Hazard. Mater. 2018, 359, 248–257. [Google Scholar] [CrossRef]
  239. Wang, D.; Jia, F.; Wang, H.; Chen, F.; Fang, Y.; Dong, W.; Zeng, G.; Li, X.; Yang, Q.; Yuan, X. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 2018, 519, 273–284. [Google Scholar] [CrossRef]
  240. Xiong, W.; Zeng, Z.; Li, X.; Zeng, G.; Xiao, R.; Yang, Z.; Xu, H.; Chen, H.; Cao, J.; Zhou, C.; et al. Ni-doped MIL-53(Fe) nanoparticles for optimized doxycycline removal by using response surface methodology from aqueous solution. Chemosphere 2019, 232, 186–194. [Google Scholar] [CrossRef]
  241. Seo, P.W.; Khan, N.A.; Jhung, S.H. Removal of nitroimidazole antibiotics from water by adsorption over metal—Organic frameworks modified with urea or melamine. Chem. Eng. J. 2017, 315, 92–100. [Google Scholar] [CrossRef]
  242. Xiong, W.; Zeng, Z.; Li, X.; Zeng, G.; Xiao, R.; Yang, Z.; Zhou, Y.; Zhang, C.; Cheng, M.; Hu, L.; et al. Multi-walled carbon nanotube/amino-functionalized MIL-53(Fe) composites: Remarkable adsorptive removal of antibiotics from aqueous solutions. Chemosphere 2018, 210, 1061–1069. [Google Scholar] [CrossRef] [PubMed]
  243. Xiong, W.; Zeng, G.; Yang, Z.; Zhou, Y.; Zhang, C.; Cheng, M.; Liu, Y.; Hu, L.; Wan, J.; Zhou, C.; et al. Adsorption of tetracycline antibiotics from aqueous solutions on nanocomposite multi-walled carbon nanotube functionalized MIL-53(Fe) as new adsorbent. Sci. Total Environ. 2018, 627, 235–244. [Google Scholar] [CrossRef] [PubMed]
  244. Sun, W.; Li, H.; Li, H.; Li, S.; Cao, X. Adsorption mechanisms of ibuprofen and naproxen to UiO-66 and UiO-66-NH 2: Batch experiment and DFT calculation. Chem. Eng. 2018, 360, 645–653. [Google Scholar] [CrossRef]
  245. Dong, W.; Wang, D.; Wang, H.; Li, M.; Chen, F.; Jia, F.; Yang, Q.; Li, X.; Yuan, X.; Gong, J.; et al. Facile synthesis of In2S3/UiO-66 composite with enhanced adsorption performance and photocatalytic activity for the removal of tetracycline under visible light irradiation. J. Colloid Interface Sci. 2019, 535, 444–457. [Google Scholar] [CrossRef]
  246. Azhar, M.R.; Abid, H.R.; Periasamy, V.; Sun, H.; Tade, M.O.; Wang, S. Adsorptive removal of antibiotic sulfonamide by UiO-66 and ZIF-67 for wastewater treatment. J. Colloid Interface Sci. 2017, 500, 88–95. [Google Scholar] [CrossRef]
  247. Yang, W.; Han, Y.; Li, C.; Zhu, L.; Shi, L.; Tang, W.; Wang, J.; Yue, T.; Li, Z. Shapeable three-dimensional CMC aerogels decorated with Ni/Co-MOF for rapid and highly efficient tetracycline hydrochloride removal. Chem. Eng. J. 2019, 375. [Google Scholar] [CrossRef]
  248. Yu, L.L.; Cao, W.; Wu, S.C.; Yang, C.; Cheng, J.H. Removal of tetracycline from aqueous solution by MOF/graphite oxide pellets: Preparation, characteristic, adsorption performance and mechanism. Ecotoxicol. Environ. Saf. 2018, 164, 289–296. [Google Scholar] [CrossRef]
  249. Sarker, M.; Song, J.Y.; Jhung, S.H. Adsorptive removal of anti-inflammatory drugs from water using graphene oxide/metal-organic framework composites. Chem. Eng. J. 2018, 335, 74–81. [Google Scholar] [CrossRef]
  250. Gautam, R.K.; Banerjee, S.; Sanroman, M.A.; Chattopadhyaya, M.C. Synthesis of copper coordinated dithiooxamide metal organic framework and its performance assessment in the adsorptive removal of tartrazine from water. J. Environ. Chem. Eng. 2017, 5, 328–340. [Google Scholar] [CrossRef]
  251. Huang, W.; Jing, C.; Zhang, X.; Tang, M.; Tang, L.; Wu, M.; Liu, N. Integration of plasmonic effect into spindle-shaped MIL-88A(Fe): Steering charge flow for enhanced visible-light photocatalytic degradation of ibuprofen. Chem. Eng. J. 2018, 349, 603–612. [Google Scholar] [CrossRef]
  252. Jiang, D.; Zhu, Y.; Chen, M.; Huang, B.; Zeng, G.; Huang, D.; Song, B.; Qin, L.; Wang, H.; Wei, W. Modified crystal structure and improved photocatalytic activity of MIL-53 via inorganic acid modulator. Appl. Catal. B Environ. 2019, 255. [Google Scholar] [CrossRef]
  253. Rasheed, H.U.; Lv, X.; Zhang, S.; Wei, W.; Ullah, N.; Xie, J. Ternary MIL-100(Fe)@Fe3O4/CA magnetic nanophotocatalysts (MNPCs): Magnetically separable and Fenton-like degradation of tetracycline hydrochloride. Adv. Powder Technol. 2018, 29, 3305–3314. [Google Scholar] [CrossRef]
  254. Li, R.; Chen, Z.; Cai, M.; Huang, J.; Chen, P.; Liu, G. Improvement of Sulfamethazine photodegradation by Fe (III) assisted MIL-53 (Fe)/percarbonate system. Appl. Surf. Sci. 2018, 457, 726–734. [Google Scholar] [CrossRef]
  255. Liu, N.; Huang, W.; Tang, M.; Yin, C.; Gao, B.; Li, Z.; Tang, L.; Lei, J.; Cui, L.; Zhang, X. In-situ fabrication of needle-shaped MIL-53(Fe) with 1T-MoS2 and study on its enhanced photocatalytic mechanism of ibuprofen. Chem. Eng. J. 2019, 359, 254–264. [Google Scholar] [CrossRef]
  256. Salimi, M.; Esrafili, A.; Jonidi Jafari, A.; Gholami, M.; Sobhi, H.R.; Nourbakhsh, M.; Akbari-Adergani, B. Photocatalytic degradation of cefixime with MIL-125(Ti)-mixed linker decorated by g-C3N4 under solar driven light irradiation. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 582. [Google Scholar] [CrossRef]
  257. Sun, J.; Feng, S.; Feng, S. composite with enhanced photocatalytic performance for ketoprofen. Inorg. Chem. Commun. 2020, 111, 107669. [Google Scholar] [CrossRef]
  258. Huo, Q.; Qi, X.; Li, J.; Liu, G.; Ning, Y.; Zhang, X.; Zhang, B.; Fu, Y.; Liu, S. Preparation of a direct Z-scheme A-Fe2O3/MIL-101(Cr) hybrid for degradation of carbamazepine under visible light irradiation. Appl. Catal. B Environ. 2019, 255. [Google Scholar] [CrossRef]
  259. Gao, Y.; Yu, G.; Liu, K.; Deng, S.; Wang, B.; Huang, J.; Wang, Y. Integrated adsorption and visible-light photodegradation of aqueous clofibric acid and carbamazepine by a Fe-based metal-organic framework. Chem. Eng. J. 2017, 330, 157–165. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram for the formation of the metal-organic framework (MOF) from metal ion and organic linker as precursors. Reproduced with permission from Reference [57].
Figure 1. The schematic diagram for the formation of the metal-organic framework (MOF) from metal ion and organic linker as precursors. Reproduced with permission from Reference [57].
Polymers 12 02648 g001
Figure 2. Interactions in adsorption of a contaminant (acid orange 7) onto the pores of MOFs. Reproduced with permission from Reference [76].
Figure 2. Interactions in adsorption of a contaminant (acid orange 7) onto the pores of MOFs. Reproduced with permission from Reference [76].
Polymers 12 02648 g002
Figure 3. Publications on the adsorption and photocatalytic degradation of some emerging pollutants using MOFs from 2010–2020. Data were obtained from science direct using the keywords; MOFs; adsorption; photocatalytic degradations; dyes, phenols; pesticides and herbicides; and pharmaceuticals and personal care products PPCPs.
Figure 3. Publications on the adsorption and photocatalytic degradation of some emerging pollutants using MOFs from 2010–2020. Data were obtained from science direct using the keywords; MOFs; adsorption; photocatalytic degradations; dyes, phenols; pesticides and herbicides; and pharmaceuticals and personal care products PPCPs.
Polymers 12 02648 g003
Figure 4. The mechanism for photocatalytic degradation of methyl orange and 4-nitrophenol using composite photocatalyst (MOF-199-NH2/BaWO4). Reproduced with permission from Reference [93].
Figure 4. The mechanism for photocatalytic degradation of methyl orange and 4-nitrophenol using composite photocatalyst (MOF-199-NH2/BaWO4). Reproduced with permission from Reference [93].
Polymers 12 02648 g004
Figure 5. Publications from 2010–2020 on the adsorption and photocatalytic degradation of dyes using MOFs. Data was obtained from the science direct using keywords MOFs; adsorption, and photocatalytic degradations dyes.
Figure 5. Publications from 2010–2020 on the adsorption and photocatalytic degradation of dyes using MOFs. Data was obtained from the science direct using keywords MOFs; adsorption, and photocatalytic degradations dyes.
Polymers 12 02648 g005
Figure 6. (a) Adsorption and photocatalytic degradations spectra of methyl orange dye using Zn and Co2+/Zn2+ metal-doped MOFs (M(tpbpc)(bdc)0.5·H2O) and (b) photographs of photocatalytic degradation of the dye using the MOFs under visible light irradiations. Reproduced with permission from Reference [97].
Figure 6. (a) Adsorption and photocatalytic degradations spectra of methyl orange dye using Zn and Co2+/Zn2+ metal-doped MOFs (M(tpbpc)(bdc)0.5·H2O) and (b) photographs of photocatalytic degradation of the dye using the MOFs under visible light irradiations. Reproduced with permission from Reference [97].
Polymers 12 02648 g006
Figure 7. Molecular structures of (a) Perfluorooctanoic acid (PFOA) and (b) Perfluorooctane sulfonates (PFOS).
Figure 7. Molecular structures of (a) Perfluorooctanoic acid (PFOA) and (b) Perfluorooctane sulfonates (PFOS).
Polymers 12 02648 g007
Figure 8. Diagram for the molecular docking simulation for adsorption of chrysene onto UiO-66(Zr) and NH2-UiO-66(Zr) MOFs (showing the pollutant in the inner pores of the UiO-66(Zr) and the outer pores of the NH2-UiO-66(Zr)). Reproduced with permission from Reference [187].
Figure 8. Diagram for the molecular docking simulation for adsorption of chrysene onto UiO-66(Zr) and NH2-UiO-66(Zr) MOFs (showing the pollutant in the inner pores of the UiO-66(Zr) and the outer pores of the NH2-UiO-66(Zr)). Reproduced with permission from Reference [187].
Polymers 12 02648 g008
Figure 9. Adsorption capacities of UiO-66(Zr), MOF-808(Fe), and MOF-802(Fe) for the removal of pharmaceutical drugs from water. Reproduced with permission from Reference [224].
Figure 9. Adsorption capacities of UiO-66(Zr), MOF-808(Fe), and MOF-802(Fe) for the removal of pharmaceutical drugs from water. Reproduced with permission from Reference [224].
Polymers 12 02648 g009
Figure 10. (a) Mechanism for photocatalytic degradation of ibuprofen using MIL-88(Fe) and Ag/AgCl@MIL-88(Fe) and (b) the reusability of the composites. Reproduced with permission from Reference [251].
Figure 10. (a) Mechanism for photocatalytic degradation of ibuprofen using MIL-88(Fe) and Ag/AgCl@MIL-88(Fe) and (b) the reusability of the composites. Reproduced with permission from Reference [251].
Polymers 12 02648 g010
Figure 11. Patents granted from 2010 to 2020 on the adsorption and photocatalytic degradation using MOFs-based materials of (a) some emerging pollutants and (b) dyes. Data obtained from the lens.org using keywords MOFs, adsorption, photocatalytic degradation, dyes, phenols, PPCPs, pesticides, and herbicides.
Figure 11. Patents granted from 2010 to 2020 on the adsorption and photocatalytic degradation using MOFs-based materials of (a) some emerging pollutants and (b) dyes. Data obtained from the lens.org using keywords MOFs, adsorption, photocatalytic degradation, dyes, phenols, PPCPs, pesticides, and herbicides.
Polymers 12 02648 g011
Table 1. MOFs reported for the adsorption of dyes.
Table 1. MOFs reported for the adsorption of dyes.
Type of MOFSynthesis MethodSurface Area (m2 g−1)PollutantsConcentration (mg L−1)% RemovalQe (mg g−1)Equilibrium TimeReusedRef
Fe-BTCSolvothermal877Orange II509220780 min4[52]
MIL-53(Fe)Solvothermal53Methyl orange100777760 min3[100]
MOF-235(Fe)Solvothermal-Methyl orange
Methylene blue
30-477
187
250 min-[101]
MIL-125(Ti)Solvothermal1108Crystal violet40-130180 min-[102]
MIL-101(Cr)Hydrothermal3514Methylene blue
Methyl red
30
300
-11
247
30 min
30 min
-
-
[103]
MIL-101(Cr)Microwave2410Reactive yellow
Reactive black
Reactive red
Reactive blue
300100386
377
390
397
24 h-[104]
MIL-100(Fe)
MIL-100(Cr)
Hydrothermal
Hydrothermal
17701760Methyl orange
Methylene blue
Methyl orange
Methylene blue
30
30
85
100
8
100
1045
736
212
645
3 days
22 days
-
-
[99]
MIL-101(Cr)
MIL-101(Cr)-SO3H
Hydrothermal
Hydrothermal
3016
1546
Fluorescein sodium Safranine Fluorescein sodium Safranine100
100
-
-
-
-
280
701
114
425
700 min
700 min
700 min
700 min
4
4
[105]
[105]
Cu-BTCHydrothermal521Methylene blue200 -96 40 min4 [106]
Cu-BTC MOF
Cu-BTC@GO
Cu-BTC@CNT
Fe3O4/Cu-BTC@GO
Solvothermal856508123176Methylene blue100-
-
-
-
67
152
172
136
12 h-[107]
Ce(III)-doped UiO-67Solvothermal1911Methylene blue
Congo red
Methyl orange
10095
96
399
800
401
80 min4
4
[98]
AlF-MOF
AlF-GO
AlF-rGO
Hydrothermal973
918
952
Congo red509993
102
179
30 min-[108]
NH2 -MIL-125(Ti)Solvothermal1350Basic blue
Methylene blue
Basic red
2093
97
99
1257
862
1296
30 min3[109]
NH2-UiO-66(Zr)Solvothermal954Methylene blue2008832115 min6[30]
UiO-66(Zr)Solvothermal1244Rhodamine Blue209190200 min5[110]
Zn-MOFRoom temp1046Methylene blue10 98326 60 min4 [111]
CPM-97(Fe)Solvothermal1397Congo red40 10083130 min3 [112]
MIL-53(Fe)Solvothermal23Methyl red100 787660 min3 [100]
MIL-101(Cr)Hydrothermal2664Xylenol orange4009015930 min3 [113]
BTB-MnSolvothermal3143Methylene blue1589308120 min6 [114]
NOTT-102(Cu)Solvothermal3006Methylene blue209785024 h3[115]
Ni-Zn-MOFSolvothermal57Congo red30-461300 min5[116]
Cu-MOF/Fe3O4Solvothermal34Malachite green509011460 min5[117]
Ni-MOF/GOBall milling70Congo red200-2489300 min-[118]
PEI-modified Cu-BTCHydrothermal785Congo red
Acid blue
1200
100
100
100
2578
132
200 min6
6
[78]
PED-MIL-101(Cr)
PED-MIL-101(Cr)
Hydrothermal3491
3296
Methyl orange
Methyl orange
50
50
NA
NA
160
194
250 min
250 min
3
3
[119]
Ac-HKUST-1Solvothermal-Crystal violet
Disulfine blue
Quinoline yellow
10
10
10
100
91
133
130
65
4 min-[120]
MIL-101(Fe)@PDopa@Fe3O4Solvothermal-Methyl red
Malachite green
100
100
92
100
833
1250
30 min
60 min
4
4
[121]
H6P2W18O62 /MOF-5Hydrothermal395Methylene blue20975210 min-[122]
Fe3O4@MIL-100(Fe)Solvothermal730Methylene blue208322124 h4 [123]
NENU/GOSolvothermal380Basic red 465881306 min-[124]
MIL-68(Al)/PVDFCasting-Methylene blue109661360 min6[125]
NH2-UiO-66(Zr)Solvothermal247Safranin13510039480 min4[126]
MIL-101(Cr)
TiO2-MIL-101(Cr)
Hydrothermal2361
531
Methylene blue20 -9
21
50 min-[127]
Table 2. MOFs and MOF composites for the photocatalytic degradation of dyes.
Table 2. MOFs and MOF composites for the photocatalytic degradation of dyes.
MOFSynthesis MethodSurface Area (m2 g−1)Bandgap (eV)PollutantsConcentration (mg L−1)Light Source(%) RemovalIrradiation TimeReusedRef
MIL-88(Fe)Hydrothermal-2.05Methylene blue32Visible-50 min4[132]
NH2-MIL-88(Fe)Microwave164-Methylene blue20Visible9860 min5[50]
MIL-100(Fe)Hydrothermal5-Basic blue 15Ultraviolet99180 min3[133]
MIL-125(Ti)Microwave-3.14Methylene blue-Visible97360 min-[134]
MIL-101(Fe)
MIL-100(Fe)
MIL-53(Fe)
MIL-88B(Fe)
Solvothermal
Solvothermal
Solvothermal
Solvothermal
2986
1798
965
19
-
-
-
-
Acid orange80 Visible95
88
62
23
120 min3[135]
MIL-53(Fe)
Ni-MIL-53(Fe)
Solvothermal300
480
2.59
2.24
Rhodamine blue14.4Visible81
91
180 min-[136]
MIL-101(Cr)
TiO2-MIL-101(Cr)
Hydrothermal2361
531
2.3
2.59
Methylene blue20Ultraviolet43
100
30 min-[127]
NH2-MIL-88B(Fe)Microwave164-Methylene blue20 Visible9845 min5[50]
NT/MIL-100(Fe)Hydrothermal1414-
-
Methylene blue
Rhodamine blue
-Visible99
94
180 min4[137]
PCN/MIL-100(Fe)Hydrothermal1252-Methylene blue
Rhodamine blue
10 Visible75
80
200 min-[131]
TiO2@MIL-101(Fe)Hydrothermal1919-Methyl orange150Ultraviolet9950 min-
NH2-MIL-125(Ti)
CQDs/NH2-MIL-125(Ti)
Hydrothermal487
198
2.43
2.33
Rhodamine blue10
10
Visible64
100
120 min
120 min
7
7
[90]
NH2-MIL-53(Al)
NH2-MIL-53(Al)/ RGO/PS
Hydrothermal 1051
95
2.7
2.4
Methylene blue30 Visible41
59
210 min3[138]
MIL-100(Fe)@Bi2S3Microwave7021.75Rhodamine blue10Visible9860 min4[139]
MOF-199Solvothermal3435.43Basic blue 20Ultraviolet-180 min-[140]
MOF-199
MOF-199-NH2/BaWO4
Hydrothermal-
-
3.2
3
Methyl orange10
10
Ultraviolet38
98
50 min-[141]
MOF-1Solvothermal-3.0Methyl violet10Ultraviolet74100 min-[142]
HU11(Pr)Solvothermal-3.3Crystal blue220Visible10024 h-[143]
UiO-66/g-C3N4Mechanical3842.72Methylene blue10Visible-180 min6[144]
Bi2MoO6/UiO-66(Zr)Hydrothermal7262.45Rhodamine blue10 Visible96120 min3[145]
In2S3/UiO-66(Zr)Solvothermal8021.4Methyl orange15 Visible9640 min5[146]
CdS@NH2-MIL-125(Ti)Solvothermal12472.36Rhodamine blue180Visible97120 min-[147]
Ag3VO4/Cu-MOF/GORoom temperature6-Acid blue10Visible100120 min3 [148]
BiVO4/Fe-MOF/GOMicrowave332.18Rhodamine blue15 Visible-60 min4 [1]
AgBr@HPU-4Room temperature--Methylene blue
Methyl orange
12.75
12.75
Visible95
92
60 min
120 min
5
5
[149]
BiVO4/MIL-53(Fe)Solvothermal332.18Rhodamine blue15 -60 min4[1]
Ag3PO4/AgBr/Ag-HKUST-1Solvothermal1 Methylene blue
Acid orange
Eosin red
15 Visible92
90
90
80 min3[150]
Ag3PO4/Bi2S3-HKUST-1Solvothermal-2.07Trypan blue
vesuvine
25 Visible98
99
25 min-[151]
MOF/CuWO4Hydrothermal8012.4Methylene blue10 Visible98135 min6[152]
QD/Eu-MOFRoom temperature-2.29Rhodamine blue2Ultraviolet9050 min-[153]
Resin/FeBTCHydrothermal-2.31Rhodamine blue
Methylene blue
400Visible99
67
30 min5[154]
MIL-53(Fe)Solvothermal-2.43Rhodamine blue1580 Visible85120 min5[155]
MIL-53(Fe)Solvothermal-3.87Methylene blue128 Visible9920 min5[156]
MIL-53(Fe)Solvothermal382.69Rhodamine blue10Visible-180 min-[157]
MIL-53(Fe)Solvothermal89-Orange green0.2Visible9890 min5[158]
MIL-100(Fe)@MIL-53(Fe)Sonochemical3151.84Methyl orange10Visible98180 min5[159]
[CoNi(m3-tp)2 (m2-pyz)2]
MOF/CuWO4
Hydrothermal1054
801
2.5
2.4
Methylene blue10Visible32
98
135 min6 [152]
UiO-66(Zr)
α-Fe2O3@UiO-66(Zr)
Solvothermal1487
1204
-Methylene blue128Visible-50 min3 [160]
UiO-66(Zr)
CuS/UiO-66(Zr)
Solvothermal-
-
3.5
2.01
Rhodamine blue10 Visible50
90
60 min3 [161]
NiFe2O4/MIL-53(Fe)Solvothermal43-Rhodamine blue4.7Visible95180 min-[162]
MIL-88(Fe)
TiO2NS@MIL-100(Fe)
Hydrothermal1670
725
2.6
2.87
Methylene blue50Visible-60 min4[163]
Table 5. Adsorptions of PPCPs onto MOFs and their composites.
Table 5. Adsorptions of PPCPs onto MOFs and their composites.
Type of MOFSynthesis MethodSurface Area (m2 g−1)PollutantsConcentration (mg L−1)% RemovalQe (mg g−1)Equilibrium TimeReusedRef
A100(Al) MOFCommercial630Carbamazepine
Ibuprofen
2
2
95
75
65
50
2 h
2 h
4[225]
NH2-MIL-68(In)Hydrothermal655p-arsanilic acid2077784 h4[226]
Fe3O4@MIL-100(Fe)Microwave1245Diclofenac100 2484 h-[227]
MIL-101 (Cr)
ED-MIL-101(Cr)
AMSA-MIL-101(Cr)
Hydrothermal3014
2322
2255
Naproxen
Clofibric
Naproxen
Clofibric
Naproxen
Clofibric
13
100
-131
315
93
105
154
347
2 h4[228]
PCN-134(Zr)Solvothermal756Diclofenac30--20 min-[229]
[Cu(BTTA)]n.2DMFSolvothermal Diclofenac
Chlorpromazine
Amodiaquine
1200
1000
1000
-
-
-
650
67
72
7.5 h
5 h
5 h
3[230]
[Zn2(fum)2(bpy)]
[Zn4O(bdc)3]
Mechanical
Solvothermal
-Amodiaquine25-0.5
48
3 h-[231]
[Zn6(IDC)4(OH)2(Hprz)2]nHydrothermal889Ampicillin
Amoxicillin
Cloxacillin
6093
88
89
-4 h4[232]
PCN-222(Zr)Solvothermal2917Chloramphenicol5009937058 sec-[233]
PCN-128Y(Zr)Solvothermal Tetracycline445640030 min-[234]
MIL-53(Al)Hydrothermal1401Dimetridazole409046710 min5 [3]
MOF-5Room temperature2510Tetracycline5097233 45 min-[235]
MIL-53(Cr)
MIL-53(Al)
Solvothermal500
500
Sulfonamide2099
98
0.4
0.4
1 h3
3
[236]
MIL-53(Fe)/Fe3O4.Solvothermal76Doxycycline300100320 30 min5[237]
MIL-101(Cr)
MIL-53(Cr)
Hydrothermal
Hydrothermal
2810
398
Clofibric acid
Carbamazepine
Clofibric acid
Carbamazepine
20 -144
35
137
31
1 h-[238]
MIL-101(Fe)
MIL-100(Fe)
MIL-53(Fe)
Hydrothermal
Hydrothermal
Solvothermal
253
1203
21
Tetracycline5055.1
44
11
52
43
12
40 min4[239]
Ni-MIL-53(Fe)Solvothermal-Doxycycline1508868412 h5[240]
MIL-101(Cr)
Urea-MIL-101(Cr)
Hydrothermal3030
1970
Dimetridazole10 -141
185
4 h4 [241]
Pd@MIL-100(Fe)Hydrothermal2102
MWCNT/NH2-MIL-53(Fe)Solvothermal126Tetracycline
Chlortetracycline
20 -
-
368
254
12 h4[242]
MWCNT/MIL-53(Fe)Solvothermal60Tetracycline
Oxytetracycline
Chlortetracycline
20 -364
326
181
10 h4[243]
UiO-66(Zr)
NH2-UiO-66(Zr)
Solvothermal1171
646
Ibuprofen
Naproxen
Ibuprofen
naproxen
9-
-
127
89
51
40
4 h
4 h
-[244]
UiO-66(Zr)
In2S3/UiO-66(Zr)
Solvothermal389
75
Tetracycline40-51
61
1 h3[245]
UiO-66(Zr)Solvothermal1155Sulfonamide100 -41710 min4[246]
Fe3O4/HKUST-1(Cu)Solvothermal328CiprofloxacinNorfloxacin2098
99
538
513
30 min10[46]
Zn(TDC)(4- BPMH)]n·n(H2O)Sonochemical235Dichlorophenol
Amoxicillin
50 99
99
-
-
3 h-
-
[196]
Ni/Co-MOF@CMCMicrowave-Tetracycline30 806255 min-[247]
MIL-68(Al)/GOHydrothermal1267Tetracycline50 -173 6 h3[248]
MIL-101(Cr)
GnO/MIL-101(Cr)
Hydrothermal-
3308
Naproxen
Ketoprofen
Naproxen
Ketoprofen
50 -112
80
171
140
12 h4[249]
Cu-DTORoom temperature120Tartrazine200 98255 40 min7[250]
Table 6. MOFs and composites employed for photocatalytic degradations of pharmaceutical drugs from wastewater.
Table 6. MOFs and composites employed for photocatalytic degradations of pharmaceutical drugs from wastewater.
MOFSynthesis MethodSurface Area (m2 g−1)Bandgap (eV)PollutantsConcentration (mg L−1)Light Source(%) RemovalIrradiation TimeReusedRef
MIL-53(Fe)Solvothermal18902.75Tetracycline10Visible972 h4[252]
MIL-101(Fe)
MIL-100(Fe)
MIL-53(Fe)
Hydrothermal
Hydrothermal
Solvothermal
253
1203
21
1.88
2.06
1.97
Tetracycline50Visible97
57
41
3 h [239]
MIL-100(Fe)@Fe3O4
MIL-100(Fe)@Fe3O4/CA
Hydrothermal725
389
2.49
1.76
Tetracycline10Visible42
85
3 h7[253]
MIL-88(Fe)
Ag/AgCl@MIL-88(Fe)
Solvothermal26
139
2.51
2.23
Ibuprofen10Visible45
93
3.5 h4[251]
CdS@NH2-MIL-125(Ti)Solvothermal13752.36Oxytetracycline180Visible-2 h 5[147]
MIL-101(Fe)
Pd@MIL-100(Fe)
Hydrothermal2006
2102
-Theophylline
Ibuprofen
Theophylline
Ibuprofen
20Visible88
92
100
100
2.5 h4[218]
UiO-66(Zr)
In2S3/UiO-66(Zr)
Solvothermal389
75
3.70
1.92
Tetracycline40Visible56
79
1 h3[245]
In2S3/UiO-66(Zr)Solvothermal482.2Tetracycline30Visible851 h5[146]
MIL-100(Fe)
Fe3O4@MIL-100(Fe)
Hydrothermal
Microwave
1766
1245
-
-
Diclofenac60visible100
99
-
-
[227]
Vis/MIL-53(Fe)/Fe(III)/SPCSolvothermal-2.91Sulfamethazine0.2Visible901 h-[254]
1T- MoS2@MIL-53(Fe)Solvothermal3370.7Ibuprofen10Visible1002 h5[255]
MIL-68(In)-NH2
g-C3N4/MIL-68(In)-NH2
Solvothermal659
537
2.81
2.65
Ibuprofen20Visible93
68
2 h-[248]
MIL-125ML
MIL-125ML/gCN
Solvothermal1001
725
2.86
2.68
Cefixime20Visible48
74
2 h4[256]
UiO-66-NH2
CNT/N-TiO2/UiO-66-NH2
Hydrothermal708
288
2.17Ketoprofen50Visible41
96
2 h-[257]
MIL-101(Cr)
α-Fe2O3/MIL-101(Cr)
Hydrothermal2518
949
3.25
3.62
Carbamazepine30Visible-
-
3 h4[258]
MIL-53(Fe)Solvothermal184-Clofibric acid
Carbamazepine
40Visible98
90
4 h4[259]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zango, Z.U.; Jumbri, K.; Sambudi, N.S.; Ramli, A.; Abu Bakar, N.H.H.; Saad, B.; Rozaini, M.N.H.; Isiyaka, H.A.; Jagaba, A.H.; Aldaghri, O.; et al. A Critical Review on Metal-Organic Frameworks and Their Composites as Advanced Materials for Adsorption and Photocatalytic Degradation of Emerging Organic Pollutants from Wastewater. Polymers 2020, 12, 2648. https://doi.org/10.3390/polym12112648

AMA Style

Zango ZU, Jumbri K, Sambudi NS, Ramli A, Abu Bakar NHH, Saad B, Rozaini MNH, Isiyaka HA, Jagaba AH, Aldaghri O, et al. A Critical Review on Metal-Organic Frameworks and Their Composites as Advanced Materials for Adsorption and Photocatalytic Degradation of Emerging Organic Pollutants from Wastewater. Polymers. 2020; 12(11):2648. https://doi.org/10.3390/polym12112648

Chicago/Turabian Style

Zango, Zakariyya Uba, Khairulazhar Jumbri, Nonni Soraya Sambudi, Anita Ramli, Noor Hana Hanif Abu Bakar, Bahruddin Saad, Muhammad Nur’ Hafiz Rozaini, Hamza Ahmad Isiyaka, Ahmad Hussaini Jagaba, Osamah Aldaghri, and et al. 2020. "A Critical Review on Metal-Organic Frameworks and Their Composites as Advanced Materials for Adsorption and Photocatalytic Degradation of Emerging Organic Pollutants from Wastewater" Polymers 12, no. 11: 2648. https://doi.org/10.3390/polym12112648

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

Article Metrics

Back to TopTop