Next Article in Journal
Near-Infrared Light Driven ZnIn2S4-Based Photocatalysts for Environmental and Energy Applications: Progress and Perspectives
Next Article in Special Issue
Advanced Stimuli-Responsive Structure Based on 4D Aerogel and Covalent Organic Frameworks Composite for Rapid Reduction in Tetracycline Pollution
Previous Article in Journal
A Series of Metal–Organic Frameworks with 2,2′-Bipyridyl Derivatives: Synthesis vs. Structure Relationships, Adsorption, and Magnetic Studies
Previous Article in Special Issue
Interfacial Synthesis of an Ultrathin Two-Dimensional Polymer Film via [2 + 2] Photocycloaddition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research on Improved MOF Materials Modified by Functional Groups for Purification of Water

School of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225002, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(5), 2141; https://doi.org/10.3390/molecules28052141
Submission received: 8 February 2023 / Revised: 18 February 2023 / Accepted: 22 February 2023 / Published: 24 February 2023
(This article belongs to the Special Issue Porous Organic Materials: Design and Applications)

Abstract

:
With the rapid development of urbanization and industrialization, water contamination has gradually become a big problem. Relevant studies show that adsorption is an efficient strategy to treat pollutants in water. MOFs are a class of porous materials with a three-dimensional frame structure shaped by the self-assembly of metal centers and organic ligands. Because of its unique performance advantages, it has become a promising adsorbent. At present, single MOFs cannot meet the needs, but the introduction of familiar functional groups on MOFs can promote the adsorption performance of MOFs on the target. In this review, the main advantages, adsorption mechanism, and specific applications of various functional MOF adsorbents for pollutants in water are reviewed. At the end of the article, we summarize and discuss the future development direction.

1. Introduction

After a protracted battle, humans have triumphantly changed the course of nature and created social economics. In the meantime, environmental destruction has posed a realistic threat to mankind’s existence and progression. Water is the source of life. Oceans cover about 70 percent of the Earth. However, the freshwater resources that can be directly used by people contain less than 1% of the total water in the world. It is worth noting that the quality of available freshwater resources for humans is deteriorating due to demographic and environmental factors. [1,2,3]. At present, every year in some poor and backward countries, about 2 million people suffer from related diseases and eventually die because of a lack of access to clean water [4].
In water, most of the pollutants can be divided into heavy metal ions, acid ions, organic dyes, organic poisons, and drugs, which, to a certain extent, will cause allergy symptoms, nausea, headaches, heart disease, and other symptoms, and even lead to death [5]. A number of strategies have been used to treat these pollutants. They include nanofiltration [6], solvent extraction [7,8], resin use [9], photocatalytic degradation [10,11], cooling, ion exchange, chemical precipitation, electrochemical deposition, reverse osmosis, biological treatment [12,13,14,15], and others. However, most of these technologies have certain shortcomings, such as expensive equipment, complicated operations, and a long experimental period. Fortunately, adsorption, as one of the most promising pollutant treatment methods, has attracted the research interest of many researchers.
For the sake of adsorbing pollutants in water, it is necessary to choose the right material. Metal-organic frame materials were first proposed in 1999 by Omar Yaghi and his team at the University of California, Berkeley [16]. The material has the advantages of low density, a large specific surface area [17], ideal porosity [18,19], big pore size, while being adjustable [20,21,22,23]. Therefore, it presents an attractive topic in the fields of catalytic conversion [24,25,26,27], gas storage [28,29,30], electrochemical biosensing [31,32,33], drug slow-release control [34], molecular recognition [35,36,37], magnetism [38], adsorption and separation of pollutants [39,40,41], and so on. In order to fully exploit the advantages of MOFs in water purification, researchers modified the functional groups on MOFs so as to further boost the adsorption performance of MOFs. All these MOFs have achieved remarkable results in water purification.

2. Several Advantages of MOFs Modified by Functional Groups as Adsorbents

The pristine MOFs have some disadvantages, such as low catalytic activity, electrical conductivity, and stability. These shortcomings have limited their achievements in water purification to some extent. These properties are enhanced by modifying the functional groups. However, the most attractive aspect is that they maintain several advantages of the pristine MOFs, which are crucial for the design of different MOFs for water purification and the removal of pollutants from water.

2.1. Large Specific Surface Area

Specific surface area means the overall area of the material per unit mass. One of the crucial characteristics of MOFs is their huge specific surface area. The larger the specific surface area, the better the surface properties (such as catalytic capacity, surface adsorption capacity, and surface activity) of the material. For example, Prof. Omar Yaghi et al., the founders of MOF materials, synthesized MOF-5 with a simple cubic structure using terephthalic acid (PTA) as the organic ligand and Zn as the metal center [42]. The material has a large surface area. By calculating the experimental data, we can work out that the specific surface area of MOF-5 is 2900 m2·g−1. After that, no matter what method researchers used to synthesize MOF-5, their average surface area was greater than 2000 m2/g. In 2004–2005, Gerard Ferey’s research group successively synthesized MIL-100 [43] and MIL-101 [44] by using Cr as the metal center and using 1,3,5-benzenetricarboxylic acid (BTC) and 1,2-benzenedicarboxylic acid (BDC) as ligands. Their specific surface areas are 3100 m2·g−1 and 5900 m2·g−1, accordingly. Their work can be considered to have opened a new chapter in the development of MOF materials. Now, some studies have shown that the modification of MOFs with appropriate functional groups is beneficial to the increase of specific surface area to a certain extent.

2.2. Porous and Adjustable Structure

A pore is a hollow area in a porous substance where guest molecules have been eliminated. For MOFs, there are many openings on the surface of functional materials. These pores are essential for the use of MOFs. This feature enables MOFs to be successfully applied to the storage of small molecular gases (H2 [45], CO2 [46], CH4 [47], H2O [48], etc.), gas separation (H2/CO2 [49], CO2/N2 [50], CO2/CH4 [51], C3H6/C3H8 [52], etc.), and pollutant adsorption. By utilizing various metal centers and ligands, we can modify the structure of MOF materials to produce a range of materials. Yue et al. synthesized [Zn3(btca)2(OH)2](guest)n materials, for example [53]. Taking advantage of the framework’s flexibility, this material can be highly adsorbed with CO2 in the presence of N2, He, and Ar. This example shows that MOFs are both porous and flexible (Figure 1). MOFs with modified functional groups can still maintain this property. There are still openings on the surface of the material, and different ligands can be selected to adjust the structure and modify different functional groups.

2.3. Coordinated Unsaturated Metal Sites

A coordinated unsaturated metal site is one where the outer orbitals of the metal site are not filled with electrons. We can also understand them as coordinated open metal sites. Unlike noble gases, it does not achieve an electron-stable structure [54]. MOFs contain abundant unsaturated coordination metal sites, which are not only easy to exchange but also can act as the lone pair electron interaction between Lewis acid and the oxygen atom in the C=O double bond, thus reducing the bond energy of the C=O double bond and promoting the shaping of unsaturated alcohols. Chen et al. synthesized a catalyzer based on PdAg@MIL-101 to catalyze the reactants [55]. Pd has great hydrogenation activity, MIL-101 has Lewis acidity, and the target product’s selectivity is significantly increased by Ag (Figure 2). The final catalytic effect is also satisfactory. After introducing different functional groups on the surface of the material, MOFs can still have coordinated unsaturated metal sites, further improving the catalytic activity of MOFs and accelerating the interaction between pollutants and MOFs.

3. Synthesis of MOFs Modified by Functional Groups

3.1. Direct Synthesis Using Functionalized Linkers

Functionalized MOFs can be synthesized directly by simply using linkers containing various functional groups. For example, we can synthesize materials of the UiO-66(Zr)s series using Zr as the metal center and various types of BDC (benzenedicarboxylate) as the organic ligand [56]. Cu-BTC-NH2 (BTC: 1,3,5-benzenetricarboxylate) can be synthesized by using Cu as the metal center and 2-amino-1,3,5-benzenetricarboxylate (BTC-NH2) as the organic ligand [57]. The direct synthesis method has the advantages of simplicity and convenience. However, the cost of linkers containing various functional groups leads to the high price of this method. In addition, there are two problems with the use of this method. On the one hand, the synthesis of functionalized MOFs is not easy due to the coordination reaction activity of some functional groups. On the other hand, the reaction products usually have low crystallinity or porosity.

3.2. Post-Synthetic Modification

Functionalized MOFs can also be obtained by post-modification methods after the synthesis of the pristine MOFs. The first approach is to take full advantage of the open metal sites on the MOFs. For example, we can use Lewis bases like aminoethanol, aminobutane, ethylenediamine, and others to graft onto the open metal sites of MOFs like MIL-101(Cr), MIL-100(Al, Cr, Fe, V), MOF-74s, and others, thereby introducing -NH2 [58,59]. It should be noted that some well-crystallized MOFs do not have open metal sites [60,61,62]. If they are functionalized using a post-modification approach, the effect is not that significant. An alternative approach is to functionalize MOFs through a multi-step process. This involves chemical reactions between pristine MOFs and organic ligands. For example, MOF-NH2 can react with butane sulfonic acid, oxalyl chloride, and hydroquinone chloride to introduce -(CH4)2-SO3H, -CO-COOH, and -CO-C6H4-COOH, respectively [63,64].

3.3. Other Means

In addition to the two main categories mentioned above, we can also use other methods to functionalize MOFs. We’ll take MIL-101(Cr) as an example. MIL-101(Cr) can be oxidized to introduce -NO2 on the benzene ring, and then MIL-101(Cr)-NO2 can be reduced further to MIL-101(Cr)-NH2 [65]. In addition, -SH introduced onto MIL-101(Cr) by grafting can be oxidized to -SO3H [66]. MIL-101(Cr)-SO3H, which is directly synthesized by monosodium 2-sulfoterephthalic acid, could further react with AgNO3 to produce MIL-101(Cr)-SO3Ag [67,68].

4. Mechanism of Contaminant Adsorption in MOFs Modified by Functional Groups

In this part, we mainly introduce the mechanisms of adsorption of pollutants by MOFs after the introduction of functional groups. The interactions are mainly divided into the following five types: Electrostatic interaction, hydrogen-bond interaction, π–π interaction, hydrophobic interaction, and Lewis acid–base interaction.

4.1. Electrostatic Interaction

Electrostatic action is one of the most familiar principles for pollutant treatment. Electrostatic force refers to the interaction between charged bodies at rest. This interaction usually occurs via the surfaces of an object. As adsorbents, MOFs and pollutants with opposite charges may attract each other, resulting in electrostatic interaction, so as to achieve the effect of water purification. It is worth noting that zeta potential is an important factor influencing electrostatic action [69]. It depends on the solution pH value. The higher the pH value, the more negative charge will be on the surface. Oppositely, the lower the pH value, the more positive charge will be on the surface. [70]. For example, Wang et al. synthesized UiO-66-NH2 for Pb2+ and Cd2+ adsorption. The higher the pH value, the stronger the adsorption capacity of the material for Pb2+ and Cd2+ [71]. This can be explained by the point of zero charge (PZC). By studying the diagram of pH and zeta potential, we can see that the value of PZC is 3.86. When pH > PZC, the MOFs are negatively charged, which may increase the electrostatic interaction between the surface of the MOFs and the cations, resulting in increased adsorption of Pb2+ and Cd2+. On the contrary, when pH < PZC, the adsorption of Pb2+ and Cd2+ by MOFs with positive charges decreases. Additionally, Ibrahim et al. used UiO-66-NH2@SiO2 to study its adsorption capacity for Cr2O72−, methyl orange (MO), and methylene blue (MB) [72]. The relational graph of zeta potential and pH indicates that the PZC of MOF material is 5.6. When the pH was adjusted to 5.4 (Cr2O72−), 3.5 (MO), and 9.7 (MB), UiO-66-NH2@SiO2 exhibited ideal adsorption efficiency for the three pollutants. Since both Cr2O72− and MO are negatively charged, lowering the pH promotes their adsorption by MOFs. MB is a cationic dye, so increasing the pH value can make MOFs negatively charged so that MB can be adsorbed by electrostatic interaction. Therefore, we need to reasonably regulate the solution’s pH value according to the properties of MOFs by introducing corresponding functional groups so as to achieve effective adsorption of pollutants.

4.2. Hydrogen-Bond Interaction

Hydrogen bonding is widely regarded as a vital principle in biochemical or organic communities [73]. The hydrogen-bond is a bond that is weaker than a chemical bond and stronger than intermolecular force. The intermolecular force, also known as van der Waals’ force, is a weak interaction that exists between molecules. It can be mainly divided into dispersion force, orientation force and induction force. (ΔG hydrogen-bond: 25–40 KJ·mol−1, ΔG chemical bond: >100 KJ·mol−1, ΔG van der Waals’ force: 2–20 KJ·mol−1). Hydrogen-bond is caused by the interaction of hydrogen atoms with highly electronegative atoms (O, N, S, F, etc.) [74]. This is usually done when dealing with organic pollutants. Functional groups containing hydrogen atoms modified on MOFs (such as -NH2, -OH, etc.) can, for example, hydrogen-bond with some heterocyclic compounds (such as pyridine) [75].

4.3. π–π. Interaction

The π–π interaction is related to the interaction between π orbitals in molecules. The π orbital is the molecular orbital produced by the formation of the π bond. This is a weak interaction that usually exists between aromatic rings. It often occurs between two molecules that are electron-deficient and electron-rich and is not a covalent bond interaction as important as hydrogen bonding [76]. Most of the organic ligands used to synthesize MOFs are compounds containing aromatic rings (such as terephthalic acid, phthalic acid, homophenic acid, etc.). They can achieve π–π interaction with pollutants in water containing aromatic rings to form a tight, layered structure that can effectively resist the influence of external factors, allowing the two molecules to stay together.

4.4. Hydrophobic Interaction

Hydrophobicity is a physical property of the surface of advanced materials. Hydrophobic materials are made up of molecules with long carbon chains and low water solubility. MOFs are a typical class of hydrophobic materials, most of which tend to be non-polar solvents (such as saturated hydrocarbons and benzene compounds). In the presence of MOFs, non-polar substances aggregate in the water solution and tend to exclude water molecules [77]. Using this effect, MOFs can be successfully applied to pollutant removal. For example, Lin et al. used HKUST-1 to remove oil droplets from solution, and the interpretation of the result showed that the adsorption capacity of MOF material to oil droplets reached 4000 mg·g−1 [78]. This is primarily due to the strong hydrophobic action of the benzene ring in the oil droplets and HKUST-1. Yang et al. reported a particularly hydrophobic porous material (FMOF-1) with high adsorption capacities for benzene, cyclohexane, n-hexane, p-xylene, and toluene. At 100% relative humidity, no adsorption of water was detected [79].

4.5. Lewis Acid–Base Interaction

Lewis’ theory of acid and base was first proposed by American chemist Gilbert N. Lewis. This theory defines that any substance that can accept electron pairs is called an acid, and any substance that can give them is called a base [80]. A Lewis acid–base reaction is a reaction in which a coordination covalent bond is formed between an electron-pair acceptor and an electron-pair donor. This theory plays a key function in the MOF-based adsorption course. For instance, we can modify some groups with Lewis basic properties (such as -NH2, -SH, etc.) on the original MOF material and capture heavy metal ions from water using the soft and hard acid–base theories [81,82].

5. Application of MOFs by Modified Functional Groups in Water Purification

In this part, we classify MOFs according to different functional groups and introduce the MOFs’ specific application in the purification of water after introducing common functional groups in detail. The corresponding adsorption results and adsorption mechanism are arranged in Table 1 for the convenience of readers.

5.1. N-Containing Group

5.1.1. -NH2

Liu et al. synthesized UiO-66-NH2-CS aerogel monomers with hierarchical structure by covalent cross-linking, which showed effective and stable adsorption of Pb2+ [83]. The aerogel was characterized by many pores and a low density. The adsorption of Pb2+ by the material was 102.03 mg·g−1. Among them, N and Pb2+ play a coordination role, and O plays a synergistic adsorption role to a certain extent. After three cycles, the material’s adsorption capacity remained at 90.12%. This strategy provides an effective and universal approach for using MOFs in the field of pollutant treatment (Figure 3).
Azmi et al. proposed a new synthesis method to prepare MIL-96, which has different crystal habits and a larger particle size [84]. After synthesizing the material, they added hydrolyzed polyacrylamide (HPAM) to increase the size of the material. Activated carbon (AC) was used as a reference adsorbent to examine the removal impact of MIL-96-RHPAM2 on perfluorooctanoic acid (PFOA). The results showed that MIL-96-RHPAM2 had a good PFOA adsorption capability of 340 mg·g−1. The amine group and the anionic carboxylate of PFOA in the HPAM skeleton generate electrostatic bonds, which contribute to the high adsorption of PFOA by MIL-96-RHPAM2. Samuel et al. synthesized Cu-BTC@NH2 composites and investigated their adsorption capacity for two drug contaminants, ibuprofen (IBF) and acetaminophen (ACE), in aqueous solutions [85]. The course of adsorption accords with the pseudo-second-order kinetic model and Langmuir adsorption, which reveal that chemisorption is the prime adsorption process. Based on hydrophobic and electrostatic effects, the adsorption capacities of IBF and ACE were 187.97 mg·g−1 and 125.45 mg·g−1, accordingly. The laboratory findings display that the approach can be applied to eliminate harmful drugs from toxic wastewater. In a study, Park et al. used MIL-101 to adsorb bisphenol S (BPS) in water and investigated the effectiveness of the introduction of -NH2 on adsorption [86]. The elimination capacity of BPS was greatly enhanced, and the adsorption capacity reached 513 mg·g−1 when the -NH2 group was added to MIL-101. The research reveals that hydrogen bonds exist in MIL-101-NH2, as well as O=S=O and -OH in BPS. The material can still have good adsorption characteristics after routine ethanol washing. The method also provides a novel horizon for the adsorption of organics containing sulfonyl in wastewater. Wu et al. compounded an adsorbent (UiO-66-NH2) that can efficiently remove Cr6+ from water [87]. According to the pseudo-second-order kinetic model, the initial stage of Cr6+ adsorption is rapid, and the subsequent adsorption phase is slow. The adsorption capacity of Cr6+ was determined by the Langmuir model to be 32.36 mg·g−1, and it occurred spontaneously at pH = 6.5. The electrostatic attraction between Cr6+ and UiO-66-NH2 may be the principle of Cr6+ adsorption by UiO-66-NH2. Roushani et al. demonstrated in an experiment that TMU-16-NH2 is a resultful adsorbent for the treatment of Cd2+ from water solutions [88]. The results show that the initial concentration and pH of the solution have a great influence on the adsorption, which is spontaneous. The adsorbent’s maximal Cd2+ adsorption capacity was 126.6 mg·g−1. Mechanism studies show that the amino group of TMU-16-NH2 is a good binding site for Cd2+, and the adsorption of Cd2+ is facilitated by the creation of coordination bonds between -NH2 groups and Cd2+. The performance of the adsorbent has been greatly improved compared with the results of research. Lv et al. used two MOFs (original MIL-68(In) and NH2-MIL-68(In)) to investigate the adsorption efficiency of typical organic arsenic compounds, arsine acid (p-ASA). [89]. The adsorption quantity of NH2-MIL-68(In) for p-ASA (401.6 mg·g−1) is greater than that of pure MIL-68(In), which can be attributed to hydrogen bonding and the π–π effect (Figure 4). In addition, the result of the adsorption of p-ASA is exothermic and spontaneous, conforming to the pseudo-second-order intraparticle diffusion model. The addition of -NH2 to MIL-68(In) accelerates adsorption performance by increasing hydrogen bonding. This research offers a crucial theoretical and experimental foundation for the use of MOFs in the elimination of organic arsenic compounds like p-ASA.
Roushani et al. investigated the adsorption and removal of the anionic dye methyl orange (MO) by two MOF materials (TMU-16-NH2 and TMU-16) [90]. Thermodynamic and kinetic outcomes display that the removal of MO by the two MOFs is a spontaneous pseudo-second-order adsorption, and the removal rate of MO by TMU-16-NH2 (393.7 mg·g−1) is superior to that of the original TMU-16 because the hydrogen-bond between MO and -NH2 is formed and the material itself is stable. The adsorption properties of these materials for anionic dyestuffs are stronger than those of other previously reported porous materials. Electrostatic interaction and hydrogen bonding both contribute to the understanding of the adsorption mechanism. Haque et al. prepared two metal-organic frameworks, NH2-MIL-101(Al) and MIL-101(Al) [91]. These two materials have been applied to dislodge methylene blue (MB) from aqueous solutions. The adsorption capacity of NH2-MIL-101(Al) on MB at 30 °C was up to 762 mg·g−1 due to the electrostatic interaction between the cationic dye MB and the amino group of MOFs. The original MIL-101(Al) exhibited lower adsorption capability, which explains why -NH2 was introduced. Liu et al. put forward a magnetic MOF substance for the adsorption of several cationic and anionic dye types [92]. Adsorption isotherm and thermodynamic studies showed that the maximum adsorption capabilities of NH2-MIL-101(Al) for indigo carmine (IC) and malachite green (MG) were 135 mg·g−1 and 274.4 mg·g−1, respectively. This is primarily the contribution of π–π interaction, electrostatic interaction, hydrogen bonding and hydrophobic interaction between magnetic NH2-MIL-101(Al) and organic dyes. The method has the characteristics of a short separation time and a high reuse rate, which is expected to make it a new adsorbent for adsorbing and removing dyes from aqueous solutions. Oveisi et al. used Ti as a metal source, 2-amino-1,4-phthalate ester (NH2-BDC), and 1,4-phthalate ester (BDC) as organic connectors to synthesize five kinds of materials in the MIL series, which are MIL-X1, MIL-X2, MIL-X3, NH2-MIL-125(Ti), and MIl-125(Ti) (the first three materials are obtained by adjusting the ratio of BDC to NH2-BDC, respectively) [93] (Figure 5). These materials were used to study the removal of the cationic dyes methylene blue (MB), basic blue 41 (BB41), and basic red 46 (BR46) from water. The adsorption of pollutants by the five materials of the MIL series conforms to the pseudo-second-order kinetic model and the Langmuir adsorption, and the synthesized MIL has high reuse and stability within the three cycle periods. Most importantly, the maximum electron density was found in NH2-MIL-125(Ti), so the adsorption performance is particularly ideal. The adsorption of MB, BB41, and BR46 is respectively 862, 1257, and 1296 mg·g−1.

5.1.2. -NMe3+

Wu et al. created a novel quaternary amine anion-exchange MOF (UiO-66-NMe3+) for the adsorption of 2,4-dichlorophenoxyacetic acid (2,4-d), a toxic pesticide that is extensively utilized [94]. UiO-66-NMe3+ had a maximum adsorption capacity for 2,4-d of 279 mg·g−1, which was much higher than UiO-66 and UiO-66-NH2. This is mainly attributed to the following two reasons. For one thing, the functionality of quaternary amine groups can effectively increase the electrostatic effect. For another, the π–π conjugation between 2,4-d and MOFs improves the adsorption properties of UiO-66-NMe3+. After seven cycles, the adsorption performance decreased only slightly. This technique offers the best alternative for effectively removing 2,4-d from complex water conditions by adsorption. Wei et al. created the functionalized quaternary ammonium salt MIL-101(Cr)-NMe3+ and took advantage of it as a novel adsorbent to wipe off diclofenac sodium (DCF) from wastewater [95]. The results showed that DCF quickly adsorbed onto MIL-101(Cr)-NMe3+. At 20 °C for 30 min, DCF can be adsorbed in large quantities (310.6 mg·g−1), which conforms to the pseudo-second-order kinetic model and the Langmuir adsorption. According to the adsorption mechanism analysis, the electrostatic interaction and π–π interaction between MIL-101(Cr)-NMe3+ and DCF are the reasons for the efficacious adsorption. Based on the results of the regeneration experiment, MIL-101(Cr)-NMe3+ could be recovered at least five times without suffering a substantial decrease in adsorption ability.
Based on the preceding discussion, we can conclude that MOFs modified with -NH2 or -NMe3+ can be useful in water purification. For organic matter in water (such as organic poisons, dyes, and drugs), the general adsorption mechanism focuses on hydrogen bonding and electrostatic interaction. This is very understandable. Because N is a highly electronegative atom, it can be used as the acceptor of H, thus forming hydrogen bonds and realizing the adsorption of pollutants. For heavy metal ions in water, -NH2 successfully acts as a Lewis base. Because metal ions are Lewis acids, they can interact with -NH2 in Lewis acid–base reactions to produce the adsorption effect.

5.2. S-Containing Group

5.2.1. -SH

Ke et al. selected the classic 3D Cu-MOF (HKUST-1) and used a simple post-synthesis strategy to achieve sulfhydryl functionalization of MOFs to investigate their application in removing Hg2+ from water [96]. They prepared a sequence of sulfhydryl-modified HKUST-1 by ligand-bonding the coordination unsaturated metal center in HKUST-1 to the -SH group in disulfide glycol. Under the same conditions, sulfhydryl-functionalized HKUST-1 showed very high adsorption capacity (714.29 mg·g−1) and fine adsorption affinity for Hg2+ in water. Nevertheless, unfunctionalized HKUST-1 demonstrated no adsorption for Hg2+. This method is considered to be a new and effective method for the treatment of heavy metal ions in water. Li et al. created sulfhydryl-modified MOFs (UiO-66-SH) by a simple tactic and used them for selectively extracting Hg2+ in aqueous solutions [97]. After the introduction of -SH, the rate of UiO-66-SH’s adsorption to Hg2+ was extremely rapid, showing high adsorption performance (3.91 mmol·g−1) in a broad pH range (2.3–8.0), and the adsorption efficiency remained high (> 90%) after seven regenerations. Even though there are other heavy metal ions (Cu2+, Mn2+, Ni2+, Cd2+, Ba2+, and Co2+), the adsorbent exhibited selective adsorption of Hg2+ (Figure 6). Density functional theory (DFT) calculations revealed that this was primarily due to the strong coordination between Hg2+ and -SH. It is worth mentioning that UiO-66-SH possesses a unique attraction for organic mercury forms as well. (PhHg+, EtHg+, and MeHg+). This method can provide a new idea for removing mercury from ecological environments.
By functionalizing MIL-88A with mercaptoethanol, Singh et al. created MIL-88A-SH, a new-style MOF-based adsorbent, and tested its removal of Hg2+ from water and Hg from air [98]. The absorption is constant in the pH range of 5.0–9.0, and the adsorption capacity of Hg2+ is very high, about 1111.1 mg·g−1. In addition, when there are other interfering metal ions (Cu2+, As3+, Zn2+, Cr6+, Pb2+, and Cd2+), MIL-88A-SH still shows good adsorption properties for Hg2+. For Hg in the air, about 45.6 mg·g−1 of Hg2+ is adsorbed. This is mainly because Hg was oxidized to Hg2+ and complexed with sulfhydryl groups during the adsorption process. These data indicate that MIL-88A-SH can be used as a highly effective adsorbent to solve mercury pollution. Zhang et al. made use of the hydrothermal method to create MOF-5. On this basis, mercaptan-functionalized MOFs (HS-mSi@MOF-5) were invented and applied to research the adsorption effects of Pb2+ and Cd2+ in water [99]. In view of the capacity and rate of adsorption, the adsorption performance of HS-mSi@MOF-5 (312.5 mg·g−1, 65.2 mg·g−1) is preferable to that of the original material (211 mg·g−1, 4.2 mg·g−1). In the water stability test, the unmodified MOF-5 will gradually dissolve in an acidic solution, but its stability increases after the introduction of -HS. Based on the mechanism of adsorption, the coordination interaction and electrostatic interaction between -HS and toxic metal ions are critical for the disposal of Cd2+ and Pb2+ (Figure 7).

5.2.2. -SO3H

Moradi et al. synthesized and used sulfonated MOFs supported on Fe3O4 NPs as a Fenton-like catalyst to remove methyl orange (MO) from water [100]. When the initial concentration is 100 mg·L−1, the incipient concentration of H2O2 is 40 mg·L−1, the microwave power is 500 W, and the pH value is 3.0, the rate of the elimination of MO is very fast. Only when the microwave radiation time is 6 min, as much as 99.9% of MO can be removed. Finally, microwave-induced Fe3O4@MIL-100(Fe)-SO3H degradation will be a wastewater dye removal technique with great promise. Hasan et al. first synthesized three Zr-MOF series materials (UiO-66-SO3H, UiO-66-NH2, and UiO-66) for the removal and adsorption of diclofenac sodium (DCF) in water and contrasted these three materials with activated carbon (AC) [101]. These three materials showed better adsorption performance than AC. The adsorption kinetics between UiO-66 and DCF are faster due to possible π–π interaction and electrostatic interaction. After the introduction of -SO3H, the kinetics and capacity (263 mg·g−1) of adsorption were significantly enhanced, which may result from the acid–base attraction of -SO3H in UiO-66-SO3H and -NH2 in DCF. Interestingly, however, the opposite trend was observed when UiO-66-NH2 was utilized as a sorbent. This is primarily due to the alkali-base rejection of -NH2 in DCF and UiO-66-NH2, on account of relevant results. Wang et al. immobilized -SO3H on the surface of MOFs to synthesize sulfonic acid-functionalized MOFs to remove Cd2+ from aqueous solutions [102]. Cu3(BTC)2-SO3H showed a relatively good absorption capacity of 88.7 mg·g−1, exceeding that of the reference adsorbent. Additionally, it has quick kinetics for the adsorption of Cd2+ from aqueous solutions that are 1–3 orders of magnitude higher than those of current adsorbent materials. Furthermore, despite the existence of other interfering ions, it indicates a strong selectivity for Cd2+ and is simple to regenerate and recycle without suffering a major reduction in Cd2+ adsorption capacity. Liu et al. focused on the adsorption behavior of UiO-66@mSi-SO3H, UiO-66@mSi-SH, and UiO-66 on Cd2+ [103]. In consideration of the data of the Elovich and the Langmuir models, the adsorption capacity can be UiO-66 < UiO-66@mSi-SH < UiO-66@mSi-SO3H. The theoretical maximum adsorption capacities of UiO-66@mSi-SO3H and UiO-66@mSi-SH for Cd2+ are 409.96 and 212 mg·g−1, respectively. In the presence of -SH and -SO3H, Cd2+ in water was replaced by adsorbents to accomplish the aim of getting rid of Cd2+. -SO3H has diverse and complex coordination forms, allowing UiO-66-mSi-SO3H to have an excellent adsorption capacity for removing Cd2+. Lv et al. synthesized a late-model magnetic nanocomposite with a core-shell structure of Fe3O4@UiO-66-SO3H by means of a viable stepwise assembly tactic and then made the most of it as an adsorbent to wipe off methylene blue (MB) from aqueous solutions [104]. A considerable affinity for MB may be shown in the synthesized Fe3O4@UiO-66-SO3H, which has a maximum adsorption capacity of 297.3 mg·g−1. The adsorption process is basically completed within 15 min. Because of the effect of charge selectivity, Fe3O4@UiO-66-SO3H has good reusability and can selectively adsorb MB from the mixed solution. The π–π interaction and electrostatic interaction are important during the adsorption process. (Figure 8).

5.2.3. -SO4

Kang et al. selected Zr-BTC-NH2-SO4 as an adsorbent to capture radioactive Ba2+ in nuclear wastewater [105] (Figure 9). By reason of the plentiful -SO4 groups in the skeleton, which are strong Ba2+ chelating groups, and the fact that the binding site is completely exposed, in comparison to most adsorbents, Zr-BTC-NH2-SO4 has a greater adsorption capacity of 181.8 mg·g−1. When the concentration of interfering ions is ten times that of Ba2+, the material still shows excellent selectivity. Most crucially, the course of adsorption is irreversible for Ba2+, which can availably avert secondary pollution. This work has contributed to the development of adsorbents for the handling of radioactive Ba2+ from nuclear effluent.
Peng et al. proposed a Ba2+ defect concept based on MOFs by introducing -SO4 functional groups into the pore structure of MOFs [106]. The functionalized MOFs can remove more than 90% of Ba2+ within 5 min, and the removal efficiency reaches 99% after equilibrium. Notably, the sulfate-based functionalized material showed a high Ba2+ absorption capacity of 131.1 mg·g−1, exceeding that of most reportorial adsorbents, and could selectively capture Ba2+ from outlet water. This paper offers a fresh viewpoint on environmental remediation and the removal of radioactive Ba2+ from nuclear effluent.
MOFs modified with functional groups containing S atoms are generally used in the catalytic field, but due to some special properties, they can also be used to remove pollutants from water sources [117]. For S-containing groups, -SH, -SO3H, and -SO4 are the most representative. Since these three are nucleophiles, they can be regarded as Lewis bases, which explains their ability to adsorb metal ions from water through the Lewis acid–base process. For -SO3H, when it is modified into organic molecules, it can increase the acidity of an aqueous solution, making the pH drop, thus promoting the electrostatic interaction between it and pollutants. In addition, -SO3H is a hydrophilic group, which can increase the water solubility of MOFs. For -SH, it can also produce the same effect as -SO3H, but the acidity is not that strong, so the electrostatic effect of -SH is weaker than that of -SO3H. In addition, thanks to the presence of S atoms, MOFs modified by functional groups can also form hydrogen bonds with organic matter, but the effect is less obvious than the previous two.

5.3. -OH and -COOH

5.3.1. -OH

Wu et al. synthesized MIL-88A nanorods by the hydrothermal method, which was first used to deal with As5+ in water [107]. The chemical process conforms to the Langmuir adsorption model and the pseudo-second-order kinetic model, indicating that the adsorption of MIL-88A on As5+ is chemical adsorption and monolayer adsorption. In consideration of its distinctive structure, which contains several -OH groups and has a high swelling to water, MIL-88A has a maximum adsorption capacity of 145 mg·g−1. These results indicate that MIL-88A can be regarded as a fantastic candidate for arsenic removal (Figure 10).
Seo et al. used MIL-101-(OH)2, MIL-101-OH, MIL-101-NH2, and MIL-101 to study the adsorption of ibuprofen, naproxen, and oxybenone in water [108]. Among them, naproxen adsorption is the most representative. The fact that MIL-101 with H donor functional groups like -OH and -NH2 is particularly effective in adsorbing naproxen may be because the O atom on the drug and the H atom in the adsorbent have formed hydrogen bonds. The more hydrogen bonds contained, the more significant the adsorption effect, which also explained why MIL-101-(OH)2 had the best adsorption effect on naproxen (185 mg·g−1). In addition, MIL-101-(OH)2 can be recovered multiple times by simply washing it with ethyl alcohol, suggesting its potential for adsorption-based removal of PPCPs in water. Song et al. explored the effectiveness of hydroxyl groups on the adsorption of 5 PPCPs (bisphenol A, p-chloro-m-xylenol, naproxen, ketoprofen, and triclosan) on MIL-101 [109]. The hydrogen-bond is considered to be the main mechanism of adsorption of PPCPs on MOFs, in which MOFs and PPCPs can be H donors and H acceptors, respectively. This result is based on the fact that the greater the number of H receptors in PPCPs and the greater the number of -OH groups in MOFs, then the higher the adsorption capacity. This phenomenon is particularly significant for the adsorption of MIL-101-(OH)3 to chloroxylenol, ketoprofen, and naproxen. Sun et al. adopted a hydrothermal approach to create a novel adsorbent UiO-66-(OH)2/GO nanomaterial [110]. The excellent adsorption feature of UiO-66-(OH)2/GO on tetracycline hydrochloride (TC) and methylene blue (MB) pollutants (37.96 mg·g−1, 96.69 mg·g−1) confirmed the excellent adsorption performance of UiO-66-(OH)2/Go. The Freundlich isothermal model and the pseudo-second-order kinetic model demonstrate that the adsorption is chemisorption and single-layer adsorption, and these adsorption processes are completely spontaneous. These findings imply that UiO-66-(OH)2/GO offers valuable information on the production of sorbent materials for purifying water.

5.3.2. -COOH

Zhu et al. produced a recyclable MOF-based nanoparticle, UiO-66-2COOH, and created biochar from readily available waste biomass to remediate Sb3+ contamination in water by combining the two [111]. The adsorption capacity was 56.49 mg·g−1 when pH = 9.1, at 70 °C for 4 h, the initial concentration was 100 mg·L−1, and the mass ratio of UiO-66-2COOH to biochar was 4:1. The Freundlich isotherm model and the pseudo-second-order kinetic model were also studied. In the presence of NO3, SO42−, Cl, and PO43−, the adsorbent has no influence on the adsorption of Sb3+. Ren et al. reported the stable double-group MOF (Zr-BTC-COOH-SO4), a skeleton containing strontium chelating groups (-SO4, -COOH) and consumingly ionized groups (-COOH) [112]. The greatest sorption was 67.5 mg·g−1 for less than 5 min. It is worth mentioning that solution pH and adsorption temperature have no obvious effect on the adsorption of Sr2+ by Zr-BTC-COOH-SO4. The adsorption mechanism is that -COOH results in prompt adsorption based on electrostatic interaction, and the introduction of -SO4 improves the sorption capacity to a great extent. These results indicate that Zr-BTC-COOH-SO4 is a probable candidate for Sr2+ removal. Soltani et al. prepared the functionalized nanocomposite LDH/UiO-66-(Zr)-(COOH)2 by a simple “green” synthesis scheme and employed it as a powerful adsorbent to dislodge Hg2+ and Ni2+ from aqueous media [113] (Figure 11). It was found that the optimum Langmuir sorption capacity was 509.8 mg·g−1 (Hg2+) and 439 mg·g−1 (Ni2+) under constant conditions. The existence of -COOH makes the surface of the material have ample active sites, which improves the adsorption performance to a certain extent. This strategy has contributed to the design of more environmentally friendly and safer heavy metal ion adsorbents.
From the above discussion, we know that -OH and -COOH are both Lewis bases, which can be combined with metal ions by Lewis acid–base action to form coordination covalent bonds. In addition, both -OH and -COOH contain O, which is a highly electronegative element and thus can form hydrogen bonds with many antibiotics (many of which contain O, N, F, Cl, etc.). Since the bond energy of the hydrogen-bond can be regarded as electrostatic interaction to some extent, electrostatic interaction is also the adsorption mechanism of pollutants adsorbed by -OH and -COOH in water systems. The presence of -OH and -COOH increased the active sites of chemical reactions and further increased the adsorption capacity.

5.4. Halogen Groups

Sini et al. prepared two Zr-MOF-based materials (UiO-66-(F4) and UiO-66) for selective adsorption of PFOA and PFOS from aqueous medium [114]. By characterizing MOFs, they determined how efficiently they absorbed PFOX. In comparison with UiO-66 without modifying the functional groups, UiO-66-(F4) has stronger F-F interaction in the cavity, which increases its adsorption affinity for fluorinated pollutants. Under certain conditions, the saturated adsorption capacity is approximately 470 mg·g−1. These findings imply that fluorinated organic contaminants can be pretty much eliminated from aqueous solutions by using Zr-MOFs as a platform. Amador et al. synthesized UiO-66-F4 and UiO-66 using Zr as the metal center and TFBDC (2, 3, 5, 6-tetrafluoro-1, 4-phenyl dicarboxylic acid) and BDC (1, 4-phenyl dicarboxylic acid) as organic joints [115]. These two materials showed excellent adsorption capacity for organic pollutants in water. The hydrophobic material can be obtained by adding -F to the ligand structure of MOFs, which changes the interaction between adsorbate and adsorbent from π–π interaction in the original UiO-66 to hydrophobic interaction in UiO-66-F4. For benzene, UiO-66-F4 (76.3 mg·g−1) displayed a more superior adsorption feature than UiO-66 (34.1 mg·g−1). Deng et al. synthesized a tubular adsorbent for capturing harmful pollutants in outlet water [116]. SCNU-Z1-Cl has a large specific surface area and high hydrolytic stability in the pH range of 4–10. Because SCNU-Z1-Cl has a large tubular channel and unaligned anions in the skeleton, it can adsorb some anions and anionic dyes by ion exchange. SCNU-Z1-Cl had adsorption capacities of 292, 318, 126, and 241 mg·g−1 on MnO4, ReO4, CrO42−, and Cr2O72−, respectively, which were the highest in the field of MOF/COF. The selectivity of the material remains high even in the presence of other anion species. For organic dyes, the adsorption capacities of Congo red, methyl blue, methyl orange, and acid orange A with different sizes were 585, 262, 285, and 180 mg·g−1, within 1 h (Figure 12). SCNU-Z1-Cl can separate organic coloring matter in three disparate patterns: kinetics-dependent selective adsorption, charge-dependent, and size-dependent. The above data fully demonstrate that SCNU-Z1-Cl is an ideal adsorbent that can make outstanding contributions to environmental materials.
At present, the modification of halogen groups on MOFs is rarely studied. Halogens have a strong electronegativity, which can produce an electrical induction effect and increase the hydrophobicity of molecules so as to obtain hydrophobic materials. When the material comes into contact with the solution, it can interact with the contaminants in the water and separate them.

6. Summary and Expectation

In summary, MOFs are very effective at adsorbing pollutants from water. On the one hand, the original MOFs themselves possess the features of a large specific surface area, a large and adjustable pore size, and well-coordinated unsaturated metal sites. For another, after modifying MOFs with functional groups (-NH2, -NMe3+, -SH, -SO3H, -SO4, -OH, -COOH, -F, -Cl), they give full play to the function of functional groups on the basis of retaining the original characteristics. The above two points further enhance the adsorption performance of MOFs on heavy metal ions, organic dyestuffs, organic poisons, drugs, and acid ions. The main adsorption mechanisms, which represent a critical effect in the specific adsorption process, are electrostatic interaction, π–π interaction, hydrogen bonding, Lewis acid–base interaction, and hydrophobic interaction.
Although there are satisfactory results in pollutant disposal by functional MOFs, we still have several expectations. It is hoped that these opinions will encourage the use of MOF materials in the water purification industry. Firstly, MOFs modified with functional groups require specific binding between specific groups and target objects, which requires priority to be given to group functionalization of reactants. However, there are many conditions for functionalization, and the original framework structure is easily destroyed after functionalization. The adsorption capacity of MOFs’ material on pollutants decreased or even disappeared. Therefore, we should find corresponding solutions. Secondly, the metal centers of MOFs (Cr, Co, Cu, etc.), the organic ligands (imidazoles, carboxylic acids, etc.), and the reaction solvents (DMF, DMAC, etc.) are toxic chemical reagents regardless of the introduction of functional groups. These chemicals are not only expensive but also prone to returning to the environment and causing secondary pollution. Therefore, it is necessary to find green synthesis methods to improve the problem. Thirdly, most of the MOF materials with functional groups need to undergo high temperature and high pressure reactions in the reactor, and the synthesis time is long. We should try to find a method that can synthesize MOFs under mild conditions. Fourthly, at present, the research is in the preliminary stage, and there are not many MOFs applied in water purification. Relevant researchers should continue to explore new MOF materials. Finally, treatment of pollutants with low concentrations is still difficult.

Author Contributions

Conceptualization, data curation, writing—original draft, J.L.; writing-review and editing, funding acquisition, project administration, resources, and supervision, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

Y.W. acknowledges the financial support of the National Natural Science Foundation of China (22174124), the Top-notch Academic Programs Project of Jiangsu Higher Education Institutions (TAPP), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, N.; Gevenga, C.; Echeverria, J. Managing Water for People and Nature. Science 2001, 292, 1071–1072. [Google Scholar] [CrossRef] [PubMed]
  2. Vorosmarty, C.J.; Green, P.; Salisbury, J.; Lammers, R.B. Global Water Resources: Vulnerability from Climate Change and Population Growth. Science 2000, 289, 284–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rolston, A. Social changes affect water quality too. Nature 2016, 536, 396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. 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]
  5. Shannon, M.A.; Bohn, P.W.; Elimelech, M.; Georgiadis, J.G.; Mariñas, B.J.; Mayes, A.M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301–310. [Google Scholar] [CrossRef]
  6. Nthunya, L.N.; Bopape, M.F.; Mahlangu, O.T.; Mamba, B.B.; Van der Bruggen, B.; Quist-Jensen, C.A.; Richards, H. Fouling, performance and cost analysis of membrane-based water desalination technologies: A critical review. J. Environ. Manag. 2022, 301, 113922. [Google Scholar] [CrossRef]
  7. Lv, Z.; Liang, C.; Cui, J.; Zhang, Y.; Xu, S. A facile route for the synthesis of mesoporous melamine-formaldehyde resins for hexavalent chromium removal. RSC Adv. 2015, 5, 18213–18217. [Google Scholar] [CrossRef]
  8. Li, L.L.; Feng, X.Q.; Han, R.P.; Zang, S.Q.; Yang, G. Cr(VI) removal via anion exchange on a silver-triazolate MOF. J. Hazard. Mater. 2017, 321, 622–628. [Google Scholar] [CrossRef]
  9. 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]
  10. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef]
  11. Tanhaei, M.; Mahjoub, A.R.; Safarifard, V. Energy-efficient sonochemical approach for the preparation of nanohybrid composites from graphene oxide and metal-organic framework. Inorg. Chem. Commun. 2019, 102, 185–191. [Google Scholar] [CrossRef]
  12. Shayegan, H.; Ali, G.A.M.; Safarifard, V. Amide-Functionalized Metal-Organic Framework for High Efficiency and Fast Removal of Pb(II) from Aqueous Solution. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3170–3178. [Google Scholar] [CrossRef]
  13. Sadegh, H.; Ali, G.A.M.; Gupta, V.K.; Makhlouf, A.S.H.; Shahryari-ghoshekandi, R.; Nadagouda, M.N.; Sillanpaa, M.; Megiel, E. The role of nanomaterials as effective adsorbents and their applications in wastewater treatment. J. Nanostruct. Chem. 2017, 7, 1–14. [Google Scholar] [CrossRef] [Green Version]
  14. Nthunya, L.; Masheane, M.; George, M.; Kime, M.-B.; Mhlanga, S. Removal of Fe and Mn From Polluted Water Sources in Lesotho Using Modified Clays. J. Water Chem. Technol. 2019, 41, 81–86. [Google Scholar] [CrossRef]
  15. Makgabutlane, B.; Nthunya, L.N.; Musyoka, N.; Dladla, B.S.; Nxumalo, E.N.; Mhlanga, S.D. Microwave-assisted synthesis of coal fly ash-based zeolites for removal of ammonium from urine. RSC Adv. 2020, 10, 2416–2427. [Google Scholar] [CrossRef]
  16. Furukawa, H.; Cordova, K.E.; O’ keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 974–987. [Google Scholar] [CrossRef] [Green Version]
  17. Kennedy, R.D.; Krungleviciute, V.; Clingerman, D.J.; Mondloch, J.E.; Peng, Y.; Wilmer, C.E.; Sarjeant, A.A.; Snurr, R.Q.; Hupp, J.T.; Yildirim, T.; et al. Carborane-Based Metal-Organic Framework with High Methane and Hydrogen Storage Capacities. Chem. Mater. 2013, 25, 3539–3543. [Google Scholar] [CrossRef] [Green Version]
  18. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, E.; Yazaydin, A.O.; Snurr, R.Q.; O’ keeffe, M.; Kim, J.; Yaghi, O.M. Ultrahigh Porosity in Metal-Organic Frameworks. Science 2010, 329, 424–428. [Google Scholar] [CrossRef] [Green Version]
  19. Slater, A.G.; Cooper, A.I. Function-led design of new porous materials. Science 2015, 348, 988–997. [Google Scholar] [CrossRef]
  20. Chen, J.W.; Li, K.; Yang, J.; Gu, J.L. Hierarchical large-pore MOFs templated from poly(ethylene oxide)-b-polystyrene diblock copolymer with tuneable pore sizes. Chem. Commun. 2022, 58, 10028–10031. [Google Scholar] [CrossRef]
  21. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’ keeffe, M.; Yaghi, O.M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 2002, 295, 469–472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Cui, X.L.; Chen, K.J.; Xing, H.B.; Yang, Q.W.; Krishna, R.; Bao, Z.B.; Wu, H.; Zhou, W.; Dong, X.L.; Han, Y.; et al. Pore chemistry and size control in hybrid porous materials for acetylene capture from ethylene. Science 2016, 353, 141–144. [Google Scholar] [CrossRef] [PubMed]
  23. Hwang, Y.K.; Hong, D.Y.; Chang, J.S.; Jhung, S.H.; Seo, Y.K.; Kim, J.; Vimont, A.; Daturi, M.; Serre, C.; Férey, G. Amine Grafting on Coordinatively Unsaturated Metal Centers of MOFs: Consequences for Catalysis and Metal Encapsulation. Angew. Chem. Int. Edit. 2010, 47, 4144–4148. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, M.H. Precise Construction of MOFs with Catalytic Functions: Selective Regulators for Hydrogenation Reactions. Acta Phys.-Chim. Sin. 2016, 32, 2645–2646. [Google Scholar]
  25. Ji, P.F.; Manna, K.; Lin, Z.K.; Feng, X.Y.; Urban, A.; Song, Y.; Lin, W.B. Single-Site Cobalt Catalysts at New Zr123-O)83-OH)82-OH)6 Metal-Organic Framework Nodes for Highly Active Hydrogenation of Nitroarenes, Nitriles, and Isocyanides. J. Am. Chem. Soc. 2017, 12, 12234–12242. [Google Scholar] [CrossRef]
  26. Lai, Q.; Zheng, L.R.; Liang, Y.Y.; He, J.P.; Zhao, J.X.; Chen, J.H. Metal-Organic-Framework-Derived Fe-N/C Electrocatalyst with Five-Coordinated Fe-Nx Sites for Advanced Oxygen Reduction in Acid Media. ACS Catal. 2017, 7, 1655–1663. [Google Scholar] [CrossRef]
  27. Shen, J.Q.; Liao, P.Q.; Zhou, D.D.; He, C.T.; Wu, J.X.; Zhang, W.X.; Zhang, J.P.; Chen, X.M. Modular and Stepwise Synthesis of a Hybrid Metal-Organic Framework for Efficient Electrocatalytic Oxygen Evolution. J. Am. Chem. Soc. 2017, 139, 1778–1781. [Google Scholar] [CrossRef]
  28. Qasem, N.A.A.; Ben-Mansour, R.; Habib, M.A. Efficient CO2 adsorptive storage using MOF-5 and MOF-177. Appl. Energ. 2017, 210, 317–326. [Google Scholar] [CrossRef]
  29. Zou, L.F.; Zhou, H.C. Hydrogen Storage in Metal-Organic Frameworks. In Nanostructured Materials for Next-Generation Energy Storage and Conversion; Chen, Y.P., Bashir, S., Liu, J.L., Eds.; Springer: Berlin, Heidelberg, Germany, 2017; Volume 22, pp. 143–170. [Google Scholar]
  30. Liu, D.M.; Wu, H.H.; Wang, S.Z.; Xie, Z.G.; Li, J.; Lin, W.B. A high connectivity metal-organic framework with exceptional hydrogen and methane uptake capacities. Chem. Sci. 2012, 3, 3032–3037. [Google Scholar] [CrossRef]
  31. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Duyne, R.P.V.; Hupp, J.T. Metal-Organic Framework Materials as Chemical Sensors. Chem. Rev. 2015, 112, 1105–1125. [Google Scholar] [CrossRef]
  32. Chen, B.; Yang, Y.; Zapata, F.; Lin, G.; Qian, G.; Lobkovsky, E.B. Luminescent Open Metal Sites within a Metal-Organic Framework for Sensing Small Molecules. Adv. Mater. 2010, 19, 1693–1696. [Google Scholar] [CrossRef]
  33. Liu, T.Z.; Rong, H.; Zhang, X.; Zhang, K.L.; Liu, Y.; Zhang, X.B.; Bai, R.Y.; Li, D.; Yang, Y.H. Metal-Organic Framework Nanomaterials as Novel Signal Probes for Electron Transfer Mediated Ultrasensitive Electrochemical Immunoassay. Anal. Chem. 2016, 88, 12516–12523. [Google Scholar] [CrossRef]
  34. Chen, Y.J.; Li, P.; Modica, J.A.; Drout, R.J.; Farha, O.K. Acid-Resistant Mesoporous Metal-Organic Framework toward Oral Insulin Delivery: Protein Encapsulation, Protection, and Release. J. Am. Chem. Soc. 2018, 140, 5678–5681. [Google Scholar] [CrossRef]
  35. Cychosz, K.A.; Wong-Foy, A.G.; Matzger, A.J. Liquid Phase Adsorption by Microporous Coordination Polymers: Removal of Organosulfur Compounds. J. Am. Chem. Soc. 2008, 130, 6938–6939. [Google Scholar] [CrossRef]
  36. Herm, Z.R.; Wiers, B.M.; Mason, J.A.; Baten, J.M.V.; Hudson, M.R.; Zajdel, P.; Brown, C.R.; Masciocchi, N.; Krishna, R.; Long, J.R. Separation of Hexane Isomers in a Metal-Organic Framework with Triangular Channels. Science 2013, 340, 960–964. [Google Scholar] [CrossRef] [Green Version]
  37. Das, S.; Xu, S.X.; Ben, T.; Qiu, S.L. Chiral Recognition and Separation by Chirality-Enriched Metal-Mrganic Frameworks. Angew. Chem. Int. Edit. 2018, 57, 8629–8633. [Google Scholar] [CrossRef]
  38. Ryu, D.W.; Lee, W.R.; Lee, J.W.; Yoon, J.H.; Kim, H.C.; Koh, E.K.; Hong, C.S. Magnetic metal-organic framework constructed from a paramagnetic metalloligand exhibiting a significant sorption and reversible magnetic conversions. Chem. Commun. 2009, 46, 8779–8781. [Google Scholar] [CrossRef]
  39. Lin, Y.C.; Kong, C.L.; Zhang, Q.J.; Chen, L. Metal-Organic Frameworks for Carbon Dioxide Capture and Methane Storage. Adv. Energy Mater. 2017, 7, 1601296. [Google Scholar] [CrossRef]
  40. Murray, L.J.; Dincǎ, M.; Long, J.R. Hydrogen storage in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1294–1314. [Google Scholar] [CrossRef]
  41. Yang, Q.Y.; Ma, L.L.; Zhong, C.L.; An, X.H.; Liu, D.H. Enhancement of CO2/N2 Mixture Separation Using the Thermodynamic Stepped Behavior of Adsorption in Metal-Organic Frameworks. J. Phys. Chem. C 2011, 115, 2790–2797. [Google Scholar] [CrossRef]
  42. Li, H.L.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O.M. Design and Synthesis of an Exceptionally Stable and Highly Porous Metal-Organic Framework. Nature 1999, 402, 276–279. [Google Scholar] [CrossRef] [Green Version]
  43. Férey, G.; Serre, C.; Draznieks, C.M.; Millange, F.; Surblé, S.; Dutour, J.; Margiolaki, I. A Hybrid Solid with Giant Pores Prepared by a Combination of Targeted Chemistry, Simulation, and Powder Diffraction. Angew. Chem. Int. Edit. 2004, 116, 6456–6461. [Google Scholar] [CrossRef]
  44. Férey, G.; Mellot-draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  45. Vitillo, J.G.; Regli, L.; Chavan, S.; Ricchiardi, G.; Spoto, G.; Dietzel, P.D.C.; Bordiga, S.; Zecchina, A. Role of Exposed Metal Sites in Hydrogen Storage in MOFs. J. Am. Chem. Soc. 2008, 130, 8386–8396. [Google Scholar] [CrossRef]
  46. Gautam, S.; Cole, D. CO2 Adsorption in Metal-Organic Framework Mg-MOF-74: Effects of Inter-Crystalline Space. Nanomaterials 2020, 10, 2274. [Google Scholar] [CrossRef]
  47. Lee, Y.R.; Cho, S.M.; Ahn, W.S. Effects of polydimethylsiloxane coating of Ni-MOF-74 on CH4 storage. Korean J. Chem. Eng. 2018, 35, 1542–1546. [Google Scholar] [CrossRef]
  48. Wang, G.F.; Graham, E.; Zheng, S.L.; Zhu, J.X.; Zhu, R.L.; He, H.P.; Sun, Z.M.; Mackinnon, I.D.R.; Xi, Y.F. Diatomite-Metal-Organic Framework Composite with Hierarchical Pore Structures for Adsorption/Desorption of Hydrogen, Carbon Dioxide and Water Vapor. Materials 2020, 13, 4700. [Google Scholar] [CrossRef]
  49. Schulte, Z.M.; Kwon, Y.H.; Han, Y.; Liu, C.; Li, L.; Yang, Y.H.; Jarvi, A.G.; Saxena, S.; Veser, G.; Johnson, J.K.; et al. H2/CO2 separations in multicomponent metal-adeninate MOFs with multiple chemically distinct pore environments. Chem. Sci. 2020, 11, 12807. [Google Scholar] [CrossRef]
  50. Habib, N.; Durak, O.; Zeeshan, M.; Uzun, A.; Keskin, S. A novel IL/MOF/polymer mixed matrix membrane having superior CO2/N2 selectivity. J. Membr. Sci. 2022, 658, 120712. [Google Scholar] [CrossRef]
  51. Altintas, C.; Keskin, S. Molecular Simulations of MOF Membranes and Performance Predictions of MOF/Polymer Mixed Matrix Membranes for CO2/CH4 Separations. ACS Sustain. Chem. Eng. 2019, 7, 2739–2750. [Google Scholar] [CrossRef]
  52. Yoon, J.W.; Kim, A.R.; Kim, M.J.; Yoon, T.U.; Kim, J.H.; Bae, Y.S. Low-temperature Cu(I) loading on a mesoporous Metal-Organic framework for adsorptive separation of C3H6/C3H8 mixtures. Microporous Mesoporous Mater. 2019, 279, 271–277. [Google Scholar] [CrossRef]
  53. Yue, Y.F.; Rabone, J.A.; Liu, H.J.; Mahurin, S.M.; Li, M.R.; Wang, H.L.; Lu, Z.L.; Chen, B.L.; Wang, J.H.; Fang, Y.X.; et al. A Flexible Metal-Organic Framework: Guest Molecules Controlled Dynamic Gas Adsorption. J. Phys. Chem. C 2015, 119, 9442–9449. [Google Scholar] [CrossRef]
  54. Liang, H.; Liu, R.P.; An, X.Q.; Hu, C.Z.; Zhang, X.W.; Liu, H.J. Bimetal-organic frameworks with coordinatively unsaturated metal sites for highly efficient Fenton-like catalysis. Chem. Eng. J. 2021, 414, 128669. [Google Scholar] [CrossRef]
  55. Chen, Y.Z.; Zhou, Y.X.; Wang, H.W.; Lu, J.L.; Uchida, T.; Xu, Q.; Yu, S.H.; Jiang, H.L. Multifunctional PdAg@MIL-101 for One-Pot Cascade Reactions: Combination of Host-Guest Cooperation and Bimetallic Synergy in Catalysis. ACS Catal. 2015, 5, 2062–2069. [Google Scholar] [CrossRef]
  56. Hu, Z.; Zhao, D. De facto methodologies toward the synthesis and scale-up production of UiO-66-type metal-organic frameworks and membrane materials. Dalton Trans. 2015, 44, 19018–19040. [Google Scholar] [CrossRef]
  57. Peikert, K.; Hoffmann, F.; Fröba, M. Amino substituted Cu3(BTC)2: A new metal-organic framework with a versatile functionality. Chem. Commun. 2012, 48, 11196–11198. [Google Scholar] [CrossRef]
  58. Hong, D.Y.; Hwang, Y.K.; Serre, C.; Ferey, G.; Chang, J.S. Porous chromium terephthalate MIL- 101 with coordinatively unsaturated sites: Surface functionalization, encapsulation, sorption and catalysis. Adv. Funct. Mater. 2009, 19, 1537–1552. [Google Scholar] [CrossRef]
  59. Chung, Y.G.; Haldoupis, E.; Bucior, B.J.; Haranczyk, M.; Lee, S.; Zhang, H.; Vogiatzis, K.D.; Milisavljevic, M.; Ling, S.; Camp, J.S. Advances, updates, and analytics for the computation-ready, experimental metal-organic framework database: CoRE MOF 2019. J. Chem. Eng. Data 2019, 64, 5985–5998. [Google Scholar] [CrossRef]
  60. Jeong, W.; Lim, D.-W.; Kim, S.; Harale, A.; Yoon, M.; Suh, M.P.; Kim, J. Modeling adsorption properties of structurally deformed metal-organic frameworks using structure-property map. Proc. Natl. Acad. Sci. USA 2017, 114, 7923–7928. [Google Scholar] [CrossRef] [Green Version]
  61. Ahmed, I.; Khan, N.A.; Jhung, S.H. Adsorptive denitrogenation of model fuel by functionalized UiO-66 with acidic and basic moieties. Chem. Eng. J. 2017, 321, 40–47. [Google Scholar] [CrossRef]
  62. Sarker, M.; Shin, S.; Jhung, S.H. Synthesis and functionalization of porous Zr-diaminostilbenedicarboxylate metal-organic framework for storage and stable delivery of ibuprofen. ACS Omega 2019, 4, 9860–9867. [Google Scholar] [CrossRef] [PubMed]
  63. Mondol, M.M.H.; Bhadra, B.N.; Jhung, S.H. Removal of nitrogen-containing compounds from microalgae derived biofuel by adsorption over functionalized metal organic frameworks. Fuel 2020, 280, 118622. [Google Scholar] [CrossRef]
  64. Sarker, M.; Song, J.Y.; Jhung, S.H. Carboxylic-acid-functionalized UiO-66-NH2: A promising adsorbent for both aqueous-and non-aqueous-phase adsorptions. Chem. Eng. J. 2018, 331, 124–131. [Google Scholar] [CrossRef]
  65. Bernt, S.; Guillerm, V.; Serre, C.; Stock, N. Direct covalent post-synthetic chemical modification of Cr-MIL-101 using nitrating acid. Chem. Commun. 2011, 47, 2838–2840. [Google Scholar] [CrossRef] [PubMed]
  66. Chung, Y.-M.; Kim, H.-Y.; Ahn, W.-S. Friedel-Crafts acylation of p-xylene over sulfonated zirconium terephthalates. Catal. Lett. 2014, 144, 817–824. [Google Scholar] [CrossRef]
  67. Huang, M.; Chang, G.; Su, Y.; Xing, H.; Zhang, Z.; Yang, Y.; Ren, Q.; Bao, Z.; Chen, B. A metal-organic framework with immobilized Ag(I) for highly efficient desulfurization of liquid fuels. Chem. Commun. 2015, 51, 12205–12207. [Google Scholar] [CrossRef]
  68. She, H.; Ma, X.; Chang, G. Highly efficient and selective removal of N-heterocyclic aromatic contaminants from liquid fuels in a Ag(I) functionalized metal-organic framework: Contribution of multiple interaction sites. J. Colloid Interface Sci. 2018, 518, 149–155. [Google Scholar] [CrossRef]
  69. Valadi, F.M.; Ekramipooya, A.; Gholami, M.R. Selective separation of Congo Red from a mixture of anionic and cationic dyes using magnetic-MOF: Experimental and DFT study. J. Mol. Liq. 2020, 318, 114051. [Google Scholar] [CrossRef]
  70. Amin, N.K. Removal of direct blue-106 dye from aqueous solution using new activated carbons developed from pomegranate peel: Adsorption equilibrium and kinetics. J. Hazard. Mater. 2009, 1–3, 52–62. [Google Scholar] [CrossRef]
  71. Wang, K.; Gu, J.W.; Yin, N. Efficient Removal of Pb(II) and Cd(II) Using NH2-Functionalized Zr-MOFs via Rapid Microwave-Promoted Synthes. Ind. Eng. Chem. Res. 2017, 56, 1880–1887. [Google Scholar] [CrossRef]
  72. Ibrahim, A.H.; El-Mehalmey, W.A.; Haikal, R.R.; Safy, M.E.A.; Amin, M.; Shatla, H.R.; Karakalos, S.G.; Alkordi, M.H. Tuning the Chemical Environment within the UiO-66-NH2 Nanocages for Charge-Dependent Contaminant Uptake and Selectivity. Inorg. Chem. 2019, 58, 15078–15087. [Google Scholar] [CrossRef]
  73. Abdollahi, N.; Moussavi, G.; Giannakis, S. A review of heavy metals’ removal from aqueous matrices by Metal-Organic Frameworks (MOFs): State-of-the art and recent advances. J. Environ. Chem. Eng. 2022, 10, 107394. [Google Scholar] [CrossRef]
  74. Planas, N.; Dzubak, A.L.; Poloni, R.; Lin, L.-C.; McManus, A.; McDonald, T.M.; Neaton, J.B.; Long, J.R.; Smit, B.; Gagliardi, L. The Mechanism of Carbon Dioxide Adsorption in an Alkylamine-Functionalized Metal-Organic Framework. J. Am. Chem. Soc. 2013, 135, 7402–7405. [Google Scholar] [CrossRef]
  75. Hasan, Z.; Tong, M.; Jung, B.K.; Ahmed, I.; Zhong, C.; Jhung, S.H. Adsorption of Pyridine over Amino-Functionalized Metal-Organic Frameworks: Attraction via Hydrogen Bonding versus Base–Base Repulsion. J. Phys. Chem. C 2014, 118, 21049–21056. [Google Scholar] [CrossRef]
  76. Ahmed, I.; Jhung, S.H. Adsorptive denitrogenation of model fuel with CuCl-loaded metal-organic frameworks (MOFs). Chem. Eng. J. 2014, 251, 35–42. [Google Scholar] [CrossRef]
  77. Huo, S.-H.; Yan, X.-P. Metal-organic framework MIL-100(Fe) for the adsorption of malachite green from aqueous solution. J. Mater. Chem. 2012, 22, 7449–7455. [Google Scholar] [CrossRef]
  78. Lin, K.-Y.A.; Yang, H.; Petit, C.; Hsu, F.-K. Removing oil droplets from water using a copper-based metal organic frameworks. Chem. Eng. J. 2014, 249, 293–301. [Google Scholar] [CrossRef]
  79. Yang, C.; Kaipa, U.; Mather, Q.Z.; Wang, X.; Nesterov, V.; Venero, A.F.; Omary, M.A. Fluorous Metal-Organic Frameworks with Superior Adsorption and Hydrophobic Properties toward Oil Spill Cleanup and Hydrocarbon Storage. J. Am. Chem. Soc. 2011, 133, 18094–18097. [Google Scholar] [CrossRef]
  80. Ania, C.O.; Bandosz, T.J. Importance of Structural and Chemical Heterogeneity of Activated Carbon Surfaces for Adsorption of Dibenzothiophene. Langmuir 2005, 21, 7752–7759. [Google Scholar] [CrossRef]
  81. Xiong, J.; Zhu, W.; Li, H.; Ding, W.; Chao, Y.; Wu, P.; Xun, S.; Zhang, M.; Li, H. Few-layered graphene-like boron nitride induced a remarkable adsorption capacity for dibenzothiophene in fuels. Green Chem. 2015, 17, 1647–1656. [Google Scholar] [CrossRef]
  82. Khan, N.A.; Hasan, Z.; Jhung, S.H. Ionic liquids supported on metal- organic frameworks: Remarkable adsorbents for adsorptive desulfurization. Chem. Eur. J. 2014, 20, 376–380. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, Q.; Li, S.S.; Yu, H.H.; Zeng, F.M.; Li, X.; Su, Z.M. Covalently crosslinked zirconium-based metal-organic framework aerogel monolith with ultralow-density and highly efficient Pb(II) removal. J. Colloid Interface Sci. 2020, 561, 211–219. [Google Scholar] [CrossRef] [PubMed]
  84. Azmi, L.H.M.; Williams, D.R.; Ladewig, B.P. Polymer-Assisted Modification of Metal-Organic Framework MIL-96(Al): Influence on Particle Size, Crystal Morphology and Perfluorooctanoic Acid (PFOA) Removal. Chemosphere 2020, 262, 128072. [Google Scholar] [CrossRef] [PubMed]
  85. Samuel, M.S.; Savunthari, K.V.; Chandrasekar, N.; Balaji, R.; Selvarajan, E. Removal of environmental contaminants of emerging concern using metal-organic framework composite. Environ. Technol. Innov. 2022, 25, 102216. [Google Scholar] [CrossRef]
  86. Park, J.M.; Jhung, S.H. A remarkable adsorbent for removal of bisphenol S from water: Aminated metal-organic framework, MIL-101-NH2. Chem. Eng. J. 2020, 396, 125224. [Google Scholar] [CrossRef]
  87. Wu, S.; Ge, Y.; Wang, Y.; Chen, X.; Li, F.; Xuan, H.; Li, X. Adsorption of Cr (VI) on nano UiO-66-NH2 MOFs in water. Environ. Technol. 2018, 39, 1937–1948. [Google Scholar] [CrossRef]
  88. Roushani, M.; Saedi, Z.; Baghelani, Y.M. Removal of cadmium ions from aqueous solutions using TMU-16-NH2 metal organic framework. Environ. Nanotechnol. Monit. Manag. 2017, 7, 89–96. [Google Scholar]
  89. Lv, Y.C.; Zhang, R.S.; Zeng, S.L.; Liu, K.Y.; Huang, S.Y.; Liu, Y.F.; Xu, P.F.; Lin, C.X.; Cheng, Y.J.; Liu, M.H. 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]
  90. Roushani, M.; Saedi, Z.; Beygi, T.M. Anionic dyes removal from aqueous solution using TMU-16 and TMU-16-NH2 as isoreticular nanoporous metal organic frameworks. J. Taiwan Inst. Chem. Eng. 2016, 66, 164–171. [Google Scholar] [CrossRef]
  91. Haque, E.; Lo, V.; Minett, A.I.; Harris, A.T.; Church, T.L. Dichotomous adsorption behaviour of dyes on an amino-functionalised metal-organic framework, amino-MIL-101(Al). J. Mater. Chem. A 2014, 2, 193. [Google Scholar] [CrossRef]
  92. Liu, H.C.; Chen, L.G.; Ding, J. Adsorption behavior of magnetic amino-functionalized metal-organic framework for cationic and anionic dyes from aqueous solution. RSC Adv. 2016, 6, 48884. [Google Scholar] [CrossRef]
  93. 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]
  94. Wu, G.G.; Ma, J.P.; Li, S.; Wang, S.S.; Jiang, B.; Luo, S.Y.; Li, J.H.; Wang, X.Y.; Guan, Y.F.; Chen, L.X. Cationic metal-organic frameworks as an efficient adsorbent for the removal of 2,4-dichlorophenoxyacetic acid from aqueous solutions. Environ. Res. 2020, 186, 109542. [Google Scholar] [CrossRef]
  95. Wei, Y.; Wei, C.; Xia, Y. Efficient adsorption and removal of diclofenac sodium from water with quaternary ammonium functionalized metal-organic frameworks. Chin. J. Chromatogr. 2018, 36, 222–229. [Google Scholar]
  96. Ke, F.; Qiu, L.G.; Yuan, Y.P.; Peng, F.M.; Jiang, X.; Xie, A.J.; Shen, Y.H.; Zhu, J.F. Thiol-functionalization of metal-organic framework by a facile coordination-based postsynthetic strategy and enhanced removal of Hg2+ from water. J. Hazard. Mater. 2011, 196, 36–43. [Google Scholar] [CrossRef]
  97. Li, J.; Liu, Y.; Ai, Y.J.; Alsaedi, A.; Hayat, T.; Wang, X.K. Combined experimental and theoretical investigation on selective removal of mercury ions by metal organic frameworks modified with thiol groups. Chem. Eng. J. 2018, 354, 790–801. [Google Scholar] [CrossRef]
  98. Singh, N.; Srivastava, I.; Dwivedi, J.; Sankararamakrishnan, L. Ultrafast removal of ppb levels of Hg(II) and volatile Hg(0) using post modified metal organic framework. Chemosphere 2020, 270, 129490. [Google Scholar] [CrossRef]
  99. Zhang, J.M.; Xiong, Z.H.; Li, C.; Wu, C.S. Exploring a thiol-functionalized MOF for elimination of lead and cadmium from aqueous solution. J. Mol. Liq. 2016, 221, 43–50. [Google Scholar] [CrossRef]
  100. Moradi, S.E.; Dadfarnia, S.; Shabani, A.M.H.; Emami, S. Microwave-enhanced Fenton-like degradation by surface-modified metal-organic frameworks as a promising method for removal of dye from aqueous samples. Turk. J. Chem. 2017, 41, 426–439. [Google Scholar] [CrossRef]
  101. Hasan, Z.; Khan, N.A.; Jhung, S.H. Adsorptive removal of diclofenac sodium from water with Zr-based metal-organic frameworks. Chem. Eng. J. 2016, 284, 1406–1413. [Google Scholar] [CrossRef]
  102. Wang, Y.; Ye, G.Q.; Chen, H.H.; Hu, X.Y.; Niu, Z.; Ma, S.Q. Functionalized metal-organic framework as a new platform for efficient and selective removal of cadmium(II) from aqueous solution. J. Mater. Chem. A 2015, 3, 15292–15298. [Google Scholar] [CrossRef]
  103. Liu, J.M.; Cui, H.H.; Li, J.H.; Chen, M.C. A research on the cadmium ions adsorption of Sulfhydryl- and sulfo-functionalized UiO-66 with silica layer from water. J. Environ. Chem. Eng. 2021, 9, 104621. [Google Scholar] [CrossRef]
  104. Lv, S.W.; Liu, J.M.; Li, C.Y.; Ma, H.; Wang, Z.H.; Zhao, N.; Wang, S. Fabrication of Fe3O4@UiO-66-SO3H core-shell functional adsorbents for highly selective and efficient removal of organic dyes. New J. Chem. 2019, 43, 7770. [Google Scholar] [CrossRef]
  105. Kang, C.F.; Peng, Y.G.; Tang, Y.Z.; Huang, H.L.; Zhong, C.L. Sulfate-Rich Metal-Organic Framework for High Efficiency and Selective Removal of Barium from Nuclear Wastewater. Ind. Eng. Chem. Res. 2017, 56, 13866–13873. [Google Scholar] [CrossRef]
  106. Peng, Y.G.; Huang, H.L.; Liu, D.H.; Zhong, C.L. Radioactive Barium Ion Trap Based on Metal-Organic Framework for Efficient and Irreversible Removal of Barium from Nuclear Wastewater. ACS Appl. Mater. Interfaces 2016, 8, 8527–8535. [Google Scholar] [CrossRef]
  107. Wu, H.; Ma, M.D.; Gai, W.Z.; Yang, H.X.; Zhou, J.G.; Cheng, Z.X.; Xu, P.G.; Deng, Z.Y. Arsenic removal from water by metal-organic framework MIL-88A microrods. Environ. Sci. Pollut. Res. 2018, 25, 27196–27202. [Google Scholar] [CrossRef]
  108. 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]
  109. Song, J.Y.; Jhung, S.H. Adsorption of pharmaceuticals and personal care products over metal-organic frameworks functionalized with hydroxyl groups: Quantitative analyses of H-bonding in adsorption. Chem. Eng. J. 2017, 322, 366–374. [Google Scholar] [CrossRef]
  110. Sun, Y.Z.; Chen, M.; Liu, H.; Zhu, Y.; Wang, D.B.; Yan, M. Adsorptive removal of dye and antibiotic from water with functionalized zirconium-based metal organic framework and graphene oxide composite nanomaterial UiO-66-(OH)2/GO. Appl. Surf. Sci. 2020, 525, 146614. [Google Scholar] [CrossRef]
  111. Zhu, G.C.; Lin, J.L.; Yuan, Q.; Wang, X.F.; Zhao, Z.L.; Hursthouse, A.S.; Wang, Z.H.; Li, Q.B. A biochar supported magnetic metal organic framework for the removal of trivalent antimony. Chemosphere 2021, 282, 131068. [Google Scholar] [CrossRef]
  112. Ren, L.; Zhao, X.D.; Liu, B.S.; Huang, H.L. Synergistic effect of carboxyl and sulfate groups for effective removal of radioactive strontium ion in a Zr-metal-organic framework. Water Sci. Technol. 2021, 83, 2001–2011. [Google Scholar] [CrossRef]
  113. Soltani, R.; Pelalak, R.; Pishnamazi, M.; Marjani, A.; Shirazian, S. A water-stable functionalized NiCo-LDH/MOF nanocomposite: Green synthesis, characterization, and its environmental application for heavy metals adsorption. Arab. J. Chem. 2021, 14, 103052. [Google Scholar] [CrossRef]
  114. Sini, K.; Bourgeois, D.; Idouhar, M.; Carboni, M.; Meyer, D. Metal-organic framework sorbents for the removal of perfluorinated compounds in an aqueous environment. New J. Chem. 2018, 42, 17889–17894. [Google Scholar] [CrossRef]
  115. Amador, R.N.; Cirre, L.; Carboni, M.; Meyer, D. BTEX removal from aqueous solution with hydrophobic Zr metal organic frameworks. J. Environ. Manag. 2018, 214, 17–22. [Google Scholar] [CrossRef]
  116. Deng, S.-Q.; Mo, X.-J.; Zheng, S.-R.; Jin, X.; Gao, Y.; Cai, S.-L.; Fan, J.; Zhang, W.-G. Hydrolytically Stable Nanotubular Cationic Metal-Organic Framework for Rapid and Efficient Removal of Toxic Oxo-Anions and Dyes from Water. Inorg. Chem. 2019, 58, 2899–2909. [Google Scholar] [CrossRef]
  117. Chen, J.; Li, K.; Chen, L.; Liu, R.; Huang, X.; Ye, D. Conversion of fructose into 5-hydroxymethylfurfural catalyzed by recyclable sulfonic acid-functionalized metal-organic frameworks. Green Chem. 2014, 16, 2490–2499. [Google Scholar] [CrossRef]
Figure 1. Selective adsorption of CO2 by [Zn3(btca)2(OH)2](guest)n. Reprinted with permission from Ref. [53]. Copyright 2015, American Chemical Society.
Figure 1. Selective adsorption of CO2 by [Zn3(btca)2(OH)2](guest)n. Reprinted with permission from Ref. [53]. Copyright 2015, American Chemical Society.
Molecules 28 02141 g001
Figure 2. Schematic illustration showing the multi-step reaction over PdAg@MIL-101 involving Pd/Ag sites and Lewis acid. Reprinted with permission from Ref. [55]. Copyright 2015, American Chemical Society.
Figure 2. Schematic illustration showing the multi-step reaction over PdAg@MIL-101 involving Pd/Ag sites and Lewis acid. Reprinted with permission from Ref. [55]. Copyright 2015, American Chemical Society.
Molecules 28 02141 g002
Figure 3. (ac) Synthesis route of UiO-66-NH2-CS aerogel (d) Adsorption of Pb2+ by UiO-66-NH2-CS (e) Synthesis equation of UiO-66-NH2-CS. Reprinted with permission from Ref. [83]. Copyright 2019, Elsevier.
Figure 3. (ac) Synthesis route of UiO-66-NH2-CS aerogel (d) Adsorption of Pb2+ by UiO-66-NH2-CS (e) Synthesis equation of UiO-66-NH2-CS. Reprinted with permission from Ref. [83]. Copyright 2019, Elsevier.
Molecules 28 02141 g003
Figure 4. Structural formula of NH2-MIL-68(In) and main adsorption mechanism of p-ASA by NH2-MIL-68(In). Reprinted with permission from Ref. [89]. Copyright 2018, Elsevier.
Figure 4. Structural formula of NH2-MIL-68(In) and main adsorption mechanism of p-ASA by NH2-MIL-68(In). Reprinted with permission from Ref. [89]. Copyright 2018, Elsevier.
Molecules 28 02141 g004
Figure 5. Synthesis and structure diagram of NH2-MIL-125(Ti). Reprinted with permission from Ref. [93]. Copyright 2017, Elsevier.
Figure 5. Synthesis and structure diagram of NH2-MIL-125(Ti). Reprinted with permission from Ref. [93]. Copyright 2017, Elsevier.
Molecules 28 02141 g005
Figure 6. Schematic diagram of UiO-66-SH selective adsorption of Hg2+ at different pH values. Reprinted with permission from Ref. [97]. Copyright 2018, Elsevier.
Figure 6. Schematic diagram of UiO-66-SH selective adsorption of Hg2+ at different pH values. Reprinted with permission from Ref. [97]. Copyright 2018, Elsevier.
Molecules 28 02141 g006
Figure 7. Schematic diagram of the strength of HS-mSi@MOF-5 on metal ions at different pH values. Reprinted with permission from Ref. [99]. Copyright 2016, Elsevier.
Figure 7. Schematic diagram of the strength of HS-mSi@MOF-5 on metal ions at different pH values. Reprinted with permission from Ref. [99]. Copyright 2016, Elsevier.
Molecules 28 02141 g007
Figure 8. Schematic diagram of selective adsorption of MB by Fe3O4@UiO-66-SO3H and corresponding adsorption mechanism. Reprinted with permission from Ref. [104]. Copyright 2019, Royal Society of Chemistry.
Figure 8. Schematic diagram of selective adsorption of MB by Fe3O4@UiO-66-SO3H and corresponding adsorption mechanism. Reprinted with permission from Ref. [104]. Copyright 2019, Royal Society of Chemistry.
Molecules 28 02141 g008
Figure 9. Schematic diagram of experimental setup used for breakthrough experiments. Reprinted with permission from Ref. [105]. Copyright 2017, American Chemical Society.
Figure 9. Schematic diagram of experimental setup used for breakthrough experiments. Reprinted with permission from Ref. [105]. Copyright 2017, American Chemical Society.
Molecules 28 02141 g009
Figure 10. Synthesis method and structure diagram of MIL-88A nanorods. Reprinted with permission from Ref. [107]. Copyright 2018, Springer Nature.
Figure 10. Synthesis method and structure diagram of MIL-88A nanorods. Reprinted with permission from Ref. [107]. Copyright 2018, Springer Nature.
Molecules 28 02141 g010
Figure 11. The overall synthesis process of the LDH/MOF NC. Reprinted with permission from Ref. [113]. Copyright 2021, Elsevier.
Figure 11. The overall synthesis process of the LDH/MOF NC. Reprinted with permission from Ref. [113]. Copyright 2021, Elsevier.
Molecules 28 02141 g011
Figure 12. Schematic diagram of adsorption effect of SCNU-Z1-Cl on five organic dyes. Reprinted with permission from Ref. [116]. Copyright 2019, American Chemical Society.
Figure 12. Schematic diagram of adsorption effect of SCNU-Z1-Cl on five organic dyes. Reprinted with permission from Ref. [116]. Copyright 2019, American Chemical Society.
Molecules 28 02141 g012
Table 1. Application of MOFs modified by functional groups in water purification.
Table 1. Application of MOFs modified by functional groups in water purification.
MOFIntroduced GroupContaminantAdsorbing
Capacity(mg·g−1)
Main Adsorption MechanismRef.
UiO-66-NH2Pb2+102.03Lewis acid–base interaction[83]
MIL-96-NH2PFOA 340.00Electrostatic interaction[84]
Cu-BTC-NH2IBF
ACE
187.97
125.45
Electrostatic interaction, Hydrophobic interaction[85]
MIL-101-NH2BPS513.00Hydrogen-bond interaction[86]
UiO-66-NH2Cr6+32.36Electrostatic interaction[87]
TMU-16-NH2Cd2+126.60Lewis acid–base interaction[88]
MIL-68(In)-NH2p-ASA401.60π–π interaction,
Hydrogen-bond interaction
[89]
TMU-16-NH2MO393.70Electrostatic interaction, Hydrogen-bond interaction[90]
MIL-101(Al)-NH2MB762.00Electrostatic interaction[91]
MIL-101(Al)-NH2MG
IC
274.40
135.00
Electrostatic interaction, Hydrogen-bond interaction,
π–π interaction,
Hydrophobic interaction
[92]
MIL-125(Ti)-NH2BR46
BB41
MB
1296.00
1257.00
862.00
Hydrogen-bond interaction[93]
UiO-66-NMe3+2,4-d279.00Electrostatic interaction,
π–π interaction
[94]
MIL-101(Cr)-NMe3+DCF310.60Electrostatic interaction,
π–π interaction
[95]
Cu-MOF-SHHg2+714.29Lewis acid–base interaction[96]
UiO-66-SHHg2+784.30Lewis acid–base interaction[97]
MIL-88A-SHHg2+1111.10Lewis acid–base interaction[98]
MOF-5-SHPb2+
Cd2+
312.50
65.20
Electrostatic interaction,
Lewis acid–base interaction
[99]
MIL-100(Fe)-SO3HMO99.9%Hydrogen-bond interaction,
π–π interaction
[100]
UiO-66-SO3HDCF263.00Electrostatic interaction,
π–π interaction
[101]
Cu-BTC-SO3HCd2+88.70Lewis acid–base interaction, Electrostatic interaction[102]
UiO-66-SO3HCd2+409.96Lewis acid–base interaction, Electrostatic interaction[103]
UiO-66-SO3HMB297.30Lewis acid–base interaction,
π–π interaction
[104]
Zr-BTC-SO4Ba2+181.80Lewis acid–base interaction[105]
MOF-808-SO4Ba2+131.10Lewis acid–base interaction[106]
MIL-88A-OHAs5+145.00Lewis acid–base interaction[107]
MIL-101-OHNAP185.00Hydrogen-bond interaction[108]
MIL-101-OHPCMX
KET
NAP
79.00
80.00
152.00
Electrostatic interaction, Hydrogen-bond interaction[109]
UiO-66-OHMB
TC
96.69
37.96
Lewis acid–base interaction, Hydrogen-bond interaction[110]
UiO-66-COOHSb3+56.49Electrostatic interaction[111]
Zr-BTC-COOHSr2+67.50Electrostatic interaction[112]
UiO-66-(Zr)-COOHHg2+
Ni2+
509.80
439.00
Electrostatic interaction[113]
UiO-66-FPFOA470.00Hydrophobic interaction[114]
UiO-66-FB76.30π–π interaction,
Hydrophobic interaction
[115]
SCNU-Z1-ClCrO42−
Cr2O72−
MnO4
ReO4
MO
AO
CR
MB
126.00
241.00
292.00
318.00
285.00
180.00
585.00
262.00
Hydrophobic interaction[116]
PFOA: perfluorooctanoic acid; IBF: ibuprofen; ACE: acetaminophen; BPS: bisphenol S; p-ASA: arsine acid; MO: methyl orange; MB: methylene blue; MG: malachite green; IC: indigo carmine; BR46: basic Red 46; BB41: basic blue 41; 2,4-d: 2,4-dichlorophenoxyacetic acid; DCF: diclofenac sodium; NAP: naproxen; PCMX: p-chloro-m-xylenol; KET: ketoprofen; TC: tetracycline hydrochloride; B: benzene; AO: acid orange A; CR: Congo red.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, J.; Wang, Y. Research on Improved MOF Materials Modified by Functional Groups for Purification of Water. Molecules 2023, 28, 2141. https://doi.org/10.3390/molecules28052141

AMA Style

Liu J, Wang Y. Research on Improved MOF Materials Modified by Functional Groups for Purification of Water. Molecules. 2023; 28(5):2141. https://doi.org/10.3390/molecules28052141

Chicago/Turabian Style

Liu, Junyan, and Yang Wang. 2023. "Research on Improved MOF Materials Modified by Functional Groups for Purification of Water" Molecules 28, no. 5: 2141. https://doi.org/10.3390/molecules28052141

Article Metrics

Back to TopTop