Next Article in Journal
Approaches for Selective Oxidation of Methane to Methanol
Next Article in Special Issue
Fast Immobilization of Human Carbonic Anhydrase II on Ni-Based Metal-Organic Framework Nanorods with High Catalytic Performance
Previous Article in Journal
Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy
Previous Article in Special Issue
Enhancing the Water Resistance of Mn-MOF-74 by Modification in Low Temperature NH3-SCR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Various Metal-Organic Frameworks (MOFs) as Catalysts for Air and Water Pollution Environmental Remediation

1
Department of Chemistry, Pusan National University, Busandaehak-ro 63beon-gil, Geumjeong-gu, Busan 46241, Korea
2
Department of Physics, Pusan National University, Busandaehak-ro 63beon-gil, Geumjeong-gu, Busan 46241, Korea
3
Department of Organic Material Science and Engineering, Pusan National University, Geumjeong-gu, Busan 46241, Korea
4
Center for Innovative Materials and Architectures (INOMAR), Vietnam National University, Ho Chi Minh City 721337, Vietnam
5
Department of Korean Medical Science, Healthy Aging Korea Medical Research Center, Pusan National University, Yangsan 50612, Gyeongsangnam-do, Korea
*
Authors to whom correspondence should be addressed.
Submission received: 11 December 2019 / Revised: 3 February 2020 / Accepted: 5 February 2020 / Published: 6 February 2020
(This article belongs to the Special Issue MOFs for Advanced Applications)

Abstract

:
The use of metal-organic frameworks (MOFs) to solve problems, like environmental pollution, disease, and toxicity, has received more attention and led to the rapid development of nanotechnology. In this review, we discuss the basis of the metal-organic framework as well as its application by suggesting an alternative of the present problem as catalysts. In the case of filtration, we have developed a method for preparing the membrane by electrospinning while using an eco-friendly polymer. The MOFs were usable in the environmental part of catalytic activity and may provide a great material as a catalyst to other areas in the near future.

1. Introduction

Recently, environmental pollution is increasing due to toxic waste and hazardous organic compounds [1]. The amount of poisonous compounds that are released into the environment is a serious problem for human life. In the pragmatic situation, air and organic pollutants are highly involved they can be commonly expressed as particulates, acidic substances, gases, or mixtures [2,3].
Nitroaromatic compounds are found in soil, air, and water samples due to wastewater sources from the plastic, pesticide, pharmaceutical, and dye industry [4]. In addition, the development of hydrogen and methane storage systems is necessary for the widespread use of green energy. Moreover, the separation and selective gas adsorption of poisonous gases, such as nitrogen dioxide and ammonia gas, are significant in the field of air pollution [5,6,7,8,9].
It is essential that sensors should be developed as the way to prevent toxic materials, one of the leading causes of environment pollution, from being released into the environment. Sulfur dioxide and nitroaromatic compounds are considered to be poisonous and harmful to human life [10]. Our previous work has reported a PdAg nanoparticle infused metal-organic framework (MOF) used for the detection of 4-nitrophenol while using an electrochemical sensor [11]. In addition, Salama et al., reported an SO2 gas sensor while using an MOF [12].
Metal-organic frameworks (MOFs) have a wide range of interesting properties such as high specific surface area and facile modification [13,14,15,16,17,18,19,20,21,22,23,24,25,26]. MOFs are microporous materials that form three-dimensional (3D) crystalline networks, which are prepared by combining various metal ions with organic linkers in an appropriate solvent [27,28,29,30,31,32]. Over the past years, MOFs have been used as catalysts, absorbents, and filters. MOFs have many advantages due to their modifiable properties, such as high specific surface area and porosity [30,33,34,35,36]. The exceptional characteristics of MOFs have led to their possible application in a wide range of technological areas, including gas sorption, separation, storage [5,6,7,8,9], sensing [10,11,12,37], and heterogeneous catalysis [38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53].
In this review, we focus on various MOFs and their applications in different fields, such as: (1) controlled gas uptake of toxic gases, including ammonia, nitrogen dioxide, and sulfur dioxide [6,7,12]; (2) sensors using PdAg nanoparticle infused MOFs and Cd-MOFs [11,54]; (3) Hydrogen gas storage [55,56,57]; and, (4) filtration for air and water pollution control while using nanofibrous MOFs [58,59] (Scheme 1).

2. Basic of Stable MOFs

2.1. Characterization of MOFs

MOFs are an arising group of porous materials that were synthesized from metal ions and organic linkers [60,61]. The ever-expanding improvement in the performance of MOFs and facile control over their properties, MOFs have attracted the interest of engineers and scientists [62,63,64,65,66,67]. Earlier MOFs were synthesized from divalent metals, which showed superior properties and a diverse range of applications [60], such as gas sorption, separation, storage [5,6,7,8,9], sensing [10,11,12,37], and heterogeneous catalysis [38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53]. In particular, stable MOFs can be predicted by the strength of the metal-organic linker bond that formed in their framework. We have summarized some of the typical stable MOFs prepared from divalent, trivalent, and tetravalent metal ions in Table 1 [68].

2.2. Tetravalent Metal-Carboxylate Based MOFs

Tetravalent metals, such as Ce4+, Zr4+, and Ti4+, and carboxylate linker based MOFs are a comparatively new field of study. Lillerud et al. and Férey et al. have reported on Zr-MOFs and Ti-MOFs, respectively [116,121]. Both Zr- and Ti-MOFs have been applied in various fields because of their high stability [68]. On the other hand, Ce-MOFs are fascinating materials due to their redox properties and possible catalytic activity. For example, a Ce-MOF, which is composed of Ce3+ and Ce4+, exhibits unique oxidase-like catalytic performance [158].

2.3. Trivalent Metal-Carboxylate Based MOFs

MOFs that are composed of trivalent metal cations and carboxylate linkers have two main secondary building units (SBUs): (1) The [M33-O)(COO)6] cluster, which includes a μ3-oxo-centered trimer of MO6 octahedra and (2) the [M(OH)(COO)2]n chain, which has a μ2-hydroxo corner sharing MO6 octahedral unit [68].

2.4. Divalent Metal-Azolate Based MOFs

Another type of stable MOFs is composed of soft M2+ ions and azolate-based ligands while using hard soft acid base (HSAB) theory. Some of the organic reagents are in the form of azolate-based linkers (Table 2) [159]. Azoles generally release a proton to coordinate with the M ions, similar to carboxylic acids [68]. In addition, azoles display well-known coordination properties and the sp2 nitrogen donors in pyridines and azoles are essentially alike, but different to carboxylic acids [68].

3. Toxic Gas Sensors

The detection of toxic gases is important in environmental remediation and human health problems. Accordingly, many groups have studied new sensing materials for latent modification. Schröder et al. have reported several MOFs that are used for the reversible adsorption of nitrogen dioxide [6]. The Dincă group have reported the use of microporous triazolate-based MOFs for the detection of ammonia gas [7]. Salma et al. have studied the synthesis of a foremost chemical sensor for the identification of sulfur dioxide at room temperature (RT) [12].

3.1. Robust Porous MOF for Reversible Adsorption of NO2

As one of the major air pollutants, nitrogen dioxide is fatal to the environment and it causes serious health problems [75,76,77,78]. Decreasing NOx contamination is a difficult issue due to the highly active atomic bond with oxygen and corrosive nature [79]. Therefore, various materials, including metal oxides, mesoporous silica, zeolites, and activated carbons, have been studied as NO2 adsorbents. However, these materials show low adsorption capacities and irreversible uptake due to the disproportionation of NO+ and NO3−. MOFs have been used as solid adsorbents, but an isothermal adsorption study on NO2 has not been conducted to date. Therefore, Han et al. studied the isothermal adsorption of MOFs and confirmed that MFM-300 (Al) can interact with highly reactive NO2. Consequently, MFM-300 (Al) has great potential as a practical solid absorbent.
Figure 1 shows the adsorption isotherms that were obtained for MFM-300 (Al) in various gases, including NO2, CO2, SO2, CO, CH4, N2, H2, O2, and Ar at room temperature and pressure. The maximum NO2 isotherm uptake was ~4.1 mmol g–1 at room temperature and pressure. This value was much higher than modified Y zeolites [160], mixed oxides, such as Ce1–xZrxO2 [161], NH3 functionalized SBA-15 [162], urea-modified mesoporous carbons [163], and activated carbons [164]. Furthermore, the crystallinity and sorption capacity were not changed after the cycling of the sorption and desorption steps.

3.2. Ammonia Sorption in MOFs

Ammonia (NH3) is present in the atmosphere due to global agriculture and industry. The current industrial standard sorbents show a low affinity and limited capacity for NH3. The heterogeneous pore size in carbon-based materials has caused a fundamental problem in studies on ammonia sorption. Recent studies have focused on sorbents containing Lewis or Brønsted acid sites that show a higher affinity toward NH3 molecules to solve this problem. In this study, Dinca et al. showed the static and dynamic ammonia capacities of various microporous triazolate-based MOFs.
Dynamic breakthrough measurements using 0.1% NH3 showed that Co2Cl2BBTA and Co2Cl2BTDD have a NH3 breakthrough capacity of 8.56 and 4.78 mmol g–1, respectively (Table 3). This is equal to 1.48 and 1.08 molecules of NH3 per Co atom, respectively. The NH3 breakthrough capacity value is reduced in 80% relative humidity (RH), regardless of the pore size due to the adsorption of water. The saturation value was 4.36 mol kg–1 for Co2Cl2BBTA and 3.38 mol kg–1 for Co2Cl2BTDD. These results indicate 0.76 and 0.77 NH3 molecules are absorbed per open metal site in Co2Cl2BBTA and Co2Cl2BTDD, respectively.

3.3. Highly Performance of SO2 MOF Sensor

Although sulfur dioxide is one of the most toxic and serious air pollutants, consumption for fossil fuel is increasing [10]. The main adverse health issues occur upon continuous exposure to SO2, with a primary 1-h acceptable limit of 75 ppb. Thus, a sensitive sensor, which can detect even a small amount of SO2 gas, is necessary. However, the detection of SO2 gas by chemical reaction from CaO to CaSO3 has low efficiency and irreversibility [165,166]. Therefore, it is necessary to achieve the reversible physisorption and selective interaction with SO2. Thus far, there are many studies that have been conducted based on the metal oxide (SnO2, WO3, and TiO2) that show high- sensitivity, recover time, and selectivity. However, a sensor based on metal oxide requires the high temperature (200–600 °C), which means that it requires high energy and power.
Recently, MOFs are attractive because of satisfying these requirements mentioned above. However, one of the issues using MOFs in sensing devices is directly related to the fabrication as thin films form. In this study, Salama et al. showed the fabrication of a MOF thin film on various supports and its gas-sensing properties.
Among the various MOFs, such as MFM-300 (Al), MFM-202-a, Zn3[Co(CN)6]2, Co3[Co(CN)6]2, Mg-MOF-74, and Ni(bdc)(ted)0.5, they have chosen the MFM-300 (In) MOF because of its high sorption capacity. To confirm the sensing properties of MFM-300 (In) and measured the changing the capacitance. In particular, MFM-30 (In) MOF was grown on a prefunctionallized IDE with an OH-terminated self-assembled monolayer (SAM) under optimized conditions [167]. X-ray diffraction (XRD) and scanning electron microscopy (SEM) confirmed the structural properties of the film.
The MFM-300 (In) MOF sensor showed exceptional performance. In addition, it can detect SO2 in the ppb range down to 75 ppb (Figure 2). The remarkable detection properties are related to the change in the permittivity of the thin film, depending on the adsorption of SO2 molecules. There are two types of interaction in the adsorption process: (1) Analyte-framework interactions and (2) analyte-analyte interactions. Changing these two interactions induce a change in the capacitance of the thin film. The MOF sensor exhibited good stability over three weeks of operation.
They also measured the sensing performance with relative humidity (RH) at 350 and 1000 ppb of SO2 gas. However, it does not show distinctive signals as compared to the “dry” condition (Figure 3a). Therefore, it confirmed that the practical applicability of MFM-300 (in) MOF as SO2 sensor in humidity condition. The temperature dependent sensing properties from 22 °C to 80 °C showed that the performance decrease up to 35% with increasing temperature (Figure 3b). Finally, selectivity performance was conducted in various gases of MFM-300 (In) MOF. As a result, it showed good selectivity for SO2 gas when compared to others (Figure 3c).

4. Detection and Reduction of Toxic Water via MOFs

4.1. Detection of Toxic 4-Nitrophenol via AgPd Nanoparticles on Functionalized MOFs

Nitroaromatic compounds are continuous organic contaminants that originate from industrial waste. 4-Nitrophenol (4-NP) is one of the harmful phenolic pollutants found in chemical waste [54], which is due to its high polarity and subsequent high solubility in water.

4.1.1. Synthesis of UiO-66-L and AgPd Nanoparticles Embedded on UiO-66-L (L=NH2 and NO2)

The synthesis of UiO-66-L was previously reported in the literature [168]. ZrCl4 was dispersed in DMF and the resulting mixture activated with acetic acid at 55 °C. 2-NH2-H2BDC and 2-NO2-H2BDC were used to functionalize UiO-66, respectively.
AgPd@UiO-66-L MOF was prepared via a reduction method while using sodium borohydride. UiO-66-L MOFs were homogeneously dispersed in water, and AgNO3 and PdCl2 were dispersed in the resulting MOFs dispersion, respectively. An aqueous solution of NaBH4 was added to the Ag+, Pd2+, and MOFs mixture to reduce the Ag and Pd. The product was isolated via centrifugation, washing, and drying (Scheme 2).

4.1.2. Characterization of UiO-66-L and AgPd@UiO-66-L

In this review, the characterization of UiO-66-L and AgPd@UiO-66-L while using SEM, TEM, and XRD is described. The FE-SEM and TEM images show the good dispersion of metal NPs on the AgPd@UiO-66-NH2 MOF and its octahedral morphology (Figure 4a,b). The elemental distribution of UiO-66-NH2 was investigated while using HAADF and elemental mapping (Figure 4c–j), which confirmed the bimetallic AgPd nanoparticles were loaded into both the bulk MOF and on its surface.
The XRD spectra showed the crystallinity and structural aspects of the as-synthesized materials (Figure 5). The XRD spectra that were recorded for UiO-66-NH2 and UiO-66-NO2 were similar and in good agreement with AgPd and bare UiO-66 (Figure 5a–c). This shows that the materials crystallinity was not changed after functionalization of the -NH2 and -NO2 groups (Figure 5d,e). The characteristic peaks for Ag and Pd were observed at 2θ = 38.03° and 40.01°, respectively. Moreover, the existence of both Ag and Pd peaks in Figure 5f,g, respectively, show the obvious crystallinity of the AgPd alloy.

4.1.3. A Comparison of the Catalytic Performance by the Detection and Reduction of 4-Nitrophenol

In this paper, 0.5 mM of 4-nitrophenol was detected while using AgPd@UiO-66-L by cyclic voltammetry. The sharp reduction peak at –0.7 V vs. Ag/AgCl shows the reduction of 4-nitrophenol to 4-hydroxyaminophenol by AgPd@UiO-66-NH2/GCE (Figure 6a). In addition, a quasi-reversible anodic peak was observed at ~0.1 V vs. Ag/AgCl, due to the oxidation of 4-hydroxyaminophenol to 4-nitrosophenol. In this result, AgPd@UiO-66-NH2 showed improved performance when compared to the other materials studied. Figure 6b shows the cyclic voltammograms for the sensing of various concentrations of 4-nitrophenol (0.25–400 μm) by the AgPd@UiO-66-NH2/GCE electrode in 0.1 M phosphate buffered saline (PBS) at pH 7. Increasing the concentration of 4-nitrophenol from 0.25 μM to 400 μM exhibited a linear increase in the cathodic reduction peak due to the beneficial electrocatalytic reduction of 4-nitrophenol to 4-hydroxyaminophenol.
The catalytic activities of the as-synthesized products were also studied while using this catalytic reduction reaction using UV-Vis spectroscopy. 4-Nitrophenol shows an absorption peak at ~317 nm in an aqueous medium. Upon the addition of NaBH4 powder to the solution, a new peak was detected at 400 nm (Figure 6c). The resulting thick yellow mixture and shift in the absorption peak was due to the generation of p-nitrophenolate [169]. Figure 6d showed catalytic activity by AgPd@UiO-66-NH2 in the mixed solution of sodium borohydride and 4-nitrophenol. The color of mixed solution altered from yellow to colourless because of the activation of 4-nitrophenolate ion to 4-aminophenol.

4.2. Detection of Antibiotics in Water Using Cd-MOF as a Fluorescent Probe

Antibiotics are used to treat bacterial infections in animals and humans, but they are an important organic pollutant [170,171]. The abuse of antibiotics has led to extreme residues in subsoil water and surface water [172,173,174,175,176,177]. Therefore, luminescent MOFs have been synthesized and applied in the detection of antibiotics in water as an alternative to liquid chromatography (LC) combined with UV-Vis spectroscopy, capillary electrophoresis, Raman spectroscopy, mass spectroscopy, and ion mobility spectroscopy [141,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193]. In this paper, the reported Cd-MOF material was used toward the detection of the antibiotic, ceftriaxone sodium (CRO).

4.2.1. Synthesis and Detection of Cd-MOF

The synthesis of Cd-MOF while using 1,4-bis(2-methyl-imidazole-1-yl)butane (bbi) was carried out while using a literature process [194]. The detection of antibiotics was achieved using UV spectroscopy in the wavelength range of 270–350 nm.

4.2.2. Characterization of Cd-MOF

The Cd-MOF was analyzed in terms of its thermal and chemical stability under harsh conditions. Figure 7a exhibited no weight loss in the temperature range of 35–300 °C due to the absence of water in the Cd-MOF. The framework started to decompose at temperatures >300 °C. Meanwhile, the Cd-MOF was stored in aqueous solutions of the antibiotic and distilled water under various pH conditions for 24 h (Figure 7b). The PXRD spectra of the treated samples are in good agreement with the base simulated data. In these results, the Cd-MOF exhibited high physical and chemical stability in alkaline, acidic, and antibiotic solutions, respectively.
The MOF candidates used as luminescent materials are often constructed from conjugated organic ligand linkers and d10 metal ions [195,196,197,198,199,200]. Therefore, the Cd-MOF shows enhanced fluorescence intensity when compared to those that are constructed from organic ligands, such as bbi and H2L (Figure 8).

4.2.3. Chemical Sensors for Antibiotic Detection

The application of Cd-MOF as a fluorescent sensor for detecting antibiotics in water was investigated because of its robust luminescence properties in water. Cd-MOF was dispersed in antibiotic solutions containing LIN, AZL, PEN, AMK, ERY, AMX, AZM, GEN, ATM, FOX, CSU, CEC, CED, CFM, MTR, SXT, and CRO (LIN: Lincomycin hydrochoride, AZL: Azithromycin latobionate, PEN: Penicillin, AMK: Amikacin, ERY: Erythromycin ethylsuccinate, AMX: Amoxicillin, AZM: Azithromycin, GEN: Gentamicin, ATM: Aztreonam, FOX: Cefoxitin, CSU: Cefathiamidine, CEC: Cefaclor, CED: Cefradine, CFM: Cefixime, MTR: metronidazole, SXT: Sulfamethoxazole, and CRO: Ceftriaxone sodium, respectively). As a result, the effective fluorescence quenching of Cd-MOF by these antibiotics follows the order of AZL < GRN < LIN < AMK < ERY < PEN < AZM < AMX < FOX < CEC < CSU < MTR < CFM < ATM < SXT < CRO (Figure 9a). In particular, the efficient detection of CRO via fluorescence quenching was ~90% in this study. Figure 9b shows the distinction of the Cd-MOF sensor for quantitative analysis while using a fluorescence titration experiment. This graph shows the straightforward and dramatic trend in the fluorescence intensity of Cd-MOF that was detected from 0 to 70 μL. Cd-MOF is an effective probe used to detect CRO. The Cd-MOF sensor in an aqueous solution of CRO exhibits highly sensitive fluorescence intensity under various pH conditions (Figure 9c, pH = 4–11). From this graph, Cd-MOF can be used under various pH conditions with no effect on the experimental results. In particular, Cd-MOF shows high performance and stability at pH = 6–7. Figure 9d shows that the rapid detection of various concentrations of CRO can be achieved when an aqueous solution of CRO was added to the Cd-MOF suspension.

5. Filtration of Water and Air Pollutants Using Nanofibrous MOFs Prepared via Electrospinning Methods

5.1. Nanofiber MOF Filter for Particulate Matter

Serious threats to human health have been arisen due to the rapid development of the economy and industry, with the most dangerous of them being air pollution [201,202]. In practice, air pollutants are very diverse and are typically composed of particulate matter (PM) and toxic gases. PM is harmful to the environment affecting human health, air quality and the climate. Particulate matter whose aerodynamic diameters are <2.5 μm (PM2.5) and <10 μm (PM10) can penetrate the respiratory system and cause health problems upon prolonged exposure [203,204]. A lot of attention has been focused on researching PM filters to solve these problems. Among them, research on filters made using MOFs via electrospinning methods is a major factor.

5.1.1. Basic Theory of Electrospinning

Electrospinning is the most facile way to make nanofiber membranes containing organic and inorganic components while using polymer melts and solutions [205]. The basic principle of electrospinning is to apply a strong electric field using a high voltage power supply and drawing the fabricated fiber as they solidify (Scheme 3) [206]. It is easy to install and produce, so this method enables mass production at a low cost. Although the basic principle of electrospinning is simple, its mechanism is very complicated, because there are many factors that affect the process. Among them, the process parameters include the nozzle diameter, applied voltage, and tip-to-collector distance (TCD) [207]. In addition, it is also influenced by the solution characteristics and physical properties, such as the concentration, viscosity, surface tension, and vapor pressure. In particular, there are many advantages that affect performance of filter, such as specific surface area, alterable fiber diameter, and pore size.

5.1.2. Detail Mechanism of Electrospinning

Electrospinning is the simplest way for producing micro or nanoscale fibers [208,209,210]. Although the basic theory of this process is uncomplicated, its detail mechanism is so complicated. Many conditions, such as solution terms and experimental environment, affect the electrospinning process.
In the case of solution term, viscosity, surface tension, vapor pressure and solution conductivity have a major influence on electrospinning. The most important of these is viscosity, which can vary greatly with electrospinning. If all conditions are the same, except for the viscosity, it will affect the thickness and formation of the fibers. If the viscosity of the solution is too high, since the drop for electrospinning is not formed, electrospinning itself might be disrupted. On the other hand, if it is considerably low, the fibers are not able to withstand the tension caused by stretching in the process of drawing fibers, so that the fiber are broken. At the optimum viscosity for electrospinning, generally, if the viscosity increases, the stretching proceeds relatively slowly, and the fibers may become thick and vice versa. Similarly, vapor pressure also affects the thickness of the fiber. The vapor pressure of the solution has an intuitive effect on the evaporation rate of the interpolation solvent, so that adjusting the vapor pressure can control the thickness of the fiber.
Surface tension is thought to influence jet formation after drop formation. If the surface tension is too high, drops are formed, but it is difficult to form a jet. This is because the surface tension of the solution is higher than the high voltage applied and, thus, the jet is not formed. In general, it is a good idea to consider the surface tension of the solution if a drop is formed but no jet is formed.
Electrospinning is basically a process for very low conductivity solutions. Electrospinning on conductive solutions is a very difficult process. The reason is that when the conductivity of the solution is high, the drop itself is not formed, and charge is directly discharged from the solution to the collector, which makes it difficult to form electrospinning. It might be possible to spinning a conductive solution by applying a very high voltage, so that the amount of charge applied is greater than that discharged, in order to spinning a conductive solution. The experimental environmental factors affecting electrospinning include TCD, voltage, nozzle size, and temperature and humidity. In the case of temperature and humidity, it is difficult to control these, which is one of the main reasons why the reproducibility of electrospinning result is different every day. Conditions that are related to electrospinning settings, such as TCD voltage and nozzle size, have a relatively small impact. During the electrospinning experiment, the TCD and voltage can be changed in real time, so it is relatively easy to know the optimal conditions for spinning. In this context, the detailed mechanisms and conditions of electrospinning are very diverse and complex, but, if the optimum conditions are found, then a fiber that has high propulsion can be obtained.

5.1.3. Characterization of MOF@PAN, PS (Polystyrene)

A high ratio of MOFs can be loaded into polymer composites without agglomeration by adjusting the morphology and particle size of dissimilar MOF crystals. Figure 10 shows the SEM images of polymer and various MOF non-woven fabrics. By controlling the electrospinning conditions, such as voltage, flow rate of the solution, and TCD, four MOFs can be formed into fiber materials. Figure 10 shows that the MOF nanoparticles are well dispersed in the polymer fibers without any apparent agglomeration, despite a high loading of 60 wt%.

5.1.4. Filtration of Particle Matter Using ZIF-8@PAN

Generally, MOF filters can capture pollutants, including particulate matter via three mechanisms: (1) Pollutants can be bound to the OMSs, (2) interaction with the functional groups in the MOF filters, and (3) electrostatic interactions with the MOF filter. Figure 11 shows the particulate matter removal efficiency that was observed for MOFs@PAN filters. Figure 11a shows that the ZIF-8@PAN filter has the highest removal efficiency for PM2.5 and PM10 among the various MOFs studied. Particulate matter is very polar because of the presence of water vapor and various ions. MOFs, which have unbalanced defects and metal ions on the surface, offer positive charge. This is why the surface of PM can be polarized enhancing the electrostatic interactions. The zeta potential represents these electrostatic interactions. Among the MOFs studied, ZIF-8 exhibited the highest zeta potential of 47.5 mV. In this context, the ZIF-8@PAN filter displayed higher removal efficiency than the other filters studied and its efficiency was maintained after 48 h of exposure to polluted air (Figure 11b).

5.2. Nanofiber MOF Filter for Water Pollutants

Water contamination has become an important issue in environmental remediation due to the increase of urban areas and industrialization. Domestic wastewater is continuously discharged into the environment. Generally, the major pollutants in food wastewater are soluble organic food additives and insoluble organic compounds. Many methods have been used to treat these types of pollutants, such as advanced oxidation, adsorption, and photocatalytic membrane technology [211,212,213,214,215]. Among these methods, membrane technology is preferred due to its facile operation. However, research studies mainly concentrate on the removal of one type of pollutant, either soluble or insoluble pollutants. Recently, porous materials, such as MOFs, have been applied to water filtration to treat these contaminants [211,212,213,214,215,216,217,218,219]. It is necessary to use an electrospinning process to obtain superhydrophilic-underwater superoleophobic properties. In this context, an electrospun polyacrylonitrile (PAN) and MIL-100 (Fe) composite filter (PAN@ MIL-100(Fe)) have been fabricated to treat domestic wastewater.

5.2.1. Schematic of the PAN@MIL-100 (Fe) Filter

Scheme 4 shows the process that is used to prepare the PAN@MIL-100 (Fe) filter. In view of the facile electrospinning process, a H3BTC/PAN electrospun fiber filter was prepared as the precursor to load MIL-100 (FE), where the PAN fiber is used as a polymer frame and trimesic acid used as the initial reaction site for MIL-100 (FE) growth. As the hydrothermal reaction proceeds, trimesic acid acts as a nucleation site for the growth of MIL-100 (Fe) on the PAN fibers.

5.2.2. Characterization of the PAN@MIL-100 (Fe) Filter

Figure 12a and b show that the H3BTC/PAN fiber filter has a smooth surface without any beads. The average diameter is 110 nm. After the growth process, the PAN@MIL-100 (Fe) filter has a rough fiber surface with lots of particles. Many particles are covered on the PAN fibers with an average diameter of 211 nm, which is increased during the coating process. This indicates that the MOFs are successfully coated onto the PAN fibers without any aggregation.

5.2.3. Filtration of the Wastewater with Soluble Pollutants Using PAN@MIL-100 (Fe) Filter

The electrospun fiber filter that was prepared without MOFs has macro-size pores, so it is difficult to effectively separate the pollutants. The adsorption interactions between the fibers and pollutants is major factor in the filtration performance of the filter. In this context, Figure 13 shows the results of filtering amaranth red (AR) and vanillin (VA) as soluble pollutants. AR and VA are approximately 99% removed and the removal efficiencies were both >95% after 10 adsorption-desorption cycles.

5.2.4. Filtration of Wastewater Containing Insoluble Pollutants Using the PAN@MIL-100 (Fe) Filter

In the removal of oil, an important parameter is the surface wettability. To treat insoluble (oil) pollutants, it is essential that the filter has the property of selective wettability (superhydrophilicity and underwater superoleophobicity), which allows for water to pass through filter, but not oil. The selective wettability is that the filter can wet both water and oil in air, but in water has only hydrophilicity. Its basic mechanism is that water around the filter acts as a barrier to prevent oil from passing through. Figure 14a,b show that the PAN@MIL-100 (Fe) filter has selective wettability and the underwater oil pollutants contact angles are 151° and 154°. After five cycles, the oil removal efficiency is only slightly changed.

6. Conclusions

The application of various MOFs materials was reviewed with a focus on toxic sensor, reduction catalyst, hydrogen storage, and filter, which act as successful functional materials. We reviewed all of the materials while using MOFs that exhibited good performance and various application. The application of MOF material for toxic sensor, such as NO2, SO2, and ammonia gas, showed high performance via MFM-300 (Al), Co based MOFs (Co2Cl2BTDD and Co2Cl2BBTA), and MFM-300 (In). Our previous catalyst e.g., PdAg@UiO-66-L for the detection and reduction of 4-nitrophenol also showed good catalytic activity. Moreover, a new Cd-MOF as a fluorescent detect exhibited high sensitivity and selectivity in CRO. Various MOFs for hydrogen adsorption, such as MOF-177, SNU-6, and MOF-74 composited by Co/Ni mixed-material, exhibited good performance while using physisorption analysis method. Filtration for particle matters and wastewater using organic fiber modified as MOF-based material showed good adsorption in polluted condition. In summary, MOFs can be expected one of the best candidate to solve environmental pollution and energy storage in the near future.

Author Contributions

K.H.P. provided academic direction and Discussion of the results and revising the full manuscript. S.J. collected materials and wrote the introduction, basic of stable MOFs, and conclusion part. S.S. and J.H.L. contributed to the materials and methods and results and discussion equally. H.S.K., B.T.P., K.-T.H., and S.P. participated in discussion of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2017R1A4A1015533, 2017R1D1A1B03036303 and 2018R1D1A1B07045663). This research was partially supported by VietNam National University Ho Chi Minh City (NCM2019-50-01).

Acknowledgments

This research was supported by PNU-RENovation (2018–2019).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BDCterephthalate
FUMfumarate
2,6-NDCnaphthalene-2,6-dicarboxylate
BTCbenzene-1,3,5-tricarboxylate
BDC-NH22-aminoterephthalate
BTEC1,2,4,5-benzenetetracarboxylate
NTC1,4,5,8-naphthalenetetracarboxylate
BPDCbiphenyl-4,4′-dicarboxylate
BPTAbiphenyl-3,3′,5,5′-tetracarboxylate
BTB1,3,5-benzenetrisbenzoate
BeDC4,4′-benzophenonedicarboxylate
1,3-BDCisophthalate
BTTB4,4′,4′′-[benzene-1,3,5-triyl-tris(oxy)]tribenzoate
TCPPmeso-tetrakis(4-carboxylatephenyl)porphyrin
TATB4,4′,4″-s-triazine-2,4,6-triyl-tribenzoate
TCPT3,3″,5,5″-tetrakis(4-carboxyphenyl)-p-terphenyl
TMQPTC2′,3′′,5′′,6′-tetramethyl-[1,1′:4′,1′′:4′′,1′′′-quaterphenyl]-3,3′′′,5,5′′′-tetracarboxylate
ABDC4,4-azobenzenedicarboxylate
1,3-ADC1,3-adamantanedicarboxylate
FTZB2-fluoro-4-(tetrazol-5-yl)benzoate
CDCtrans-1,4-cyclohexanedicarboxylate
AB4-aminobenzoate
BDAbenzene-1,4-dialdehyde
BPDA4,4′-biphenyldicarboxaldehyde
TPDC[1,1′:4′,1″-terphenyl]-4,4″-dicarboxylate
ETTC4′,4″,4′″,4″″-(ethene-1,1,2,2-tetrayl)tetrabiphenyl-4-carboxylate
MTBC4′,4″,4′″,4″″-methanetetrayltetrabiphenyl-4-carboxylate
PZDC1H-pyrazole-3,5-dicarboxylate
MTB4,4′,4″,4′″-methanetetrayltetrabenzoate
XF4,4′-((1E,1′E)-(2,5-bis((4-carboxylatephenyl)ethynyl)-1,4-phenylene)bis(ethene-2,1- diyl))dibenzoate
DTTDCdithieno[3, 2-b;2′,3′-d]-thiophene-2,6- dicarboxylate
TDC2,5-thiophenedicarboxylate
TBAPy1,3,6,8-tetrakis(p-benzoate)pyrene
PTBA4-[2-[3,6,8-tris[2-(4-carboxylatephenyl)-ethynyl]-pyren-1-yl]ethynyl]-benzoate
Py-XP4′,4′″,4′″″,4′″″″-(pyrene-1,3,6,8-tetrayl) tetrakis(2′,5′-dimethyl-[1,1′-biphenyl]-4-carboxylate
Por-PPmeso-tetrakis-(4-carboxylatebiphenyl)- porphyrin
Py-PTP4,4′,4″,4′″-((pyrene-1,3,6,8-tetrayltetrakis(benzene-4,1-diyl))tetrakis(ethyne-2,1-diyl))tetrabenzoate
Por-PTPmeso-tetrakis-(4-((phenyl)ethynyl)benzoate)porphyrin
EDDB4,4′-(ethyne-1,2-diyl)dibenzoate
CTTA5′-(4-carboxyphenyl)-2′,4′,6′-trimethyl-[1,1′:3′,1″-terphenyl]-4,4″-dicarboxylate
TTNA6,6′,6″- (2,4,6-trimethylbenzene-1,3,5-triyl)tris(2-naphthoate))
PEDC4,4′-(1,4-phenylenebis- (ethyne-2,1-diyl))dibenzoate
BTDC2,2′-bithiophene-5,5′-dicarboxylate
BTBA4,4′,4″,4′″-(biphenyl-3,3′,5,5′-tetrayltetrakis(ethyne-2,1-diyl))tetrabenzoate
PTBA4-[2-[3,6,8-tris[2-(4-carboxylatephenyl)-ethynyl]-pyren-1-yl]ethynyl]-benzoate
BTE4,4′,4″-(benzene-1,3,5-triyl-tris(ethyne-2,1-diyl))tribenzoate
BTPP1,3,5-Tris((1H-pyrazol-4-yl)phenyl)benzene
BTP1,3,5-tris(1H-pyrazol-4-yl)benzene
1,4-BDP1,4-benzenedi(4′-pyrazolyl)
1,3-BDP1,3-benzenedi(4′-pyrazolyl)
TPP10,15,20-tetra(1H-pyrazol-4-yl)-porphyrin
mIM2-methylimidazolate
bIMbenzimidazolate
nIM2-nitroimidazolate
5cbIM5-chlorobenzimidazolate
ICAimidazolate-2-carboxyaldehyde
5-mTz5-methyltetrazolate
2-mbIM2-methylbenzimidazolate

References

  1. Sandoval, A.; Hernández-Ventura, C.; Klimova, T.E. Titanate nanotubes for removal of methylene blue dye by combined adsorption and photocatalysis. Fuel 2017, 198, 22–30. [Google Scholar] [CrossRef]
  2. Kampa, M.; Castanas, E. Human health effect of air polution. Environ. Pollut. 2008, 151, 362–367. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, J.; Tan, S.; Lund, P.D.; Zhou, H. Impact of H2O on organic-inorganic hybrid perovskite solar cells. Energy Environ. Sci. 2017, 10, 2284–2311. [Google Scholar] [CrossRef] [Green Version]
  4. Liu, Z.; Yang, C.; Qiao, C. Biodegradation of p-nitrophenol and 4-chlorophenol by Stenotrophomonas sp. FEMS Microbiol. Lett. 2007, 277, 150–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Li, H.; Wang, K.; Sun, Y.; Lollar, C.T.; Li, J.; Zhou, H.C. Recent advances in gas storage and separation using metal–organic frameworks. Mater. Today 2018, 21, 108–121. [Google Scholar] [CrossRef]
  6. Han, X.; Godfrey, H.G.W.; Briggs, L.; Davies, A.J.; Cheng, Y.; Daemen, L.L.; Sheveleva, A.M.; Tuna, F.; McInnes, E.J.L.; Sun, J.; et al. Reversible adsorption of nitrogen dioxide within a robust porous metal–organic framework. Nat. Mater. 2018, 17, 691–696. [Google Scholar] [CrossRef]
  7. Rieth, A.J.; Dincă, M. Controlled Gas Uptake in Metal-Organic Frameworks with Record Ammonia Sorption. J. Am. Chem. Soc. 2018, 140, 3461–3466. [Google Scholar] [CrossRef]
  8. Wang, C.; Zhang, H.; Feng, C.; Gao, S.; Shang, N.; Wang, Z. Multifunctional Pd@MOF core-shell nanocomposite as highly active catalyst for p-nitrophenol reduction. Catal. Commun. 2015, 72, 29–32. [Google Scholar] [CrossRef]
  9. Li, Y.Z.; Wang, H.H.; Yang, H.Y.; Hou, L.; Wang, Y.Y.; Zhu, Z. An Uncommon Carboxyl-Decorated Metal–Organic Framework with Selective Gas Adsorption and Catalytic Conversion of CO2. Chem. A Eur. J. 2018, 24, 865–871. [Google Scholar] [CrossRef]
  10. Galloway, J.N. Acid deposition: Perspectives in time and space. Water Air Soil Pollut. 1995, 85, 15–24. [Google Scholar] [CrossRef]
  11. Hira, S.A.; Nallal, M.; Park, K.H. Fabrication of PdAg nanoparticle infused metal-organic framework for electrochemical and solution-chemical reduction and detection of toxic 4-nitrophenol. Sens. Actuators B Chem. 2019, 298, 126861. [Google Scholar] [CrossRef]
  12. Chernikova, V.; Yassine, O.; Shekhah, O.; Eddaoudi, M.; Salama, K.N. Highly sensitive and selective SO2 MOF sensor: The integration of MFM-300 MOF as a sensitive layer on a capacitive interdigitated electrode. J. Mater. Chem. A 2018, 6, 5550–5554. [Google Scholar] [CrossRef] [Green Version]
  13. Yaqoob, L.; Noor, T.; Iqbal, N.; Nasir, H.; Zaman, N. Development of Nickel-BTC-MOF-derived nanocomposites with rGO towards electrocatalytic oxidation of methanol and its product analysis. Catalysts 2019, 9, 856. [Google Scholar] [CrossRef] [Green Version]
  14. Liu, C.; Wang, W.; Liu, B.; Qiao, J.; Lv, L.; Gao, X.; Zhang, X.; Xu, D.; Liu, W.; Liu, J.; et al. Recent advances in MOF-based nanocatalysts for photo-promoted CO2 reduction applications. Catalysts 2019, 9, 658. [Google Scholar] [CrossRef] [Green Version]
  15. Padmanaban, S.; Yoon, S. Surface modification of a MOF-based catalyst with Lewis metal salts for improved catalytic activity in the fixation of CO2 into polymers. Catalysts 2019, 9, 892. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, S.; Gao, Q.; Dong, X.; Wang, Q.; Niu, Y.; Chen, Y.; Jinag, H. Enhancing the Water Resistance of Mn-MOF-74 by Modification in Low Temperature NH3-SCR. Catalysts 2019, 9, 1004. [Google Scholar] [CrossRef] [Green Version]
  17. Xie, X.; Huang, M. Enzymatic Production of Biodiesel Using Immobilized Lipase on Core-Shell Structured Fe3O4@MIL-100(Fe) Composites. Catalysts 2019, 9, 850. [Google Scholar] [CrossRef] [Green Version]
  18. Kim, A.; Muthuchamy, N.; Yoon, C.; Joo, S.H.; Park, K.H. MOF-derived Cu@Cu2O nanocatalyst for oxygen reduction reaction and cycloaddition reaction. Nanomaterials 2018, 8, 138. [Google Scholar] [CrossRef] [Green Version]
  19. Sun, W.; Li, X.; Sun, C.; Huang, Z.; Xu, H.; Shen, W. Insights into the pyrolysis processes of Ce-MOFs for preparing highly active catalysts of toluene combustion. Catalysts 2019, 9, 682. [Google Scholar] [CrossRef] [Green Version]
  20. Maksimchuk, N.; Lee, J.S.; Ayupov, A.; Chang, J.S.; Kholdeeva, O. Cyclohexene oxidation with H2O2 over metal-organic framework MIL-125(Ti): The effect of protons on reactivity. Catalysts 2019, 9, 324. [Google Scholar] [CrossRef] [Green Version]
  21. Ren, R.; Zhao, H.; Sui, X.; Guo, X.; Huang, X.; Wang, Y.; Dong, Q.; Chen, J. Exfoliated Molybdenum Disulfide Encapsulated in a Metal Organic Framework for Enhanced Photocatalytic Hydrogen Evolution. Catalysts 2019, 9, 89. [Google Scholar] [CrossRef] [Green Version]
  22. Huh, S. Direct Catalytic Conversion of CO2 to Cyclic Organic Carbonates under Mild Reaction Conditions by Metal—Organic Frameworks. Catalysts 2019, 9, 34. [Google Scholar] [CrossRef] [Green Version]
  23. Martínez, F.M.; Albiter, E.; Alfaro, S.; Luna, A.L.; Colbeau-Justin, C.; Barrera-Andrade, J.M.; Remita, H.; Valenzuela, M.A. Hydrogen production from glycerol photoreforming on TiO2/HKUST-1 composites: Effect of preparation method. Catalysts 2019, 9, 338. [Google Scholar] [CrossRef] [Green Version]
  24. Stawowy, M.; Róziewicz, M.; Szczepańska, E.; Silvestre-Albero, J.; Zawadzki, M.; Musioł, M.; Łuzny, R.; Kaczmarczyk, J.; Trawczyński, J.; Łamacz, A. The impact of synthesis method on the properties and CO2 sorption capacity of UiO-66(Ce). Catalysts 2019, 9, 309. [Google Scholar] [CrossRef] [Green Version]
  25. Bedia, J.; Muelas-ramos, V.; Peñas-garzón, M.; Almudena, G.; Rodríguez, J.J.; Belver, C. A Review on the Synthesis and Characterization of Metal Organic Frameworks for Photocatalytic Water Purification. Catalysts 2019, 9, 52. [Google Scholar] [CrossRef] [Green Version]
  26. Singh, H.; Zhuang, S.; Nunna, B.B.; Lee, E.S. Thermal stability and potential cycling durability of nitrogen-doped graphene modified by metal-organic framework for oxygen reduction reactions. Catalysts 2018, 8, 607. [Google Scholar] [CrossRef] [Green Version]
  27. He, L.; Liu, Y.; Liu, J.; Xiong, Y.; Zheng, J.; Liu, Y.; Tang, Z. Core-shell noble-metal@metal-organic-framework nanoparticles with highly selective sensing property. Angew. Chem. Int. Ed. 2013, 52, 3741–3745. [Google Scholar] [CrossRef]
  28. Lin, X.; Gao, G.; Zheng, L.; Chi, Y.; Chen, G. Encapsulation of strongly fluorescent carbon quantum dots in metal-organic frameworks for enhancing chemical sensing. Anal. Chem. 2014, 86, 1223–1228. [Google Scholar] [CrossRef]
  29. Wu, C.M.; Rathi, M.; Ahrenkiel, S.P.; Koodali, R.T.; Wang, Z. Facile synthesis of MOF-5 confined in SBA-15 hybrid material with enhanced hydrostability. Chem. Commun. 2013, 49, 1223–1225. [Google Scholar] [CrossRef]
  30. Gu, Z.G.; Li, D.J.; Zheng, C.; Kang, Y.; Wöll, C.; Zhang, J. MOF-Templated Synthesis of Ultrasmall Photoluminescent Carbon-Nanodot Arrays for Optical Applications. Angew. Chem. Int. Ed. 2017, 56, 1–7. [Google Scholar] [CrossRef]
  31. Wei, Y.; Han, S.; Walker, D.A.; Fuller, P.E.; Grzybowski, B.A. Nanoparticle core/shell architectures within mof crystals synthesized by reaction diffusion. Angew. Chem. Int. Ed. 2012, 51, 7435–7439. [Google Scholar] [CrossRef] [PubMed]
  32. Li, Z.; Zeng, H.C. Surface and bulk integrations of single-layered Au or Ag nanoparticles onto designated crystal planes {110} or {100} of ZIF-8. Chem. Mater. 2013, 25, 1761–1768. [Google Scholar] [CrossRef]
  33. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef] [PubMed]
  34. Kitagawa, S.; Kitaura, R.; Noro, S.I. Functional porous coordination polymers. Angew. Chemie Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef]
  35. Férey, G. Hybrid porous solids: Past, present, future. Chem. Soc. Rev. 2008, 37, 191–214. [Google Scholar] [CrossRef]
  36. O’Keeffe, M.; Yaghi, O.M. Deconstructing the Crystal Structures of Metal À Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2003, 31, 474–484. [Google Scholar]
  37. Masih, D.; Chernikova, V.; Shekhah, O.; Eddaoudi, M.; Mohammed, O.F. Zeolite-like Metal-Organic Framework (MOF) Encaged Pt(II)-Porphyrin for Anion-Selective Sensing. ACS Appl. Mater. Interfaces 2018, 10, 11399–11405. [Google Scholar] [CrossRef]
  38. Sonkar, P.K.; Prakash, K.; Yadav, M.; Ganesan, V.; Sankar, M.; Gupta, R.; Yadav, D.K. Co(II)-porphyrin-decorated carbon nanotubes as catalysts for oxygen reduction reactions: An approach for fuel cell improvement. J. Mater. Chem. A 2017, 5, 6263–6276. [Google Scholar] [CrossRef]
  39. Song, E.; Shi, C.; Anson, F.C. Comparison of the behavior of several cobalt porphyrins as electrocatalysts for the reduction of O2 at graphite electrodes. Langmuir 1998, 14, 4315–4321. [Google Scholar] [CrossRef]
  40. El-Deab, M.S.; Okajima, T.; Ohsaka, T. Electrochemical reduction of oxygen on gold nanoparticle-electrodeposited glassy carbon electrodes. J. Electrochem. Soc. 2003, 150, 851–857. [Google Scholar] [CrossRef]
  41. Liao, L.; Bian, X.; Xiao, J.; Liu, B.; Scanlon, M.D.; Girault, H.H. Nanoporous molybdenum carbide wires as an active electrocatalyst towards the oxygen reduction reaction. Phys. Chem. Chem. Phys. 2014, 16, 10088–10094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sui, S.; Wang, X.; Zhou, X.; Su, Y.; Riffat, S.; Liu, C.J. A comprehensive review of Pt electrocatalysts for the oxygen reduction reaction: Nanostructure, activity, mechanism and carbon support in PEM fuel cells. J. Mater. Chem. A 2017, 5, 1808–1825. [Google Scholar] [CrossRef]
  43. Ma, T.Y.; Zheng, Y.; Dai, S.; Jaroniec, M.; Qiao, S.Z. Mesoporous MnCo2O4 with abundant oxygen vacancy defects as high-performance oxygen reduction catalysts. J. Mater. Chem. A 2014, 2, 8676–8682. [Google Scholar] [CrossRef]
  44. Huang, T.; Mao, S.; Zhou, G.; Wen, Z.; Huang, X.; Ci, S.; Chen, J. Hydrothermal synthesis of vanadium nitride and modulation of its catalytic performance for oxygen reduction reaction. Nanoscale 2014, 6, 9608–9613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhang, X.; Chen, Y.; Wang, J.; Zhong, Q. Nitrogen and fluorine dual-doped carbon black as an efficient cathode catalyst for oxygen reduction reaction in neutral medium. ChemistrySelect 2016, 4, 696–702. [Google Scholar] [CrossRef]
  46. Kjaergaard, C.H.; Rossmeisl, J.; Nørskov, J.K. Enzymatic versus inorganic oxygen reduction catalysts: Comparison of the energy levels in a free-energy scheme. Inorg. Chem. 2010, 49, 3567–3572. [Google Scholar] [CrossRef]
  47. Cheng, F.; Chen, J. Metal-air batteries: From oxygen reduction electrochemistry to cathode catalysts. Chem. Soc. Rev. 2012, 41, 2172–2192. [Google Scholar] [CrossRef]
  48. He, G.; Qiao, M.; Li, W.; Lu, Y.; Zhao, T.; Zou, R.; Li, B.; Darr, J.A.; Hu, J.; Titirici, M.M.; et al. S, N-Co-Doped Graphene-Nickel Cobalt Sulfide Aerogel: Improved Energy Storage and Electrocatalytic Performance. Adv. Sci. 2017, 4, 1600214. [Google Scholar] [CrossRef]
  49. Yang, J.; Sun, H.; Liang, H.; Ji, H.; Song, L.; Gao, C.; Xu, H. A Highly Efficient Metal-Free Oxygen Reduction Electrocatalyst Assembled from Carbon Nanotubes and Graphene. Adv. Mater. 2016, 28, 4606–4613. [Google Scholar] [CrossRef]
  50. Mukerjee, S.; Srinivasan, S. Enhanced electrocatalysis of oxygen reduction on platinum alloys in proton exchange membrane fuel cells. J. Electroanal. Chem. 1993, 357, 201–224. [Google Scholar] [CrossRef]
  51. Kundu, S.; Nagaiah, T.C.; Xia, W.; Wang, Y.; Van Dommele, S.; Bitter, J.H.; Santa, M.; Grundmeier, G.; Bron, M.; Schuhmann, W.; et al. Electrocatalytic activity and stability of nitrogen-containing carbon nanotubes in the oxygen reduction reaction. J. Phys. Chem. C 2009, 113, 14302–14310. [Google Scholar] [CrossRef]
  52. Su, B.; Hatay, I.; Trojánek, A.; Samec, Z.; Khoury, T.; Gros, C.P.; Barbe, J.M.; Daina, A.; Carrupt, P.A.; Girault, H.H. Molecular electrocatalysis for oxygen reduction by cobalt porphyrins adsorbed at liquid/liquid interfaces. J. Am. Chem. Soc. 2010, 132, 2655–2662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Gotti, G.; Fajerwerg, K.; Evrard, D.; Gros, P. Electrodeposited gold nanoparticles on glassy carbon: Correlation between nanoparticles characteristics and oxygen reduction kinetics in neutral media. Electrochim. Acta 2014, 128, 412–419. [Google Scholar] [CrossRef] [Green Version]
  54. Xing, P.; Wu, D.; Chen, J.; Song, J.; Mao, C.; Gao, Y.; Niu, H. A Cd-MOF as a fluorescent probe for highly selective, sensitive and stable detection of antibiotics in water. Analyst 2019, 144, 2656–2661. [Google Scholar] [CrossRef]
  55. Park, H.J.; Suh, M.P. Mixed-ligand metal-organic frameworks with large pores: Gas sorption properties and single-crystal-to-single-crystal transformation on guest exchange. Chem. A Eur. J. 2008, 14, 8812–8821. [Google Scholar] [CrossRef]
  56. Villajos, J.A.; Orcajo, G.; Martos, C.; Botas, J.Á.; Villacañas, J.; Calleja, G. Co/Ni mixed-metal sited MOF-74 material as hydrogen adsorbent. Int. J. Hydrogen Energy 2015, 40, 5346–5352. [Google Scholar] [CrossRef]
  57. Wong-Foy, A.G.; Matzger, A.J.; Yaghi, O.M. Exceptional H2 saturation uptake in microporous metal-organic frameworks. J. Am. Chem. Soc. 2006, 128, 3494–3495. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Yuan, S.; Feng, X.; Li, H.; Zhou, J.; Wang, B. Preparation of nanofibrous metal-organic framework filters for efficient air pollution control. J. Am. Chem. Soc. 2016, 138, 5785–5788. [Google Scholar] [CrossRef]
  59. Zhao, R.; Tian, Y.; Li, S.; Ma, T.; Lei, H.; Zhu, G. An electrospun fiber based metal–organic framework composite membrane for fast, continuous, and simultaneous removal of insoluble and soluble contaminants from water. J. Mater. Chem. A 2019, 7, 22559–22570. [Google Scholar] [CrossRef]
  60. Li, H.; Eddaoudi, M.; O’Keefe, 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]
  61. Zhou, H.C.; Long, J.R.; Yaghi, O.M. Introduction to metal-organic frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef] [PubMed]
  62. 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]
  63. Férey, C.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. Chemistry: A chromium terephthalate-based solid with unusually large pore volumes and surface area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  64. Banerjee, R.; Phan, A.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O.M. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 939, 939–944. [Google Scholar] [CrossRef] [PubMed]
  65. Lu, W.; Wei, Z.; Gu, Z.Y.; Liu, T.F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle, T.; et al. Tuning the structure and function of metal-organic frameworks via linker design. Chem. Soc. Rev. 2014, 43, 5561–5593. [Google Scholar] [CrossRef]
  66. Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): Routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933–969. [Google Scholar] [CrossRef]
  67. Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O.M. Topological analysis of metal-organic frameworks with polytopic linkers and/or multiple building units and the minimal transitivity principle. Chem. Rev. 2014, 114, 1343–1370. [Google Scholar] [CrossRef]
  68. Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; Sun, Y.; Qin, J.; Yang, X.; Zhang, P.; et al. Stable Metal–Organic Frameworks: Design, Synthesis, and Applications. Adv. Mater. 2018, 30, 1704303. [Google Scholar] [CrossRef] [Green Version]
  69. Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Férey, G. A Rationale for the Large Breathing of the Porous Aluminum Terephthalate (MIL-53) Upon Hydration. Chem. A Eur. J. 2004, 10, 1373–1382. [Google Scholar] [CrossRef]
  70. Alvarez, E.; Guillou, N.; Martineau, C.; Bueken, B.; Van De Voorde, B.; Le Guillouzer, C.; Fabry, P.; Nouar, F.; Taulelle, F.; De Vos, D.; et al. The Structure of the Aluminum Fumarate Metal–Organic Framework A520. Angewandte 2015, 3664–3668. [Google Scholar] [CrossRef]
  71. Gaab, M.; Trukhan, N.; Maurer, S.; Gummaraju, R.; Müller, U. The progression of Al-based metal-organic frameworks–From academic research to industrial production and applications. Microporous Mesoporous Mater. 2012, 157, 131–136. [Google Scholar] [CrossRef]
  72. Loiseau, T.; Mellot-draznieks, C.; Muguerra, H.; Férey, G.; Haouas, M.; Taulelle, F. Hydrothermal synthesis and crystal structure of a new three-dimensional aluminum-organic framework MIL-69 with 2,6-naphthalenedicarboxylate(ndc),Al(OH)(ndc)·H2O. C. R. Chim. 2005, 8, 765–772. [Google Scholar] [CrossRef]
  73. Loiseau, T.; Lecroq, L.; Volkringer, C.; Marrot, J.; Fèrey, G.; Haouas, M.; Taulelle, F.; Bourrelly, S.; Llewellyn, P.L.; Latroche, M. MIL-96, a Porous Aluminum Trimesate 3D Structure Constructed from a Hexagonal Network of 18-Membered Rings and µ3 -Oxo-Centered Trinuclear Units. J. Am. Chem. Soc. 2006, 128, 10223–10230. [Google Scholar] [CrossRef] [PubMed]
  74. Volkringer, C.; Popov, D.; Loiseau, T.; Fèrey, G.; Burghammer, M.; Riekel, C.; Haouas, M.; Taulelle, F. Synthesis, Single-Crystal X-ray Microdiffraction and NMR Characterizaion of the Giant Pore Metal-Organic Framework Aluminum Trimesate MIL-100. Chem. Mater. 2009, 21, 5695–5697. [Google Scholar] [CrossRef]
  75. Serra-Crespo, P.; Ramos-Fernandez, E.V.; Gascon, J.; Kapteijn, F. Synthesis and characterization of an amino functionalized MIL-101(Al): Separation and catalytic properties. Chem. Mater. 2011, 23, 2565–2572. [Google Scholar] [CrossRef]
  76. Volkringer, C.; Popov, D.; Loiseau, T.; Guillou, N.; Fèrey, G.; Haouas, M.; Taulelle, F.; Mellot-Draznieks, C.; Burghammer, M.; Riekel, C. A microdiffraction set-up for nanoporous metal-organic-framework-type solids. Nat. Mater. 2007, 6, 760–764. [Google Scholar] [CrossRef]
  77. Volkringer, C.; Loiseau, T.; Guillou, N.; Fèrey, G.; Haouas, M.; Taulelle, F.; Audebrand, N.; Margiolaki, I.; Popov, D.; Burghammer, M.; et al. Structural transitions and flexibility during dehydration-Rehydration process in the MOF-type aluminum pyromellitate A12(OH)2[C10O8H2](MIL-118). Cryst. Growth Des. 2009, 9, 2927–2936. [Google Scholar] [CrossRef]
  78. Volkringer, C.; Loiseau, T.; Haouas, M.; Taulelle, F.; Popov, D.; Burghammer, M.; Riekel, C.; Zlotea, C.; Cuevas, F.; Latroche, M.; et al. Occurrence of uncommon infinite chains consisting of edge-sharing octahedra in a porous metal organic framework-type aluminum pyromellitate Al4(OH)8[C10O8H2](MIL-120): Synthesis, Structure, and Gas Sorption Properties. Chem. Mater. 2009, 21, 5783–5791. [Google Scholar] [CrossRef]
  79. Volkringer, C.; Loiseau, T.; Guillou, N.; Férey, G.; Haouas, M.; Taulelle, F.; Elkaim, E.; Stock, N. High-throughput aided synthesis of the porous metal-organic framework-type aluminum pyromellitate, MIL-121, with extra carboxylic acid functionalization. Inorg. Chem. 2010, 49, 9852–9862. [Google Scholar] [CrossRef]
  80. Volkringer, C.; Loiseau, T.; Guillou, N.; Férey, G.; Elkaïm, E. Syntheses and structures of the MOF-type series of metal 1,4,5,8,-naphthalenetetracarboxylates M2(OH)2[C14O8H4](Al, Ga, In) with infinite trans-connected M-OH-M chains (MIL-122). Solid State Sci. 2009, 11, 1507–1512. [Google Scholar] [CrossRef]
  81. Senkovska, I.; Hoffmann, F.; Fröba, M.; Getzschmann, J.; Böhlmann, W.; Kaskel, S. New highly porous aluminium based metal-organic frameworks: Al(OH)(ndc) (ndc = 2,6-naphthalene dicarboxylate) and Al(OH)(bpdc) (bpdc = 4,4′-biphenyl dicarboxylate). Microporous Mesoporous Mater. 2009, 122, 93–98. [Google Scholar] [CrossRef]
  82. Yang, S.; Sun, J.; Ramirez-Cuesta, A.J.; Callear, S.K.; David, W.I.; Anderson, D.P.; Newby, R.; Blake, A.J.; Parker, J.E.; Tang, C.C.; et al. Selectivity and direct visualization of carbon dioxide and sulfur dioxide in a decorated porous host. Nat. Chem. 2012, 4, 887–894. [Google Scholar] [CrossRef] [PubMed]
  83. Ahnfeldt, T.; Guillou, N.; Gunzelmann, D.; Margiolaki, I.; Loiseau, T.; Férey, G.; Senker, J.; Stock, N. [Al4(OH)2(OCH3)4(H2N-Bdc)3xH2O: A 12-connected porous metal-organic framework with an unprecedented aluminum-containing brick. Angew. Chem. Int. Ed. 2009, 48, 5163–5166. [Google Scholar] [CrossRef] [PubMed]
  84. Reinsch, H.; Feyand, M.; Ahnfeldt, T.; Stock, N. CAU-3: A new family of porous MOFs with a novel Al-based brick: [Al2(OCH3)4(O2C-X-CO2)] (X = aryl). Dalton Trans. 2012, 41, 4164–4171. [Google Scholar] [CrossRef]
  85. Reinsch, H.; Krüger, M.; Wack, J.; Senker, J.; Salles, F.; Maurin, G.; Stock, N. A new aluminium-based microporous metal-organic framework: Al(BTB) (BTB = 1,3,5-benzenetrisbenzoate). Microporous Mesoporous Mater. 2012, 157, 50–55. [Google Scholar] [CrossRef]
  86. Reinsch, H.; Krüger, M.; Marrot, J.; Stock, N. First keto-functionalized microporous Al-based metal-organic framework: [Al(OH)(O2C-C6H4-CO-C6H4-CO2)]. Inorg. Chem. 2013, 52, 1854–1859. [Google Scholar] [CrossRef] [Green Version]
  87. Barpanda, P.; Liu, G.; Ling, C.D.; Tamaru, M.; Avdeev, M.; Chung, S.C.; Yamada, Y.; Yamada, A. Na2FeP2O7: A safe cathode for rechargeable sodium-ion batteries. Chem. Mater. 2013, 25, 3480–3487. [Google Scholar] [CrossRef]
  88. Wang, Z.W.; Chen, M.; Liu, C.S.; Wang, X.; Zhao, H.; Du, M. A Versatile AlIII-Based Metal-Organic Framework with High Physicochemical Stability. Chem. Eur. J. 2015, 21, 17215–17219. [Google Scholar] [CrossRef]
  89. Fateeva, A.; Chater, P.A.; Ireland, C.P.; Tahir, A.A.; Khimyak, Y.Z.; Wiper, P.V.; Darwent, J.R.; Rosseinsky, M.J. A water-stable porphyrin-based metal-organic framework active for visible-light photocatalysis. Angew. Chem. Int. Ed. 2012, 51, 7440–7444. [Google Scholar] [CrossRef]
  90. Feng, D.; Liu, T.F.; Su, J.; Bosch, M.; Wei, Z.; Wan, W.; Yuan, D.; Chen, Y.P.; Wang, X.; Wang, K.; et al. Stable metal-organic frameworks containing single-molecule traps for enzyme encapsulation. Nat. Commun. 2015, 6, 1–8. [Google Scholar] [CrossRef] [Green Version]
  91. Lian, X.; Chen, Y.P.; Liu, T.F.; Zhou, H.C. Coupling two enzymes into a tandem nanoreactor utilizing a hierarchically structured MOF. Chem. Sci. 2016, 7, 6969–6973. [Google Scholar] [CrossRef] [Green Version]
  92. Alezi, D.; Belmabkhout, Y.; Suyetin, M.; Bhatt, P.M.; Weseliński, L.J.; Solovyeva, V.; Adil, K.; Spanopoulos, I.; Trikalitis, P.N.; Emwas, A.H.; et al. MOF Crystal Chemistry Paving the Way to Gas Storage Needs: Aluminum-Based soc -MOF for CH4, O2, and CO2 Storage. J. Am. Chem. Soc. 2015, 137, 13308–13318. [Google Scholar] [CrossRef]
  93. Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.; Férey, G. Very large breathing effect in the first nanoporous chromium(III)-based solids: MIL-53 or CrIII(OH)·{O2C-C6H4-CO2}·{HO2C-C6H4 -CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124, 13519–13526. [Google Scholar] [CrossRef]
  94. Serre, C.; Mellot-Draznieks, C.; Surlbé, S.; Audebrand, N.; Filinchuk, Y.; Férey, G. Role of Solvent-Host Interactions That Lead to Very Large Swelling of Hybrid Frameworks. Science 2007, 315, 1828–1832. [Google Scholar] [CrossRef] [Green Version]
  95. Long, P.; Wu, H.; Zhao, Q.; Wang, Y.; Dong, J.; Li, J. Solvent effect on the synthesis of MIL-96(Cr) and MIL-100(Cr). Microporous Mesoporous Mater. 2011, 142, 489–493. [Google Scholar] [CrossRef]
  96. Férey, G.; Serre, C.; Mellot-Draznieks, C.; 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. Ed. 2004, 43, 6296–6301. [Google Scholar] [CrossRef]
  97. Sonnauer, A.; Hoffmann, F.; Fröba, M.; Kienle, L.; Duppel, V.; Thommes, M.; Serre, C.; Férey, G.; Stock, N. Giant pores in a chromium 2,6-Naphthalenedicarboxylate Open- Framework structure with MIL-101 topology. Angew. Chem. Int. Ed. 2009, 48, 3791–3794. [Google Scholar] [CrossRef]
  98. Lian, X.; Feng, D.; Chen, Y.P.; Liu, T.F.; Wang, X.; Zhou, H.C. The preparation of an ultrastable mesoporous Cr(III)-MOF via reductive labilization. Chem. Sci. 2015, 6, 7044–7048. [Google Scholar] [CrossRef] [Green Version]
  99. Liu, T.F.; Zou, L.; Feng, D.; Chen, Y.P.; Fordham, S.; Wang, X.; Liu, Y.; Zhou, H.C. Stepwise synthesis of robust metal-organic frameworks via postsynthetic metathesis and oxidation of metal nodes in a single-crystal to single-crystal transformation. J. Am. Chem. Soc. 2014, 136, 7813–7816. [Google Scholar] [CrossRef]
  100. Whitfield, T.R.; Wang, X.; Liu, L.; Jacobson, A.J. Metal-organic frameworks based on iron oxide octahedral chains connected by benzenedicarboxylate dianions. Solid State Sci. 2005, 7, 1096–1103. [Google Scholar] [CrossRef]
  101. Fateeva, A.; Horcajada, P.; Devic, T.; Serre, C.; Marrot, J.; Grenèche, J.M.; Morcrette, M.; Tarascon, J.M.; Maurin, G.; Férey, G. Synthesis, structure, characterization, and redox properties of the porous MIL-68(Fe) solid. Eur. J. Inorg. Chem. 2010, 24, 3789–3794. [Google Scholar] [CrossRef]
  102. Fateeva, A.; Clarisse, J.; Pilet, G.; Grenèche, J.M.; Nouar, F.; Abeykoon, B.K.; Guegan, F.; Goutaudier, C.; Luneau, D.; Warren, J.E.; et al. Iron and porphyrin metal-organic frameworks: Insight into structural diversity, stability, and porosity. Cryst. Growth Des. 2015, 15, 1819–1826. [Google Scholar] [CrossRef]
  103. Horcajada, P.; Surblé, S.; Serre, C.; Hong, D.Y.; Seo, Y.K.; Chang, J.S.; Grenèche, J.M.; Margiolaki, I.; Férey, G. Synthesis and catalytic properties of MIL-100(Fe), an iron(III) carboxylate with large pores. Chem. Commun. 2007, 27, 2820–2822. [Google Scholar] [CrossRef]
  104. Lupu, D.; Ardelean, O.; Blanita, G.; Borodi, G.; Lazar, M.D.; Biris, A.R.; Ioan, C.; Mihet, M.; Misan, I.; Popeneciu, G. Synthesis and hydrogen adsorption properties of a new iron based porous metal-organic framework. Int. J. Hydrogen Energy 2011, 36, 3586–3592. [Google Scholar] [CrossRef]
  105. Feng, D.; Wang, K.; Wei, Z.; Chen, Y.P.; Simon, C.M.; Arvapally, R.K.; Martin, R.L.; Bosch, M.; Liu, T.F.; Fordham, S.; et al. Kinetically tuned dimensional augmentation as a versatile synthetic route towards robust metal-organic frameworks. Nat. Commun. 2014, 5, 1–8. [Google Scholar] [CrossRef]
  106. Wang, K.; Feng, D.; Liu, T.F.; Su, J.; Yuan, S.; Chen, Y.P.; Bosch, M.; Zou, X.; Zhou, H.C. A series of highly stable mesoporous metalloporphyrin Fe-MOFs. J. Am. Chem. Soc. 2014, 136, 13983–13986. [Google Scholar] [CrossRef]
  107. Reineke, T.M.; Eddaoudi, M.; Fehr, M.; Kelley, D.; Yaghi, O.M. From condensed lanthanide coordination solids to microporous frameworks having accessible metal sites. J. Am. Chem. Soc. 1999, 121, 1651–1657. [Google Scholar] [CrossRef]
  108. Serre, C.; Férey, G. Hydrothermal synthesis, thermal behaviour and structure determination from powder data of a porous three-dimensional europium trimesate: Eu3(H2O)(OH)6[C6H3(CO2)3]·3H2O or MIL-63. J. Mater. Chem. 2002, 12, 3053–3057. [Google Scholar] [CrossRef]
  109. Millange, F.; Serre, C.; Marrot, J.; Garbdant, N.; Pellé, F.; Férey, G. Synthesis, structure and properties of a three-dimensional porous rare-earth carboxylate MIL-83(Eu): Eu2(O2C-C10H14-CO2)3. J. Mater. Chem. 2004, 14, 642–645. [Google Scholar] [CrossRef]
  110. Devic, T.; Serre, C.; Audebrand, N.; Marrot, J.; Férey, G. MIL-103, a 3-D lanthanide-based metal organic framework with large one-dimensional tunnels and a high surface area. J. Am. Chem. Soc. 2005, 127, 12788–12789. [Google Scholar] [CrossRef]
  111. Jiang, H.L.; Tsumori, N.; Xu, Q. A series of (6,6)-connected porous lanthanide-organic framework enantiomers with high thermostability and exposed metal sites: Scalable syntheses, structures, and sorption properties. Inorg. Chem. 2010, 49, 10001–10006. [Google Scholar] [CrossRef]
  112. Xue, D.X.; Cairns, A.J.; Belmabkhout, Y.; Wojtas, L.; Liu, Y.; Alkordi, M.H.; Eddaoudi, M. Tunable rare-earth fcu-MOFs: A platform for systematic enhancement of CO2 adsorption energetics and uptake. J. Am. Chem. Soc. 2013, 135, 7660–7667. [Google Scholar] [CrossRef]
  113. Assen, A.H.; Belmabkhout, Y.; Adil, K.; Bhatt, P.M.; Xue, D.; Jiang, H.; Eddaoudi, M. Ultra-Tuning of the Rare-Earth fcu-MOF Aperture Size for Selective Molecular Exclusion of Branched Paraffins. Angew. Chem. Int. Ed. 2015, 54, 14353–14358. [Google Scholar] [CrossRef]
  114. Lammert, M.; Wharmby, M.T.; Smolders, S.; Bueken, B.; Lieb, A.; Lomachenko, K.A.; Vos, D.D.; Stock, N. Cerium-based metal organic frameworks with UiO-66 architecture: Synthesis, properties and redox catalytic activity. Chem. Commun. 2015, 51, 12578–12581. [Google Scholar] [CrossRef] [Green Version]
  115. Dalapati, R.; Sakthivel, B.; Dhakshinamoorthy, A.; Buragohain, A.; Bhunia, A.; Janiak, C.; Biswas, S. A highly stable dimethyl-functionalized Ce(IV)-based UiO-66 metal-organic framework material for gas sorption and redox catalysis. CrystEngComm 2016, 18, 7855–7864. [Google Scholar] [CrossRef]
  116. Dan-Hardi, M.; Serre, C.; Frot, T.; Rozes, L.; Maurin, G.; Sanchez, C.; Férey, G. A new photoactive crystalline highly porous titanium(IV) dicarboxylate. J. Am. Chem. Soc. 2009, 131, 10857–10859. [Google Scholar] [CrossRef]
  117. Yuan, S.; Liu, T.F.; Feng, D.; Tian, J.; Wang, K.; Qin, J.; Zhang, Q.; Chen, Y.P.; Bosch, M.; Zou, L.; et al. A single crystalline porphyrinic titanium metal-organic framework. Chem. Sci. 2015, 6, 3926–3930. [Google Scholar] [CrossRef] [Green Version]
  118. Bueken, B.; Vermoortele, F.; Vanpoucke, D.E.P.; Reinsch, H.; Tsou, C.; Valvekens, P.; De Baerdemaeker, T.; Ameloot, R.; Kirschhock, C.E.A.; Van Speybroeck, V.; et al. A Flexible Photoactive Titanium Metal–Organic Framework Based on a [TiIV33-O)(O)2(COO)6] Cluster. Angew. Chem. Int. Ed. 2015, 54, 13912–13917. [Google Scholar] [CrossRef]
  119. Furukawa, H.; Doan, T.L.H.; Cordova, K.E.; Yaghi, O.M. A Titanium–Organic Framework as an Exemplar of Combining the Chemistry of Metal–and Covalent–Organic Frameworks. J. Am. Chem. Soc. 2016, 138, 4330–4333. [Google Scholar]
  120. Nguyen, H.L.; Vu, T.T.; Le, D.; Doan, T.L.H.; Nguyen, V.Q.; Phan, N.T.S. A Titanium–Organic Framework: Engineering of the Band-Gap Energy for Photocatalytic Property Enhancement. ACS Catal. 2017, 7, 338–342. [Google Scholar] [CrossRef]
  121. Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. [Google Scholar] [CrossRef]
  122. Wei, Z.; Gu, Z.; Arvapally, R.K.; Chen, Y.; Mcdougald, R.N.; Ivy, J.F.; Yakovenko, A.A.; Feng, D.; Omary, M.A.; Zhou, H. Rigidifying Fluorescent Linkers by Metal–Organic Framework Formation for Fluorescence Blue Shift and Quantum Yield Enhancement. J. Am. Chem. Soc. 2014, 136, 8269–8276. [Google Scholar] [CrossRef]
  123. Feng, D.; Gu, Z.; Li, J.; Jiang, H.; Wei, Z.; Zhou, H. Zirconium-Metalloporphyrin PCN-222: Mesoporous Metal–Organic Frameworks with Ultrahigh Stability as Biomimetic Catalysts. Angew. Chem. Int. Ed. 2012, 51, 10307–10310. [Google Scholar] [CrossRef]
  124. Feng, D.; Gu, Z.; Chen, Y.; Park, J.; Wei, Z.; Sun, Y.; Bosch, M.; Yuan, S.; Zhou, H. A Highly Stable Porphyrinic Zirconium Metal–Organic Framework with shp-a Topology. J. Am. Chem. Soc. 2014, 136, 17714–17717. [Google Scholar] [CrossRef]
  125. Feng, D.; Chung, W.; Wei, Z.; Gu, Z.; Jiang, H.; Chen, Y.; Darensbourg, D.J.; Zhou, H. Construction of Ultrastable Porphyrin Zr Metal–Organic Frameworks through Linker Elimination. J. Am. Chem. Soc. 2013, 135, 17105–17110. [Google Scholar] [CrossRef]
  126. Jiang, H.; Feng, D.; Wang, K.; Gu, Z.; Wei, Z.; Chen, Y.; Zhou, H. An Exceptionally Stable, Porphyrinic Zr Metal–Organic Framework Exhibiting pH-Dependent Fluorescence. J. Am. Chem. Soc. 2013, 135, 13934–13938. [Google Scholar] [CrossRef]
  127. Liu, T.; Feng, D.; Chen, Y.; Zou, L.; Bosch, M.; Yuan, S.; Wei, Z.; Fordgam, S.; Wang, K.; Zhou, H. Topology-Guided Design and Syntheses of Highly Stable Mesoporous Porphyrinic Zirconium Metal–Organic Frameworks with High Surface Area. J. Am. Chem. Soc. 2015, 137, 413–419. [Google Scholar] [CrossRef] [Green Version]
  128. Zhang, M.; Chen, Y.; Bosch, M.; Gentle, T.; Wang, K.; Feng, D.; Wang, Z.U.; Zhou, H. Symmetry-Guided Synthesis of Highly Porous Metal–Organic Frameworks with Fluorite Topology. Angew. Chem. Int. Ed. 2014, 53, 815–818. [Google Scholar] [CrossRef]
  129. Yuan, S.; Lu, W.; Chen, Y.; Zhang, Q.; Liu, T.; Feng, D.; Wang, X.; Qin, J.; Zhou, H. Sequential Linker Installation: Precise Placement of Functional Groups in Multivariate Metal–Organic Frameworks. J. Am. Chem. Soc. 2015, 137, 3177–3180. [Google Scholar] [CrossRef]
  130. Feng, D.; Wang, K.; Su, J.; Liu, T.; Park, J.; Wei, Z.; Bosch, M.; Yakovenko, A.; Zou, X.; Zhou, H. A Highly Stable Zeotype Mesoporous Zirconium Metal–Organic Framework with Ultralarge Pores. Angew. Chem. Int. Ed. 2015, 54, 149–154. [Google Scholar] [CrossRef]
  131. Yuan, S.; Qin, J.; Zou, L.; Chen, Y.; Wang, X.; Zhang, Q.; Zhou, H. Thermodynamically Guided Synthesis of Mixed-Linker Zr-MOFs with Enhanced Tunability. J. Am. Chem. Soc. 2016, 138, 6636–6642. [Google Scholar] [CrossRef] [PubMed]
  132. Furukawa, H.; Gándara, F.; Zhang, Y.; Jiang, J.; Queen, W.L.; Hudson, M.R.; Yaghi, O.M. Water Adsorption in Porous Metal–Organic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, 136, 4369–4381. [Google Scholar] [CrossRef] [PubMed]
  133. Morris, W.; Volosskiy, B.; Demir, S.; Gándara, F.; Mcgrier, P.L.; Furukawa, H.; Cascio, D.; Stoddart, F.; Yaghi, O.M. Synthesis, Structure, and Metalation of Two New Highly Porous Zirconium Metal–Organic Frameworks. Inorg. Chem. 2012, 51, 6443–6445. [Google Scholar] [CrossRef] [PubMed]
  134. Bon, V.; Senkovskyy, V.; Senkovska, I.; Kaskel, S. Zr(IV) and Hf(IV) based metal-organic frameworks with reo-topology. Chem. Commun. 2012, 48, 8407–8409. [Google Scholar] [CrossRef] [Green Version]
  135. Bon, V.; Senkovska, I.; Weiss, M.S.; Kaskel, S. Tailoring of network dimensionality and porosity adjustment in Zr- and Hf-based MOFs. CrystEngComm 2013, 15, 9572–9577. [Google Scholar] [CrossRef]
  136. Bon, V.; Senkovska, I.; Baburin, I.A.; Kaskel, S. Zr- and Hf-Based Metal–Organic Frameworks: Tracking Down the Polymorphism. Cryst. Growth Des. 2013, 13, 1231–1237. [Google Scholar] [CrossRef]
  137. Deria, P.; Mondloch, J.E.; Tylianakis, E.; Ghosh, P.; Bury, W.; Snurr, R.Q.; Hupp, J.T.; Farha, O.K. Perfl uoroalkane Functionalization of NU-1000 via Solvent-Assisted Ligand Incorporation: Synthesis and CO2 Adsorption Studies. J. Am. Chem. Soc. 2013, 135, 16801–16804. [Google Scholar] [CrossRef]
  138. Gutov, O.V.; Bury, W.; Gomez-gualdron, D.A.; Krungleviciute, V.; Fairen-Jimenez, D.; Mondloch, J.E.; Sarjeant, A.A.; Al-Juaid, S.S.; Snurr, R.Q.; Hupp, J.T.; et al. Water-Stable Zirconium-Based Metal–Organic Framework Material with High-Surface Area and Gas-Storage Capacities. Chem. Eur. J. 2014, 20, 12389–12393. [Google Scholar] [CrossRef] [Green Version]
  139. Wang, T.C.; Bury, W.; Gómez-Gualdrón, D.A.; Vermeulen, N.A.; Mondloch, J.E.; Deria, P.; Zhang, K.; Moghadam, P.Z.; Sarjeant, A.A.; Snurr, R.Q.; et al. Ultrahigh Surface Area Zirconium MOFs and Insights into the Applicability of the BET Theory. J. Am. Chem. Soc. 2015, 137, 3585–3591. [Google Scholar] [CrossRef]
  140. Guillerm, V.; Ragon, F.; Devic, T.; Vishnuvarthan, M.; Campo, B.; Vimont, A.; Clet, G.; Yang, Q.; Maurin, G.; Férey, G.; et al. A Series of Isoreticular, Highly Stable, Porous Zirconium Oxide Based Metal–Organic Frameworks. Angew. Chem. Int. Ed. 2012, 51, 9267–9271. [Google Scholar] [CrossRef]
  141. Wang, B.; Lv, X.; Feng, D.; Xie, L.; Zhang, J.; Li, M.; Xie, Y.; Li, J.; Zhou, H. Highly Stable Zr(IV)-Based Metal–Organic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204–6216. [Google Scholar] [CrossRef] [PubMed]
  142. Schaate, A.; Dühnen, S.; Platz, G.; Lilienthal, S.; Schneider, A.M.; Behrens, P. A Novel Zr-Based Porous Coordination Polymer Containing Azobenzenedicarboxylate as a Linker. Eur. J. Inorg. Chem. 2012, 2012, 790–796. [Google Scholar] [CrossRef]
  143. Lv, X.; Tong, M.; Huang, H.; Wang, B.; Gan, L.; Yang, Q.; Zhong, C.; Li, J. Journal of Solid State Chemistry A high surface area Zr (IV) -based metal–organic framework showing stepwise gas adsorption and selective dye uptake. J. Solid State Chem. 2015, 223, 104–108. [Google Scholar] [CrossRef]
  144. Schaate, A.; Roy, P.; Preuße, T.; Lohmeier, S.J.; Godt, A.; Behrens, P. Porous Interpenetrated Zirconium–Organic Frameworks (PIZOFs): A Chemically Versatile Family of Metal–Organic Frameworks. Chem. Eur. J. 2011, 17, 9320–9325. [Google Scholar] [CrossRef] [PubMed]
  145. Yoon, M.; Moon, D. New Zr (IV) based metal-organic framework comprising a sulfur-containing ligand: Enhancement of CO2 and H2 storage capacity. Microporous Mesoporous Mater. 2015, 215, 116–122. [Google Scholar] [CrossRef]
  146. Kalidindi, S.B.; Nayak, S.; Briggs, M.E.; Jansat, S.; Katsoulidis, A.P.; Miller, G.J.; Warren, J.E.; Antypov, D.; Corà, F.; Slater, B.; et al. Chemical and Structural Stability of Zirconium-based Metal–Organic Frameworks with Large Three-Dimensional Pores by Linker Engineering. Angew. Chem. Int. Ed. 2015, 54, 221–226. [Google Scholar] [CrossRef] [Green Version]
  147. Wang, R.; Wang, Z.; Xu, Y.; Dai, F.; Zhang, L.; Sun, D. Porous Zirconium Metal–Organic Framework Constructed from 2D → 3D Interpenetration Based on a 3,6-Connected kgd Net. Inorg. Chem. 2014, 53, 7086–7088. [Google Scholar] [CrossRef]
  148. Cliffe, M.J.; Castillo-Martinez, E.; Wu, Y.; Lee, J.; Forse, A.C.; Firth, F.C.N.; Moghadam, P.Z.; Fairen-Jimenez, D.; Gaultois, M.W.; Hill, J.A.; et al. Metal–Organic Nanosheets Formed via Defect-Mediated Transformation of a Hafnium Metal–Organic Framework. J. Am. Chem. Soc. 2017, 139, 5397–5404. [Google Scholar] [CrossRef]
  149. Ji, P.; Manna, K.; Lin, Z.; Feng, X.; Urban, A.; Song, Y.; Lin, W. 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, 139, 7004–7011. [Google Scholar] [CrossRef]
  150. Cao, L.; Lin, Z.; Shi, W.; Wang, Z.; Zhang, C.; Hu, X.; Wang, C.; Lin, W. Exciton Migration and Amplified Quenching on Two-Dimensional Metal–Organic Layers. J. Am. Chem. Soc. 2017, 139, 7020–7029. [Google Scholar] [CrossRef]
  151. Tăbăcaru, A.; Galli, S.; Pettinari, C.; Masciocchi, N.; McDonald, T.M.; Long, J.R. Nickel(II) and copper (I,II)-based metal-organic frameworks incorprating an extended tris-pyrazolate linker. CrystEngComm 2015, 17, 4992–5001. [Google Scholar] [CrossRef]
  152. Colombo, V.; Galli, S.; Choi, H.J.; Han, G.D.; Maspero, A.; Palmisano, G.; Masciocchi, N.; Long, J.R. High thermal and chemical stability in pyrazolate-bridged metal–organic frameworks with exposed metal sites. Chem. Sci. 2011, 2, 1311–1319. [Google Scholar] [CrossRef] [Green Version]
  153. Choi, H.J.; Dinc, M.; Dailly, A.; Long, J.R. Hydrogen storage in water-stable metal–organic frameworks incorporating. Energy Environ. Sci. 2010, 3, 117–123. [Google Scholar] [CrossRef]
  154. Wang, K.; Lv, X.; Feng, D.; Li, J.; Chen, S.; Sun, J.; Song, L.; Xie, Y.; Li, J.; Zhou, H. Pyrazolate-Based Porphyrinic Metal–Organic Framework with Extraordinary Base-Resistance. J. Am. Chem. Soc. 2016, 138, 914–919. [Google Scholar] [CrossRef]
  155. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; Keeffe, M.O.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 10, 10186–10191. [Google Scholar] [CrossRef] [Green Version]
  156. Wu, H.; Qian, X.; Zhu, H.; Ma, S.; Zhu, G.; Long, Y. Controlled synthesis of highly stable zeolitic imidazolate framework-67 dodecahedra and their use towards the templated formation of a hollow Co3O4 catalyst for CO oxidation. RSC Adv. 2016, 6, 6915–6920. [Google Scholar] [CrossRef]
  157. Morris, W.; Doonan, C.J.; Furukawa, H.; Banerjee, R.; Yaghi, O.M. Crystals as Molecules: Postsynthesis Covalent Functionalization of Zeolitic Imidazolate Frameworks. J. Am. Chem. Soc. 2008, 130, 12626–12627. [Google Scholar] [CrossRef]
  158. Xiong, Y.; Chen, S.; Ye, F.; Su, L.; Zhang, C.; Shen, S.; Zhao, S. Synthesis of a mixed valence state Ce-MOF as an oxidase mimetic for the colorimetric detection of biothiols. Chem. Commun. 2015, 51, 4635–4638. [Google Scholar] [CrossRef]
  159. Zhang, J.P.; Zhang, Y.B.; Lin, J.B.; Chen, X.M. Metal azolate frameworks: From crystal engineering to functional materials. Chem. Rev. 2012, 112, 1001–1033. [Google Scholar] [CrossRef]
  160. Goupil, J.M.; Hemidy, J.F.; Cornet, D. Adsorption of NO2 on modified Y zeolites. Zeolites 1982, 2, 47–50. [Google Scholar] [CrossRef]
  161. Levasseur, B.; Ebrahim, A.M.; Bandosz, T.J. Role of Zr4+ cations in NO2 adsorption on Ce1-xZrxO2 mixed oxides at ambient conditions. Langmuir 2011, 27, 9379–9386. [Google Scholar] [CrossRef] [PubMed]
  162. Levasseur, B.; Ebrahim, A.M.; Bandosz, T.J. Interactions of NO2 with amine-functionalized SBA-15: Effects of synthesis route. Langmuir 2012, 28, 5703–5714. [Google Scholar] [CrossRef] [PubMed]
  163. Florent, M.; Tocci, M.; Bandosz, T.J. NO2 adsorption at ambient temperature on urea-modified ordered mesoporous carbon. Carbon 2013, 63, 283–293. [Google Scholar] [CrossRef]
  164. Belhachemi, M.; Jeguirim, M.; Limousy, L.; Addoun, F. Comparison of NO2 removal using date pits activated carbon and modified commercialized activated carbon via different preparation methods: Effect of porosity and surface chemistry. Chem. Eng. J. 2014, 253, 121–129. [Google Scholar] [CrossRef]
  165. Ryu, H.J.; Grace, J.R.; Lim, C.J. Simultaneous CO2/SO2 capture characteristics of three limestones in a fluidized-bed reactor. Energy Fuels 2006, 20, 1621–1628. [Google Scholar] [CrossRef]
  166. Tchalala, M.R.; Bhatt, P.M.; Chappanda, K.N.; Tavares, S.R.; Adil, K.; Belmabkhout, Y.; Shkurenko, A.; Cadiau, A.; Heymans, N.; De Weireld, G.; et al. Fluorinated MOF platform for selective removal and sensing of SO2 from flue gas and air. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef] [Green Version]
  167. Savage, M.; Cheng, Y.; Easun, T.L.; Eyley, J.E.; Argent, S.P.; Warren, M.R.; Lewis, W.; Murray, C.; Tang, C.C.; Frogley, M.D.; et al. Selective Adsorption of Sulfur Dioxide in a Robust Metal–Organic Framework Material. Adv. Mater. 2016, 28, 8705–8711. [Google Scholar] [CrossRef]
  168. Luan, Y.; Qi, Y.; Gao, H.; Zheng, N.; Wang, G. Synthesis of an amino-functionalized metal-organic framework at a nanoscale level for gold nanoparticle deposition and catalysis. J. Mater. Chem. A 2014, 2, 20588–20596. [Google Scholar] [CrossRef]
  169. Vinoth, R.; Babu, S.G.; Bahnemann, D.; Neppolian, B. Nitrogen doped reduced graphene oxide hybrid metal free catalyst for effective reduction of 4-nitrophenol. Sci. Adv. Mater. 2015, 7, 1443–1449. [Google Scholar] [CrossRef]
  170. Firestone, M.; Kavlock, R.; Zenick, H.; Kramer, M. The U.S. environmental protection agency strategic plan for evaluating the toxicity of chemicals. J. Toxicol. Environ. Health Part B 2010, 13, 139–162. [Google Scholar] [CrossRef]
  171. Tacconelli, E.; Carrara, E.; Savoldi, A.; Harbarth, S.; Mendelson, M.; Monnet, D.L.; Pulcini, C.; Kahlmeter, G.; Kluytmans, J.; Carmeli, Y.; et al. Discovery, research, and development of new antibiotics: The WHO priority list of antibiotic-resistant bacteria and tuberculosis. Lancet Infect. Dis. 2018, 18, 318–327. [Google Scholar] [CrossRef]
  172. Yan, C.; Yang, Y.; Zhou, J.; Liu, M.; Nie, M.; Shi, H.; Gu, L. Antibiotics in the surface water of the Yangtze Estuary: Occurrence, distribution and risk assessment. Environ. Pollut. 2013, 175, 22–29. [Google Scholar] [CrossRef] [PubMed]
  173. Papadopoulos, F.; Parissopoulos, G.; Papadopoulos, A.; Zdragas, A.; Ntanos, D.; Prochaska, C.; Metaxa, I. Assessment of reclaimed municipal wastewater application on rice cultivation. Environ. Manag. 2009, 43, 135–143. [Google Scholar] [CrossRef] [PubMed]
  174. Pan, M.; Chu, L.M. Fate of antibiotics in soil and their uptake by edible crops. Sci. Total Environ. 2017, 599–600, 500–512. [Google Scholar] [CrossRef]
  175. Chen, F.; Ying, G.G.; Kong, L.X.; Wang, L.; Zhao, J.L.; Zhou, L.J.; Zhang, L.J. Distribution and accumulation of endocrine-disrupting chemicals and pharmaceuticals in wastewater irrigated soils in Hebei, China. Environ. Pollut. 2011, 159, 1490–1498. [Google Scholar] [CrossRef]
  176. Segura, P.A.; Takada, H.; Correa, J.A.; Saadi, K.E.; Koike, T.; Onwona-Agyeman, S.; Ofosu-Anim, J.; Sabi, E.B.; Wasonga, O.V.; Mghalu, J.M.; et al. Global occurrence of anti-infectives in contaminated surface waters: Impact of income inequality between countries. Environ. Int. 2015, 80, 89–97. [Google Scholar] [CrossRef]
  177. Martinez, J.L. Environmental pollution by antibiotics and by antibiotic resistance determinants. Environ. Pollut. 2009, 157, 2893–2902. [Google Scholar] [CrossRef]
  178. Nandi, S.; Banesh, S.; Trivedi, V.; Biswas, S. A dinitro-functionalized metal-organic framework featuring visual and fluorogenic sensing of H2S in living cells, human blood plasma and environmental samples. Analyst 2018, 143, 1482–1491. [Google Scholar] [CrossRef]
  179. Yao, Z.Q.; Li, G.Y.; Xu, J.; Hu, T.L.; Bu, X.H. A Water-Stable Luminescent ZnII Metal-Organic Framework as Chemosensor for High-Efficiency Detection of CrVI-Anions (Cr2O72− and CrO42−) in Aqueous Solution. Chem. A Eur. J. 2018, 24, 3192–3198. [Google Scholar] [CrossRef]
  180. Hou, S.L.; Dong, J.; Jiang, X.L.; Jiao, Z.H.; Wang, C.M.; Zhao, B. Interpenetration-Dependent Luminescent Probe in Indium-Organic Frameworks for Selectively Detecting Nitrofurazone in Water. Anal. Chem. 2018, 90, 1516–1519. [Google Scholar] [CrossRef] [Green Version]
  181. Håkansson, K.; Coorey, R.V.; Zubarev, R.A.; Talrose, V.L.; Håkansson, P. Low-mass ions observed in plasma desorption mass spectrometry of high explosives. J. Mass Spectrom. 2000, 35, 337–346. [Google Scholar] [CrossRef]
  182. Moros, J.; Laserna, J.J. New Raman-laser-induced breakdown spectroscopy identity of explosives using parametric data fusion on an integrated sensing platform. Anal. Chem. 2011, 83, 6275–6285. [Google Scholar] [CrossRef] [PubMed]
  183. Benito-Peña, E.; Urraca, J.L.; Moreno-Bondi, M.C. Quantitative determination of penicillin V and amoxicillin in feed samples by pressurised liquid extraction and liquid chromatography with ultraviolet detection. J. Pharm. Biomed. Anal. 2009, 49, 289–294. [Google Scholar] [CrossRef] [PubMed]
  184. Samanta, P.; Desai, A.V.; Sharma, S.; Chandra, P.; Ghosh, S.K. Selective Recognition of Hg2+ ion in Water by a Functionalized Metal-Organic Framework (MOF) Based Chemodosimeter. Inorg. Chem. 2018, 57, 2360–2364. [Google Scholar] [CrossRef]
  185. Wang, C.X.; Xia, Y.P.; Yao, Z.Q.; Xu, J.; Chang, Z.; Bu, X.H. Two luminescent coordination polymers as highly selective and sensitive chemosensors for CrVI -anions in aqueous medium. Dalton Trans. 2019, 48, 387–394. [Google Scholar] [CrossRef]
  186. Liu, X.J.; Zhang, Y.H.; Chang, Z.; Li, A.L.; Tian, D.; Yao, Z.Q.; Jia, Y.Y.; Bu, X.H. A Water-Stable Metal-Organic Framework with a Double-Helical Structure for Fluorescent Sensing. Inorg. Chem. 2016, 55, 7326–7328. [Google Scholar] [CrossRef]
  187. Xing, K.; Fan, R.; Du, X.; Song, Y.; Chen, W.; Zhou, X.; Zheng, X.; Wang, P.; Yang, Y. A windmill-like Zn3L2 cage exhibiting conformational change imparted sensing for DMA and highly selective naked-eye detection of Co2+ ion by dynamic quenching. Sens. Actuators B Chem. 2018, 257, 68–76. [Google Scholar] [CrossRef]
  188. Du, T.; Zhang, H.; Ruan, J.; Jiang, H.; Chen, H.Y.; Wang, X. Adjusting the Linear Range of Au-MOF Fluorescent Probes for Real-Time Analyzing Intracellular GSH in Living Cells. ACS Appl. Mater. Interfaces 2018, 10, 12417–12423. [Google Scholar] [CrossRef]
  189. Wang, M.; Guo, L.; Cao, D. Amino-Functionalized Luminescent Metal-Organic Framework Test Paper for Rapid and Selective Sensing of SO2 Gas and Its Derivatives by Luminescence Turn-On Effect. Anal. Chem. 2018, 90, 3608–3614. [Google Scholar] [CrossRef]
  190. Tabrizchi, M.; ILbeigi, V. Detection of explosives by positive corona discharge ion mobility spectrometry. J. Hazard. Mater. 2010, 176, 692–696. [Google Scholar] [CrossRef]
  191. Moreno-González, D.; Lara, F.J.; Jurgovská, N.; Gámiz-Gracia, L.; García-Campaña, A.M. Determination of aminoglycosides in honey by capillary electrophoresis tandem mass spectrometry and extraction with molecularly imprinted polymers. Anal. Chim. Acta 2015, 891, 321–328. [Google Scholar] [CrossRef] [PubMed]
  192. Blasco, C.; Corcia, A.D.; Picó, Y. Determination of tetracyclines in multi-specie animal tissues by pressurized liquid extraction and liquid chromatography-tandem mass spectrometry. Food Chem. 2009, 116, 1005–1012. [Google Scholar] [CrossRef]
  193. Zhou, Z.; Han, M.L.; Fu, H.R.; Ma, L.F.; Luo, F.; Li, D.S. Engineering design toward exploring the functional group substitution in 1D channels of Zn-organic frameworks upon nitro explosives and antibiotics detection. Dalton Trans. 2018, 47, 5359–5365. [Google Scholar] [CrossRef] [PubMed]
  194. Huang, X.Y.; Yue, K.F.; Jin, J.C.; Liu, J.Q.; Wang, C.J.; Wang, Y.Y.; Shi, Q.Z. Three-dimensional fivefold interpenetrating microporous metal-organic framework based on mixed flexible ligands. Inorg. Chem. Commun. 2010, 13, 338–341. [Google Scholar] [CrossRef]
  195. Maspoch, D.; Ruiz-Molina, D.; Veciana, J. Old materials with new tricks: Multifunctional open-framework materials. Chem. Soc. Rev. 2007, 36, 770–818. [Google Scholar] [CrossRef]
  196. Allendorf, M.D.; Bauer, C.A.; Bhakta, R.K.; Houk, R.J.T. Luminescent metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1330–1352. [Google Scholar] [CrossRef]
  197. Dai, J.C.; Wu, X.T.; Fu, Z.Y.; Cui, C.P.; Hu, S.M.; Du, W.X.; Wu, L.M.; Zhang, H.H.; Sun, R.Q. Synthesis, structure, and fluorescence of the novel cadmium(II)-Trimesate coordination polymers with different coordination architectures. Inorg. Chem. 2002, 41, 1391–1396. [Google Scholar] [CrossRef]
  198. Tian, D.; Li, Y.; Chen, R.Y.; Chang, Z.; Wang, G.Y.; Bu, X.H. A luminescent metal-organic framework demonstrating ideal detection ability for nitroaromatic explosives. J. Mater. Chem. A 2014, 2, 1465–1470. [Google Scholar] [CrossRef]
  199. Yam, V.W.-W.; Lo, K.K.-W. Luminescent polynuclear d10 metal complexes. Chem. Soc. Rev. 1999, 28, 323–334. [Google Scholar] [CrossRef]
  200. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal-organic frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef]
  201. Fiore, A.M.; Naik, V.; Spracklen, D.V.; Steiner, A.; Unger, N.; Prather, M.; Bergmann, D.; Cameron-Smith, P.J.; Cionni, I.; Collins, W.J.; et al. Global air quality and climate. Chem. Soc. Rev. 2012, 41, 6663–6683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Lelieveld, J.; Evans, J.S.; Fnais, M.; Giannadaki, D.; Pozzer, A. The contribution of outdoor air pollution sources to premature mortality on a global scale. Nature 2015, 525, 367–371. [Google Scholar] [CrossRef] [PubMed]
  203. Hoek, G.; Krishnan, R.M.; Beelen, R.; Peters, A.; Ostro, B.; Brunekreef, B.; Kaufman, J.D. Long-term air pollution exposure and cardio-respiratory mortality: A review. Environ. Health 2013, 12, 43–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Seaton, A.; MacNee, W.; Donaldson, K.; Godden, D. Particulate air pollution and acute health effects. Lancet 1995, 345, 176–178. [Google Scholar] [CrossRef]
  205. Choi, S.; Kim, H.R.; Kim, H.S. Fabrication of superabsorbent nanofibers based on sodium polyacrylate/poly(vinyl alcohol) and their water absorption characteristics. Polym. Int. 2019, 68, 764–771. [Google Scholar] [CrossRef]
  206. Carroll, C.P.; Joo, Y.L. Axisymmetric instabilities of electrically driven viscoelastic jets. J. Non-Newton. Fluid Mech. 2008, 153, 130–148. [Google Scholar] [CrossRef]
  207. Hwang, Y.J.; Choi, S.; Kim, H.S. Structural deformation of PVDF nanoweb due to electrospinning behavior affected by solvent ratio. E-Polymers 2018, 18, 339–345. [Google Scholar] [CrossRef]
  208. Li, D.; Xia, Y. Fabrication of titania nanofibers by electrospinning. Nano Lett. 2003, 3, 555–560. [Google Scholar] [CrossRef]
  209. Kong, C.S.; Lee, S.G.; Lee, S.H.; Lee, K.Y.; Noh, H.W.; Yoo, W.S.; Kim, H.S. Electrospinning Instabilities in the Drop Formation and Multi-Jet Ejection Part I: Various Concentrations of PVA (Polyvinyl Alcohol) Polymer Solution. J. Macromol. Sci. Part B 2011, 50, 517–527. [Google Scholar] [CrossRef]
  210. Kong, C.S.; Yoo, W.S.; Jo, N.G.; Kim, H.S. Electrospinning mechanism for producing nanoscale polymer fibers. J. Macromol. Sci. Part B Phys. 2010, 49, 122–131. [Google Scholar] [CrossRef]
  211. Dai, X.H.; Fan, H.X.; Yi, C.Y.; Dong, B.; Yuan, S.J. Solvent-free synthesis of a 2D biochar stabilized nanoscale zerovalent iron composite for the oxidative degradation of organic pollutants. J. Mater. Chem. A 2019, 7, 6849–6858. [Google Scholar] [CrossRef]
  212. Song, J.Y.; Bhadra, B.N.; Khan, N.A.; Jhung, S.H. Adsorptive removal of artificial sweeteners from water using porous carbons derived from metal azolate framework-6. Microporous Mesoporous Mater. 2018, 260, 1–8. [Google Scholar] [CrossRef]
  213. Liu, N.; Chen, Y.; Zhang, W.; Qu, R.; Zhang, Q.; Feng, L.; Wei, Y. A versatile CeO2/Co3O4 coated mesh for food wastewater treatment: Simultaneous oil removal and UV catalysis of food additives. Water Res. 2018, 137, 144–152. [Google Scholar] [CrossRef] [PubMed]
  214. Pintor, A.M.A.; Vilar, V.J.P.; Botelho, C.M.S.; Boaventura, R.A.R. Oil and grease removal from wastewaters: Sorption treatment as an alternative to state-of-the-art technologies. A critical review. Chem. Eng. J. 2016, 297, 229–255. [Google Scholar] [CrossRef]
  215. Zhu, Z.; Wang, W.; Qi, D.; Luo, Y.; Liu, Y.; Xu, Y.; Cui, F.; Wang, C.; Chen, X. Calcinable Polymer Membrane with Revivability for Efficient Oily-Water Remediation. Adv. Mater. 2018, 30, 1–8. [Google Scholar] [CrossRef] [PubMed]
  216. Chatzisymeon, E.; Stypas, E.; Bousios, S.; Xekoukoulotakis, N.P.; Mantzavinos, D. Photocatalytic treatment of black table olive processing wastewater. J. Hazard. Mater. 2008, 154, 1090–1097. [Google Scholar] [CrossRef] [PubMed]
  217. Dong, J.; Xu, F.; Dong, Z.; Zhao, Y.; Yan, Y.; Jin, H.; Li, Y. Fabrication of two dual-functionalized covalent organic polymers through heterostructural mixed linkers and their use as cationic dye adsorbents. RSC Adv. 2018, 8, 19075–19084. [Google Scholar] [CrossRef] [Green Version]
  218. Zhang, L.; Sun, J.S.; Sun, F.; Chen, P.; Liu, J.; Zhu, G. Facile Synthesis of Ultrastable Porous Aromatic Frameworks by Suzuki–Miyaura Coupling Reaction for Adsorption Removal of Organic Dyes. Chem. A Eur. J. 2019, 25, 3903–3908. [Google Scholar] [CrossRef]
  219. Sun, D.T.; Gasilova, N.; Yang, S.; Oveisi, E.; Queen, W.L. Rapid, Selective Extraction of Trace Amounts of Gold from Complex Water Mixtures with a Metal-Organic Framework (MOF)/Polymer Composite. J. Am. Chem. Soc. 2018, 140, 16697–16703. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Various applications by metal-organic frameworks (MOFs).
Scheme 1. Various applications by metal-organic frameworks (MOFs).
Catalysts 10 00195 sch001
Figure 1. Gas uptake properties of MFM-300 (Al) in various gas environment such as NO2, CO2, SO2, CH4, CO, N2, H2, O2, and Ar (open symbol: adsorption, solid symbol: desorption). Reproduced and adapted from Ref. [6]; Copyright (2018), Springer Nature.
Figure 1. Gas uptake properties of MFM-300 (Al) in various gas environment such as NO2, CO2, SO2, CH4, CO, N2, H2, O2, and Ar (open symbol: adsorption, solid symbol: desorption). Reproduced and adapted from Ref. [6]; Copyright (2018), Springer Nature.
Catalysts 10 00195 g001
Figure 2. (a) Detection of SO2 in the 75 to 1000 ppb concentration range (b) linear response for MFM-300 (In) MOF-sensor upon exposure to SO2 for a 24-day (c) reproducibility cycles for the detection. Reproduced and adapted from Ref. [12]; Copyright (2018), RSC Journal of Materials Chemistry A.
Figure 2. (a) Detection of SO2 in the 75 to 1000 ppb concentration range (b) linear response for MFM-300 (In) MOF-sensor upon exposure to SO2 for a 24-day (c) reproducibility cycles for the detection. Reproduced and adapted from Ref. [12]; Copyright (2018), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 g002
Figure 3. Environment effect of sensing performance with (a) relative humidity (b) temperature and (c) selectivity performance in various gases of MFM-300 (In) MOF sensor. Reproduced and adapted from Ref. [12]; Copyright (2018), RSC Journal of Materials Chemistry A.
Figure 3. Environment effect of sensing performance with (a) relative humidity (b) temperature and (c) selectivity performance in various gases of MFM-300 (In) MOF sensor. Reproduced and adapted from Ref. [12]; Copyright (2018), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 g003
Scheme 2. Synthesis of AgPd@UiO-66-L and the electrochemical reduction mechanism of 4-nitrophenol.
Scheme 2. Synthesis of AgPd@UiO-66-L and the electrochemical reduction mechanism of 4-nitrophenol.
Catalysts 10 00195 sch002
Figure 4. Scanning electron microscopy (SEM) image (a), TEM image (b), HAADF image (c), total element (d), and elemental mapping (ej) of AgPd@UiO-66-NH2. The bar indicate 100 nm (a), 50 nm (b), and 40 nm (cj). Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Figure 4. Scanning electron microscopy (SEM) image (a), TEM image (b), HAADF image (c), total element (d), and elemental mapping (ej) of AgPd@UiO-66-NH2. The bar indicate 100 nm (a), 50 nm (b), and 40 nm (cj). Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Catalysts 10 00195 g004
Figure 5. PXRD pattern of UiO-66 (a), UiO-66-NH2 (b), UiO-66-NO2 (c), Ag@UiO-66-NH2 (d), Pd@UiO-66-NH2 (e), AgPd@UiO-66-NH2 (f), and AgPd@UiO-66-NO2 (g). Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Figure 5. PXRD pattern of UiO-66 (a), UiO-66-NH2 (b), UiO-66-NO2 (c), Ag@UiO-66-NH2 (d), Pd@UiO-66-NH2 (e), AgPd@UiO-66-NH2 (f), and AgPd@UiO-66-NO2 (g). Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Catalysts 10 00195 g005
Figure 6. (a) CVs of the identical electrodes in 0.1 M PBS solution including 0.5 mM of 4-nitrophenol, (b) CVs of AgPd@UiO-66-NH2/GCE in 0.1 M PBS at various concentration 4-nitrophenol (0.25 μM–400 μM), (c) UV-VIS absorption spectra of 4-nitrophenol before addition of NaBH4 solution as compared to after, and (d) Intensity of absorbance peak change detection via UV-VIS spectra for the reduction of 4-nitrophenol within NaBH4 in AgPd@UiO-66-NH2 during the time-dependent. Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Figure 6. (a) CVs of the identical electrodes in 0.1 M PBS solution including 0.5 mM of 4-nitrophenol, (b) CVs of AgPd@UiO-66-NH2/GCE in 0.1 M PBS at various concentration 4-nitrophenol (0.25 μM–400 μM), (c) UV-VIS absorption spectra of 4-nitrophenol before addition of NaBH4 solution as compared to after, and (d) Intensity of absorbance peak change detection via UV-VIS spectra for the reduction of 4-nitrophenol within NaBH4 in AgPd@UiO-66-NH2 during the time-dependent. Reproduced and adapted from Ref. [11]; Copyright (2018), Elsevier Sensors and Actuators B: Chemical.
Catalysts 10 00195 g006
Figure 7. (a) Thermal gravity analysis graph of Cd-MOF and (b) PXRD spectrum of Cd-MOF neglected in various pH aqueous solution and water for 24 h. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Figure 7. (a) Thermal gravity analysis graph of Cd-MOF and (b) PXRD spectrum of Cd-MOF neglected in various pH aqueous solution and water for 24 h. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Catalysts 10 00195 g007
Figure 8. The fluorescence spectra of solid samples of Cd-MOF, bbi, and H2L organic ligand. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Figure 8. The fluorescence spectra of solid samples of Cd-MOF, bbi, and H2L organic ligand. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Catalysts 10 00195 g008
Figure 9. (a) Comparison of various antibiotics quenching efficiency using Cd-MOF at room temperature. (b) Dispersed in aqueous solution of CRO for fluorescence spectra (c) CRO aqueous solution tested in various pH aqueous solution using Cd-MOF, and (d) Fluorescence intensity of CRO solution in Cd-MOF suspensions in time dependent. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Figure 9. (a) Comparison of various antibiotics quenching efficiency using Cd-MOF at room temperature. (b) Dispersed in aqueous solution of CRO for fluorescence spectra (c) CRO aqueous solution tested in various pH aqueous solution using Cd-MOF, and (d) Fluorescence intensity of CRO solution in Cd-MOF suspensions in time dependent. Reproduced and adapted from Ref. [54]; Copyright (2019), RSC Analyst.
Catalysts 10 00195 g009
Scheme 3. Electrospinning Setup.
Scheme 3. Electrospinning Setup.
Catalysts 10 00195 sch003
Figure 10. SEM images of polymer and various MOFs loaded nonwoven fabrics. Reproduced and adapted from Ref. [58]; Copyright (2016), ACS Journal of American Society.
Figure 10. SEM images of polymer and various MOFs loaded nonwoven fabrics. Reproduced and adapted from Ref. [58]; Copyright (2016), ACS Journal of American Society.
Catalysts 10 00195 g010
Figure 11. (a) Particulate matter removal efficiencies of polyacrylonitrile (PAN) filter, Al2O3@PAN filter and MOF@PAN filter (b) Long term PM2.5 removal efficiencies of PAN filter and ZIF-8@PAN filter. Reproduced and adapted from Ref. [58]; Copyright (2016), ACS Journal of American Society.
Figure 11. (a) Particulate matter removal efficiencies of polyacrylonitrile (PAN) filter, Al2O3@PAN filter and MOF@PAN filter (b) Long term PM2.5 removal efficiencies of PAN filter and ZIF-8@PAN filter. Reproduced and adapted from Ref. [58]; Copyright (2016), ACS Journal of American Society.
Catalysts 10 00195 g011
Scheme 4. (a) Schematic illustration of fabricating the PAN@MIL-100 (Fe) filter. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Scheme 4. (a) Schematic illustration of fabricating the PAN@MIL-100 (Fe) filter. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 sch004
Figure 12. SEM image of (a) H3BTC/PAN filter and (b) PAN@MIL-100 (Fe) filter. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Figure 12. SEM image of (a) H3BTC/PAN filter and (b) PAN@MIL-100 (Fe) filter. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 g012
Figure 13. (a) Removal efficiency and (b) adsorption kinetic curves toward AR and VA by PAN@MIL-100 (Fe)filter. (c) Adsorption-desorption cycles. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Figure 13. (a) Removal efficiency and (b) adsorption kinetic curves toward AR and VA by PAN@MIL-100 (Fe)filter. (c) Adsorption-desorption cycles. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 g013
Figure 14. (a) Contact angles of water and oil in air and under water. (b) Separation efficiency versus the recycling number. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Figure 14. (a) Contact angles of water and oil in air and under water. (b) Separation efficiency versus the recycling number. Reproduced and adapted from Ref. [59]; Copyright (2019), RSC Journal of Materials Chemistry A.
Catalysts 10 00195 g014
Table 1. Abridgment of some stable MOFs [68].
Table 1. Abridgment of some stable MOFs [68].
MOFs a)Clusters/CoresLinkers b)BET Surface Area (m2 g−1)Ref.
MIL-53 (Al)[Al(OH)(COO)2]nBDC1181[69]
Al-FUM[Al(OH)(COO)2]nFUM1080[70,71]
MIL-69[Al(OH)(COO)2]n2,6-NDCNA[72]
MIL-96 (Al)[Al33-O)(COO)6]
[Al(OH)(COO)2]n
BTCNA[73]
MIL-100 (Al)[Al33-O)(COO)6]BTC2152[74]
MIL-101 (Al)[Al33-O)(COO)6]BDC-NH22100[75]
MIL-110[Al8(OH)15(COO)9]BTC1400[76]
MIL-118[Al(OH)(COO)2(COOH)2]nBTECNA[77]
MIL-120[Al(OH)(COO)2]nBTEC308[78]
MIL-121[Al(OH)(COO)2]nBTEC162[79]
MIL-122[Al(OH)(COO)2]nNTCNA[80]
DUT-5[Al(OH)(COO)2]nBPDC1613[81]
NOTT-300[Al(OH)(COO)2]nBPTA1370[82]
CAU-1[Al8(OH)4(OCH3)8(COO)12]BDC-NH21700 c)[83]
CAU-3-BDC[Al12(OCH3)24(COO)12]BDC1550[84]
CAU-3-BDC-NH2[Al12(OCH3)24(COO)12]BDC-NH21250[84]
CAU-3-NDC[Al12(OCH3)24(COO)12]2,6-NDC2320[84]
CAU-4[Al(OH)(COO)2]nBTB1520[85]
CAU-8[Al(OH)(COO)2]nBeDC600[86]
CAU-10[Al(OH)(COO)2]n1,3-BDC635[87]
467-MOF[Al(OH)(COO)2]nBTTB725[88]
Al-PMOF[Al(OH)(COO)2]nTCPP1400[89]
PCN-333 (Al)[Al33-O)(COO)6]TATB4000[90]
PCN-888 (Al)[Al33-O)(COO)6]HTB3700[91]
Al-soc-MOF-1[Al33-O)(COO)6]TCPT5585[92]
MIL-53 (Cr)[Cr(OH)(COO)2]nBDCNA[93]
MIL-88A (Cr)[Cr33-O)(COO)6]FUMNA[94]
MIL-88B (Cr)[Cr33-O)(COO)6]BDCNA[94]
MIL-88C (Cr)[Cr33-O)(COO)6]2,6-NDCNA[94]
MIL-88D (Cr)[Cr33-O)(COO)6]BPDCNA[94]
MIL-96 (Cr)[Cr33-O)(COO)6]
[Cr(OH)(COO)2]n
BTCNA[95]
MIL-100 (Cr)[Cr33-O)(COO)6]BTC3100 c)[96]
MIL-101 (Cr)[Cr33-O)(COO)6]BDC4100[63]
MIL-101-NDC (Cr)[Cr33-O)(COO)6]2,6-NDC2100[97]
PCN-333 (Cr)[Cr33-O)(COO)6]TATB2548[98]
PCN-426 (Cr)[Cr33-O)(COO)6]TMQPTC3155[99]
MIL-53 (Fe)[Fe(OH)(COO)2]nBDCNA[100]
MIL-68 (Fe)[Fe(OH)(COO)2]nBDC665[101]
MIL-141 (Fe)[Fe(OH)(COO)2]nTCPP420[102]
FepzTCPP (FeOH)2[Fe(OH)(COO)2]nPyrazine, TCPP760[102]
MIL-88A (Fe)[Fe33-O)(COO)6]FUMNA[94]
MIL-88B (Fe)[Fe33-O)(COO)6]BDCNA[94]
MIL-88C (Fe)[Fe33-O)(COO)6]2,6-NDCNA[94]
MIL-88D (Fe)[Fe33-O)(COO)6]BPDCNA[94]
MIL-100 (Fe)[Fe33-O)(COO)6]BTC2800 c)[103]
MIL-101 (Fe)[Fe33-O)(COO)6]BDC2823[104]
PCN-250 (Fe)[Fe33-O)(COO)6]ABDC1486[105]
PCN-250 (Fe2Co)[Fe2Co(μ3-O)(COO)6]ABDC1400[105]
PCN-333 (Fe)[Fe33-O)(COO)6]TATB2427[90]
PCN-600 (Fe)[Fe3(μ3-O)(COO)6]TCPP2270[106]
Tb2(BDC)3[Tb(H2O)2(COO)3]nBDCNA[107]
MIL-63[Eu23-OH)7(COO)]nBTC15[108]
MIL-83[Eu(μ3-O)3(COO)3(COOH)3]n1,3-ADCNA[109]
MIL-103[Tb(H2O)(COO)4]nBTB930[110]
Y-BTC[Y(H2O)(COO)3]nBTC1080[111]
Tb-BTC[Tb(H2O)(COO)3]nBTC786[111]
Y-FTZB[Y6(μ3-OH)8(COO)6(CN4)6]FTZB1310[112]
Tb-FTZB[Tb6(μ3-OH)8(COO)6(CN4)6]FTZB1220[112]
Y-FUM[Y6(μ3-OH)8(COO)12]FUM691[113]
Tb-FUM[Tb6(μ3-OH)8(COO)12]FUM503[113]
Ce-UiO-66[Ce63-O)43-OH)4(COO)12]BDC1282[114]
Ce-UiO-66-(CH3)2[Ce6(μ3-O)43-OH)4(COO)12]BDC-(CH3)2845[115]
MIL-125[Ti8O8(OH)4(COO)12]BDC1550[116]
PCN-22[Ti7O6(COO)12]TCPP1284[117]
COK-69[Ti3O3(COO)6]CDCNA[118]
MOF-901[Ti6O6(OMe)6(COO)6]AB, BDA550[119]
MOF-902[Ti6O6(OMe)6(COO)6]AB, BPDA400[120]
UiO-66[Zr63-O)43-OH)4(COO)12]BDC1187[121]
UiO-67[Zr63-O)43-OH)4(COO)12]BPDC3000[121]
UiO-68[Zr63-O)43-OH)4(COO)12]TPDC4170[121]
PCN-94[Zr63-O)43-OH)4(COO)12]ETTC3377[122]
PCN-222[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TCPP2223[123]
PCN-223[Zr63-O)43-OH)4(COO)12]TCPP1600[124]
PCN-224[Zr63-O)43-OH)4(OH)6(H2O)6(COO)6]TCPP2600[125]
PCN-225[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TCPP1902[126]
PCN-228[Zr63-O)43-OH)4(COO)12]TCP-14510[127]
PCN-229[Zr63-O)43-OH)4(COO)12]TCP-24619[127]
PCN-230[Zr63-O)43-OH)4(COO)12]TCP-34455[127]
PCN-521[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]MTBC3411[128]
PCN-700[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]Me2BPDC1807[129]
PCN-777[Zr63-O)43-OH)4(OH)6(H2O)6(COO)6]TATB2008[130]
PCN-133[Zr63-O)43-OH)4(COO)12]BTB, DCDPS1462[131]
PCN-134[Zr63-O)43-OH)4(OH)2(H2O)2(COO)10]BTB, TCPP1946[131]
MOF-801[Zr63-O)43-OH)4(COO)12]FUM990[132]
MOF-802[Zr63-O)43-OH)4(OH)2(H2O)2(COO)10]PZDCNA[132]
MOF-808[Zr63-O)43-OH)4(OH)6(H2O)6(COO)6]BTC2060[132]
MOF-812[Zr63-O)43-OH)4(COO)12]MTB2335[132]
MOF-841[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]MTB1390[132]
MOF-525[Zr63-O)43-OH)4(COO)12]TCPP2620[133]
MOF-535[Zr63-O)43-OH)4(COO)12]XF1120[133]
MOF-545[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TCPP2260[133]
DUT-51[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]DTTDC2335[134]
DUT-52[Zr63-O)43-OH)4(COO)12]2,6-NDC1399[135]
DUT-84[Zr63-O)43-OH)4(OH)6(H2O)6(COO)6]2,6-NDC637[135]
DUT-67[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TDC1064[136]
DUT-68[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TDC891[136]
DUT-69[Zr63-O)43-OH)4(OH)2(H2O)2(COO)10]TDC560[136]
NU-1000[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TBAPy2320[137]
NU-1100[Zr63-O)43-OH)4(COO)12]PTBA4020[138]
NU-1101[Zr63-O)43-OH)4(COO)12]Py-XP4422[139]
NU-1102[Zr63-O)43-OH)4(COO)12]Por-PP4712[139]
NU-1103[Zr63-O)43-OH)4(COO)12]Py-PTP5646[139]
NU-1104[Zr63-O)43-OH)4(COO)12]Por-PTP5290[139]
MIL-140A[ZrO(COO)2]nBDC415[140]
MIL-140B[ZrO(COO)2]n2,6-NDC460[140]
MIL-140C[ZrO(COO)2]nBPDC670[140]
MIL-140D[ZrO(COO)2]nCl2ABDC701[140]
BUT-12[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]CTTA3387[141]
BUT-13[Zr63-O)43-OH)4(OH)4(H2O)4(COO)8]TTNA3948[141]
Zr-ABDC[Zr63-O)43-OH)4(COO)12]ABDC3000[142]
BUT-30[Zr63-O)43-OH)4(COO)12]EDDB3940[143]
PIZOF[Zr63-O)43-OH)4(COO)12]PEDC2080[144]
Zr-BTDC[Zr63-O)43-OH)4(COO)12]BTDC2207[145]
Zr-BTBA[Zr63-O)43-OH)4(COO)12]BTBA4342[146]
Zr-PTBA[Zr63-O)43-OH)4(COO)12]PTBA4116[146]
Zr-BTB[Zr63-O)43-OH)4(OH)6(H2O)6(COO)6]BTB613[147]
hcp UiO-67[Hf123-O)83-OH)82-OH)6(COO)18]BPDC1424[148]
Zr12-TPDC[Zr123-O)83-OH)82-OH)6(COO)18]TPDC1967[149]
Hf12-BTE[Hf123-O)83-OH)82-OH)6(COO)18]BTENA[150]
Cu-BTPP[Cu33-OH)(PZ)3]BTPP660[151]
Ni3(BTP)2[Ni4(PZ)8]BTP1650[152]
Zn (1,4-BDP)[Zn(PZ)2]n1,4-BDP1710[153]
Zn (1,3-BDP)[Zn(PZ)2]n1,3-BDP820[153]
PCN-601[Ni8(OH)4(H2O)2(PZ)12]TPP1309[154]
ZIF-8[ZnN4]mIM1947[155]
ZIF-11[ZnN4]bIM1676[155]
ZIF-67[CoN4]mIM1587[156]
ZIF-90[ZnN4]ICA1270[157]
ZIF-68[ZnN4]nIM, bIM1220[64]
ZIF-69[ZnN4]nIM, 5cbIM1070[64]
ZIF-70[ZnN4]IM, nIM1970[64]
a) These MOFs can be modified by functional organic compounds such as amino, nitro, methyl, halogen, or hydroxyl groups. These MOFs are not explained in this paper; b) All linkers name are abbreviations [68]; c) Langmuir surface area.
Table 2. Structure and typical coordination modes of azolates.
Table 2. Structure and typical coordination modes of azolates.
LinkerStructureTypical Coordination Modes
Imidazole (HIM) Catalysts 10 00195 i001 Catalysts 10 00195 i002
Pyrazole (HPZ) Catalysts 10 00195 i003 Catalysts 10 00195 i004
Triazole (HTZ and HVTZ) Catalysts 10 00195 i005 Catalysts 10 00195 i006
Catalysts 10 00195 i007 Catalysts 10 00195 i008
Tetrazole (HTTZ) Catalysts 10 00195 i009 Catalysts 10 00195 i010
Table 3. Saturation NH3 breakthrough capacities value at 0.1% of MOFs.
Table 3. Saturation NH3 breakthrough capacities value at 0.1% of MOFs.
Dry (0% RH)Wet (80% RH)
Co2Cl2BTDD4.783.38
Co2Cl2BBTA8.564.36
Cu2Cl2BBTA7.525.73

Share and Cite

MDPI and ACS Style

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. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10020195

AMA Style

Jang S, Song S, Lim JH, Kim HS, Phan BT, Ha K-T, Park S, Park KH. Application of Various Metal-Organic Frameworks (MOFs) as Catalysts for Air and Water Pollution Environmental Remediation. Catalysts. 2020; 10(2):195. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10020195

Chicago/Turabian Style

Jang, Sanha, Sehwan Song, Ji Hwan Lim, Han Seong Kim, Bach Thang Phan, Ki-Tae Ha, Sungkyun Park, and Kang Hyun Park. 2020. "Application of Various Metal-Organic Frameworks (MOFs) as Catalysts for Air and Water Pollution Environmental Remediation" Catalysts 10, no. 2: 195. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10020195

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