| Home | E-Submission | Sitemap | Contact Us |  
Environ Eng Res > Volume 29(5); 2024 > Article
Lalawmpuia, Lalhruaitluangi, Lalhmunsiama, and Tiwari: Metal organic framework (MOF): Synthesis and fabrication for the application of electrochemical sensing

Abstract

Metal-organic frameworks (MOF) are a class of porous hybrid materials of metal ions linked by organic bridging ligands. These materials consist of different families of crystalline frameworks with metal ions and metal-ion clusters linked by the coordination bonds with suitable organic ligands. These porous materials possess high surface area, variable pore sizes, various functions, and excellent thermal stability. As a result, these unique and tailored MOF materials found varied applications in electrochemical sensor development. The review critically examines the approach to synthesizing a wide range of MOF compounds. The utilization of MOF in fabricating a small-scale sensor device demonstrates the ability to detect various water contaminants in aquatic environments. Moreover, it showcased an in-depth advancement in sensor and device development utilizing MOFs materials. The efficient use of electrochemical sensors is discussed critically and presents these studies’ challenges and future perspectives.

1. Introduction

Metal-organic framework (MOF) is a category of porous materials that has drawn the interest of researchers in the past decades. The MOFs are primarily hybrid materials having organic and inorganic moieties. The organic linker comprises organic ligands in its structure, with the metal cluster acting as inorganic components [1]. The target MOF is distinguished or embedded by the functional groups among metal nodes and organic linkers, supporting the selectivity and reliability of materials for specific purposes, including electrochemical sensing [2]. The MOF showed varied applications in the area of gas adsorption [3], catalysis [4], optical storage media [5], drug delivery [6], sensing, separation [7], redox-active electrode materials [8], etc. due to its micro-to-meso porous structure, high surface area, and flexible structures. Increasing the electrocatalytic signal by immobilizing the metal nanoparticles within the pores of MOFs is advantageous, resulting in suitable and sensitive sensing [9,10].
The focus of research has gradually shifted from the studies on the structure and properties of MOFs and moved toward the studies on the interaction between physical morphology, structure, and applications of these intriguing materials. Most MOFs are polydisperse microcrystalline powders, and these materials possibly possess intrinsic limitations such as improper handling properties, slow mass transfer, and mechanical instability [11,12]. For instance, a MOF packed in powder form within an adsorption column requires high pressure due to the gradual compression of MOF with pressure and, consequently, an increased mass resistance inside the column [13]. Moreover, using MOFs in powder form in catalysis frequently results in problems related to catalyst recycling [14]. The MOFs showed numerous advantages over conventional porous materials because of their large surface area, pore working capabilities, regulated shape, and pore size due to using suitable organic linkers in material synthesis [15]. Due to their distinctive properties, MOFs are known hybrid materials with possible applicability in electrochemical sensing techniques for determining or detecting various micro-pollutants and heavy metals in aquatic environments [10]. Initially, the MOFs were derived using the divalent cations of transition metals viz, zinc (Zn2+), cobalt (Co2+), copper (Cu2+), etc., which included MOF-5, HKUST-1, ZIF-8, etc. Alkaline earth metals (such as Mg, Ca, and Al), transition metals (Ti, Cr, Mn, Fe, Co, Ni, Cu, Zn, Zr, and Cd), lanthanides (such as Pr, La), and other metal ions (Bi) provide inorganic nodes in deriving a variety of MOFs. Organic ligands (POLs) such as imidazole, triazoles, and bi- and tridentate carboxylic acids synthesize the porous MOFs. Metal ions bind with the oxo, carboxylate, and non-metallic sites, known as secondary building units (SBUs), when characterizing MOF structures [16]. According to the kind of building units in the MOF frame, the coordination geometry of MOFs defines the square, pyramidal, trigonal, bipyramidal, tetrahedral, and octahedral geometry [17,18]. The use of divalent transition ions makes the crystallization of synthesized MOF easy; however, divalent cations showed several disadvantages, such as low hydrothermal stability that impacts the annealing of materials and even some applications [19]. Therefore, the selection of metal cations in the synthesis of MOF is crucial.
Further, the thermodynamic studies indicated that higher-charged metal cations possess strong metal-ligand bonds, significantly enhancing MOF’s hydrothermal stability [20]. The stability of the MOF increases with higher valency ions such as iron (Fe), aluminum (Al), zirconium (Zr), and titanium (Ti), which indicates that metal ions with greater valency are more stable than those of divalent metal cations such as zinc (Zn), copper (Cu) and cobalt (Co). In addition, aluminum is abundant on the Earth’s crust, cheap, and lightweight, enabling large-scale production of MOF utilizing aluminium for varied applications such as adsorbent for pollutants removal, sensors, and storage material for gases [21].
Voltammetry and amperometry are the most reliable and efficient electroanalytical techniques for the qualitative and quantitative determination of several analytes at trace levels due to their reasonably high sensitivity, robust instrumentation, and relatively cost-effective compared to other spectroscopic or chromatographic techniques. In recent years, trace-level detection and measurements have drawn more attention using robust electrochemical techniques. With inexpensive instrumentation, it offers quick and reliable detection down to the ppb (parts per billion) levels [22]. Literature shows that these methods offer fresh perspectives on analytical techniques, breaking through previous barriers by producing a straightforward device with many benefits like flexibility, biocompatibility, biodegradability, ease of use, a high surface-to-volume ratio, and affordability [23]. The core component of the electroanalytical techniques is the working electrode at which the analytes react electrochemically. The electrodes fabricated with suitable materials enabled a selective oxidation or reduction of the analyte to be obtained [24,25]. The suitable and strategic electrode alterations enhance the sensing capabilities of the working electrodes [26,27]. Many typical electrodes show poor surface kinetics and significantly decrease selectivity and sensitivity. However, it causes a variety of key difficulties. Therefore, preparing an electrode with high repeatability in its chemical and electrical properties and reproducible pretreatment for a given target analyte at a trace level is a requirement for an electrochemical measurement [28].
Similarly, the MOFs have a practical choice because they selectively analyze the analytes of interest, enabling them to detect them efficiently. The application of MOFs in electroanalytical methods is a relatively newer approach. However, designing an attractive and intriguing framework for detecting the target analytes or biomolecules in aqueous media is challenging in this research area [29]. The specific applications enable the design of suitable MOFs; hence, they are futuristic engineered materials with varied applications in diverse research areas. Kajal et al. recently published a publication demonstrating MOFs’ applications in electrochemical sensing. Their work focuses on applying and fabricating MOF composites for sensing various compounds [30].
Furthermore, MOFs generally possess unique properties such as large specific surface area, pore volume, molecular pore sizes, and flexibility of the framework that is accessible for the target ions/molecules. The application of MOFs in modifying carbon-based electrode studies for sensing and detecting several analytes in the literature. However, designing and fabricating the miniaturized device development utilizing the engineered metal-organic frameworks is the futuristic application of MOFs. The review aims to highlight the synthesis of various kinds of MOFs and focus on the application of the synthesized MOFs in sensing using electroanalytical techniques to detect potential water contaminants at trace levels.

2. Synthesis of MOF

MOFs built from inorganic nodes and organic linkers attracted much interest because of their structural variety, rarity of properties, and capacity to customize for specific functions. The organic linkers and metal centers design the structure of MOFs appropriately; the organic linkers participate in shear connections, and the metal centers act as joints [31]. There are several methods to synthesize MOFs, which are briefly discussed below.

2.1. Hydrothermal/Solvothermal Synthesis

Hydrothermal synthesis is typically adopted to produce sustainable metal-ligand bonds across the framework, which yields the most comprehensive dynamical products [32]. Long-term heating of the reaction mixture at high pressure and temperature enables the hydrothermal synthesis of MOF [33]. Based on the material’s dissolution rate in hot water under intense vapor tension upheld at a temperature variation allying, the reactor’s opposite nodes yield single crystals. Moreover, the ‘solvothermal’ process utilizes suitable solvents besides water in the material synthesis. The method results from the growth of high-quality crystals, but generally, Hydrothermal/Solvothermal processes gradually and necessitate elevated performing temperatures [34]. Fig. 1 illustrates a schematic of the steps in the synthetic approach using hydrothermal and solvothermal processes.
The hydrothermal method synthesizes the Cu (4, 4′-bpy)NO3 (H20) crystals with rectangular parallelepiped-shaped, utilizing the 4, 4′-bipyridine as a nitrogen donor aromatic ligand [35]. Similarly, the hydrothermal method synthesizes the [Cu3(TMA)2(H2O)3]n complex known as HKUST-1 using the carboxylic as a functional group [19]. The hydrothermal route by simple metal ions self-assembling at the adaptive bis-(imidazole) binding sites using the transition metals (cobalt, nickel, and zinc) synthesizes the MOFs having various applications [36]. The hydrothermal reaction yielded a polycrystalline Fe-MIL-100 powder with a significantly large attainable and enduring porosity, demonstrating an intriguing Friedel-Crafts reaction with catalytic activity employing the redox properties of Fe(III) [37]. Ni (II) is a typical transition metal ion with varied interest because of its inexpensive cost, high abundance, superior catalytic activity, and electrochemical characteristics [38,39]. Ni-MOFs are synthesized by the hydrothermal process using Ni(II) and the 1,3,5-benzene tricarboxylate as a ligand [40]. Nickel-based compounds with the proper design showed considerable potential for usage in the electrochemical industry as electrode surface modifications and electrocatalysts [4143]. The solvothermal method synthesizes the Mg-MOF-nNH2 with double ligands. The metal center is magnesium, and the ligands are 2,5-dihydroxyterephthalic acid and 2-amino terephthalic acid. The micropore volume of Mg-MOF-1/8NH2 measures 0.46 cm3g−1 for these MOF samples. Additionally, the synthesized materials have a high specific surface area of 924.19 m2g−1 [44]. Similarly, a single-step solvothermal method obtains the Mn/Fe-MOF@Pd1.0. The Mn2+ and Fe3+ are the metal sources of MOF, and Pd is doped within the MOF. The catalytic activity of Fe-MOF resembles that of an enzyme. Bimetallic active sites enhance the catalytic activity, and the presence of Pd further synergizes the acid tolerance and stability of Mn/Fe-MOF solid [45].

2.2. Ultrasonic Methods

Ultrasonic-assisted synthesis offers a relatively environment-friendly technique for MOF synthesis in an enclosing reaction condition (i.e., enclosing temperature and atmospheric pressure) with a shorter reaction time. Furthermore, the ultrasonication synthesis approaches avoid safety concerns, providing an opportunity to expand on the twelve principles of green chemistry [46]. Amongst some of the diverse MOF synthetic techniques, the ultrasonication method is affordable and environmentally benign. It could produce a high yield while operating under ambient temperature and pressure in a solvent-free reaction [47].
Compared to other methods, ultrasonic-aided MOF synthesis is practically viable and containable and quickly produces the product with a significant yield [48]. Ultrasonic cavitation for producing MOFs is relatively new and has recently attracted greater attention [49]. In 2009, Khan et al. [50] conducted the first study on the impact of ultrasonic irradiation on the [Cu3 (TMA)2(H2O)3]n, also known as Cu-BTC MOF. The ultrasonically assisted synthesis of MOFs significantly reduces the reaction time compared to the electronically controlled and microwave irradiation methods. Additionally, a limited sonication enabled the shrinking of the size of the MOF particles; however, an increased sonication time from 6 to 45 min caused aggregation of the MOFs [51].
Sonochemical irradiation technique synthesizing the urea-containing metal-organic frameworks, i.e., TMU-31 and TMU-32. Further, various parameters, such as the initial reagent concentrations and irradiation periods, determine the morphology of MOF. The results reveal that these MOFs can achieve homogeneous plate shape at 0.005 M concentration with a maximum power supply of 360 W. Furthermore, phenol sensing was assessed and compared using these MOFs. The results indicate that hydrogen-bonding and packing interactions have a substantial role in detecting the phenol. Incorporating urea into the MOF structure is crucial for the framework’s ability to detect phenol [52]. The synthesis of MOFs containing embedded porphyrin units still needs to be done. Under well-regulated reaction conditions, it takes a long time to synthesize, and a combined phase with diverse crystal shapes is nevertheless often observed. The sonochemical synthesis produces the high-purity uniform-sized Zr-based porphyrinic MOF-525 and MOF-545 using the tetrakis (4-carboxyphenyl) porphyrin and zirconyl chloride octahydrate.
Moreover, benzoic acid modulated MOF-525, while trifluoroacetic acid modulated the formation of MOF-545 [53]. Solvothermal and sonochemical techniques synthesize the zinc-based MOF using adipic acid as an aliphatic ditopic linker. Ultrasonic irradiation enables MOF formation efficiency; also, the particle sizes shrunk due to ultrasonic irradiation. The samples’ catalytic effectiveness was assessed using the electrochemical reduction of CO2, with the only byproducts being carbon monoxide and hydrogen [54].
On the other hand, the possible role of solvent in the yield of MOF production using the ultrasound method was reported [55]. The study investigates the effects of solvents, both binary and ternary combinations, on the textural features of Cu-BTC. Additionally, various ultrasonic parameters, such as sonication duration and power supply, determine the formation of MOF (Cf Fig. 2). The sonication process was carried out for 2 h at 750 watts using a mixture of three solvents, namely water (H2O), ethanol (C2H5OH), and dimethylformamide (DMF). This process enhances the yield of copper metal-organic framework (Cu-MOF) [55].

2.3. Microwave-Aided Synthesis

Microwave-aided synthesis is the most efficient approach for carrying out many reactions. Compared to conventional solvothermal synthesis, microwave irradiation reduces the reaction time and enhances the crystal growth of porous materials that take several days or weeks in conventional methods [56]. The microwave irradiation energy, time of exposure, solvent concentration, and solvent systems are the key parameters regulating the yield and crystal growth of the MOFs, and microwave irradiation showed a positive impact on material characteristics and properties [57]. Microwave-assisted synthesis acknowledges rapid heating, fast kinetics, phase purity, increased yield, improved dependability, and repeatability over hydrothermal synthesis [5861]. Additionally, it offers an effective method to regulate the distribution of macroscopic morphology, particle size, and phase selectivity during the synthesis of inorganic solids and nanocomposite materials. Although the synthesis is much faster, the properties of the crystals produced by the microwave-assisted technique are comparable to those produced by the conventional solvothermal process [6264]. Microwave-assisted approaches, on the other hand, have sparked more interest in studying the effects of irradiation period, power, temperature, solvent concentration, and metal ion/organic linker ratio, for example, on the synthesis of MOF-5 by employing the microwave-assisted technique [65]. The nanocrystal crystallization process optimizes for several parameters: time, temperature, and power as 1 h at 130°C and 600–1000 W, respectively. This study showed that the crystal formation occurred under microwave irradiation of 15 min and produced a high-quality crystal between 30 min and 24 h. Similarly, microwave-assisted synthesis produces the Zr-based MOF (Zr-fum-fcu-MOF), which has an octahedral shape at the reaction temperature of 100°C. Fig. 3 depicts an overview of the synthesis procedure [66].

2.4. Electrochemical Synthesis

The electrochemical process of MOF synthesis possesses several advantages over conventional MOF synthesis, including relatively short reaction time, relatively simple equipment setup, real-time MOF structure modification, quick synthesis, no necessity of precursor metal salts, and direct accumulation of MOFs on the preferred substrates, etc. [67,68]. In addition to its simple process, the electrochemical synthesis of MOFs has provided numerous favorable circumstances, such as random and quick synthesis with lesser use of linker and solvent, excellent yield, and low energy consumption [69]. The mild reaction conditions, which perform at ambient temperature and pressure, are the most attractive feature of electrochemical synthesis. Despite these advantages, it is a less utilized approach, particularly in synthesizing functionalized framework materials [70]. This approach also constantly altered the real-time response, enabling the direct output of crack-free nanostructures without a pre-treated surface at high temperatures.
On the other hand, relatively high reaction temperatures, longer reaction times, and thermal-induced cracking on the films are shown in the solvothermal or hydrothermal techniques [71]. Electrochemically synthesized the Zn-based MOF, and the physicochemical parameters, viz., reaction time, electrolyte quantity, current, and voltage, are optimized for greater solid yield. The results indicated that both reaction time and current density significantly impacted the purity and yield, and at an applied current of 60 mA and a reaction time of 2 h, yielded 87% of the product [72]. Fig. 4 shows the instrumentation of a simple electrochemical synthesis of Zn3(BTC)2 MOF [73].

2.5. Mechanochemical Synthesis

Mechanochemical synthesis is one of the most exciting chemical modifications employed to obtain high purity with enhanced yield of several MOFs [74]. Most coordination polymerization processes involving multisite ligands with metal ions proceed readily in a suitable solution environment. The solvent-free or solid-state formation of MOFs without any toxic or hazardous solvents has progressively received attention in recent years due to significant advances in mechanochemical synthesis [50]. Coordination polymerization, broadly, involves the reaction in the presence of solution, multisite ligands, and metal ions. However, the mechanochemical synthesis of MOFs utilizes minimal solvents, is solvent-free, or is a solid-state organic process without unfavorable and toxic solvents [75].
Furthermore, compared to the diffusion and solvothermal methods, this technique is advantageous for large-scale MOF manufacturing in a shorter reaction time and at room temperature [76,77]. The solid-solid reaction has the potential to synthesize a large-scale production of materials and provides simplicity in handling since it directly produces the products in powdered forms. Although mechanochemical synthesis is solvent-free or solvent-less, a solvent-based purification step is required. Despite this, the mechanochemical synthesis is reasonably environmentally friendly and commercially exciting for MOF production [78]. The liquid-assisted mechanochemical synthesis produced the MOF-5, showing that the solid possessed a relatively low BET surface area and contained many amide precursor by-products [79]. A copper-based MOF with a product yield of 97% was synthesized in the mechanochemical process using the Cu(OAc)2.H2O and H3BTC and at 30 Hz and 20 min of reaction time produced a dark blue color solid [55]. The framework structures produced by the mechanochemical method are easily separable from the host molecules, providing repeatable free pore access for additional uses [80]. The mechanochemical synthesis of HKUST-1 and MOF-14 showed the method’s applicability in efficient MOF production [74]. Fig. 5 demonstrates the mechanochemical method of synthesizing MOF.
The different types of MOF have different routes of synthesis. The organic linker should be chosen wisely depending on the central metal atom. Table 1 displays the different types of MOF synthesized using different routes under varying conditions.

2.6. Slow Evaporation Method

In the slow evaporation approach, reagent solutions are blended and allowed to undergo gradual evaporation. A critical concentration produces the crystals, facilitating deposition and subsequent crystal development. Since slow evaporation requires no external energy, it is a straightforward process for preparing MOF compounds. Nonetheless, this approach occasionally designs the MOF, but the primary disadvantage is that the crystal takes a long time to form. Slow evaporation may involve using a solvent combination to increase the starting material’s solubility or a low-boiling solvent to hurry up slow evaporation [81,82].
A two-dimensional coordination polymer, denoted as [Cu(SCN) (hmp)]n CP (1), was synthesized using the slow evaporation technique. The reaction of CuSO4·5H2O, 2-pyridine-methanol (hmpH), sodium thiocyanate, and sodium hydroxide in water are involved, aiming for specific practical uses. The CP (1) explores the fluorescence sensing capabilities of various nitroaromatic compounds and toxic metal complexes. Notably, CP (1) exhibited exceptional sensitivity and selectivity to nitrobenzene, also in competing nitroaromatics. Additionally, CP (1) displayed exceptional sensitivity and selectivity towards Fe3+ compared to many metal ions. CP (1) demonstrated impressive recyclability for both analytes, nearly restoring its initial intensity after multiple washings. Furthermore, CP (1) is an excellent material for the adsorption of organic dyes with different sizes and charges, such as methylene blue (MB), methyl orange (MO), and Rhodamine-B. The CP (1) is designed by bridging the coordination chemistry, having various intriguing functional applications [83].

2.7. Diffusion Method

In the diffusion technique, reagent solutions are either stacked atop one another with a solvent layer in between or undergo a gradual diffusion process facilitated by physical barriers. Occasionally, gels serve as both the medium for crystallization and diffusion. At the interface of these layers, crystals emerge following the gradual dissipation of the precipitating solvent into the subsequent layer [84]. In particular, the diffusion approach is practical when the products have limited solubility. A technique combining ultrasonic and vapor phase diffusion was used for rapidly synthesizing [Tb(1,3,5-BTC)]n, a terbium-based metal-organic framework (MOF).
Compared with existing solvothermal and ultrasound-assisted approaches, this combined ultrasound-vapour phase diffusion strategy produces nanoscale MOF crystals at a significantly higher rate. Notably, the resultant [Tb(1,3,5-BTC)]n nanocrystals showed exceptional selectivity in detecting picric acid (PA), with no interference from other nitroaromatic chemicals such as nitrobenzene, 2-nitrotoluene, 4-nitrotoluene, 2,4-dinitrotoluene, and 2,6-dinitrotoluene. As a result, they are considered attractive candidates for developing novel luminescence sensors designed for the exact detection of PA [85].

2.8. Template Strategy

An alternative approach for synthesizing unique MOFs involves incorporating template molecules into the reaction mixture, a practice seldom seen in conventional synthetic methods, and this template synthesis technique is an effective tool for studying hierarchical porous materials [86]. Organic solvents are one of the template molecules that have been used frequently [87,88], small organic ligands, organic amines [89,90], carboxylic acids or carboxylate groups [9193], and heterocyclic aromatic compounds [9496]. Each template molecule exerts distinct influences on MOFs’ crystallization and synthesis processes. The solubility and polarity of the solvent are the primary factors influencing the crystallization of MOF [87]. Organic amines influence the reaction’s pH and make it easier for organic ligands to deprotonate. [97]. Using carboxylate groups as ligands allows the filling of MOF pores, while heterocyclic compounds can act as counterions and weak organic bases when protonated. Surfactants produce micelles within solvents, controlling the shape and size of the final MOFs, while ionic liquids serve as solvents and counterions [98].

2.9. Microemulsion Method

This technique is widely employed in the production of nanoparticles and has recently found application in synthesizing MOFs [99]. Water microemulsions having nanometer-sized water droplets sustained within the organic phase by a surfactant. The micelles in these microemulsions act as nanoreactors, regulating nucleation and crystal growth rates. The water-surfactant ratio and the surfactant type regulate the microemulsion’s size and number of micelles. This approach provides exact control over the size of nanoscale materials. However, it comes with notable drawbacks, including high costs and the use of surfactants that are often environmental pollutants [100102]. The microemulsion method synthesizes the zeolitic imidazolate framework-8 (ZIF-8), employing mild reaction conditions at standard temperature and pressure and a short reaction time. ZIF-8’s amine groups reacted with the epoxy group, producing exceptional adhesion between ZIF-8 and the epoxy matrix. This prevented the overall ZIF-8 clustering and resulted in increased cross-linking densities. Leveraging these upper hands, incorporating ZIF-8 substantially enhanced the composite coating’s mechanical properties and anti-corrosion efficacy compared to an epoxy cover alone. This study delves into the influence of ZIF-8 content on the coating’s protective performance and identifies the optimal ZIF-8 bulk in the epoxy coating. This finding is a pivotal reference for future applications of metal-organic framework materials in anti-corrosive coatings [103]. The advantages and limitations of various methods employed to synthesize MOFs are summarized and given in Table 2.

3. Fabrication and Design of MOF-Based Sensors

Owing to their high porosity, enormous surface area, and cavity structures, MOFs are valuable in electrochemical sensor applications. These unique properties provide MOFs with a high-level catalytic feature. Moreover, the robust use of these materials on the electrode surface enables efficient development of electrochemical sensors. MOF’s high porosity and surface area enhance these sensors’ detection sensitivity [138]. MOFs show low electronic conduction properties due to their poor overlapping between the electronic states and frontier orbitals of ligand and metal ions [139,140]. Therefore, the direct use of pure MOFs on electrodes or other electroanalytical methods is limited [141143]. A recent report showed the uses of carbon-based materials viz., fullerenes (C60), graphene, and multi-wall carbon nanotubes (MWCNTs) for the modification of nickel-based MOF, which resulted in an increased electrical conductivity with decreased charge transfer resistance, high porosity, and large surface area. These composite materials exhibited enhanced applications in sensor developments [144].

3.1. Carbon-Based Electrode Modification by MOF

Among several electrode modifiers used in electrochemical processes, MOFs have shown potential due to their large surface area, copious adsorption sites, and versatile functionality that helps the detection efficiently for several analytes, including heavy metal ions [145,146]. In general, the carbon allotropes deployed in the fabrication of electrodes are graphite or derived materials intended to modify the carbon paste electrode (CPE) and glassy carbon (GC) for solid carbon electrodes. The conductivity of carbon-based materials allotropes is enhanced primarily due to sp2 hybridized bonds and six-membered aromatic rings [147]. It is essential to pre-modify carbon-based electrodes to follow a specific protocol for introducing MOF modifiers at the electrode surface. In the case of glassy carbon electrodes, the modification involves a direct coating of the polished surface by MOF. In contrast, the modifier fabricates the carbon paste electrode by mixing graphite powder with MOF using a specific organic binder [148].

3.1.1. MOF modified carbon paste electrode (CPE)

The carbon paste electrode is one of the most often utilized electrodes in sense because of its easy and versatile fabrication [149]. The carbon paste electrodes are inexpensive and made by blending graphite with MOF along with a suitable organic binder. High-purity graphite powder with a particle size of 1 to 10 micrometers fabricates electrodes [150]. Further, the graphite powder’s purity greatly influences the electrodes’ electrochemical performance [151]. In addition, the organic binder helps the modifier (MOF) bind at the electrode surface, which must be stable, insoluble, and impurities-free. Paraffin oils, aliphatic and aromatic hydrocarbons, silicone oils and greases, halogenated hydrocarbons, and similar derivatives efficiently bind the materials in the CPEs [152]. Graphite powder has high electrical conductivity; hence, it is easy to fabricate the MOF-modified carbon paste electrodes in the electrochemical processes [153]. In typical CPEs, the modifier (MOF) acts as a Lewis acid in catalysis, which further helps to reduce the sensor’s overpotential and charge transfer resistance. Also, the distinct pore size enhanced the selective determination of various analytes in the complex matrix [154].
Fabricate the carbon paste electrodes using graphite powder with MOF as a modifier and organic binder. The mixture is blended in an agate mortar to obtain a well-distributed suspension placed into an electrode frame. The electrical connections generally utilized copper, titanium, or platinum wires. Teflon tube having a diameter of Ca. 0.5 cm is an electrode frame. The graphite powder and paraffin binder were mixed at 7:3 and then introduced into the electrode frame, and a copper wire was used as electrical contact [155]. The duplicating paper polishes the fabricated electrode surface.
Additionally, two approaches demonstrate the carbon paste electrode fabrication, viz., the in-situ and ex-situ approaches [156]. The in-situ fabrication utilizes the modifier (MOF) with graphite powder. The disadvantage of the in-situ approach is excess graphite powder that adversely impacts the specific properties of the MOFs or even destroys their physical structure [157]. On the other hand, the ex-situ approach utilizes the modifier to mix the MOF with the graphite powder, resulting in composite material [158]. The ex-situ approach is reliable since it provides a more accessible and efficient modification of CPE for various electrochemical applications. Fig. 6 shows a simple schematic of an electrochemical approach.

3.1.2. MOF modified glassy carbon electrode (GCE)

Glassy carbon electrodes showed outstanding physical properties withstanding higher temperatures, hardness, small density, little electrical resistance, low thermal resistance, extreme chemical resistance, and gas and liquid impermeability. Glassy carbon, a non-graphitizing carbon, has physical properties that depend on the 873 to 3273 K annealing temperature. In addition to its excellent thermal stability, glassy carbon possesses exceptional chemical resistance, and its oxidation rates in oxygen, carbon dioxide, and water vapor are lower than those of any other carbon. Glassy carbons are low-density (~1.5 g cm−3) and have limited surface area. The cavities created by randomly orientated and interwoven graphite layers form 1–5 nm pores. The stiff structure produces glassy carbons with higher tensile and compressive strengths than graphite. Moreover, glassy carbon is ideal for electrocatalysis due to its low electrical resistivity (~3–8 ×10−4 Ω cm) [159,160]. A restrained pyrolysis of phenol-formaldehyde resin in an inert atmosphere produces a glassy carbon electrode. Pyrolysis is the thermal decomposition of an organic precursor or hydrocarbons. Glassy carbon is usually obtained at a temperature >2000°C since pyrolysis composites have low thermal conduction, resulting in a thermal drop within the sample [161]. On the other hand, the nanoscale fabrication of glassy carbon is achieved relatively at a lower pyrolysis temperature, i.e., Ca. 900°C associated with oxygen impurities [162].
For the fabrication of the glassy carbon electrode, the glassy carbon electrode was initially polished with an alumina slurry on a polishing cloth in the construction of the working electrode, followed by dipping of the electrode tip in deionized water and ethanol, followed by sonication for 10 min, and the electrode was dried slightly above the room temperature. The modifier (MOF) was mixed with ethanol and then sonicated for 20 min to get a fine solution mixture. It was then applied to the tip of the glassy carbon electrode and dried at room temperature. Ultimately, the Nafion solution was incorporated into the MOF-modified electrode by drop casting method and dried at ambient temperature [163].

4. Application of MOF in Electrochemical Sensing

In electrochemical sensors, using MOF-modified electrodes in amperometric and voltammetric methods gained attention, particularly for detecting several pharmaceuticals, heavy metals, biomolecules, and other micro-pollutants. The MOF-based sensors showed a significantly low detection limit (LOD) and enhanced selectivity compared to the sensors based on other materials.

4.1. Biomolecules Sensing

Biomolecules are the essentials of living organisms that control various biochemical functions of the body. Understanding the essential physiological roles that biomolecules play in guiding the healthy growth and development of the human body is made possible by studying these molecules [164]. Therefore, the quantitative determination of biomolecules is necessary for chemical pathology or even food chemistry. The electrochemical technique has been prominent in quantifying various biomolecules in different matrices [165]. Various methods were employed to detect biomolecules; however, the electrochemical methods are promising and show potential in the trace detection of several biomolecules [166]. Although there is a substantial amount of literature on MOFs, there needs to be more research focused on comprehending the impact of MOF particle characteristics on their biocompatibility [167].
Electroanalytical sensing methods are essential for quantitatively or qualitatively detecting or identifying target analytes in an electrochemical cell by evaluating electrochemical signals such as current or potential [168]. The electrochemical detection of biomolecules utilizing fabricated electrodes operates in two ways: i) Enzyme-modified electrodes and ii) Non-enzymatic or electrodes modified with novel materials (MOF, nanomaterials, metal oxides, etc.). Electrochemical techniques are widely recognized, uncomplicated, economical, and convenient for detecting different biomolecules. It also holds the potential for downsizing healthcare devices [169]. MOFs have the potential to be efficient electrochemical sensors for detecting biomolecules due to their large surface area, regulated varied pore structure, increased functionality, and unique catalytic activity [170]. In order to obtain an enhanced selectivity for an analyte, the methods primarily utilize a variety of enzymes to modify the carbon-based electrodes.
Nevertheless, enzyme-modified electrodes showed several limitations in the sensing process since some enzymes are often less mobile on the electrode surface and are stable only at specific temperatures and pH levels [171,172]. These limiting factors encourage the use of non-enzymatic modified sensors. The enzyme replaces this electrode type with novel materials, especially MOF, which has an exceptional physicochemical property with high sensitivity and selectivity towards various analytes [155].
The MOFs synthesized using transition metals like Fe2+, Cu2+, Co3+, and Ni2+ are promising in low-level electrochemical detection of several biomolecules, viz., glucose, ascorbic acid, urea, and H2O2 [173]. The MOF-based biosensors have garnered significant interest among researchers in developing tools to sense different diseases and health disorders [174]. The detection of glucose is a widely accepted electrochemical method; Cu-MOF@Pt was effectively used to measure the glucose level in human serum samples with excellent recovery and repeatability [175]. The glassy carbon electrode was modified with the nickel-based MOF and employed in cyclic voltammetry (CV) to detect glucose. NiO and Ni/NiO/CNTs display well-defined NiO redox peaks in the absence of glucose and increased peak currents in the addition of glucose.
Furthermore, in the presence and absence of glucose, Ni/NiO/CNTs exhibit significantly greater peak current than NiO, indicating enhanced electrochemical performance. As glucose concentrations rise, the oxidation peak currents rise, with a minor positive shift in peak potential and a drop in the reduction peak currents. The results indicate that Ni/NiO/CNTs have outstanding electrocatalytic activity on glucose oxidation. The Eq. (1 & 2) demonstrate the mechanism of glucose oxidation [176]:
(1)
Ni(OH)2+OH-NiOOH+H2O+e-
(2)
NiOOH+glucoseNi(OH)2+glucolactone
On the other hand, the copper and nickel MOF-based electrodes were fabricated and utilized in the electrochemical detection of glucose. A redox couple of Cu2+/Cu3+ and Ni2+/Ni3+ occurs in an alkaline medium, which in turn leads to the oxidation of transition metals used (i.e., copper or nickel) as shown in Eq. (1) and Eq. (2), which further catalyzes the oxidation of glucose. Fig. S1 shows the reaction mechanism [177].
The mechanism of glucose oxidation on the electrode surface proceeded as [178,179]:
(3)
M(OH)2+OH-MOOH+H2O+e-
(4)
MOOH+glucoseM(OH)2+glucolactone
(5)
Glucolactone (Hydrolysis)Gluconic Acid
The first step oxidizes the M2+ to M3+, and these species take part at the electrode surface in glucose oxidation. The most crucial species for glucose oxidation is M3+, which also serves as the primary electron transfer mediator. M2+ in the MOF undergoes oxidation to produce M3+ throughout the potential scan between 0.3 and 0.5 V. M3+ preferentially oxidizes the glucose to produce gluconolactone, which produces gluconic acid by hydrolysis [179]. Fig. S2(A) shows the differential pulse voltammogram produced when 50.0 μM of glucose in 0.10 M NaOH. Adding 50.0 μM glucose to 0.10 M NaOH gives an oxidation peak current at +0.380 V. Fig. S2(B) demonstrates the amperometric i-t curves for glucose produced by GC/CuO electrode in 0.10 M NaOH agitated solution at an applied voltage of 0.500 V. The initial current response of the GC/CuO electrode was due to 5.0 μM glucose, and the continued supply of 5.0 μM glucose in every step with a sample interval of 50.0 sec enhances the response of the current [177].
A chromium-based metal-organic framework (MIL-101) modified with platinum nanoparticles (PtNPs) simultaneously detects the Xanthine, uric acid, and dopamine in spiked serum samples using the differential pulse voltammetry (DPV) [180]. The sensor has an extensive linear range (0.50 – 162.0 μM), a low detection limit (0.420 μM), and excellent selectivity, according to differential pulse voltammetry. The simultaneous presence of dopamine, uric acid, and hypoxanthine in the detection of Xanthine at working potentials of 0.130, 0.280, 0.680, and 1.050 V (vs. 3 M KCl saturated Ag/AgCl), estimating the concentration of Xanthine in spiked blood specimens. Fig. S3 shows the DPV curves of the simultaneous detection of dopamine (DA), uric acid (UA), Xanthine (XA), and hypoxanthine (HXA) [173]. Further, Table S1 comprises several MOF-based materials utilized in the electrochemical detection of various biomolecules in real and complex samples. Moreover, Table S1 includes the LOD and the linear range of detection of these biologically important molecules.

4.2. Detection of Hydrogen Peroxide

Hydrogen peroxide (H2O2) has physical and chemical significance in different research fields, viz., pharmaceutical industry, food and chemical industry, environment, and biological samples [153,181,182]. Additionally, it plays a significant role as an intermediate in biological and environmental interactions [183,184]. In the food industry, bleaching or disinfection utilizes hydrogen peroxide during food processing. However, higher concentrations of hydrogen peroxide residues in food products negatively affect human health. Excessive levels of retained hydrogen peroxide in food showed several harmful effects, which include cancer, rapid aging, coronary artery disease, severe gastrointestinal issues, and neurological illnesses [185]. Hydrogen peroxide (H2O2) is an active oxygen compound with antimicrobial characteristics and a significant molecular signaling molecule in many biological tissues [186]. The necessity of detecting hydrogen peroxide is manifold, and the miniaturized device for on-site detection is a prerequisite for various biological or environmental samples. The utilization of natural enzymes in electrochemical sensors is highly effective. The voltammetric response of these sensors demonstrates excellent charge transport through the MOF composite and quick electron exchange between the MOF and electrode. These MOF nanosheets have enzyme-like properties that catalyze them to oxidize hydrogen peroxide. This characteristic allows for developing an electrochemical sensor with fast and accurate quantitative detection. The sensor also demonstrates exceptional sensitivity, selectivity, stability, and durability at the desired pH level. Consequently, these sensors offer an alternative approach for detecting H2O2 [187]. The electrochemical methods are promising because of their robustness in operation, selectivity, dependability, and sensitivity towards the analyte, and the on-site detection with a miniaturized device added to the method’s suitability [188].
A novel electrochemical sensing platform detects hydrogen peroxide (H2O2) using the Pt/MoSe2 nanomesh, which offers an increased active site and a larger specific surface area. The Pt/MoSe2 nanomesh sensors show exceptional sensitivity and specificity towards H2O2. The correlation coefficient (R2) and the limit of detection (LOD) are 0.9983 and 2.56 μM, respectively, at a signal-to-noise (S/N) ratio of 3. In addition, the Pt/MoSe2 nanomesh exhibits excellent resistance to interference, even when exposed to potentially interfering tiny molecules found within cells, such as ascorbic acid (AA), glucose (Glc), and others [189]. A poly(amidoamine)-dendrimer/poly (PAMAM/PNR) modified electrode is utilized to concurrently detect the electrochemical oxidation of L-dopa and the reduction of H2O2. The sensor electrode modified with PAMAM/PNR shows a linear response between the L-dopa concentration (0.08 to 2400 μM) and the current with a lower limit of detection (LOD) of 0.026 μM [190]. Cobalt-based MOF (Co(pbda) (4,4-bpy)·2H2O]n) was incorporated on GCE and used for the detection of hydrogen peroxide (H2O2) at 0.1 M NaOH solution employing cyclic voltammetry (CV) [182]. Three reduction peaks occur on the cyclic voltammogram of the Co-MOF modified GC electrode: (a) in the absence of H2O2, (b) with the addition of 0.1 M NaOH solution containing 1 mM H2O2 and (c) with the addition of 0.1 M NaOH solution containing 2 mM H2O2 at approximately −0.40 V. The amperometric response indicates that the Co-MOF modified GC electrode shows enhanced electrocatalytic response towards the H2O2 reduction current, and the calibration line obtains for a wide concentration range of H2O2 between 5.0 μM and 9.0 mM. The method provides a low detection limit of 3.76 μM and a good sensitivity of 83.100 A μmM−1 cm2 at an applied potential of −0.40 V. Additionally, the detection method demonstrates stable performance over time with excellent selectivity. Furthermore, the mechanism indicated that the Co-MOF is highly efficient in intrinsic peroxidase-like activity, which allows it to catalyze H2O2 decomposition to form hydroxyl radical, which then oxidizes the peroxidase substrate (terephthalic acid) to produce color. Fig. S4 shows the reduction of H2O2 at an applied potential of −0.4V without the presence of H2O2, along with the addition of 0.1M NaOH solution containing 1 and 2 mM H2O2. It also depicts the Co-MOF-modified GCE’s amperometric response at a potential range of −0.4 to −0.6V [118].
Similarly, a solvothermal method synthesizes the Sn-based metal-organic framework (Sn-MOF@CNT) by incorporating carbon nanotubes (CNT) with Sn-MOF. A specifically designed approach efficiently detects hydrogen peroxide. The MOF possesses a large surface area and pore volume and maintains crystallinity even after the addition of CNT onto its surface [191]. The cyclic voltammogram of the bare electrode does not exhibit any oxidation or reduction peaks within the voltage range of −0.2 V to −0.8 V. Moreover, the gold electrode modified with the Sn-MOF@CNT shows no significant reduction in peak current in the absence of H2O2. The different concentrations of H2O2 (0.01, 0.03, and 0.05 mM) are present, resulting in a significant reduction in peak current (−0.6 V). The absence of H2O2 results in no electrochemical signals on the Sn-MOF@CNT/Au electrode. However, 0.2 μM H2O2 caused the electrode to display both cathodic and anodic peak currents. The observations prove that hydrogen peroxide undergoes a disproportionate reaction, i.e., the hydrogen peroxide decomposes into two products through oxidation and reduction half-reactions. The reaction results in the formation of water and oxygen at the electrode surface, as shown in equations given below:
(6)
Cathode: H2O2+2H++2e-2H2O
(7)
Anode: 2H2O22H2O+O2+2e-
(8)
Total: H2O2H2O+O2
As a result, the amperometric response of the Sn-MOF@CNT modified Au electrode shows a low detection limit of about 4.7 ×10−3 μM of hydrogen peroxide. Table S2 comprises the use of different MOFs in the electrochemical detection of hydrogen peroxide in various samples. Table S2 includes the limit of detection and experimental parameters for various systems.

4.3. Organic Pollutant Sensing

Wastewater containing different industrial effluents impacts ecological stability because micro-pollutants can harm living organisms even at low concentrations [192]. Aromatic organic compounds such as phenol, nitrobenzene, hydrazine, chlorinated phenols, polyaromatic hydrocarbons, pesticides, nitrites, and pharmaceuticals have shown more significant environmental concerns, contaminated water bodies, and severely affecting human health. The detection and elimination of these pollutants in the aquatic environment have been a source of desperation for environmentalists for the past few years [193]. Drug and pharmaceutical residue toxicity refer to the harm inflicted on an organism by these chemical compounds. Drug toxicity can have a detrimental impact on organs such as the liver, kidney, heart, and other parts of the body. Drug toxicity is directly proportional to the dose and manifests a range of undesirable consequences when it surpasses the therapeutic level. Excessive consumption and over-administration of drugs result in acute toxicity within an individual’s body.
Similarly, the overutilization of drugs can lead to the occurrence of drug toxicity in both food and the environment since these substances migrate from the circulatory system to various organs [194,195]. Metol (N-methyl-p-aminophenol sulphate) is a prevalent phenol compound used as a chemical raw material in the production of photographic development and is a potential environmental pollutant. A cerium-based metal-organic framework shows an excellent sensing material for detecting metol by voltammetry. The detection method achieves a linear range from 0.1 to 200 μM, with a detection limit of 30 nM [196198]. Chlorophenols, an organic pollutant, are widely present in several industries and extensively utilized as pesticides, fungicides, herbicides, dyes, and solvents. Chlorophenols are extremely hazardous organic contaminants that pose serious environmental concerns. Chlorophenols are toxic to people and can accumulate in the food chain, impacting human health. The modified GCEs are excellent in detecting chlorophenol with a low limit of detection (LOD = 0.328 μM) at a wide linear detection range from 10 to 160 μM. The electrode shows strong stability and selectivity towards these pollutants. Furthermore, the results of the recovery studies show the electrochemical sensor’s suitability for detecting actual samples. The findings indicate that the HP-UiO66@Ni-BDC MOF shows potential in the low-level detection of chlorophenol [199201].
Yadav et al. demonstrate the synthesis of zinc (II) based MOF (MOF-5) decorated with the Au(NPs), and the material modifies the glassy carbon electrode and detects the nitrobenzene and nitrite in an aqueous medium [202]. Fig. S5(a) shows cyclic voltammograms of nitrite (1.0 mM) using the GC/Au-MOF-5, GC/MOF-5, and GC electrodes in 0.1 M phosphate buffer (PBS) (pH 7.0) at 20 mV/s. Without nitrite, no peaks occur at GC and GC/MOF-5 (Fig. S5(a): a & b), but well-defined peaks occur at GC/Au-MOF-5 due to the formation of surface Au oxide and their reduction (Fig. S5 (a): c′). The surface of Au oxidized at 0.85 V and reduced at 0.46 V during the reverse scan. On the other hand, the nitrite shows oxidation at 0.93 V and 1.01 V using the GC and GC/MOF-5 electrodes, respectively (Fig. S5(a): a′& b′). Similar oxidation occurs at 0.76 V at the GC/Au-MOF-5 (Fig. S5(a): c′) with enhanced anodic peak current. The higher nitrite oxidation potential at GC/MOF-5 is due to the interaction of free carboxylate groups of MOF-5 and nitrite. The results imply a significant increase in peak currents at low applied potentials using the GC/Au-MOF-5 electrode. Therefore, Au(NPs) in Au-MOF-5 showed an efficient electrocatalytic activity for nitrite oxidation. Fig. S5(b) shows the cyclic voltammograms of nitrobenzene (NB) using the GC, GC/MOF-5, and GC/Au-MOF-5 electrodes in an N2 saturated atmosphere employing 0.1 M pH 7.0 phosphate buffer. In the absence of NB, no peak current appeared in the voltammograms, and a distinct reduction of nitrobenzene is observed at an applied potential of −0.77, 0.79, and 0.77 V for the GC, GC/MOF-5 and GC/Au-MOF-5 electrodes, respectively [202].
Many countries have limited the use of patulin residues due to their severe toxicity and ubiquitous occurrence in foods. The Food and Drug Administration of the United States and China mandates a maximum amount of patulin at 50 μM in fruit juice and processing products, and the European Union mandates a limit of 10 μM in newborn and child meals [203]. Because of the widespread occurrence and high toxicity of patulin in food, a rapid and sensitive detection method averts the possible harmful effects of patulin on human health [132]. Copper-based MOF decorated with the Au(NPs) was used to modify the GCE, which electrochemically detects patulin in apple juice. The square wave voltammograms of patulin show that the cathodic signal increases with the patulin concentration rising from 0.001 nM to 1.0 nM and 1.0 nM to 250.0 nM. The LOQ was 0.001 nM, and the LOD was 3.33 × 10−4 nM. RSDs of 1.65% and 0.83% were found for six repeat observations of 1.0 nM and 250.0 nM patulin, respectively [204].
Similarly, the HKUST-1 MOF incorporated with Au(NPs)-GCE electrochemically detects the paracetamol. The MOF/AuNP modified GCE detects the paracetamol efficiently, and a wide linear range of paracetamol concentration (0.01 μM to 100 μM) provides the detection limit of 0.0011 μM [205]. Cu-BTC nanocrystals/CNTs modified GC electrode showed an improved oxidation current for the metformin detection. Peak current demonstrated good linearity with concentrations ranging from 0.5 μM to 25 μM with a detection limit of 0.12 μM under optimal conditions. Further, the detection method detects metformin in pharmaceutical samples [206]. The EPC-modified GCE showed good sensitivity and low detection limit (i.e., 2.9 μM) in detecting chloramphenicol residual in honey. Using the electrochemical detection method, an aluminum-based metal-organic framework (DUT-4) detects chloramphenicol (CAP), a commonly used broad-spectrum antibiotic. The presence of reduced graphene oxide (rGO) with the DUT-4 significantly enhances the electrocatalytic activity of the electrode in detecting the CAP. The synthesized sensor demonstrates exceptional sensitivity, covering a wide linear range of 1–1000 μM of CAP, and achieves a remarkably low detection limit of 76.97 nM [207]. Table S3 lists different types of MOFs used to detect organic contaminants.

4.4. Heavy Metals Sensing

Heavy metals that pose a significant hazard to human health and the marine environment include lead, mercury, cadmium, chromium, and arsenic. [208]. These heavy metals contaminate the environment due to many geological and human-caused phenomena, including agricultural, industrial, and domestic wastes [209]. Heavy metal ions (HMIs), a primary form of food pollutants, pose a significant risk to human health even at lower levels due to their persistence, elevated toxicity, and prolonged build-up in the body [210]. Heavy metals are metallic elements that have a density of more than 5, primarily including zinc (Zn), chromium (Cr), mercury (Hg), copper (Cu), lead (Pb), and arsenic (As). Despite not being a metal element, As exhibits a toxicity level comparable to that of heavy metals, and its relative density satisfies the criteria for heavy metals [211]. Hazardous metallic ions (HMIs) within the human body damage the skin, bones, liver, kidneys, and perhaps the brain system [212]. As an illustration, Pb is one of the most detrimental HMIs and has the potential to adversely affect the brain, kidney, neurological system, and reproductive system [213]. Excessive amounts of Zn lead to gastrointestinal discomfort and aberrant pancreatic enzyme levels. Elevated concentrations of copper cause the development of cancer and genetic abnormalities [214]. Recently, emphasis has been drawn to detecting heavy metals utilizing various MOF-based sensors in various environmental matrices [144]. Electrochemical analysis is one of the prominent and robust techniques due to its accuracy and reliability. The exceptional features of MOF, i.e., high porosity, homogeneous structures, large surface area, and ease of functionalization, received greater attention in the efficient and selective use in sensor developments for detecting several heavy metals [215]. The bulk organic linkers coordinate with the core metal atom, restricting MOFs’ electrical conductivity [216]. Therefore, MOFs are typically coupled with high-conductivity elements such as metal oxides, metal nanoparticles, and carbon compounds to overcome the limitations. Furthermore, MOFs with high conductivity are alternative materials for the trace detection of several heavy metal toxic ions. However, the repeatability of the signal due to the micro-sized MOFs is one of the challenges in the detection system [24, 217, 218].
A hydrothermal technique obtains the zinc-based metal-organic framework (Zn-MOF), and the material shows exceptional stability in water at various pH. The molecule exhibits high selectivity, sensitivity, and a low detection limit, 4.10 μM for Cr2O72− and 6.63 μM for CrO42− in aqueous solution [219]. Similarly, a hydrothermal method synthesizes the Eu-MOF, {[Eu2(L)(phen)2(ox)2(H2O)2]·10H2O·phen}n, which exhibits stability in water at varied pH values. The Eu-MOF has good sensitivity and selectivity in detecting gas/liquid benzaldehyde, Hg2+, and Cr2O72−/CrO42− by luminescence quenching effects, making it a potential luminous sensor. Eu-MOF exhibits exceptional adsorption capabilities for Hg2+ and can effectively be employed for the quantitative detection of Hg2+ in tap water, green tea, and river water. The sensor demonstrates a recovery rate ranging from 99.88% to 102.66% for detecting Hg2+ [220]. A novel electrochemical aptasensor utilizing aptamer detects the Cd2+ and Pb2+ concurrently in fruits and vegetables. The double-stranded DNA, including aptamers, obtains the electrode by forming an Au-S link. The electrochemical labeling of Cd2+ and Pb2+ aptamers with Methylene Blue and Ferrocene groups resulted in competition between the metal ions and the complementary sequences (CP), leading to variations in the electrochemical signal. Consequently, both Cd2+ and Pb2+ can be simultaneously detected. The electrochemical aptasensor demonstrates a linear response for Cd2+ and Pb2+ detection within the concentration range of 0.1 to 1000 nmol/L under ideal conditions. The detection limits for Cd2+ and Pb2+ were 89.31 and 16.44 nmol/L, respectively [216]. Cd2+ and Pb2+ were simultaneously and selectively detected in water samples using the Yb-MOF-modified electrode [221]. Under ideal circumstances, the Yb-MOF/GCE electrode detects the Cd2+ and Pb2+ independently at values ranging from 0 to 40 ppb. I (μA) = 0.2045 * C (ppb) − 0.0838 (Cd2+) and I (μA) = 0.1971 * C (ppb) + 0.0686 (Pb2+) are the recorded current responses on DPASV curves that grow linearly with increasing ion concentration. For Cd2+ and Pb2+, the detection limits (LOD) were 7.40 ppb and 2.02 ppb, respectively. The Yb-MOF/GCE detects the Cd2+ and Pb2+ simultaneously in the 0 to 50 ppb concentration range. The peak current increased linearly with the ion concentrations, as shown in Fig. S6. The sensitivities were 469 μA ppm−1 cm−2 and 397 μA ppm−1 cm−2 for Cd2+ and Pb2+, respectively. The detection limits for Cd2+ and Pb2+ were 3.0 ppb and 1.6 ppb, respectively. Pb2+ was more favorably absorbed and reduced on the porous structure of the Yb-MOF framework than Cd2+ [222]. A susceptible electrochemical sensor detects the Pb2+ ions using the Bi-MOFs. The solvothermal method synthesizes the MOF (Bi/Bi2O3) followed by the co-doping of porous carbon composite (Bi/Bi2O3@C) through one-step carbonization. The material demonstrates the ability to retain Bi-MOFs’ structure, which is essential for its functionality. Additionally, it exhibits a relatively large specific surface area, indicating its potential for increased adsorption and reactivity. Furthermore, the material shows high electron-transfer capabilities, suggesting its suitability for various electrochemical applications. The square wave anodic stripping voltammetry (SWASV) using the MOF-modified electrode efficiently detects the Pb2+. It had a wide linear range, ranging from 37.5 nM to 2.0 μM, indicating its ability to measure a wide range of concentrations accurately. Additionally, the sensor had a low detection limit of 6.3 nM. The sensor exhibits favorable stability, reproducibility, and satisfactory selectivity [223]. Table S4 shows the latest developments in the pH and potential optimization for the electrochemical detection of heavy metals utilizing MOFs.
In aqueous solutions, the Fe@YAU-101/GCE quantifies various metal ions, viz., Cd2+, Pb2+, and Hg2+. The oxidation peak current signals of the Fe@YAU-101/GCE exhibit an upward trend as the concentration of Cd2+, Pb2+, and Hg2+ increases, allowing for the identification of these ions. The Cd2+, Pb2+, and Hg2+ in solution are carried within the concentration range of 0.003 to 42 μM, 0.004 to 80 μM, and 0.045 to 66 μM, respectively. The presence of Cd2+, Pb2+, and Hg2+ in a solution does not affect the DPV response signal using the Fe@YAU-101/GCE, hence enabling the detection of these ions simultaneously. The sensor exhibits a notably greater current signal for Hg2+ and Pb2+ than Cd2+, suggesting better sensitivity for Hg2+ and Pb2+ than Cd2+. The regression equations for Cd2+, Pb2+ and Hg2+ are: I =0.0141C +0.0349 (R2 = 0.9954), y = 0.2861C-0.0554 (R2 = 0.9975), and y = 0.2701C + 0.3320 (R2 = 0.9955), respectively. The limit of detection (LOD) values for Cd2+, Pb2+, and Hg2+ were 1 ×10−9 M, 1.5 × 10−8 M, and 1.33 ×10−9 M, respectively [224].

5. Conclusion and Future Perspective

MOFs and MOF-derived materials are suitable modifiers of electrodes for detecting different pollutants in aquatic environments. MOFs have unique qualities such as a high specific surface area, the ability to deal with pores, and the ability to regulate pore shape and size. Moreover, using suitable organic linkers in the material synthesis makes the material a flexible framework for the target adsorbate ions/molecules. As a result, there is potential for advancement in creating MOFs-based sensing devices by enhancing the material’s sensitivity and stability. More research into the development of MOFs and MOF-derived materials for practical uses may open up new paths for sensing in environmental applications.
The significant limitations of MOF-based sensors include inadequate stability and electrical conductivity of MOF materials. The instability of MOFs is primarily due to the susceptibility of metal-ligand coordination bonds, seemingly breaking in the presence of water molecules. If the synthesized MOF lacks structural stability in the aqueous solution employed for electroanalysis, the thin film of MOF applied onto the electrode’s surface may undergo partial dissolution or transformation into MOF-derived oxide or hydroxide, along with residual organic linkers. Moreover, certain MOFs exhibit electrical insulating properties. However, the effective method to enhance electric charge transport in traditional MOFs is to obtain the framework using materials with inherent electrical conductivity. Though several MOF-based sensors detect two or three water pollutants simultaneously, extensive studies are required to ensure the ability of MOF-based sensors to detect the contaminants in water simultaneously. Several researchers reported ground-breaking achievements in the applications of MOFs for electrochemical sensors in the recent past; however, advancements towards the MOFs developments for targeted sensors applications. Compared to carbon-based materials, MOF-derived composites face lesser sensitivities, poor stability, and less repeatability. Therefore, synthesizing advanced MOF materials could enable viable alternatives in miniaturized device development. Similarly, there are several challenges in the development of sensors using MOF and MOF-based materials; the following are the critical issues of employing MOF-based materials in electrochemical sensors:
  1. One of the challenges is the synthesis of sophisticated materials to control the shape and size of MOFs; this might result in uniform growth of nanostructures with a significant increase in surface area.

  2. It is still challenging to distribute active metals evenly on the surface of carbons made from MOFs.

  3. Although many MOFs employed fabricating the electrodes for heavy metal detection, the insights into electrochemical sensing mechanisms could enable greater implications for device development.

  4. MOF stability in an aqueous media remains challenging; studies on coupling hydrophobic ligands with high valence metal ions could provide newer research areas for suitable applications.

  5. Since MOFs’ pore width and geometry play an essential role in the highly selective determination of food contaminants, synthesizing functionalized MOFs could enable the required selectivity towards the target analyte species in complex matrices.

Supplementary Information

Notes

Conflict-of-Interest Statement

The authors declare that they have no conflict of interest.

Author Contributions

R.L. (M.Sc.) collected literature and rough draft of manuscript prepared. M.L. (M.Sc.) collected literature and completed the rough drafting of the manuscript. L. (Assistant Professor) reviewed and modified the manuscript. D.T. (Professor) conceptualized the problem, critically evaluated the manuscript and finalized the draft manuscript.

Reference

1. Wiśniewska P, Haponiuk J, Saeb MR, Rabiee N, Bencherif SA. Mitigating metal-organic framework (MOF) toxicity for biomedical applications. Chem. Eng. J. 2023;471:144400. 10.1016/j.cej.2023.144400.
crossref

2. Ettlinger R, Lächelt U, Gref R, et al. Toxicity of metal–organic framework nanoparticles: from essential analyses to potential applications. Chem. Soc. Rev. 2022;51:464–484. 10.1039/D1CS00918D.
crossref pmid

3. Fan W, Yuan S, Wang W, et al. Optimizing Multivariate Metal–Organic Frameworks for Efficient C2H2/CO2 Separation. J. Am. Chem. Soc. 2020;142:8728–8737. 10.1021/jacs.0c00805
crossref pmid

4. Herbst A, Janiak C. MOF catalysts in biomass upgrading towards value-added fine chemicals. Cryst. Eng. Comm. 2017;19:4092–4117. 10.1039/c6ce01782g
crossref

5. Pornea AM, Kim H. Synthesis of hybrid dual-MOF encapsulated phase-changing material for improved broadband light absorption and photothermal conversion enabling efficient solar energy storage. Sol. Energy Mater Sol. Cells. 2022;244:111817. 10.1016/j.solmat.2022.111817
crossref

6. Dong Z, Sun Y, Chu J, Zhang X, Deng H. Multivariate Metal–Organic Frameworks for Dialing-in the Binding and Programming the Release of Drug Molecules. J. Am. Chem. Soc. 2017;139:14209–14216. 10.1021/jacs.7b07392
crossref pmid

7. Tchalala MR, Bhatt PM, Chappanda KN, et al. Fluorinated MOF platform for selective removal and sensing of SO2 from fluegasandair. Nat Commun. 2019;10:10.1038/s41467-019-09157-2.
crossref pmid pdf

8. Kannangara YY, Rathnayake UA, Song J-K. Redox active multi-layered Zn-pPDA MOFs as high-performance supercapacitor electrode material. Electrochim. Acta. 2019;297:145–154. 10.1016/j.electacta.2018.11.186.
crossref

9. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature. 1999;402:276–279. 10.1038/46248
crossref pdf

10. Sohrabi H, Ghasemzadeh S, Ghoreishi Z, et al. Metal-organic frameworks (MOF)-based sensors for detection of toxic gases: A review of current status and future prospects. Mater. Chem. Phys. 2023;299:127512. 10.1016/j.matchemphys.2023.127512
crossref

11. Ren J, Langmi HW, North BC, Mathe M. Review on processing of metal–organic framework (MOF) materials towards system integration for hydrogen storage. Int. J. Eng. Res. 2015;39:607–620. 10.1002/er.3255.
crossref

12. Stock N, Biswas S. Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2012;112:933–969. 10.1021/cr200304e
crossref pmid

13. Carrington ME, Rampal N, Madden DG, et al. Sol-gel processing of a covalent organic framework for the generation of hierarchically porous monolithic adsorbents. Chem. 2022;8:2961–2977. 10.1016/j.chempr.2022.07.013.
crossref

14. Jaluddin SN, Kassim K, Ibrahim WNW. Mixed Linker Metal Organic Framework as an Efficient and Reusable Catalyst for Suzuki Miyaura Coupling Reaction. Procedia Chem. 2015;16:8–14. 10.1016/j.proche.2015.12.008.
crossref

15. Yaghi OM, Li H, Eddaoudi M, O’Keeffe M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature. 1999;402:276–279. 10.1038/46248.
crossref pdf

16. Ha J, Lee JH, Moon HR. Alterations to secondary building units of metal–organic frameworks for the development of new functions. Inorg. Chem. Front. 2019;7:12–27. 10.1039/C9QI01119F.
crossref

17. Yuan S, Qin JS, Lollar CT, Zhou H. Stable Metal–Organic Frameworks with Group 4 Metals: Current Status and Trends. ACS Cent. Sci. 2018;4:440–450. 10.1021/acscentsci.8b00073.
crossref pmid pmc pdf

18. Fan WK, Sherryna A, Tahir M. Advances in Titanium Carbide (Ti3C2Tx) MXenes and Their Metal–Organic Framework (MOF)-Based Nanotextures for Solar Energy Applications: A Review. ACS Omega. 2022;7:38158–38192. 10.1021/acsomega.2c05030.
crossref pmid pmc pdf

19. Chui S, Lo SM, Charmant JP, Orpen AG, Williams ID. A Chemically Functionalizable Nanoporous. Mater. Sci. 1999;283:1148–1150.


20. Devic T, Serre C. High valence 3p and transition metal based MOFs. Chem. Soc. Rev. 2014;43:6097–6115. 10.1039/C4CS00081A.
crossref pmid

21. Loiseau T, Volkringer C, Haouas M, Taulelle F, Férey G. Crystal chemistry of aluminium carboxylates: From molecular species towards porous infinite three-dimensional networks. C R Chim. 2015;18:1350–1369. 10.1016/j.crci.2015.08.006.
crossref

22. Lee SM, Zirlianngura , Anjudikkal J, Tiwari D. Electrochemical sensor for trace determination of cadmium(II) from aqueous solutions: use of hybrid materials precursors to natural clays. Int. J. Environ. Anal. Chem. 2016;96:490–504. 10.1080/03067319.2016.1172220.
crossref

23. Sarikokba , Lalmalsawmi J, Kumar Prasad S, Tiwari D. Development of a novel sensor with high sensitivity for electroanalytical determination of bisphenol A based on chitosan-3-mercaptopropyl trimethoxysilane modified glassy carbon electrode. Microchem. J. 2022;181:107748. 10.1016/j.microc.2022.107748.
crossref

24. Cao Y, Wang L, Shen C, Wang C, Hu X, Wang G. An electrochemical sensor on the hierarchically porous Cu-BTC MOF platform for glyphosate determination. Sens. Actuators B Chem. 2019;283:487–494. 10.1016/j.snb.2018.12.064.
crossref

25. Zirlianngura , Tiwari D, Ha J-H, Lee SM. Efficient Use of Porous Hybrid Materials in a Selective Detection of Lead(II) from Aqueous Solutions: An Electrochemical Study. Metals. 2017;7:124. 10.3390/met7040124.
crossref

26. Sarikokba , Lalhmunsiama , Lalmalsawmi J, Tiwari D, Lee SM, Yang JK. Greener synthesis of novel nanocomposite material for efficient and selective trace detection of Pb2+:Insights of stability and real implications of fabricated electrode. Environ Eng Res. 2024;29. 10.4491/eer.2023.023.


27. Tiwari D, Jamsheera A, Zirlianngura , Lee SM. Use of hybrid materials in the trace determination of As(V) from aqueous solutions: An electrochemical study. Environ. Eng. Res. 2017;22:186–192. 10.4491/eer.2016.045.
crossref

28. Rana A, Baig N, Saleh TA. Electrochemically pretreated carbon electrodes and their electroanalytical applications – A review. J. Electroanal. Chem. 2019;833:313–332. 10.1016/j.jelechem.2018.12.019.
crossref

29. Jung K-W, Choi BH, Lee SY, Ahn KH, Lee YJ. Green synthesis of aluminum-based metal organic framework for the removal of azo dye Acid Black 1 from aqueous media. J. Ind. Eng. Chem. 2018;67:316–325. 10.1016/j.jiec.2018.07.003.
crossref

30. Kajal N, Singh V, Gupta R, Gautam S. Metal organic frameworks for electrochemical sensor applications: A review. Environ. Res. 2022;204:112320. 10.1016/j.envres.2021.112320.
crossref pmid

31. Hu C, Xiao JD, Mao XD, Song LL, Yang XY, Liu SJ. Toughening mechanisms of epoxy resin using aminated metal-organic framework as additive. Mater. Lett. 2019;240:113–116. 10.1016/j.matlet.2018.12.123.
crossref

32. Hu Z, Wang Y, Zhao D. Modulated Hydrothermal Chemistry of Metal–Organic Frameworks. Acc. Mater. 2022;3:1106–1114. 10.1021/accountsmr.2c00104
crossref pdf

33. Abdelkareem MA, Abbas Q, Mouselly M, Alawadhi H, Olabi AG. High-performance effective metal–organic frameworks for electrochemical applications. J. Sci-Adv. Mater. Dev. 2022;7:100465. 10.1016/j.jsamd.2022.100465.
crossref

34. Liu B, Vellingiri K, Jo S-H, Kumar P, Ok YS, Kim KH. Recent advances in controlled modification of the size and morphology of metal-organic frameworks. Nano. Res. 2018;11:4441–4467. 10.1007/s12274-018-2039-3.
crossref pdf

35. Yaghi OM, Li H. Hydrothermal Synthesis of a Metal-Organic Framework Containing Large Rectangular Channels. J. Am. Chem. Soc. 1995;117:10401–10402. 10.1021/ja00146a033.
crossref

36. Qi Y, Luo F, Che Y, Zheng J. Hydrothermal synthesis of metal-organic frameworks based on aromatic polycarboxylate and flexible bis(imidazole) ligands. Cryst. Growth Des. 2008;8:606–611. 10.1021/cg700758c.
crossref

37. Horcajada P, Surblé S, Serre C, et al. Synthesis and catalytic properties of MIL-100(Fe), an iron(III) carboxylate with large pores. Chem Commun. 2007;2820–2822. 10.1039/b704325b.


38. Chen L, Guo H, Fujita T, et al. Nanoporous PdNi bimetallic catalyst with enhanced electrocatalytic performances for electro-oxidation and oxygen reduction reactions. Adv. Funct. Mater. 2011;21:4364–4370. 10.1002/adfm.201101227.
crossref

39. Wu H, Li M, Wang Z, et al. Highly stable Ni-MOF comprising triphenylamine moieties as a high-performance redox indicator for sensitive aptasensor construction. Anal. Chim. Acta. 2019;1049:74–81. 10.1016/j.aca.2018.10.022.
crossref pmid

40. Liu X, Yang H, Diao Y, et al. Recent advances in the electrochemical applications of Ni-based metal organic frameworks (Ni-MOFs) and their derivatives. Chemosphere. 2022;307:135729. 10.1016/j.chemosphere.2022.135729.
crossref pmid

41. Ahmadi A, Nezamzadeh-Ejhieh A. A comprehensive study on electrocatalytic current of urea oxidation by modified carbon paste electrode with Ni(II)-clinoptilolite nanoparticles: Experimental design by response surface methodology. J. Electroanal. Chem. 2017;801:328–337. 10.1016/j.jelechem.2017.08.009.
crossref

42. Pourtaheri A, Nezamzadeh-Ejhieh A. Enhancement in photocatalytic activity of NiO by supporting onto an Iranian clinoptilolite nano-particles of aqueous solution of cefuroxime pharmaceutical capsule. Spectrochim. Acta. 2015;137:338–344. 10.1016/j.saa.2014.08.058.
crossref pmid

43. Sheikh-Mohseni MH, Nezamzadeh-Ejhieh A. Modification of carbon paste electrode with Ni-clinoptilolite nanoparticles for electrocatalytic oxidation of methanol. Electrochim. Acta. 2014;147:572–581. 10.1016/j.electacta.2014.09.123.
crossref

44. Lin X, Zeng W, Liu M, et al. Amino-modified Mg-MOF-74: Synthesis, characterization and CO2 adsorptionperformance. Environ Eng Res. 2023;28. 10.4491/eer.2021.569.
crossref

45. Deng D, Wang Y, Wen S, et al. Metal-organic framework composite Mn/Fe-MOF@Pd with peroxidase-like activities for sensitive colorimetric detection of hydroquinone. Anal. Chim. Acta. 2023;1279:341797. 10.1016/j.aca.2023.341797.
crossref pmid

46. Liu Y, Myers EJ, Rydahl SA, Wang X. Ultrasonic-Assisted Synthesis, Characterization, and Application of a Metal–Organic Framework: A Green General Chemistry Laboratory Project. J. Chem. Educ. 2019;96:2286–2291. 10.1021/acs.jchemed.9b00267.
crossref

47. Vinoth V, Wu JJ, Asiri AM, Anandan S. Sonochemical synthesis of silver nanoparticles anchored reduced graphene oxide nanosheets for selective and sensitive detection of glutathione. Ultrason. Sonochem. 2017;39:363–373. 10.1016/j.ultsonch.2017.04.035.
crossref pmid

48. Areerob Y, Cho JY, Jang WK, Oh WC. Enhanced sonocatalytic degradation of organic dyes from aqueous solutions by novel synthesis of mesoporous Fe3O4-graphene/ZnO @SiO2 nanocomposites. Ultrason. Sonochem. 2018;41:267–278. 10.1016/j.ultsonch.2017.09.034.
crossref pmid

49. Qiu LG, Li ZQ, Wu Y, Wang W, Xu T, Jiang X. Facile synthesis of nanocrystals of a microporous metal–organic framework by an ultrasonic method and selective sensing of organoamines. Chem Commun. 2008;3642–3644. 10.1039/B804126A.
crossref pmid

50. Khan NA, Jhung SH. Facile syntheses of metal-organic framework under ultrasound. Bull. Korean Chem. Soc. 2009;30:2921–2926. 10.5012/bkcs.2009.30.12.2921.


51. He S, Wu L, Li X, et al. Metal-organic frameworks for advanced drug delivery. Acta Pharm. Sin. B. 2021;11:2362–2395. 10.1016/j.apsb.2021.03.019.
crossref pmid pmc

52. Esrafili L, Azhdari TA, Morsali A. Ultrasonic assisted synthesis of two urea functionalized metal organic frameworks for phenol sensing: A comparative study. Ultrason. Sonochem. 2017;39:307–312. 10.1016/j.ultsonch.2017.04.039.
crossref pmid

53. Yu K, Lee Y, Seo JY, Baek KY, Chung YM, Ahn WS. Sonochemical synthesis of Zr-based porphyrinic MOF-525 and MOF-545: Enhancement in catalytic and adsorption properties. Microporous Mesoporous Mater. 2021;316:110985. 10.1016/j.micromeso.2021.110985.
crossref

54. Vaitsis C, Kanellou E, Pandis PK, et al. Sonochemical synthesis of zinc adipate Metal-Organic Framework (MOF) for the electrochemical reduction of CO2: MOF and circular economy potential. Sustain. Chem. Pharm. 2022;29:100786. 10.1016/j.scp.2022.100786.
crossref

55. Taghipour A, Rahimpour A, Rastgar M, Sadrzadeh M. Ultrasonically synthesized MOFs for modification of polymeric membranes: A critical review. Ultrason. Sonochem. 2022;90:106202. 10.1016/j.ultsonch.2022.106202.
crossref pmid pmc

56. Joseph J, Iftekhar S, Srivastava V, Fallah Z, Zare E, Sillanpää M. Iron-based metal-organic framework: Synthesis, structure and current technologies for water reclamation with deep insight into framework integrity. Chemosphere. 2021;284:131171. 10.1016/j.chemosphere.2021.131171.
crossref pmid

57. Ge J, Wu Z, Huang X, Ding M. An Effective Microwave-Assisted Synthesis of MOF235 with Excellent Adsorption of Acid Chrome Blue K. J. Nanomater. 2019;2019:e4035075. 10.1155/2019/4035075.
crossref pdf

58. Ahn WS, Kang KK, Kim KY. Synthesis of TS-1 by microwave heating of template-impregnated SiO2–TiO2xerogels. Catal. Letters. 2001;72:229–232. 10.1023/A:1009084011436.


59. Jin T, Hwang YK, Hong D-Y, et al. Microwave synthesis, characterization and catalytic properties of titanium-incorporated ZSM-5 zeolite. Res. Chem. Intermed. 2007;33:501–512. 10.1163/156856707782565804.
crossref

60. Komarneni S, Rajha RK, Katsuki H. Microwave-hydrothermal processing of titanium dioxide. Mater. Chem. Phys. 1999;61:50–54. 10.1016/S0254-0584(99)00113-3.
crossref

61. Tompsett GA, Conner WC, Yngvesson KS. Microwave synthesis of nanoporous materials. Chemphyschem. 2006;7:296–319. 10.1002/cphc.200500449.
crossref pmid pdf

62. Haque E, Khan NA, Kim CM, Jhung S. Syntheses of Metal–Organic Frameworks and Aluminophosphates under Microwave Heating: Quantitative Analysis of Accelerations. Cryst. Growth Des. 2011;11:4413–4421. 10.1021/cg200594e.
crossref

63. Ni Z, Masel RI. Rapid Production of Metal–Organic Frameworks via Microwave-Assisted Solvothermal Synthesis. J. Am. Chem. Soc. 2006;128:12394–12395. 10.1021/ja0635231.
crossref pmid

64. Parnham ER, Morris RE. Ionothermal Synthesis of Zeolites, Metal–Organic Frameworks, and Inorganic–Organic Hybrids. Acc. Chem. Res. 2007;40:1005–1013. 10.1021/ar700025k.
crossref pmid

65. Choi J-S, Son W-J, Kim J, Ahn WS. Metal–organic framework MOF-5 prepared by microwave heating: Factors to be considered. Microporous Mesoporous Mater. 2008;116:727–731. 10.1016/j.micromeso.2008.04.033.
crossref

66. Liu H, Zhao Y, Zhou C, Mu B, Chen L. Microwave-assisted synthesis of Zr-based metal–organic framework (Zr-fumfcu-MOF) for gas adsorption separation. Chem. Phys. Lett. 2021;780:138906. 10.1016/j.cplett.2021.138906.
crossref

67. Ameloot R, Stappers L, Fransaer J, Alaerts L, Sels BF, De Vos D. Patterned Growth of Metal-Organic Framework Coatings by Electrochemical Synthesis. Chem. Mater. 2009;21:2580–2582. 10.1021/cm900069f.
crossref

68. Bétard A, Fischer RA. Metal–Organic Framework Thin Films: From Fundamentals to Applications. Chem. Rev. 2012;112:1055–1083. 10.1021/cr200167v.
crossref pmid

69. Al-Kutubi H, Dikhtiarenko A, Zafarani HR, Sudhölter E, Gascon J, Rassaei L. Facile formation of ZIF-8 thin films on ZnO nanorods. CrystEngComm. 2015;17:5360–5364. 10.1039/c5ce00590f.
crossref

70. Asghar A, Iqbal N, Noor T, Kariuki B, Kidwell L, Easun T. Efficient electrochemical synthesis of a manganese-based metal–organic framework for H2 and CO2 uptake. Green Chem. 2021;23:1220–1227. 10.1039/D0GC03292A.
crossref

71. Guerrero VV, Yoo Y, McCarthy MC, Jeong H. HKUST-1 membranes on porous supports using secondary growth. J. Mater. Chem. 2010;20:3938–3943. 10.1039/b924536g.
crossref

72. Neto OJ, de L, Frós AC, de O, Barros BS, Monteiro A, Kulesza J. Rapid and efficient electrochemical synthesis of a zinc-based nano-MOF for Ibuprofen adsorption. New J Chem. 2019;43:5518–5524. 10.1039/C8NJ06420B.
crossref

73. Li W-J, Lü J, Gao S-Y, Li Q, Cao R. Electrochemical preparation of metal–organic framework films for fast detection of nitro explosives. J. Mater. Chem A. 2014;2:19473–19478. 10.1039/C4TA04203D.
crossref

74. Nadizadeh Z, Naimi-Jamal MR, Panahi L. Mechanochemical solvent-free in situ synthesis of drug-loaded MOFs for controlled drug delivery. J. Solid State Chem. 2018;259:35–42. 10.1016/j.jssc.2017.12.027.


75. Tanaka S. Mechanochemical synthesis of MOFs. Metal-Organic Frameworks for Biomedical Applications. Elsevier inc; 2020. p. 197–222.
crossref

76. Friščić T. Supramolecular concepts and new techniques in mechanochemistry: Cocrystals, cages, rotaxanes, open metal–organic frameworks. Chem. Soc. Rev. 2012;41:3493–3510. 10.1039/c2cs15332g.
crossref pmid

77. Lv D, Chen Y, Li Y, et al. Efficient Mechanochemical Synthesis of MOF-5 for Linear Alkanes Adsorption. J. Chem. Eng. Data. 2017;62:2030–2036. 10.1021/acs.jced.7b00049.
crossref

78. Kaur G, Anthwal A, Kandwal P, Sud D. Mechanochemical synthesis and theoretical investigations of Fe (II) based MOF containing 4,4′-bipyridine with ordained intercalated p-aminobenzoic acid: Application as fluoroprobe for detection of carbonyl group. Inorganica Chim. Acta. 2023;545:121248. 10.1016/j.ica.2022.121248.
crossref

79. Prochowicz D, Sokołowski K, Justyniak I, et al. A mechanochemical strategy for IRMOF assembly based on pre-designed oxo-zinc precursors. Chem. Commun. 2015;51:4032–4035. 10.1039/C4CC09917F.
crossref pmid

80. Kujawa J, Al-Gharabli S, Muzioł TM, et al. Crystalline porous frameworks as nano-enhancers for membrane liquid separation – Recent developments. Coord. Chem. Rev. 2021;440:213969. 10.1016/j.ccr.2021.213969.
crossref

81. Du M, Li C-P, Zhao X-J. Metal-Controlled Assembly of Coordination Polymers with the Flexible Building Block 4-Pyridylacetic Acid (Hpya). Cryst Growth Des. 2006;6:335–341. 10.1021/cg0502542.
crossref

82. Halper SR, Do L, Stork JR, Cohen S. Topological Control in Heterometallic Metal–Organic Frameworks by Anion Templating and Metalloligand Design. J. Am. Chem. Soc. 2006;128:15255–15268. 10.1021/ja0645483.
crossref pmid

83. Khan MS, Khalid M, Ahmad MS, Shahid M, Ahmad M. Three-in-one is really better: exploring the sensing and adsorption properties in a newly designed metal–organic system incorporating a copper(II) ion. Dalton Trans. 2019;48:12918–12932. 10.1039/C9DT02578B.
crossref pmid

84. Lazarou KN, Psycharis V, Terzis A, Raptopoulou C. Network diversity and supramolecular isomerism in copper(II)/1,2-bis(4-pyridyl)ethane coordination polymers. Polyhedron. 2011;30:963–970. 10.1016/j.poly.2010.12.030.
crossref

85. Xiao J-D, Qiu L-G, Ke F, et al. Rapid synthesis of nanoscale terbium-based metal–organic frameworks by a combined ultrasound-vapour phase diffusion method for highly selective sensing of picric acid. J. Mater. Chem. A. 2013;1:8745–8752. 10.1039/C3TA11517H.
crossref

86. Zhang W, Liu Y, Lu G, et al. Mesoporous Metal–Organic Frameworks with Size-, Shape-, and Space-Distribution-Controlled Pore Structure. Adv. Mater. 2015;27:2923–2929. 10.1002/adma.201405752.
crossref pmid

87. Chen M, Zhao H, Liu C-S, Wang X, Shi HZ, Du M. Template-directed construction of conformational supramolecular isomers for bilayer porous metal–organic frameworks with distinct gas sorption behaviors. Chem. Commun. 2015;51:6014–6017. 10.1039/C5CC00018A.
crossref pmid

88. Leszczyński MK, Justyniak I, Gontarczyk K, Lewiński J. Solvent Templating and Structural Dynamics of Fluorinated 2D Cu-Carboxylate MOFs Derived from the Diffusion-Controlled Process. Inorg. Chem. 2020;59:4389–4396. 10.1021/acs.inorgchem.9b03472.
crossref pmid pmc pdf

89. Fang Q, Zhu G, Xue M, Sun J, Qiu S. Amine-Templated Assembly of Metal–Organic Frameworks with Attractive Topologies. Cryst. Growth Des. 2008;8:319–329. 10.1021/cg070604f.
crossref

90. Pan J, Ma Y-J, Han S-D, Hu J, Mu Y, Wang G. Construction of the Lanthanide Diphosphonates via a Template-Synthesis Strategy: Structures, Proton Conduction, and Magnetic Behavior. Cryst. Growth Des. 2019;19:3045–3051. 10.1021/acs.cgd.9b00293.
crossref

91. de Lill DT, Gunning NS, Cahill CL. Toward Templated Metal–Organic Frameworks: Synthesis, Structures, Thermal Properties, and Luminescence of Three Novel Lanthanide–Adipate Frameworks. Inorg. Chem. 2005;44:258–266. 10.1021/ic048755k.
crossref pmid

92. Xu Z-X, Ma Y-L, Zhang J. Enantiopure anion templated synthesis of a zeolitic metal–organic framework. Chem. Commun. 2016;52:1923–1925. 10.1039/C5CC09308B.
crossref pmid

93. Zhao N, Sun F, He H, Jia J, Zhu G. Solvent-Induced Single Crystal to Single Crystal Transformation and Complete Metal Exchange of a Pyrene-Based Metal–Organic Framework. Cryst. Growth Des. 2014;14:1738–1743. 10.1021/cg401887b.
crossref

94. Cai B, Ren Y, Jiang H, et al. A mixed-valence lanthanide metal–organic framework, templated by 2,2′-bipyridine formed in situ reaction: synthesis, structure, and luminescent properties. Cryst. Eng. Comm. 2012;14:5285–5288. 10.1039/C2CE25160D.
crossref

95. Panda T, Pachfule P, Banerjee R. Template induced structural isomerism and enhancement of porosity in manganese(II) based metal–organic frameworks (Mn-MOFs). Chem. Commun. 2011;47:7674–7676. 10.1039/C1CC12278A.
crossref pmid

96. Zhang D-J, Yu Y-Z, Wang J-J, et al. Lanthanide(III) based metal organic frameworks constructed from 5-nitroisophthalate ligand with 2,4,5-tri(4-pyridyl)-imidazole templates: Syntheses, structures, and luminescent properties. Polyhedron. 2016;117:703–709. 10.1016/j.poly.2016.07.006.
crossref

97. Wang X, Liu L, Conato M, Jacobson A. Amine-Templated Anionic Metal–Organic Frameworks with the 4,4′-(Hexafluoroisopropylidene) Bis(benzoic acid) Ligand. Cryst. Growth Des. 2011;11:2257–2263. 10.1021/cg1016752.
crossref

98. Feng L, Wang K-Y, Lv X-L, Yan T-H, Zhou H. Hierarchically porous metal–organic frameworks: synthetic strategies and applications. Natl. Sci. Rev. 2020;7:1743–1758. 10.1093/nsr/nwz170.
crossref pmid pmc pdf

99. Ye R, Ni M, Xu Y, Chen H, Li S. Synthesis of Zn-based metal–organic frameworks in ionic liquid microemulsions at room temperature. RSC Adv. 2018;8:26237–26242. 10.1039/C8RA04573A.
crossref pmid pmc

100. Pileni M-P. The role of soft colloidal templates in controlling the size and shape of inorganic nanocrystals. Nat. Mater. 2003;2:145–150. 10.1038/nmat817.
crossref pmid pdf

101. Zhang L, Wang Y, Tong L, Xia Y. Synthesis of Colloidal Metal Nanocrystals in Droplet Reactors: The Pros and Cons of Interfacial Adsorption. Nano Lett. 2014;14:4189–4194. 10.1021/nl501994q.
crossref pmid

102. Zhang R, Liu J, He J, et al. Organic Reactions and Nanoparticle Preparation in CO2-InducedWater/P104/p-XyleneMicroemulsions. Chem. Eur. J. 2003;9:2167–2172. 10.1002/chem.200204496.
crossref pmid

103. Duan S, Dou B, Lin X, et al. Influence of active nanofiller ZIF-8 metal-organic framework (MOF) by microemulsion method on anticorrosion of epoxy coatings. Colloids Surf. 2021;624:126836. 10.1016/j.colsurfa.2021.126836.
crossref

104. Mu X, Jiang J, Chao F, Chen J. Ligand modification of UiO-66 with an unusual visible light photocatalytic behavior for RhB degradation. Dalton Trans. 2018;47:1895–1902. 10.1039/C7DT04477A.
crossref pmid

105. Srinivasan R, Elaiyappillai E, Nixon EJ, Sharmila I, Johnson P. Enhanced electrochemical behaviour of Co-MOF/PANI composite electrode for supercapacitors. Inorg. Chim. Acta. 2020;502:119393. 10.1016/j.ica.2019.119393.
crossref

106. Li N, Li Y, Li Q, Zhao Y, Liu C-S, Pang H. NiO nanoparticles decorated hexagonal Nickel-based metal-organic framework: Self-template synthesis and its application in electrochemical energy storage. J. Colloids Interface Sci. 2021;581:709–718. 10.1016/j.jcis.2020.07.134.
crossref

107. Ouyang B, Chen Q, Yuan H, et al. Reversible environmental impacts of iron-based metal-organic framework MIL-53(Fe) on nitrogen-fixing bacterium Azotobacter vinelandii. J. Environ. Chem. Eng. 2022;10:107794. 10.1016/j.jece.2022.107794.
crossref

108. Pu C, Zhao H, Hong Y, Zhan Q, Lan M. Elution-free ultra-sensitive enrichment for glycopeptides analyses: Using a degradable, post-modified Ce-metal–organic framework. Anal. Chim. Acta. 2019;1045:123–131. 10.1016/j.aca.2018.09.013.
crossref pmid

109. Rezki M, Septiani NLW, Iqbal M, et al. Amine-functionalized Cu-MOF nanospheres towards label-free hepatitis B surface antigen electrochemical immunosensors. J. Mater. Chem. B. 2021;9:5711–5721. 10.1039/D1TB00222H.
crossref pmid

110. Yin Y, Shao J, Zhang L, Cui Q, Wang H. Study on heat conduction and adsorption/desorption characteristic of MIL-101/few layer graphene composite. J. Porous Mater. 2021;28:1197–1213. 10.1007/s10934-021-01074-4.
crossref pdf

111. Burgaz E, Erciyes A, Andac M, Andac O. Synthesis and characterization of nano-sized metal organic framework-5 (MOF-5) by using consecutive combination of ultrasound and microwave irradiation methods. Inorg. Chim. Acta. 2019;485:118–124. 10.1016/j.ica.2018.10.014.
crossref

112. Cho H-Y, Kim J, Kim S-N, Ahn W. High yield 1-L scale synthesis of ZIF-8 via a sonochemical route. Microporous Mesoporous Mater. 2013;169:180–184. 10.1016/j.micromeso.2012.11.012.
crossref

113. Yang D-A, Cho H-Y, Kim J, Yang S-T, Ahn W. CO2capture and conversion using Mg-MOF-74 prepared by a sonochemical method. Energy Environ. Sci. 2012;5:6465–6473. 10.1039/C1EE02234B.
crossref

114. Brainer NS, dos Santos TV, Barbosa CDES, Meneghetti S. Simple and fast ultrasound-assisted synthesis of Sn-MOFs and obtention of SnO2. Mater. Lett. 2020;280:128512. 10.1016/j.matlet.2020.128512.
crossref

115. Azad FN, Ghaedi M, Dashtian K, Hajati S, Pezeshkpour V. Ultrasonically assisted hydrothermal synthesis of activated carbon–HKUST-1-MOF hybrid for efficient simultaneous ultrasound-assisted removal of ternary organic dyes and antibacterial investigation: Taguchi optimization. Ultrason. Sonochem. 2016;31:383–393. 10.1016/j.ultsonch.2016.01.024.
crossref pmid

116. Wu X, Bao Z, Yuan B, et al. Microwave synthesis and characterization of MOF-74 (M=Ni, Mg) for gas separation. Microporous Mesoporous Mater. 2013;180:114–122. 10.1016/j.micromeso.2013.06.023.
crossref

117. Yoo Y, Lai Z, Jeong H-K. Fabrication of MOF-5 membranes using microwave-induced rapid seeding and solvothermal secondary growth. Microporous Mesoporous Mater. 2009;123:100–106. 10.1016/j.micromeso.2009.03.036.
crossref

118. Jung D-W, Yang D-A, Kim J, Ahn W. Facile synthesis of MOF-177 by a sonochemical method using 1-methyl-2-pyrrolidinone as a solvent. Dalton Trans. 2010;39:2883–2887. 10.1039/B925088C.
crossref pmid

119. Minh TT, Phong NH, Van Duc H, Khieu D. Microwave synthesis and voltammetric simultaneous determination of paracetamol and caffeine using an MOF-199-based electrode. J. Mater. Sci. 2018;53:2453–2471. 10.1007/s10853-017-1715-0.
crossref pdf

120. Han Y, Cui J, Yu Y, et al. Efficient Metal-Oriented Electrodeposition of a Co-Based Metal-Organic Framework with Superior Capacitive Performance. ChemSusChem. 2022;15:e202200644. 10.1002/cssc.202200644.
crossref pmid pmc pdf

121. Worrall SD, Bissett MA, Hill PI, et al. Metal-organic framework templated electrodeposition of functional gold nanostructures. Electrochim. Acta. 2016;222:361–369. 10.1016/j.electacta.2016.10.187.
crossref

122. Tian N, Gao Y, Wu J, Luo S, Dai W. Water-resistant HKUST-1 functionalized with polydimethylsiloxane for efficient rubidium ion capture. New J. Chem. 2019;43:15539–15547. 10.1039/C9NJ03632F.
crossref

123. Naseri M, Fotouhi L, Ehsani A, Dehghanpour S. Facile electrosynthesis of nano flower like metal-organic framework and its nanocomposite with conjugated polymer as a novel and hybrid electrode material for highly capacitive pseudocapacitors. J. Colloids Interface Sci. 2016;484:314–319. 10.1016/j.jcis.2016.09.001.
crossref pmid

124. Abbasi AR, Karimi M, Daasbjerg K. Efficient removal of crystal violet and methylene blue from wastewater by ultrasound nanoparticles Cu-MOF in comparison with mechanosynthesis method. Ultrason. Sonochem. 2017;37:182–191. 10.1016/j.ultsonch.2017.01.007.
crossref pmid

125. Jeong H, Lee J. 3D-Superstructured Networks Comprising Fe-MIL-88A Metal-Organic Frameworks Under Mechanochemical Conditions. Eur. J. Inorg. Chem. 2019;2019:4597–4600. 10.1002/ejic.201900979.
crossref pdf

126. Chen Y, Wu H, Liu Z, Sun X, Xia Q, Li Z. Liquid-Assisted Mechanochemical Synthesis of Copper Based MOF-505 for the Separation of CO2 over CH4 or N2. Ind. Eng. Chem. Res. 2018;57:703–709. 10.1021/acs.iecr.7b03712.
crossref

127. Prochowicz D, Nawrocki J, Terlecki M, Marynowski W, Lewiński J. Facile Mechanosynthesis of the Archetypal Zn-Based Metal–Organic Frameworks. Inorg. Chem. 2018;57:13437–13442. 10.1021/acs.inorgchem.8b02026.
crossref pmid

128. Banerjee R, Phan A, Wang B, et al. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science. 2008;319:939–943. 10.1126/science.1152516.
crossref pmid

129. McKinstry C, Cathcart RJ, Cussen EJ, Fletcher AJ, Patwardhan SV, Sefcik J. Scalable continuous solvothermal synthesis of metal organic framework (MOF-5) crystals. Chem. Eng. J. 2016;285:718–725. 10.1016/j.cej.2015.10.023.
crossref

130. Haque E, Khan NA, Park JH, Jhung S. Synthesis of a Metal–Organic Framework Material, Iron Terephthalate, by Ultrasound, Microwave, and Conventional Electric Heating: A Kinetic Study. Chem. Eur. J. 2010;16:1046–1052. 10.1002/chem.200902382.
crossref pmid

131. Cho H-Y, Yang D-A, Kim J, Jeong S-Y, Ahn W. CO2 adsorption and catalytic application of Co-MOF-74 synthesized by microwave heating. Catal. Today. 2012;185:35–40. 10.1016/j.cattod.2011.08.019.
crossref

132. Joaristi AM, Juan-Alcañiz J, Serra-Crespo P, Kapteijn F, Gascon J. Electrochemical Synthesis of Some Archetypical Zn2+,Cu2+, and Al3+ Metal Organic Frameworks. Cryst. Growth Des. 2012;12:3489–3498. 10.1021/cg300552w.
crossref

133. Klimakow M, Klobes P, Thünemann AF, Rademann K, Emmerling F. Mechanochemical Synthesis of Metal–Organic Frameworks: A Fast and Facile Approach toward Quantitative Yields and High Specific Surface Areas. Chem. Mater. 2010;22:5216–5221. 10.1021/cm1012119.
crossref

134. Remya VR, Kurian M. Synthesis and catalytic applications of metal–organic frameworks: a review on recent literature. Int. Nano. Lett. 2019;9:17–29. 10.1007/s40089-018-0255-1.
crossref pdf

135. Lazarou KN, Psycharis V, Terzis A, Raptopoulou C. Network diversity and supramolecular isomerism in copper(II)/1,2-bis(4-pyridyl)ethane coordination polymers. Polyhedron. 2011;30:963–970. 10.1016/j.poly.2010.12.030.
crossref

136. Zhao N, Cai K, He H. The synthesis of metal–organic frameworks with template strategies. Dalton Trans. 2020;49:11467–11479. 10.1039/D0DT01879A.
crossref pmid

137. Ye R, Ni M, Xu Y, Chen H, Li S. Synthesis of Zn-based metal–organic frameworks in ionic liquid microemulsions at room temperature. RSC Adv. 2018;8:26237–26242. 10.1039/C8RA04573A.
crossref pmid pmc

138. Liu C-S, Li J, Pang H. Metal-organic framework-based materials as an emerging platform for advanced electrochemical sensing. Coord. Chem. Rev. 2020;410:213222. 10.1016/j.ccr.2020.213222.
crossref

139. Liu W, Yin X-B. Metal–organic frameworks for electrochemical applications. Trends Analyt. Chem. 2016;75:86–96. 10.1016/j.trac.2015.07.011.
crossref

140. Zhang X, Chen A, Zhong M, et al. Metal–organic frameworks (MOFs) and MOF-derived materials for energy storage and conversion. Electrochem. Energy Rev. 2019;2:29–104. 10.1021/acs.nanolett.0c04898.
crossref pdf

141. Jaouen F, Morozan A. Metal-Organic Frameworks: Electrochemical Properties. Encyclopedia of Inorganic and Bioinorganic Chemistry. Wiley online Library; 2014. p. 1–24.
crossref

142. Morozan A, Jaouen F. Metal organic frameworks for electrochemical applications. Energy Environ. Sci. 2012;5:9269–9290. 10.1039/C2EE22989G.
crossref

143. Yang J, Zhao F, Zeng B. One-step synthesis of a copper-based metal–organic framework–graphene nanocomposite with enhanced electrocatalytic activity. RSC Adv. 2015;5:22060–22065. 10.1039/C4RA16950F.
crossref

144. Shams-Eldin R, Ali AA, Hani A, et al. Metal–organic framework mediated Ni-deposition on MWCNTs for direct methanol fuel cell catalysis. SN Appl. Sci. 2023;5:166. 10.1007/s42452-023-05387-8.
crossref pdf

145. Stassen I, Styles M, Van Assche T, et al. Electrochemical Film Deposition of the Zirconium Metal–Organic Framework UiO-66 and Application in a Miniaturized Sorbent Trap. Chem. Mater. 2015;27:1801–1807. 10.1021/cm504806p.
crossref

146. Ye W, Li Y, Wang J, et al. Electrochemical detection of trace heavy metal ions using a Ln-MOF modified glass carbon electrode. J. Solid State Chem. 2020;281:121032. 10.1016/j.jssc.2019.121032.
crossref

147. Wang J. Analytical Electochemistry. 4th edJohn Wiley & Sons, Inc; New York, USA: 2000. p. 106–108.


148. Cruz-Navarro JA, Hernandez-Garcia F, Alvarez Romero GA. Novel applications of metal-organic frameworks (MOFs) as redox-active materials for elaboration of carbon-based electrodes with electroanalytical uses. Coord. Chem. Rev. 2020;412:213263. 10.1016/j.ccr.2020.213263.
crossref

149. Švancara I, Kalcher K. Electrochemistry of Carbon Electrodes. John Wiley & Sons; 2015. p. 379–424.


150. Švancara I, Vytřas K, Kalcher K, Walcarius A, Wang J. Carbon Paste Electrodes in Facts, Numbers, and Notes: A Review on the Occasion of the 50-Years Jubilee of Carbon Paste in Electrochemistry and Electroanalysis. Electroanalysis. 2009;21:7–28. 10.1002/elan.200804340.
crossref

151. Alipour Z, Hassaninejad-Darzi SK, Asadollahi-Baboli M. Construction of Pd-Ag/g-C3N4/CPE electrochemical nanosensor for simultaneous determination of clozapine and quetiapine antipsychotic drugs. J. Alloys Compd. 2023;941:168958. 10.1016/j.jallcom.2023.168958.
crossref

152. Svancara I, Kalcher K, Walcarius A. Electroanalysis with Carbon Paste Electrodes. 1st edCRC Press; 2012. p. 54–56.


153. Fazl F, Gholivand MB. High performance electrochemical method for simultaneous determination dopamine, serotonin, and tryptophan by ZrO2–CuO co-doped CeO2 modified carbon paste electrode. Talanta. 2022;239:122982. 10.1016/j.talanta.2021.122982.
crossref pmid

154. Bi W, Wang G, Hu X. Fabrication of Zn-MOF derived graphitic carbon materials with mesoporous structure for adsorptive removal of ceftazidime from aqueous solutions. Colloids Surf. A. 2022;652:129758. 10.1016/j.colsurfa.2022.129758.
crossref

155. Wang H, Hu Q, Meng Y, et al. Efficient detection of hazardous catechol and hydroquinone with MOF-rGO modified carbon paste electrode. J. Hazard. Mater. 2018;353:151–157. 10.1016/j.jhazmat.2018.02.029.
crossref pmid

156. Kumari N, Sareen S, Verma M, et al. Zirconia-based nanomaterials: recent developments in synthesis and applications. Nanoscale Adv. 4:4210–4236. 10.1039/d2na00367h.
crossref pmid

157. Chikere CO, Faisal NH, Kong-Thoo-Lin P, Fernandez C. Interaction between Amorphous Zirconia Nanoparticles and Graphite: Electrochemical Applications for Gallic Acid Sensing Using Carbon Paste Electrodes in Wine. Nanomaterials. 2020;10:537. 10.3390/nano10030537.
crossref pmid pmc

158. Petit C, Bandosz TJ. Synthesis, Characterization, and Ammonia Adsorption Properties of Mesoporous Metal–Organic Framework (MIL(Fe))–Graphite Oxide Composites: Exploring the Limits of Materials Fabrication. Adv. Funct. Mater. 2011;21:2108–2117. 10.1002/adfm.201002517.
crossref

159. Jovanović VM, Terzić S, Tripković AV, Popović K, Lović J. The effect of electrochemically treated glassy carbon on the activity of supported Pt catalyst in methanol oxidation. Electrochem. Commun. 2004;6:1254–1258. 10.1016/j.elecom.2004.10.001.
crossref

160. Wan C, Qu A, Deng L, Liu X, Wu C. Preparation of electrochemical sensor based on glassy carbon electrode and its specificity and sensitivity for directional detection of antibiotic resistance genes spreading in the water environment. Environ. Sci. Pollut. Res. 2023;30:7904–7913. 10.1007/s11356-022-22787-2.
crossref pmid pdf

161. Wu L, Lu Z, Ye J. Enzyme-free glucose sensor based on layer-by-layer electrodeposition of multilayer films of multi-walled carbon nanotubes and Cu-based metal framework modified glassy carbon electrode. Biosens. Bioelectron. 2019;135:45–49. 10.1016/j.bios.2019.03.064.
crossref pmid

162. Sharma S. Glassy Carbon: A Promising Material for Microand Nanomanufacturing. Materials. 2018;11:1857. 10.3390/ma11101857.
crossref pmid pmc

163. Arul P, John SA. Silver nanoparticles built-in zinc metal organic framework modified electrode for the selective non-enzymatic determination of H2O2. Electrochim. Acta. 2017;235:680–689. 10.1016/j.electacta.2017.03.097.
crossref

164. Dhanjai , Sinha A, Lu X, et al. Voltammetric sensing of biomolecules at carbon based electrode interfaces: A review. TrAC Trends Analyt. Chem. 2018;98:174–189. 10.1016/j.trac.2017.11.010.
crossref

165. Labib M, Sargent EH, Kelley SO. Electrochemical Methods for the Analysis of Clinically Relevant Biomolecules. Chem. Rev. 2016;116:9001–9090. 10.1021/acs.chemrev.6b00220.
crossref pmid

166. Dhanjai , Sinha A, Lu X, et al. Voltammetric sensing of biomolecules at carbon based electrode interfaces: A review. Trends Analyt. Chem. 2018;98:174–189. 10.1016/j.trac.2017.11.010.
crossref

167. Du L, Chen W, Zhu P, Tian Y, Chen Y, Wu C. Applications of Functional Metal-Organic Frameworks in Biosensors. Biotechnol. J. 2021;16:1900424. 10.1002/biot.201900424.
crossref pmid pdf

168. Molla A, Youk JH. Recent progress on electroanalytical sensing of small molecules and biomolecules using carbon dots: A review. J. Ind. Eng. Chem. 2023;127:62–81. 10.1016/j.jiec.2023.07.037.
crossref

169. Wang Z, Sun Q, Liu B, et al. Recent advances in porphyrin-based MOFs for cancer therapy and diagnosis therapy. Coord. Chem. Rev. 2021;439:213945. 10.1016/j.ccr.2021.213945.
crossref

170. Musa AM, Kiely J, Luxton R, Honeychurch K. Recent progress in screen-printed electrochemical sensors and biosensors for the detection of estrogens. Trends Analyt. Chem. 2021;139:116254. 10.1016/j.trac.2021.116254.
crossref

171. Park S, Boo H, Chung TD. Electrochemical non-enzymatic glucose sensors. Anal. Chim. Acta. 2006;556:46–57. 10.1016/j.aca.2005.05.080.
crossref pmid

172. Sarma AK, Vatsyayan P, Goswami P, Minteer S. Recent advances in material science for developing enzyme electrodes. Biosens. Bioelectron. 2009;24:2313–2322. 10.1016/j.bios.2008.09.026.
crossref pmid

173. Zhang L, Li S, Xin J, et al. A non-enzymatic voltammetric xanthine sensor based on the use of platinum nanoparticles loaded with a metal-organic framework of type MIL-101(Cr). Application to simultaneous detection of dopamine, uric acid, xanthine and hypoxanthine. Microchim. Acta. 2018;186:9. 10.1007/s00604-018-3128-4.
crossref pmid pdf

174. Pandey A, Dhas N, Deshmukh P, et al. Heterogeneous surface architectured metal-organic frameworks for cancer therapy, imaging, and biosensing: A state-of-the-art review. Coord. Chem. Rev. 2020;409:213212. 10.1016/j.ccr.2020.213212.
crossref

175. Zhou R, Zhuang X, Wu Q, et al. Cu-MOF@Pt 3D nanocomposites prepared by one-step wrapping method with peroxidase-like activity for colorimetric detection of glucose. Colloids Surf. B. 2022;216:112601. 10.1016/j.colsurfb.2022.112601.
crossref pmid

176. Yin H, Zhu J, Chen J, Gong J, Nie Q. MOF-derived in situ growth of carbon nanotubes entangled Ni/NiO porous polyhedrons for high performance glucose sensor. Mater. Lett. 2018;221:267–270. 10.1016/j.matlet.2018.03.156.
crossref

177. Arul P, John SA. Electrodeposition of CuO from Cu-MOF on glassy carbon electrode: A non-enzymatic sensor for glucose. J Electroanal Chem. 2017;799:10.1016/j.jelechem.2017.05.041.
crossref

178. Shahrokhian S, Khaki SE, Hosseini H. Direct growth of metal-organic frameworks thin film arrays on glassy carbon electrode based on rapid conversion step mediated by copper clusters and hydroxide nanotubes for fabrication of a high performance non-enzymatic glucose sensing platform. Biosens. Bioelectron. 2018;112:100–107. 10.1016/j.bios.2018.04.039.
crossref pmid

179. Wu L, Lu Z-W, Ma Y, et al. Cu(II) Metal-Organic Framework Encapsulated in Carbon Paste Electrode for High-Performance Non-Enzymatic Glucose Sensing. Chinese J. Anal. Chem. 2020;48:e20038–e20046. 10.1016/S1872-2040(20)60006-8.
crossref

180. Zhang L, Li S, Xin J, et al. A non-enzymatic voltammetric xanthine sensor based on the use of platinum nanoparticles loaded with a metal-organic framework of type MIL-101(Cr). Application to simultaneous detection of dopamine, uric acid, xanthine and hypoxanthine. Microchim. Acta. 2018;186:9. 10.1007/s00604-018-3128-4.
crossref pmid pdf

181. Feng X, Yin X, Bo X, Guo L. An ultrasensitive luteolin sensor based on MOFs derived CuCo coated nitrogen-doped porous carbon polyhedron. Sens. Actuators B Chem. 2019;281:730–738. 10.1016/j.snb.2018.11.010.
crossref

182. Zheng Y-Y, Li C-X, Ding X-T, et al. Detection of dopamine at graphene-ZIF-8 nanocomposite modified electrode. Chin. Chem. Lett. 2017;28:1473–1478. 10.1016/j.cclet.2017.03.014.
crossref

183. Hosseini H, Ahmar H, Dehghani A, Bagheri A, Tadjarodi A, Fakhari A. A novel electrochemical sensor based on metal-organic framework for electro-catalytic oxidation of L-cysteine. Biosens. Bioelectron. 2013;42:426–429. 10.1016/j.bios.2012.09.062.
crossref pmid

184. Duan D, Si X, Ding Y, et al. A novel molecularly imprinted electrochemical sensor based on double sensitization by MOF/CNTs and Prussian blue for detection of 17β-estradiol. Bioelectrochemistry. 2019;129:211–217. 10.1016/j.bioelechem.2019.04.014.
crossref pmid

185. He J, Yang H, Zhang Y, et al. Smart Nanocomposites of Cu-Hemin Metal-Organic Frameworks for Electrochemical Glucose Biosensing. Sci. Rep. 2016;6:36637. 10.1038/srep36637.
crossref pmid pmc pdf

186. Singh R, Musameh M, Gao Y, Ozcelik B, Mulet X, Doherty C. Stable MOF@enzyme composites for electrochemical biosensing devices. J. Mater. Chem. C. 2021;9:7677–7688. 10.1039/D1TC00407G.
crossref

187. Hou C, Zhao D, Wang Y, Zhang S, Li S. Preparation of magnetic Fe3O4/PPy@ZIF-8 nanocomposite for glucose oxidase immobilization and used as glucose electrochemical biosensor. J. Electroanal. Chem. 2018;822:50–56. 10.1016/j.jelechem.2018.04.067.
crossref

188. Arul P, Gowthaman NSK, John SA, et al. Tunable electrochemical synthesis of 3D nucleated microparticles like Cu-BTC MOF-carbon nanotubes composite: Enzyme free ultrasensitive determination of glucose in a complex biological fluid. Electrochim. Acta. 2020;354:136673. 10.1016/j.electacta.2020.136673.
crossref

189. Li Y, Shen Y, Zhang Y, et al. A UiO-66-NH2/carbon nanotube nanocomposite for simultaneous sensing of dopamine and acetaminophen. Anal. Chim. Acta. 2021;1158:338419. 10.1016/j.aca.2021.338419.
crossref pmid

190. Tran TQN, Das G, Yoon HH. Nickel-metal organic framework/MWCNT composite electrode for non-enzymatic urea detection. Sens. Actuators B Chem. 2017;243:78–83. 10.1016/j.snb.2016.11.126.
crossref

191. Dang W, Sun Y, Jiao H, Xu L, Lin M. AuNPs-NH2/Cu-MOF modified glassy carbonelectrodeasenzyme-free electrochemical sensor detecting H2O2. J. Electroanal. Chem. 2020;856:113592. 10.1016/j.jelechem.2019.113592.
crossref

192. Naseri M, Fotouhi L, Ehsani A. Nanostructured Metal Organic Framework Modified Glassy Carbon Electrode as a High Efficient Non-Enzymatic Amperometric Sensor for Electrochemical Detection of H2O2. J. Electrochem. Sci. Technol. 2018;9:28–36. 10.5229/JECST.2018.9.1.28.
crossref pdf

193. Meng W, Xu S, Dai L, Li Y, Zhu J, Wang L. An enhanced sensitivity towards H2O2 reduction based on a novel Cu metal–organic framework and acetylene black modified electrode. Electrochim. Acta. 2017;230:324–332. 10.1016/j.electacta.2017.02.017.
crossref

194. Zhang D, Zhang J, Zhang R, et al. 3D porous metal-organic framework as an efficient electrocatalyst for nonenzymatic sensing application. Talanta. 2015;144:1176–1181. 10.1016/j.talanta.2015.07.091.
crossref pmid

195. Cobos M, De-La-Pinta I, Quindós G, Fernández MJ, Fernández M. Graphene Oxide–Silver Nanoparticle Nanohybrids: Synthesis, Characterization, and Antimicrobial Properties. Nanomaterials. 2020;10:376. 10.3390/nano10020376.
crossref pmid pmc

196. Zhang D, Zhang J, Shi H, et al. Redox-active microsized metal-organic framework for efficient nonenzymatic H2O2 sensing. Sens. Actuators B Chem. 2015;221:224–229. 10.1016/j.snb.2015.06.079.
crossref

197. Lopa NS, Rahman MdM, Ahmed F, Chandra S, Ryu T, Kim W. A base-stable metal-organic framework for sensitive and non-enzymatic electrochemical detection of hydrogen peroxide. Electrochim. Acta. 2018;274:49–56. 10.1016/j.electacta.2018.03.148.
crossref

198. Wang M-Q, Zhang Y, Bao S-J, Yu Y-N, Ye C. Ni(II)-Based Metal-Organic Framework Anchored on Carbon Nanotubes for Highly Sensitive Non-Enzymatic Hydrogen Peroxide Sensing. Electrochim. Acta. 2016;190:365–370. 10.1016/j.electacta.2015.12.199.
crossref

199. Liu X, Chen W, Lian M, Chen X, Lu Y, Yang W. Enzyme immobilization on ZIF-67/MWCNT composite engenders high sensitivity electrochemical sensing. J. Electroanal. Chem. 2019;833:505–511. 10.1016/j.jelechem.2018.12.027.
crossref

200. Mathew G, Daniel M. Real-time electrochemical quantification of H2O2 in living cancer cells using Bismuth based MOF. J. Electroanal. Chem. 2022;914:116255. 10.1016/j.jelechem.2022.116255.
crossref

201. Liu B, Wang X, Zhai Y, et al. Facile preparation of well conductive 2D MOF for nonenzymatic detection of hydrogen peroxide: Relationship between electrocatalysis and metal center. J. Electroanal. Chem. 2020;858:113804. 10.1016/j.jelechem.2019.113804.
crossref

202. Ziad R, Columbus S, Elgamouz A, et al. Multi-functional silver nanoprism-titanium dioxide hybrid nanoarrays for trace-level SERS sensing and photocatalytic removal of hazardous organic pollutants. Spectrochim. Acta. 2023;297:122701. 10.1016/j.saa.2023.122701.
crossref pmid

203. Paz R, Viltres H, Gupta NK, et al. Covalent organic frameworks as highly versatile materials for the removal and electrochemical sensing of organic pollutants. Chemosphere. 2023;342:140145. 10.1016/j.chemosphere.2023.140145.
crossref pmid

204. Gouveia TIA, Alves A, Santos MSF. New insights on cytostatic drug risk assessment in aquatic environments based on measured concentrations in surface waters. Environ. Int. 2019;133:105236. 10.1016/j.envint.2019.105236.
crossref pmid

205. Li D, Chen H, Liu H, et al. Anticancer drugs in the aquatic ecosystem: Environmental occurrence, ecotoxicological effect and risk assessment. Environ. Int. 2021;153:106543. 10.1016/j.envint.2021.106543.
crossref pmid

206. Hasani A, Do HH, Tekalgne M, Hong SH, Jang HW, Kim S. Recent progress of two-dimensional materials and metal–organic framework-based taste sensors. J. Korean Ceram. Soc. 2020;57:353–367. 10.1007/s43207-020-00047-8.
crossref pdf

207. Xu S, Hu S, Zhu L, Wang W. Haloquinone Chloroimides as Toxic Disinfection Byproducts Identified in Drinking Water. Environ. Sci. Technol. 2021;55:16347–16357. 10.1021/acs.est.1c01690.
crossref pmid

208. Kung C-W, Chang T-H, Chou L-Y, Hupp JT, Farha OK, Ho K. Porphyrin-based metal–organic framework thin films for electrochemical nitrite detection. Electrochem. Commun. 2015;58:51–56. 10.1016/j.elecom.2015.06.003.
crossref

209. Yuan B, Zhang J, Zhang R, et al. Cu-based metal–organic framework as a novel sensing platform for the enhanced electro-oxidation of nitrite. Sens. Actuators B Chem. 2016;222:632–637. 10.1016/j.snb.2015.08.100.
crossref

210. Zhang Y, Bo X, Nsabimana A, Han C, Li M, Guo L. Electrocatalytically active cobalt-based metal–organic framework with incorporated macroporous carbon composite for electrochemical applications. J. Mater. Chem. A. 2014;3:732–738. 10.1039/C4TA04411H.
crossref

211. Li J, Xia J, Zhang F, Wang Z, Liu Q. An electrochemical sensor based on copper-based metal-organic frameworks-graphene composites for determination of dihydroxybenzene isomers in water. Talanta. 2018;181:80–86. 10.1016/j.talanta.2018.01.002.
crossref pmid

212. Wang Y, Wang L, Chen H, Hu X, Ma S. Fabrication of Highly Sensitive and Stable Hydroxylamine Electrochemical Sensor Based on Gold Nanoparticles and Metal–Metalloporphyrin Framework Modified Electrode. ACS Appl. Mater. Interfaces. 2016;8:18173–18181. 10.1021/acsami.6b04819.
crossref pmid

213. Zhang J, Xu X, Chen L. An ultrasensitive electrochemical bisphenol A sensor based on hierarchical Ce-metal-organic framework modified with cetyltrimethylammonium bromide. Sens. Actuators B Chem. 2018;261:425–433. 10.1016/j.snb.2018.01.170.
crossref

214. Xiao L, Xu R, Yuan Q, Wang F. Highly sensitive electrochemical sensor for chloramphenicol based on MOF derived exfoliated porous carbon. Talanta. 2017;167:39–43. 10.1016/j.talanta.2017.01.078.
crossref pmid

215. Cheng Y, Ma B, Tan C-P, et al. Hierarchical macro-microporous ZIF-8 nanostructures as efficient nano-lipase carriers for rapid and direct electrochemical detection of nitrogenous diphenyl ether pesticides. Sens. Actuators B Chem. 2020;321:128477. 10.1016/j.snb.2020.128477.
crossref

216. Zhang J, Xu X, Qiang Y. Ultrasensitive electrochemical aptasensor for ochratoxin A detection using AgPt bimetallic nanoparticles decorated iron-porphyrinic metal-organic framework for signal amplification. Sens. Actuators B Chem. 2020;312:127964. 10.1016/j.snb.2020.127964.
crossref

217. Mahmoudi E, Fakhri H, Hajian A, Afkhami A, Bagheri H. High-performance electrochemical enzyme sensor for organophosphate pesticide detection fcusing modified metal-organic framework sensing platforms. Bioelectrochemistry. 2019;130:107348. 10.1016/j.bioelechem.2019.107348.
crossref pmid

218. Balali-Mood M, Naseri K, Tahergorabi Z, Khazdair MR, Sadeghi M. Toxic Mechanisms of Five Heavy Metals: Mercury, Lead, Chromium, Cadmium, and Arsenic. Front Pharmacol. 2021;12. 10.3389/fphar.2021.643972.
crossref pmid

219. Shamim MA, Zia H, Zeeshan M, Khan MY, Shahid M. Metal organic frameworks (MOFs) as a cutting-edge tool for the selective detection and rapid removal of heavy metal ions from water: Recent progress. J. Environ. Chem. Eng. 2022;10:106991. 10.1016/j.jece.2021.106991.
crossref

220. Khan BN, Ullah H, Ashfaq Y, et al. Elucidating the effects of heavy metals contamination on vital organ of fish and migratory birds found at fresh water ecosystem. Heliyon. 2023;9:e20968. 10.1016/j.heliyon.2023.e20968.
crossref pmid pmc

221. Iurgenson N, Wang X, Kong L, et al. Feasibility study of multisensor systems for the assessment of water pollution index induced by heavy metal contamination. Microchem. J. 2024;197:109762. 10.1016/j.microc.2023.109762.
crossref

222. Hartiningsih D, Diana S, Yuniarti MS, Ismail MR, Sari Q. Water quality pollution indices to assess the heavy metal contamination: A case study of the estuarine waters in Cirebon City (West Java, Indonesia) Pre-and post-CARE COVID-19. Environ. Sustain. Indic. 2024;21:100318. 10.1016/j.indic.2023.100318.
crossref

223. Nasir Z, Sabir A, Salman HM, et al. Fingerprinting of heavy metal and microbial contamination uncovers the unprecedented scale of water pollution and its implication on human health around transboundary Hudiara drain in South Asia. Environ. Technol. Innov. 2023;30:103040. 10.1016/j.eti.2023.103040.
crossref

224. Haider I, Ali MA, Sanaullah M, et al. Unlocking the secrets of soil microbes: How decades-long contamination and heavy metals accumulation from sewage water and industrial effluents shape soil biological health. Chemosphere. 2023;342:140193. 10.1016/j.chemosphere.2023.140193.
crossref pmid

Fig. 1
Schematic of the hydrothermal/solvothermal synthesis of metal organic frameworks.
/upload/thumbnails/eer-2023-636f1.gif
Fig. 2
Schematic of Cu-BTC production with ultrasonic assistance and SEM images of MOF with varying DMF concentrations during sonication for 1 min; XRD patterns of Cu-BTC samples synthesized by ultrasonic irradiation for 1 min with different DMF concentrations in the solvent of water (4 mL) - ethanol (2 mL) - DMF(x mL): (1) 0.0 mL (2) 0.2 mL (3) 0.5 mL (4) 1.0 mL (5) 3.0 mL and (6) 6.0 mL [55].
/upload/thumbnails/eer-2023-636f2.gif
Fig. 3
Microwave assisted synthesis of Zr (fumarate-face centered cubic) MOF [66].
/upload/thumbnails/eer-2023-636f3.gif
Fig. 4
Electrochemical synthesis of Zn3(BTC)2-MOF [73].
/upload/thumbnails/eer-2023-636f4.gif
Fig. 5
Schematic of mechanochemical synthesis of MOF.
/upload/thumbnails/eer-2023-636f5.gif
Fig. 6
Ex-situ fabrication of modified carbon paste electrode and application in the electrochemical detection of analytes.
/upload/thumbnails/eer-2023-636f6.gif
Table 1
Synthesis of diverse MOFs obtained by distinctive methods.
Sample Metal Ligand Solvent Condition Ref.
Hydrothermal synthesis UiO-66 ZrCl4 H2BDC DMF 120°C, 24 h [104]
Co-MOF Co(NO3)2·6H2O H3BTC DMF 100°C, 24 h [105]
Ni-MOF Ni(NO3)2·6H2O H3BTC DMF 80°C, 18 h [106]
MIL-53 FeCl3·6H2O H2BDC DMF 150°C, 15 h [107]
Ce-MOF Ce(NO3)3·6H2O H3BTC DMF-Ehanol 120°C, 2 h [108]
Cu-NH2BDC Cu(NO3)2·3H2O NH2BDC DMF-Ehanol 110°C, 20 h [109]
MIL-101 Cr(NO3)3·9H2O H2BDC De-ionised water 180°C, 5 h [110]

Ultrasound MOF-5 Zn(NO3)2·6H2O H2BDC DMF 90 W, 2 min [111]
ZIF-8 Zn(NO3)2·6H2O MeIM DMF 300 W, 1 h [112]
MOF-74 Mg(NO3)2·6H2O H4dhtp DMF 500 W, 1 h [113]
Sn-BDC SnSO4 Na2BDC Deionized water 155 W, 5 min [114]
HKUST-1 Cu(NO3)2·5H2O H3BTC DMF 130 W, 1 h [115]

Microwave method Ni-MOF-74 Ni(NO3)2·6H2O DOT DMF 100°C, 90 min [116]
Mg-MOF Mg(NO3)2·6H2O DOT DMF 125°C, 90 min [116]
MOF-5 Zn(NO3)2·6H2O H2BDC DMF 300 W, 2.5 min [117]
MOF-177 Zn(NO3)2·6H2O H3BTB NMP 800 W, 35 min [118]
MOF-199 Cu(NO3)2·3H2O H3BTC DMF 250 W, 30 min [119]

Electrochemical synthesis Co-MOF Co(NO3)2·6H2O H3BTC H2O, ethanol Electrolyte (Et3NHCl)·6H2O [120]
HKUST-1 Cu foil electrode H3BTC DMSO, ethanol Electrolyte (MTBAMS) [121]
HKUST-1 Cu electrode H3BTC Methanol Electrolyte (TBATFB) [122]
Cu-MOF Cu(NO3)2·3H2O H4BTEC DMF, H2O Electrolyte (TBATFB) [123]

Mechanochemical synthesis Cu-MOF Cu(OAc)2·H2O H3BTC No solvent 15.0 min [124]
MIL-88A FeCl3·6H2O Furmarate No solvent 10.0 min [125]
MOF-505 Cu(OAc)2·H2O H4bptc DMF 40.0 Hz, 80 min [126]
IRMOF-3 Zn44O)(NHOCPh)6 NH2BDC No solvent 30.0 Hz, 30 min [127]
Table 2
Advantages and limitations of various synthesis methods of MOFs.
Synthesis method Advantages Limitations Reference
Hydrothermal
  • Large operating temperature range (i.e., between 80 and 250°C).

  • Easy transposition.

  • High yield.

  • High energy consumption.

  • Long reaction time.

  • Expensive equipment required.

[128,129]
Ultrasonic
  • Homogeneous particle size and morphology.

  • Short reaction time.

  • Formation of stable structures.

  • Suitable method for the preparation of nanosized MOFs.

  • Destruction of crystallites hindering the formation of large single crystals for X-ray diffraction studies.

[130]
Microwave
  • Reduction in crystallization time.

  • High yields.

  • Possibility to control morphology, phase selectivity, and particle distribution.

  • Simple method and easy to control the reaction parameters.

  • Difficult to isolate large single crystals.

  • Lack of a direct method for scale-up.

[131]
Electrochemical
  • Avoiding anions such as nitrates from metal salts.

  • Low temperatures of reaction and extremely quick synthesis.

  • Rarely reported method for synthesis of MOFs.

[132]
Mechanochemical
  • Solvent-free synthetic method.

  • Maintaining the high temperature and pressure is not required.

  • Difficulty in isolation of single-crystals for X-ray diffraction studies.

  • Formation of Secondary phases.

[133]
Slow evaporation
  • Synthesis does not require external energy supply.

  • Synthesis can be done at room temperature.

  • Depends on solubility of reagents.

  • Slow reaction.

[134]
Diffusion
  • Performed at room temperature and do not need energy supply.

  • Employed particularly when the products exhibit low solubility.

  • Synthesis usually takes more than three days.

[135]
Template
  • Easy and simple method

  • Regulates the morphology, particle size, and structure during the particle preparation process.

  • Selectivity in template.

  • High cost.

[136]
Microemulsion
  • Size of nanoscale materials can be controlled.

  • High cost.

  • Most of the surfactants used are harmful to the environment.

[137]
Editorial Office
464 Cheongpa-ro, #726, Jung-gu, Seoul 04510, Republic of Korea
FAX : +82-2-383-9654   E-mail : eer@kosenv.or.kr

Copyright© Korean Society of Environmental Engineers.        Developed in M2PI
About |  Browse Articles |  Current Issue |  For Authors and Reviewers