Next Article in Journal
The Inhibitory Potential of 2′-dihalo Ribonucleotides against HCV: Molecular Docking, Molecular Simulations, MM-BPSA, and DFT Studies
Previous Article in Journal
Speculation on How RIC-3 and Other Chaperones Facilitate α7 Nicotinic Receptor Folding and Assembly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Chemistry and Applications of Metal–Organic Frameworks (MOFs) as Industrial Enzyme Immobilization Systems

by
Allison R. M. Silva
1,
Jeferson Y. N. H. Alexandre
1,
José E. S. Souza
1,
José G. Lima Neto
1,
Paulo G. de Sousa Júnior
2,
Maria V. P. Rocha
1,* and
José C. S. dos Santos
1,3,*
1
Departamento de Engenharia Química, Campus do Pici, Universidade Federal do Ceará, Bloco 709, Fortaleza 60455760, Brazil
2
Centro de Ciências, Departamento de Química Orgânica e Inorgânica, Campus do Pici, Universidade Federal do Ceará, Fortaleza 60455970, Brazil
3
Instituto de Engenharias e Desenvolvimento Sustentável, Universidade da Integração Internacional da Lusofonia Afro-Brasileira, Campus das Auroras, Redencao CEP 62790970, Brazil
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(14), 4529; https://doi.org/10.3390/molecules27144529
Submission received: 26 June 2022 / Revised: 12 July 2022 / Accepted: 13 July 2022 / Published: 15 July 2022
(This article belongs to the Section Applied Chemistry)

Abstract

:
Enzymatic biocatalysis is a sustainable technology. Enzymes are versatile and highly efficient biocatalysts, and have been widely employed due to their biodegradable nature. However, because the three-dimensional structure of these enzymes is predominantly maintained by weaker non-covalent interactions, external conditions, such as temperature and pH variations, as well as the presence of chemical compounds, can modify or even neutralize their biological activity. The enablement of this category of processes is the result of the several advances in the areas of molecular biology and biotechnology achieved over the past two decades. In this scenario, metal–organic frameworks (MOFs) are highlighted as efficient supports for enzyme immobilization. They can be used to ‘house’ a specific enzyme, providing it with protection from environmental influences. This review discusses MOFs as structures; emphasizes their synthesis strategies, properties, and applications; explores the existing methods of using immobilization processes of various enzymes; and lists their possible chemical modifications and combinations with other compounds to formulate the ideal supports for a given application.

Graphical Abstract

1. Introduction

Enzymes have been widely used as natural biocatalysts in the pharmaceutical, chemical, and food industries, in addition to their well-known applications in medicine and in effluent and solid-waste treatment systems [1,2,3,4,5,6,7,8]. This is mainly due to the diversity of reactions enabled by biocatalysts, as well as their high efficiency, specificity, and selectivity [9,10,11,12,13,14,15,16,17,18]. Furthermore, enzymes are biocompatible and biodegradable structures that can be derived from renewable resources [19,20,21,22,23,24,25,26,27,28,29]. Unlike conventional organic syntheses, in enzymatic biocatalysis, the reactions of multifunctional molecules are carried out without the need for previous activation or the use of temporary protection for functional groups, resulting in more economical processes and less waste generation [29,30,31,32,33,34].
However, there are clear hurdles to the use of free enzymes, such as degradation (or denaturation) at high temperatures, the need for strict pH control during reactions, their difficult recovery and reuse, high production costs, and their instability under unfavorable environmental conditions, all of which hinder a more widespread implementation across different industries [19,20,22,30,31,35,36,37,38,39]. A suitable approach used for overcoming these problems is the immobilization of enzymes onto insoluble or solid supports [40,41,42,43,44,45,46,47,48,49,50]. Making them insoluble improves their operational characteristics under adverse conditions, which, in turn, enables their employment in media under more extreme temperatures, under comprehensive pH ranges, and in the presence of organic solvents instead of water [50]. Immobilization also allows for higher product quality and lower processing costs [51].
Another benefit to immobilization is the more efficient handling of enzymes through solid matrices in comparison to liquid-phase counterparts, which facilitates the separation of final products and reduces their contamination [52]. Additionally, immobilized enzymes show very little to no allergenicity, high recoveries, and a reuse capacity, rendering processes more economical [29,53]. To increase the stability of the enzymes during storage and make them more resistant to operational conditions, several types of support for enzyme immobilization have been studied, including magnetic nanoparticles, sol-gels, mesoporous silica, and polymers [17,27,43,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72. Immobilization technologies can also prevent subunit dissociation, aggregation, autolysis, and proteolysis, apart from delivering more suitable reaction microenvironments [73,74,75,76].
However, some challenges have also been observed in these techniques, such as low loading efficiency and enzyme denaturation due to incompatible incorporation processes [77]. In addition, conventional supports can present irregular non-uniform structures, which can impair the activity of the immobilized enzymes [78,79,80]. Among the several materials that can be used as supports for immobilized enzymes, metal–organic frameworks (MOFs) can be highlighted. These are an emerging class of porous materials built from the self-assembly of certain organic ligands and metal ions or specific clusters [81,82,83,84]. Their use as immobilization supports has been encouraged due to their inherent unique properties, such as structural flexibility, adjustable pore size, large surface area, and the possibility of post-synthetic modifications, among others [85]. The scientific relevance of using MOFs as support for enzymes can be observed by the significant increase in the number of published articles on these materials (Figure 1).
Furthermore, as observed in Figure 1, it is possible to discuss possibilities not yet evaluated for MOF applications as enzyme supports. As the figure presents and compares the number of MOF-related publications over the past 10 years, it is clear to see that the research on the topic is being carried out at an increasing pace, and it is possible to identify a vast field of present and future possibilities.
MOFs’ flexible structure size and porous environment, as well as their network of binding and interaction sites, allow for the immobilization of most enzymes and facilitates the mass transfer of substrates and products [86]. These materials have the highest surface areas ever reported for this specific application and can deliver high immobilization efficiencies due to their vast number of functional sites and pores [87,88]. Furthermore, MOFs can behave similarly to enzymes due to their inherent catalytic groups [86]. It is important to highlight, however, that for an efficient immobilization to occur, the enzyme confinement method must be carefully chosen, as any structural modification of the enzymes can lead to a significant reduction in their catalytic activity. Moreover, in case the interaction between the enzyme and the MOF is weak, enzyme leaching may occur. Therefore, it is essential to also evaluate the governing support–enzyme interactions [83].
Several authors have published reviews that discuss biocatalysts composed of enzymes and MOFs, addressing the most common methods of synthesis and immobilization of these composites, as well as their several applications [86,89,90,91,92,93,94,95,96,97,98,99]. In this scenario, this review intends to update the discussion of MOFs as highly relevant materials for a wide range of applications, as well as to discuss their roles and mechanisms of action as supports for enzymatic immobilization, and the different combinations for the formation of enzyme–MOF composites to diverse ends, such as catalysis, medicine, and in biosensor manufacturing, among others.

2. Metal–Organic Structures: Synthesis Strategies

MOFs are classes of chemicals that contain metallic ions (or coordinated metals) and organic ligands in their structures [100,101,102]. Thus, a MOF can be distinguished by a coordinated network with organic ligands (that can be mono-, di-, or trivalent, or tetravalent) containing empty spaces, or ‘pores’ [100,103,104]. The metal–ligand chemical bonds present in the composition of MOFs are predominantly of covalent nature and of the Lewis acid–base type (metal ion and ligand, in that order), given that they can generate a coordination composition [103,104]. Thus, the choice of metal and ligands ultimately determines the structure and pore size in MOFs.
There are no MOFs readily available in nature, except for stepanovite and zhemchuzhnikovite minerals [105]. Thus, the low functional stability of these materials in natural environments with characteristics of high crystallinity, microporosity (partially), high permanent surface area, and low thermal and chemical stability, in addition to their porosity and density, can substantiate both the interest in this field and the need for further studies on the use of MOFs in different areas [103,105,106,107].
As mentioned above, the unique characteristics of these materials combined with their wide range of applications reinforce the need for the development and improvement of synthesis techniques [105,106,107]. Currently, MOFs can be synthesized by different strategies, such as reticular synthesis [108], hydrothermal (solvothermal) routes [109,110,111], diffusion [112,113,114], electrochemistry [111,115,116], microwaves [117,118], mechanochemistry [119,120], heating, and ultrasound [111]. Figure 2 shows a schematic representation of the main strategies for MOF synthesis.

2.1. Methods of Synthesis

2.1.1. Reticular Synthesis

Professor Omar Yaghi et al. [121] developed a synthesis strategy based on modular chemistry, known as reticular synthesis [121]. In this methodology, polytopic organic molecules bind to transition metal ions [121,122]. Subsequently, secondary building units (SBUs) are covalently linked across the entirety of the crystal [121,122,123]. SBUs are complexes or clusters in which ligand coordination modes and metal coordination geometries can be employed to modify these fragments into extended porous networks using polytopic structures [121,122,123].
In a later work, Yaghi et al. [108] discuss that reticular chemistry refers to the arrangement of pre-established coordinate structures through rigid molecular building blocks (the SBUs) that replicate and remain united through metal–ligand bonds [108]. Furthermore, in the same work, the authors redefined the term SBU, which was initially used to characterize fragments of zeolites, but was then defined around the geometry of the units classified as extension points [108].
In this way, for the construction of a broad network, SBUs must be structured in the correct mining, as this structuring guarantees the three-dimensionality of the material, so the geometry of the binder can directly influence the structure of the material [108,121].

2.1.2. Hydrothermal–Solvothermal Synthesis

Hydrothermal synthesis was initially used for the production of zeolites. Later, it was also incorporated in the synthesis of MOFs [111]. Jarrah and Farhadi (2019) used hydrothermal synthesis to synthesize a MIL-101(Cr) and P2W18@MIL-101(Cr) nanohybrid. The nanohybrid was used in an adsorption test with the following organic dyes: methylene blue, rhodamine B, and methyl orange. The results indicated that the material obtained showed fast selective adsorption for systems with different dye concentrations [124].
In this technique, soluble impellers are used in a reactor, where the system operates under high pressures and temperatures (Figure 2a) [111,125]. The hydrothermal and solvothermal methods employed are dependent on the solvent used. In general, processes that use water as a solvent are termed hydrothermal processes [124], while those that use other solvents are classified as solvothermal processes [126].
The main advantages of this method reside in the good control of the morphology and the composition of the MOF [111]. It is worth mentioning that the rate of cooling can influence the properties of the synthesized material [109]. The main disadvantages of these methods include the processing time and the operating costs, making it difficult to reproduce them on industrial scales [111,124,125,126].

2.1.3. Diffusion Synthesis

The synthesis method via diffusion is based on the gradual transport of several interacting species [112]. Diffusion-based methods of synthesis can be subdivided into two strategies [111]. In the first strategy, liquid solvent diffusion is performed [125]. First, two layers are formed at different density levels. The precipitating solvent resides in one of these layers, and then the final product in the solvent sits in the other [111]. This way, through the contact between the two interfaces, the gradual diffusion of the precipitant between the separated layer takes place, thus facilitating the crystal development (Figure 2b) [127].
In the second strategy, gradual diffusion occurs through physical barriers [128]. In addition, gels can also be used as an environment for diffusion and crystallization. This material is used because it mitigates the slow rate of diffusion and hinders the sole precipitation of MOFs [111,127,128].

2.1.4. Electrochemical Synthesis

Electrochemical synthesis is widely used on an industrial scale to produce MOFs [129]. This methodology is based on principles of green chemistry since, when compared to the solvothermal method, for example, it imparts low costs, operating at lower pressures and temperatures, and requires shorter synthesis times, while presenting higher selectivity [130]. It is worth mentioning that during the crystallization step, issues may occur due to the development of metal ions in situ near the surface of the support, which reduces the agglomeration of crystals [131,132]. Figure 2c shows a schematic representation of an electrochemical synthesis of MOFs.
As with the hydrothermal method, the cracking process is thermally induced during temperature decay. However, as mentioned above, electrochemical synthesis occurs at milder temperatures, as compared to the former technique. According to Mueller and co-workers [111], the less abrupt cooling may favor the process of MOF formation [111,130,131].
The main disadvantage of this synthesis method compared to the hydrothermal route is the need for controlling a more significant number of variables, since parameters such as voltage and pulse, for example, need to be carefully adjusted [130,131,133].

2.1.5. Microwave-Assisted Synthesis

The microwave-assisted technique is widely used for synthesizing small particles of oxides and metals [134]. Chen et al. [135], for example, performed the synthesis of MOF-74(Ni) with different methods, such as hydrothermal and microwave-assisted methods. The researchers evaluated the performance of these materials in the adsorption of CO2/N2 and verified that the MOF-74(Ni) synthesized by microwaves presented better adsorption performance. In addition, the authors reported that the protocol studied proved to be easy to conduct, and was also faster when compared to the other methods studied [135].
Through this process, it is also possible to increase the temperature of the solution, thus facilitating the formation of nanometric metal crystals [134,136]. It is worth mentioning that this strategy apparently cannot be directly used to synthesize MOF crystals [136]. However, it can speed up the synthesis process and adequately control the size and shape of MOFs [137]. Figure 2d presents a schematic representation of the use of microwaves in MOF synthesis.
Another aspect that needs to be considered is the control of parameters for solvent evaporation. Since temperature expansion can increase the solubility of crystals in saturated solutions, the process facilitates the formation of crystals during the cooling phase [134,136,137].

2.1.6. Mechanochemical Synthesis

The mechanophysical strategy employs mechanical forces as a precursor of chemical reactions (Figure 2e). In this type of synthesis, chemical transformation is preceded by the mechanical rupture of intermolecular bonds [138,139]. Synthetic chemistry has employed mechanical activation in multicomponent reactions (ternary and higher) to form co-crystals with applications in the fields of pharmacy, organic synthesis, inorganic solid-state chemistry, and polymer science, among others [140].
Thus, several reasons are highlighted for using this strategy in the synthesis of MOFs. The main advantage of this method is the redced possible environmental impacts caused by the process. Syntheses in the absence of organic solvents can be carried out at room temperature, for example. Another positive aspect is reduced synthesis time [112,138,139,140].

2.1.7. Sonochemical Synthesis

This methodology uses frequencies between 10 MHz and 20 KHz, which are higher than those detectable by the human ear (Figure 2f) [125]. The synthesis media can be close to a solid consistency if the cavitation and the microjets emitted during the reactions have the capacity for deterioration, activation, and interface variation [141], as well as for dispersion and agglomeration [142]. Alternatively, a liquid acts under pressure, specific temperature, and homogeneous conditions, or it is the interface that acts under the pressure of the medium, in case of forcing [141].
The main advantages of using sonochemical synthesis are the speed of synthesis, energy efficiency, process simplicity, and room-temperature reaction environments [111,138,141,142]. Yu et al. [143] employed the sonochemical route in the synthesis of Zn-based porphyrins MOF-525 and MOF-545. The authors obtained both porphyrins at high purity, and processing times were of 2.5 h and 0.5 h, respectively. It is worth noting that the materials showed excellent results also in the hydrolysis of dimethyl-4-nitrophenyl phosphate (DMNP) and in the adsorption of bisphenol-A (BPA), when compared to samples obtained conventionally [143]. Table 1 lists some methods of MOFs synthesis and their characteristics.

3. Metal–Organic Frameworks

3.1. Properties

Metal–organic frameworks (MOFs) have plated several roles in many industries and have become promising materials in the areas of catalysis, drug delivery, sensors, biological markers, pesticides, and others [153]. Their wide application is linked to their key physical properties and versatility, which are evidenced by organic structures linked to a central ion and, more specifically, a metallic cation [125]. The coordination sphere has a well-defined geometry, leading to the creation of crystals originating from this spatial arrangement, allowing pores to form in a polymerized manner. A scheme of the above definition is shown in Figure 3 [84].
Metal–organic frameworks show a wide variety of physicochemical and biological properties due to the versatility of their compositions (Figure 4) [154]. The binding of a metal ion or cluster to a flexible chain of organic polymers creates excellent magnetic properties in these composites that can be widely explored. This also facilitates the removal of these nanomaterials from their respective reaction media [99]. MOFs are also excellent precursors of chemical synthesis, depending on the chemical groups present in their organic part, where they act as activators or inhibitors of reaction points [155]. Additionally, they can act as electron donors or acceptors due to the properties of these structures being associated with coordination polymers, which behave as Lewis acids [156]. Many of these structures can interact with ionic or organic membranes and selectively migrate carrying ligands or macromolecules in biological media from one region to another [107]. The semiconduction properties of these materials also enable their use in the development of cutting-edge nanotechnology materials and processes. Owing to their excellent thermal capacity, new devices that require high sensitivity, easy detection and mapping, and good thermal stability can also be produced based on these inherent characteristics [157].
The ability of MOFs to act as catalysts or supports for the immobilization of biocatalysts renders these composites widely employable in chemical syntheses [158]. The chirality of these structures also enables favorable interactions with optically active materials, allowing for enhanced selectivity for these materials when in biological media [159]. Their thermal capacity, based on the metallic components, enables MOFs to integrate structures that require rapid cooling or heating [160]. Their semiconduction properties are also associated with the metallic center or the semiconducting organic ligands of these polymers, which allows for their use in nanotechnological applications [161].
Additionally, high porosity is one of the properties that add the most value to these materials, as pore sizes can be adjusted at the time of synthesis, according to the method and chemical precursors used [162]. The pores on the contact surface can act as housings to small molecules to be carried in fluids and organisms, and even to other organic molecules responsible for a given specific activity [163]. Luminescence, another key property, is characterized by the emission of light from the excited compounds. In MOFs, this is not only associated with the type of metal present in their composition but can be potentialized by organic ligands that present ideal chromophores for this property, such as aromatic structures [164,165].

3.2. Applications

3.2.1. Adsorption

Adsorption is a fairly easy and low-cost technique that has been widely used, among other ends, to remove aquatic contaminants (Figure 5) [162]. MOFs are materials that can be successfully used in this technique due to their good adsorbent properties. More specifically, they have been employed in the removal of excess biological compounds, antibiotics, pesticides, gases, and other toxic pollutants, such as heavy metals [163]. Pan et al. [166], and Ghanem et al. [167] reported the adsorption process of organophosphate compounds used as herbicides, glufosinate (GLUF), glyphosate (GLY), and bialaphos (BIA) via MOFs. When metabolized, these compounds form derivatives that are frequently found in underground water bodies and in the soil, and that cause several environmental problems. They are also difficult to remove due to their high solubility and polarity. The adsorption process described made use of the magnetic properties of these MOFs, their high structural porosity, available surface area, and the possibility of compounds being quickly bound to the metallic center [165]. Thus, this becomes a viable technique, both from an environmental perspective and from an economic point of view, since MOFs can be reused for many cycles.
Antibiotics are drugs used to treat human and animal infections and have become an emerging environmental problem due to their excessive and incorrect disposal [168]. These compounds can be removed from aquatic systems using the MOF adsorption method, as reported in [169]. In addition to aiding the elimination of the aquatic contamination, these materials could also be used to remove polluting gases from the atmosphere via gas adsorption [170]. Many other materials are already widely used for this purpose, such as activated carbon and zeolites. However, they have shown a reduced ability to adsorb carbon dioxide [171]. Thus, materials made from metal–organic frameworks are highly promising, given their properties of adjustable pore size, easy handling and application, reuse, and selectivity [172]. In recent years, this versatility has led to a great interest in MOF, resulting in the use of these materials for different purposes. When associated with simple techniques, such as adsorption, many new options can be enabled.

3.2.2. Catalysis

There has always been high demand for cheaper and faster processes in several industries. Therefore, the use of catalysts is widely studied for the optimization of industrial processes. MOFs, for example, can be used as catalysts for chemical reactions [173]. Given the aforementioned properties, they can provide high selectivity of substrates, and can be easily separated from reaction media and vastly reused (Figure 6). In the literature, several types of chemical reactions at small and large scales have been catalyzed by MOFs, including conventional catalysis [174,175], biocatalysis [173,176,177,178], and electrocatalysis [174,179]. The development and employment of these materials at industrial scales are significant, as they are excellent catalysts. However, it is still necessary to address the stability of MOFs under various reaction conditions, such as pH, temperature, and organic solvents, which has currently been a challenge for researchers.

3.2.3. Drug Delivery

The number of biomedical applications of structures based on MOFs has been growing throughout the years due to the excellent versatility of these materials, high porosity, and large available surface area [132]. One of these key applications is in drug loading, which allows MOFs to work as carriers of the active compounds of various drugs through the body, from small organic molecules to macromolecules, such as nucleic acids and proteins (Figure 7) [180]. One issue related to this application is the toxicity of MOFs and the materials’ lack of full biocompatibility with the organism [181]. One advantage is that, due to their high loading capacity, they can be monitored in the body, allowing for the mapping of the reaction mechanism of different drugs, especially in the development of new drugs.

3.2.4. Sensors

Biosensors are promising tools which can detect quick, selective, and sensitive molecules [182]. Due to the insulating characteristics of MOFs, they show great potential in the preparation of electrochemical sensors supported by carbon, which extends their application to the detection of analytes in different industrial fields, including environmental and biomedical fields, among others [183,184,185]. MOFs are great detectors of pollutants due to their affinity for specific groups of organic molecules [186]. Organic solvents, aromatic compounds, and heavy metal ions can also be detected using MOFs made from lanthanides [153,187].
Due to their adjustable pore size and high surface area, MOFs can also provide an ideal environment to accommodate analytes, allowing them to selectively absorb and release specific substrates through size recognition, effectively increasing signal and detection capabilities [188,189]. In addition, features such as the presence of metal coordination sites and lattice structures make them superior materials for the production of electrode coatings and for analyte detection [189]. Furthermore, there is the possibility of promoting the enhancement of their sensitivity to certain analytes through functionalization by immobilizing functional sites, initiating specific coordination, or promoting hydrogen bonding interactions with the target analyte [188].
MOF composites, formed by the incorporation of active biomolecules, such as antibodies, enzymes, and nucleic acids, can improve the selectivity, sensitivity, and detection limits of electrochemical sensors [190,191]. Biomolecule–MOF composites have been designed with an innovative focus on the detection of compounds of interest depending on the application sector. Some key compounds include uric acid [192,193], glucose [194], microRNAs [195], H2O2 [196], carcinoembryonic antigens [197], acetaminophen, and dopamine [193]. The main biomacromolecules are enzymes, as they provide more ecological, economical, and sustainable processes [29].
Enzymes can be incorporated into the structure of metal–organic structures and lead to the formation of sensitive electrochemiluminescence biosensors [88,198]. Examples include the manufacture of structures responsible for the detection of oncoproteins related to tumor proliferation (Figure 8), MOF enzymes of environmental interest [199], and other applications of industrial interest (such as the immobilization of enzymes for biocatalysis and the monitoring of biochemical reactions) [200].
Wang et al. [201] developed an enzymatic sensor for the photoelectrochemical detection of hypoxanthine using a nanoscale porphyrin MOF (Al-TCPP(Zn)) modified with the xanthine oxidase enzyme. Al-TCPP(Zn) exhibited an O2-dependent cathodic photocurrent, and this signal could be used for photoelectrochemical detection. After the addition of hypoxanthine, the produced biosensor delivered better responses due to the photoreduction of the H2O2 product catalyzed by xanthine oxidase. For the photoelectrochemical detection of hypoxanthine, the proposed sensor exhibited low detection limits, which was comparable to, or even better than, previous methods in terms of linear range and limits of determination; the selectivity was tested against several interferences, showing to have only been slightly affected. The authors also pointed out the reusability of the biosensor.
In Wang et al. [202], a glucose sensor for cascade biocatalysis constructed via the double confinement of enzymes in a nanocage-based zeolite imidazole (NC-ZIF) structure was evaluated. The enzyme@NC-ZIF showed good mass transport rates and excellent enzyme conformational versatility, due to the increased mesoporosity of the structure. The produced GOx/Hemin@NC-ZIF achieved good efficiency in catalytic cascade reactions in colorimetric and electrochemical glucose biosensors, enabling long-term quantitative analysis and continuous real-time monitoring of glucose in transpiration. Although the GOx/Hemin@NC-ZIF is very promising as a sensor, the method is limited to sweat tests, requiring further studies in order for other body fluids to be applied in innovative physiological and clinical investigations.

3.2.5. Hydrogen Storage

MOFs can store hydrogen due to the large available surface of these materials [203]. Their hybrid metallic and molecular composition allows for several adjustments, such as the functionalization of possible ligands and their storage under variable temperatures [204]. MOFs have also become very promising in replacing noble metals during hydrogenation syntheses as Pt, the most commonly used metal to this end, is expensive and, even when compared to MOFs, shows lower yields in hydrogen trapping [205]. Therefore, a straightforward application of these hybrid nanomaterials is indicated, as they possess pores that serve as “gas pockets”, holding hydrogen atoms for synthesis (Figure 9).

3.2.6. Environmental Applications

The environmental applications of metal–organic structures have been widely explored in recent years, as the growing drive to minimize the impacts of chemical residues has become the focus of extensive research around the world [206]. MOFs are used as efficient removers of heavy metals in fluids and aquatic environments [191]. They have been used to remove harmful gases and pollutants [207], such as carbon dioxide [208], based on their adsorption capacity [209]. Ma and colleagues [210] synthesized a MOF compound given its application as a biosensor of organophosphate pesticides, i.e., common pollutants in the agro-industry. These nanomaterials played a substantial role in the detection and removal of organic substances and solvents [211], organic dyes [212], antibiotics [213], volatile organic compounds [210] and other contaminants of industrial effluents [214].
Another essential environmental application is the detection of ammonia levels as a result of bioaccumulation, which has drawn the attention of environmentalists. Depending on concentration ranges, this can cause serious problems in aquatic food chains [215]. Thus, metal–organic structures are an excellent alternative for identifying levels of environmental pollutants [216] and in the treatment of effluents [217]. Their easy synthesis and high reuse rates render them particularly more accessible and targeted in the environmental area, which can be noted by the increase in the number of works published in recent years on this application [218].
All the applications discussed in this work present several possibilities of exploration in the industrial sector (Table 2). The flexible topology of these materials enables new architectures and, consequently, new properties and applications for MOFs, in addition to those that already exist and are extensively studied.
Thus, it is clear that nanomaterials have been widely used in different areas, which reinforces the need to develop, synthesize, and apply MOFs. A disadvantage of their use is still the high associated costs, with processes becoming economically unfeasible depending on their chemical composition, compared to other conventional structures. However, these nanoparticles are still very promising because such costs can potentially be counterbalanced by the number of possible reuses, the ease of synthesis, the wide range of applications, and the highly flexible structure for different processes. This is reinforced by a series of previously discussed properties, and those not yet tested in association with these materials, bringing the growing use of MOFs in complex industrial processes that benefit from the advancement of nanotechnology into perspective.

4. Enzyme Immobilization with Metal–Organic Frameworks (MOFs)

The immobilization of enzymes onto nanomaterials has revolutionized the use of these macromolecules in various industrial fields, which have been more recently enhanced by the advent of metal–organic frameworks [247]. The efficient immobilization of enzymes, i.e., its support and methods, is the result of perfect matching of factors depending on the enzyme [248]. Furthermore, the choice and success of the immobilization methods in the reaction depends of the different properties of the substrates and products, as well as the diversified applications of the products obtained. In addition, all methods have advantages and limitations. Consequently, the optimal immobilization conditions for a given enzyme are determined using experimental assays.
In addition to the main factors mentioned that influence the immobilization process, other parameters are important, such as pH, temperature, ionic strength, charge, and porosity of the support. These factors have a lesser or greater effect depending on the immobilization method. As previously mentioned, MOF characteristics of structural versatility, such as the porosity, large surface area, and organic–inorganic hybridity organization, render MOFs excellent candidates for enzyme immobilization using the most diverse methods (Figure 10) [93,99,247,249,250,251]. Regarding the porosity of the support, the mesoporous MOFs have been designed and constructed to obtain a high enzyme loading capacity and to reduce the diffusion resistance of reactants and products during the reaction. According to Xia et al. [93], the size of the pore openings may allow MOFs to gain size selectivity.
In Subtopics 4.1–4.4, immobilization studies using MOFs with different methods are presented.

4.1. In Situ Synthesis

In this method, the enzymes of interest and MOF materials (metal ions and organic ligands) are mixed under mild operating conditions in a suitable solution [93]. Using this immobilization technique, Wu, Yang, and Ge [252] assessed the stability behavior of some enzymes in organic solvents and compared these results with those obtained with the same proteins in their free form. To this end, lipase B from Candida antarctica, horseradish peroxidase, and cytochrome C were immobilized on the composite ZIF-8. The results showed that, even though the enzymes had different properties, the three immobilized macromolecules showed far superior stabilities in dimethyl formaldehyde, dimethyl sulfoxide, ethanol, and methanol compared to their free counterparts. Furthermore, the immobilized enzymes preserved almost 100% of their initial activity after incubation in the organic solvent, showing that the immobilization strategy protected them against potential denaturation due to the solvents used.
Another study considering MOF parameters was performed by Gascón et al. [253]. The researchers studied the synthesis and in situ strategies used to immobilize beta-glucosidase and laccase in nanocrystalline MOF platforms which aim to increase the activity of the tested enzymes. According to the results obtained, the immobilization stages in MOF nanocrystals favored the efficiency and the specific activity of the enzymes. Derivatives formed from in situ strategies showed an enzymatic charge above 85% and a loss of enzymatic activity of around 5%. Furthermore, the studied immobilization methodology effectively preserved the enzyme activity in a non-aqueous medium (N, N-dimethylformamide). Therefore, the researchers concluded that enzymes can be effectively immobilized in MOF nanocrystals and that in situ immobilization is a viable alternative in the preparation of immobilized biocatalysts.
Even though the in situ approach to immobilizing enzymes in MOF was efficiently conducted and requires mild reaction conditions, not all MOFs are ideal for this process. This is because the mode of enzyme dispersion and their subsequent location on the support can negatively affect the immobilization reactions [252].

4.2. Covalent Bonding

Unlike the in situ strategy, immobilization by covalent bonding occurs when the already-synthesized MOF is coated with substances capable of binding to the amino groups on the enzyme surface [254]. Many MOFs are susceptible to modification with functional groups to turn them into immobilization matrices [93].
Using this strategy, Cao and collaborators [255] immobilized soy epoxide hydrolase in UiO-66-NH2 MOF with glutaraldehyde as a binding agent, later applying this derivative in the biosynthesis of a (R)-1, 2-octanediol enantiomer. The results showed that the derivative presented a remarkable enzymatic load (87.3 mg/g), and recovered activity of 88%, as well as operational stabilities related to pH, temperature, and contact with organic solvents comparable to the frozen form of the enzyme under study. In addition to the improvements in the enzymatic characteristics associated with immobilization, the protein, when tested for the synthesis of (R)-1, 2-octanediol, delivered an enantiomeric excess of 81.2%. Therefore, the authors concluded that the immobilization of soy epoxide hydrolase on MOFs via covalent bonding showed strong potential for both improving enzyme characteristics and for being applied in enantiomeric reactions.
While seeking to further optimize the preparation and reuse of enzymes immobilized in MOFs, Wang et al. [251] incorporated iron oxide during MOF synthesis and used the final support to immobilize a Candida rugosa lipase via covalent bonding. The methodology employed by the researchers is justified by the ease of separating the derivative from a given reaction medium with the aid of a simple magnetic field. The derivative obtained was tested for the hydrolysis of olive oil and delivered a conversion rate of more than 65% after 6 h of reaction at 65 °C. Furthermore, the enzyme immobilized in the composite retained about 60% of its initial activity after 10 consecutive reaction cycles. Therefore, according to the above article, the synthesized support had both a large surface area and strong magnetic characteristics, which render this specific composite a good candidate support for enzyme immobilization.

4.3. Surface Immobilization

Surface immobilization (or adsorption) is the most widely used immobilization technique [254] due to the relatively low associated costs and the easy-to-perform methodology [93]. Because it is a versatile process, adsorption can be used to immobilize different enzymes on different supports, including MOFs [93,254,256,257]. In this technique, enzymes bind to the support through weak interactions such as van der Waals forces, hydrogen bonds, or electrostatic forces; therefore, they can be easily removed from the support via variations in pH and temperature, for example [93]. However, physical adsorption is still widely used and investigated due to its simplicity and the non-requirement for complex reagents [93,254].
In an attempt to compare advantages and disadvantages of this technique, Cao et al. [257] immobilized a lipase from Bacillus subtilis in a Cu-BTC-based MOF via physical adsorption and used the obtained derivative in an esterification reaction. The researchers obtained excellent results and demonstrated that the derivatives showed high operational stability and good enzymatic activity. Even after 10 consecutive reaction cycles, the lipase retained 90.7% of its initial activity and 99.6% of its initial conversion.
Another study on surface immobilization was performed by Pang and co-workers [256]. The researchers studied the support synthesis and the subsequent laccase immobilization on mesoporous Zr-MOF. According to the results, the laccase@Zr-MOF complex exhibited an adsorption capacity of 221.83 mg/g, wide temperature and pH distributions, and better stability when compared to that of the free laccase. In addition, the immobilized enzyme was able to maintain about 50% of its activity after 10 reaction cycles of contact between the derivative and ABTS, and retained 55.4% of its initial activity after three weeks of storage. With these numbers, the authors concluded that the immobilization method was successfully employed and that the synthesized support is a potential candidate for laccase immobilization via physical adsorption.

4.4. Entrapment

The immobilization strategy using entrapment or encapsulation is based on the confinement of the enzyme to a microenvironment located inside the support [93]. Contrary to other techniques, immobilization by entrapment causes isolation of the enzyme from the reaction medium, and also gives the protein better stability against potential denaturation caused by organic solvents, high temperatures, or sudden changes in pH [93,173]. Furthermore, using a MOF as support for this type of immobilization has extra advantages compared to other matrices: (i) MOFs can be synthesized according to their most suitable pore size (supports can have specific sizes for each type of substrate to allow for the efficient insertion and binding of the immobilized enzyme, reducing diffusional limitations); (ii) large enzyme loads can be achieved using MOFs as a consequence of their pore size; and (iii) encapsulated enzymes show a lower tendency to detach from the support [173].
Making use of such advantages, Li et al. [258] encapsulated a nerve agent detoxifying enzyme (organophosphorus acid hydrolase) in a mesoporous zirconium–MOF composite. The researchers reported that the synthesized support exhibited high enzyme loading capacity (12 wt%) and considerably improved thermal and storage stabilities.
In another study, Lian and co-workers [259] immobilized two enzymes in a tandem nanoreactor using a hierarchically structured MOF (PCN-888). The immobilized enzymes were glucose oxidase (GOx) and horseradish peroxidase (HRP). For the immobilization of both proteins to be successful, the researchers had to follow an encapsulation order: GOx followed by HRP. In the described process, the largest pores of the MOF (6.2 nm) were used to accommodate glucose oxidase, the 5.0 nm cavities accommodated horseradish peroxidase, and the smallest cavities (2.0 nm) remained unobstructed and accessible for the input of substrates and the output of products. Therefore, from the results, it was possible to conclude that the MOF was able to protect both enzymes against potential denaturation and considerably increased their operational stabilities (Table 3).

5. Future Trends

The application of MOFs combined with biocatalytic agents, including natural enzymes, is relatively recent. This integration has demonstrated an interesting synergistic performance in biocatalysis, due to the increased stability and reusability of encapsulated biocatalysts and the expansion of their applications into other fields [86,260]. Since the porosity properties of MOFs were identified, their investigation has developed exponentially [83]. However, although significant progress has been made, the investigation on enzyme–MOF composites is still in early stages, with many challenges still being a hurdle to the expansion of their applications [260]. The performance of this composites is influenced by several factors, including conformation; biomolecule activity and size; morphology; and the structural irregularity of particles in the design, preparation, and analysis of functionalized MOFs [82,86].
The use of MOFs for enzyme encapsulation is a fast developing field, and a significant increase in the number of studies on their properties in a short period of time leads us to believe that new highly effective biocatalysts are on the verge of being developed [261]. Great efforts have been made to this end; however, addressing the existing obstacles and improving current strategies are necessary so that enzyme–MOF composites can be fully suitable for practical applications [86,262]. There are expectations of future investigations in this area [260]. Challenges include the low diversity of biocompatible organic ligands and the toxicity of metals, in addition to the potential application of metals and ligands that have not yet been employed to this end [261,263].
To meet enzyme requirements of high activity and stability for practical applications and to elucidate the catalytic behavior of enzyme–MOF systems, it is necessary to investigate and improve the spatial structure of enzymes in MOFs [260,262]. This includes the establishment of spatial distributions that allow the confinement of multiple enzymes in MOFs, since the effective control over the location and orientation of enzymes can contribute to an increase in catalytic efficiency and a reduction in the resistance to the mass transfer of reagents [262]. In addition, exploring the suitable pore size and distribution profiles of MOFs is certainly an essential step in the encapsulation of several enzymes. Appropriate pore sizes can be optimized to meet specific criteria of enzyme accommodation, improving catalytic properties [260].
This review aimed to gather and discuss key information on MOFs, such as their synthesis, properties, and roles in enzyme encapsulation. We believe that the discussions, methodologies, and case studies presented can be helpful to readers and researchers interested in this topic. We also believe that this work can be used as a tool in the development of MOF-based materials for diverse applications, especially those related to enzymatic biocatalysis.

6. Conclusions

This review systematically reported on the mechanisms of action, latest advances, challenges, and future perspectives of the use of MOFs as support substrates in enzyme immobilization. MOFs are considered excellent candidates to support immobilization routes. This is because they present a wide variety of physicochemical and biological properties owing to the versatility of their composition. These impart properties include structural flexibility, adjustable pore size, large surface area, and the possibility of post-synthetic modifications, among others.
The chemistry of MOFs has developed exponentially since the porosity properties of these materials were identified. However, progress still needs to be made regarding the stability of MOFs under different reaction conditions (such as pH, temperature, and organic solvents), and in the storage of this material, constituting the most challenging aspects of their research. The elucidation of the different interactions between the MOF ‘housing’ and the enzymes that reside in their microenvironments during the various encapsulation processes is also paramount, since this can guide the construction of enzyme-MOF composites of high stability and bioactivity.
As the design and synthesis of MOFs with specific functionality at predetermined pore locations improve, interactions with biomolecules become more specific, resulting in more selective structures. Additionally, the recent methodologies and technologies based on computational chemistry can contribute to the development of new versatile projects of enzyme–MOF composites of high efficiency. However, to scale up laboratory-scale processes to larger scales, a more comprehensive understanding of the nature of enzyme–MOF composites is still required.
According to the discussion presented in this article, it can be concluded that enzymes immobilized on MOF supports clearly show better catalytic activity and operational stability than when compared to those obtained with their free form. In addition, such composites show an excellent maintenance of their initial activity after incubation in organic solvents by reaching a maximum percentage, which confirms that the immobilization strategies protect these proteins against possible solvent-related denaturation. Finally, it is expected that this review article, having presented synthesis strategies, properties, and applications of both MOFs and enzyme–MOF composites, can be a significant contribution to the advancement of the research on supports for enzymatic catalysis.

Author Contributions

Writing—original draft preparation, A.R.M.S., J.Y.N.H.A., J.E.S.S., J.G.L.N., P.G.d.S.J., M.V.P.R. and J.C.S.d.S.; writing—review and editing, M.V.P.R. and J.C.S.d.S. All authors have read and agreed to the published version of the manuscript.

Funding

Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, project number 311062/2019-9). Fundação Cearense de Apoio ao Desenvolvimento Científico e Tecnológico (FUNCAP, project number PS1-00186-00216.01.00/21).

Acknowledgments

We gratefully acknowledge the financial support of the following Brazilian Agencies for Scientific and Technological Development: Fundação Cearense de Apoio ao Desenvolvimento Científico e Tecnológico (FUNCAP), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and Coordenação de Aperfeiçoamento de Ensino Superior (CAPES) (finance code 001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rocha, T.G.; Pedro, P.H.; de Souza, M.C.M.; Monteiro, R.R.C.; dos Santos, J.C.S. Lipase Cocktail for Optimized Biodiesel Production of Free Fatty Acids from Residual Chicken Oil. Catal. Lett. 2021, 151, 1155–1166. [Google Scholar] [CrossRef]
  2. Monteiro, R.R.C.; Virgen-Ortiz, J.J.; Berenguer-Murcia, Á.; da Rocha, T.N.; dos Santos, J.C.S.; Alcántara, A.R.; Fernandez-Lafuente, R. Biotechnological Relevance of the Lipase A from Candida antarctica. Catal. Today 2021, 362, 141–154. [Google Scholar] [CrossRef]
  3. Monteiro, R.R.C.; Neto, D.M.A.; Fechine, P.B.A.; Lopes, A.A.S.; Gonçalves, L.R.B.; Dos Santos, J.C.S.; de Souza, M.C.M.; Fernandez-Lafuente, R. Ethyl Butyrate Synthesis Catalyzed by Lipases a and b from Candida antarctica Immobilized onto Magnetic Nanoparticles. Improvement of Biocatalysts’ Performance under Ultrasonic Irradiation. Int. J. Mol. Sci. 2019, 20, 5807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Cavalcante, F.T.T.d.A.; Falcão, I.R.; Souza, d.S.J.E.; Rocha, T.G.; de Sousa, I.G.; Cavalcante, A.L.G.; de Oliveira, A.L.B.; de Sousa, M.C.M.; dos Santos, J.C.S. Designing of Nanomaterials-Based Enzymatic Biosensors: Synthesis, Properties, and Applications. Electrochem 2021, 2, 149–184. [Google Scholar] [CrossRef]
  5. Moreira, K.S.; Moura, L.S.; Monteiro, R.R.C.; de Oliveira, A.L.B.; Valle, C.P.; Freire, T.M.; Fechine, P.B.A.; de Souza, M.C.M.; Fernandez-Lorente, G.; Guisan, J.M.; et al. Optimization of the Production of Enzymatic Biodiesel from Residual Babassu Oil (Orbignya Sp.) via RSM. Catalysts 2020, 10, 414. [Google Scholar] [CrossRef] [Green Version]
  6. Valério, R.B.R.; Cavalcante, A.L.G.; Mota, G.F.; de Sousa, I.G.; da Silva Souza, J.E.; Cavalcante, F.T.T.; da Silva Moreira, K.; de Aguiar Falcão, I.R.; Neto, F.S.; Dos Santos, J.C.S. Understanding the Biocatalytic Potential of Lipase from Rhizopus Chinensis. Biointerface Res. Appl. Chem. 2022, 12, 4230–4260. [Google Scholar] [CrossRef]
  7. Virgen-Ortíz, J.J.; dos Santos, J.C.S.; Ortiz, C.; Berenguer-Murcia, Á.; Barbosa, O.; Rodrigues, R.C.; Fernandez-Lafuente, R. Lecitase Ultra: A Phospholipase with Great Potential in Biocatalysis. Mol. Catal. 2019, 473, 110405. [Google Scholar] [CrossRef] [Green Version]
  8. Souza, J.E.S.; Monteiro, R.R.C.; Rocha, T.G.; Moreira, K.S.; Cavalcante, F.T.T.; de Sousa Braz, A.K.; de Souza, M.C.M.; dos Santos, J.C.S. Sonohydrolysis Using an Enzymatic Cocktail in the Preparation of Free Fatty Acid. 3 Biotech 2020, 10, 1–10. [Google Scholar] [CrossRef]
  9. Velasco-Lozano, S.; Rocha-Martin, J.; Santos, J.C.S. dos Editorial: Designing Carrier-Free Immobilized Enzymes for Biocatalysis. Front. Bioeng. Biotechnol. 2022, 10, 1–5. [Google Scholar] [CrossRef]
  10. Melo, A.D.Q.; Silva, F.F.M.; Dos Santos, J.C.S.; Fernández-Lafuente, R.; Lemos, T.L.G.; Dias Filho, F.A. Synthesis of Benzyl Acetate Catalyzed by Lipase Immobilized in Nontoxic Chitosan-Polyphosphate Beads. Molecules 2017, 22, 2165. [Google Scholar] [CrossRef] [Green Version]
  11. Lima, P.J.M.; da Silva, R.M.; Neto, C.A.C.G.; Gomes e Silva, N.C.; Souza, J.E.d.S.; Nunes, Y.L.; Sousa dos Santos, J.C. An Overview on the Conversion of Glycerol to Value-Added Industrial Products via Chemical and Biochemical Routes. Biotechnol. Appl. Biochem. 2021, 0, 1–25. [Google Scholar] [CrossRef] [PubMed]
  12. da Fonseca, A.M.; de Freitas, Í.B.; Soares, N.B.; de Araújo, F.A.M.; Gaieta, E.M.; Dos Santos, J.C.S.; Sobrinho, A.C.N.; Marinho, E.S.; Colares, R.P. Synthesis, Biological Activity, and in Silico Study of Bioesters Derived from Bixin by the Calb Enzyme. Biointerface Res. Appl. Chem. 2022, 12, 5901–5917. [Google Scholar] [CrossRef]
  13. Serpa, J.d.F.; Matias, G.A.B.; Fechine, P.B.A.; da Costa, V.M.; Freire, R.M.; Denardin, J.C.; Gonçalves, L.R.B.; de Macedo, A.C.; Rocha, M.V.P. New Nanocomposite Made of Cashew Apple Bagasse Lignin and Fe3O4 for Immobilizing of Lipase B from Candida antarctica Aiming at Esterification. J. Chem. Technol. Biotechnol. 2021, 96, 2472–2487. [Google Scholar] [CrossRef]
  14. Mota, G.F.; de Sousa, I.G.; de Oliveira, A.L.B.; Cavalcante, A.L.G.; da Silva Moreira, K.; Cavalcante, F.T.T.; da Silva Souza, J.E.; de Aguiar Falcão, Í.R.; Guimarães Rocha, T.; Bussons Rodrigues Valério, B.R.R.; et al. Biodiesel Production from Microalgae Using Lipase-Based Catalysts: Current Challenges and Prospects. Algal Res. 2022, 62, 102616. [Google Scholar] [CrossRef]
  15. Cavalcante, A.L.G.; Chaves, A.V.; Fechine, P.B.A.; Holanda Alexandre, J.Y.N.; Freire, T.M.; Davi, D.M.B.; Neto, F.S.; de Sousa, I.G.; da Silva Moreira, K.; de Oliveira, A.L.B.; et al. Chemical Modification of Clay Nanocomposites for the Improvement of the Catalytic Properties of Lipase A from Candida antarctica. Process Biochem. 2022, 120, 1–14. [Google Scholar] [CrossRef]
  16. Monteiro, R.R.C.; Arana-Peña, S.; da Rocha, T.N.; Miranda, L.P.; Berenguer-Murcia, Á.; Tardioli, P.W.; dos Santos, J.C.S.; Fernandez-Lafuente, R. Liquid Lipase Preparations Designed for Industrial Production of Biodiesel. Is It Really an Optimal Solution? Renew. Energy 2021, 164, 1566–1587. [Google Scholar] [CrossRef]
  17. Moreira, K.d.S.; de Oliveira, A.L.B.; de Moura, S.L.J.; de Sousa, I.G.; Cavalcante, L.G.A.; Neto, F.S.; Valério, R.B.R.; Chaves, V.A.; de Sousa, T.F.; Cruz, M.V.D.; et al. Taguchi Design-Assisted Co-Immobilization of Lipase A and B from Candida antarctica onto Chitosan: Characterization, Kinetic Resolution Application, and Docking Studies. Chem. Eng. Res. Des. 2022, 177, 223–244. [Google Scholar] [CrossRef]
  18. Souza, J.E.d.S.; de Oliveira, G.P.; Alexandre, J.Y.N.H.; Neto, J.G.L.; Sales, M.B.; Junior, P.G.d.S.; de Oliveira, A.L.B.; de Souza, M.C.M.; dos Santos, J.C.S. A Comprehensive Review on the Use of Metal–Organic Frameworks (MOFs) Coupled with Enzymes as Biosensors. Electrochem 2022, 3, 89–113. [Google Scholar] [CrossRef]
  19. Gao, Y.; Shah, K.; Kwok, I.; Wang, M.; Rome, L.H.; Mahendra, S. Immobilized Fungal Enzymes: Innovations and Potential Applications in Biodegradation and Biosynthesis. Biotechnol. Adv. 2022, 57, 107936. [Google Scholar] [CrossRef]
  20. Betancor, L.; López-Gallego, F. Cell–Enzyme Tandem Systems for Sustainable Chemistry. Curr. Opin. Green Sustain. Chem. 2022, 34, 100600. [Google Scholar] [CrossRef]
  21. Green, J.J.; Elisseeff, J.H. Mimicking Biological Functionality with Polymers for Biomedical Applications. Nature 2016, 540, 386–394. [Google Scholar] [CrossRef] [PubMed]
  22. Aggarwal, V.; Richardson, J. In Organic Solvents. Heteroat. Analog. Aldehydes Ketones 2004, 409, 1. [Google Scholar] [CrossRef]
  23. Lin, C.; Xu, K.; Zheng, R.; Zheng, Y. Immobilization of Amidase into a Magnetic Hierarchically Porous Metal-Organic Framework for Efficient Biocatalysis. Chem. Commun. 2019, 55, 5697–5700. [Google Scholar] [CrossRef]
  24. Cui, J.D.; Jia, S.R. Optimization Protocols and Improved Strategies of Cross-Linked Enzyme Aggregates Technology: Current Development and Future Challenges. Crit. Rev. Biotechnol. 2015, 35, 15–28. [Google Scholar] [CrossRef] [PubMed]
  25. Li, J.; Wang, J.; Guo, X.; Zheng, Q.; Peng, J.; Tang, H.; Yao, S. Carbon Nanotubes Labeled with Aptamer and Horseradish Peroxidase as a Probe for Highly Sensitive Protein Biosensing by Postelectropolymerization of Insoluble Precipitates on Electrodes. Anal. Chem. 2015, 87, 7610–7617. [Google Scholar] [CrossRef] [PubMed]
  26. Schmid, A.; Doderick, J.S.; Hauer, B.; Kiener, A.; Wubbolts, M.; Witholt, B. Insight Review Articles Expansion Phase. Nature 2001, 409, 258–268. [Google Scholar] [CrossRef] [PubMed]
  27. Hartmann, M.; Jung, D. Biocatalysis with Enzymes Immobilized on Mesoporous Hosts: The Status Quo and Future Trends. J. Mater. Chem. 2010, 20, 844–857. [Google Scholar] [CrossRef]
  28. Feng, W.; Ji, P. Enzymes Immobilized on Carbon Nanotubes. Biotechnol. Adv. 2011, 29, 889–895. [Google Scholar] [CrossRef]
  29. Sheldon, R.A.; van Pelt, S. Enzyme Immobilisation in Biocatalysis: Why, What and How. Chem. Soc. Rev. 2013, 42, 6223–6235. [Google Scholar] [CrossRef] [Green Version]
  30. Kondrat, S.; Krauss, U.; von Lieres, E. Enzyme Co-Localisation: Mechanisms and Benefits. Curr. Res. Chem. Biol. 2022, 2, 100031. [Google Scholar] [CrossRef]
  31. Patti, A.; Sanfilippo, C. Stereoselective Promiscuous Reactions Catalyzed by Lipases. Int. J. Mol. Sci. 2022, 23, 2675. [Google Scholar] [CrossRef] [PubMed]
  32. Clouthier, C.M.; Pelletier, J.N. Expanding the Organic Toolbox: A Guide to Integrating Biocatalysis in Synthesis. Chem. Soc. Rev. 2012, 41, 1585–1605. [Google Scholar] [CrossRef]
  33. Muñoz Solano, D.; Hoyos, P.; Hernáiz, M.J.; Alcántara, A.R.; Sánchez-Montero, J.M. Industrial Biotransformations in the Synthesis of Building Blocks Leading to Enantiopure Drugs. Bioresour. Technol. 2012, 115, 196–207. [Google Scholar] [CrossRef] [PubMed]
  34. Wohlgemuth, R. Biocatalysis-Key to Sustainable Industrial Chemistry. Curr. Opin. Biotechnol. 2010, 21, 713–724. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, Z.; Hartmann, M. Progress in Enzyme Immobilization in Ordered Mesoporous Materials and Related Applications. Chem. Soc. Rev. 2013, 42, 3894–3912. [Google Scholar] [CrossRef] [PubMed]
  36. Franssen, M.C.R.; Steunenberg, P.; Scott, E.L.; Zuilhof, H.; Sanders, J.P.M. Immobilised Enzymes in Biorenewables Production. Chem. Soc. Rev. 2013, 42, 6491–6533. [Google Scholar] [CrossRef] [Green Version]
  37. Cantone, S.; Hanefeld, U.; Basso, A. Biocatalysis in Non-Conventional Media—Ionic Liquids, Supercritical Fluids and the Gas Phase. Green Chem. 2007, 9, 954–997. [Google Scholar] [CrossRef]
  38. Gao, C.; Zhu, H.; Chen, J.; Qiu, H. Facile Synthesis of Enzyme Functional Metal-Organic Framework for Colorimetric Detecting H2O2 and Ascorbic Acid. Chin. Chem. Lett. 2017, 28, 1006–1012. [Google Scholar] [CrossRef]
  39. Xu, C.P.; Yun, J.W. Influence of Aeration on the Production and the Quality of the Exopolysaccharides from Paecilomyces Tenuipes C240 in a Stirred-Tank Fermenter. Enzym. Microb. Technol. 2004, 35, 33–39. [Google Scholar] [CrossRef]
  40. Pinheiro, M.P.; Monteiro, R.R.C.; Silva, F.F.M.; Lemos, T.L.G.; Fernandez-Lafuente, R.; Gonçalves, L.R.B.; dos Santos, J.C.S. Modulation of Lecitase Properties via Immobilization on Differently Activated Immobead-350: Stabilization and Inversion of Enantiospecificity. Process Biochem. 2019, 87, 128–137. [Google Scholar] [CrossRef]
  41. Cavalcante, F.T.T.; Cavalcante, A.L.G.; de Sousa, I.G.; Neto, F.S.; Dos Santos, J.C.S. Current Status and Future Perspectives of Supports and Protocols for Enzyme Immobilization. Catalysts 2021, 11, 1222. [Google Scholar] [CrossRef]
  42. Monteiro, R.R.C.; Dos Santos, J.C.S.; Alcántara, A.R.; Fernandez-Lafuente, R. Enzyme-Coated Micro-Crystals: An Almost Forgotten but Very Simple and Elegant Immobilization Strategy. Catalysts 2020, 10, 891. [Google Scholar] [CrossRef]
  43. Ismail, A.R.; Baek, K.H. Lipase Immobilization with Support Materials, Preparation Techniques, and Applications: Present and Future Aspects. Int. J. Biol. Macromol. 2020, 163, 1624–1639. [Google Scholar] [CrossRef] [PubMed]
  44. Bezerra, R.M.; Monteiro, R.R.C.; Neto, D.M.A.; da Silva, F.F.M.; de Paula, R.C.M.; de Lemos, T.L.G.; Fechine, P.B.A.; Correa, M.A.; Bohn, F.; Gonçalves, L.R.B.; et al. A New Heterofunctional Support for Enzyme Immobilization: PEI Functionalized Fe3O4 MNPs Activated with Divinyl Sulfone. Application in the Immobilization of Lipase from Thermomyces Lanuginosus. Enzym. Microb. Technol. 2020, 138, 109560. [Google Scholar] [CrossRef]
  45. Rueda, N.; Dos Santos, J.C.S.; Torres, R.; Ortiz, C.; Barbosa, O.; Fernandez-Lafuente, R. Immobilization of Lipases on Heterofunctional Octyl-Glyoxyl Agarose Supports Improved Stability and Prevention of the Enzyme Desorption. Methods Enzymol. 2016, 571, 73–85. [Google Scholar]
  46. Husain, Q. Nanomaterials as Novel Supports for the Immobilization of Amylolytic Enzymes and Their Applications: A Review. Biocatalysis 2017, 3, 37–53. [Google Scholar] [CrossRef]
  47. Li, P.; Modica, J.A.; Howarth, A.J.; Vargas, L.E.; Moghadam, P.Z.; Snurr, R.Q.; Mrksich, M.; Hupp, J.T.; Farha, O.K. Toward Design Rules for Enzyme Immobilization in Hierarchical Mesoporous Metal-Organic Frameworks. Chem 2016, 1, 154–169. [Google Scholar] [CrossRef] [Green Version]
  48. Das, R.; Mishra, H.; Srivastava, A.; Kayastha, A.M. Covalent Immobilization of Β-Amylase onto Functionalized Molybdenum Sulfide Nanosheets, Its Kinetics and Stability Studies: A Gateway to Boost Enzyme Application. Chem. Eng. J. 2017, 328, 215–227. [Google Scholar] [CrossRef]
  49. Ding, C.; Sun, H.; Ren, J.; Qu, X. Immobilization of Enzyme on Chiral Polyelectrolyte Surface. Anal. Chim. Acta 2017, 952, 88–95. [Google Scholar] [CrossRef]
  50. Bilal, M.; Asgher, M.; Cheng, H.; Yan, Y.; Iqbal, H.M.N. Multi-Point Enzyme Immobilization, Surface Chemistry, and Novel Platforms: A Paradigm Shift in Biocatalyst Design. Crit. Rev. Biotechnol. 2019, 39, 202–219. [Google Scholar] [CrossRef]
  51. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of Enzyme Activity, Stability and Selectivity via Immobilization Techniques. Enzym. Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  52. Reis, C.L.B.; de Sousa, E.Y.A.; de França Serpa, J.; Oliveira, R.C.; Dos Santos, J.C.S. Design of Immobilized Enzyme Biocatalysts: Drawbacks and Opportunities. Quim. Nova 2019, 42, 768–783. [Google Scholar] [CrossRef]
  53. Bonazza, H.L.; Manzo, R.M.; dos Santos, J.C.S.; Mammarella, E.J. Operational and Thermal Stability Analysis of Thermomyces Lanuginosus Lipase Covalently Immobilized onto Modified Chitosan Supports. Appl. Biochem. Biotechnol. 2018, 184, 182–196. [Google Scholar] [CrossRef] [PubMed]
  54. da Fonseca, A.M.; Cleiton Sousa dos Santos, J.; de Souza, M.C.M.; de Oliveira, M.M.; Colares, R.P.; de Lemos, T.L.G.; Braz-Filho, R. The Use of New Hydrogel Microcapsules in Coconut Juice as Biocatalyst System for the Reaction of Quinine. Ind. Crops Prod. 2020, 145, 111890. [Google Scholar] [CrossRef]
  55. Dos Santos, J.C.S.; Garcia-Galan, C.; Rodrigues, R.C.; De Sant’Ana, H.B.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Stabilizing Hyperactivated Lecitase Structures through Physical Treatment with Ionic Polymers. Process Biochem. 2014, 49, 1511–1515. [Google Scholar] [CrossRef]
  56. de Souza, T.C.; de Sousa Fonseca, T.; de Sousa Silva, J.; Lima, P.J.M.; Neto, C.A.C.G.; Monteiro, R.R.C.; Rocha, M.V.P.; de Mattos, M.C.; dos Santos, J.C.S.; Gonçalves, L.R.B. Modulation of Lipase B from Candida antarctica Properties via Covalent Immobilization on Eco-Friendly Support for Enzymatic Kinetic Resolution of Rac-Indanyl Acetate. Bioprocess Biosyst. Eng. 2020, 43, 2253–2268. [Google Scholar] [CrossRef] [PubMed]
  57. de Oliveira, A.L.B.; Cavalcante, F.T.T.; Moreira, K.S.; Monteiro, R.R.C.; Rocha, T.G.; Souza, J.E.S.; da Fonseca, A.M.; Lopes, A.A.S.; Guimarães, A.P.; de Lima, R.K.C.; et al. Lipases Immobilized onto Nanomaterials as Biocatalysts in Biodiesel Production: Scientific Context, Challenges, and Opportunities. Rev. Virtual Quim. 2021, 13, 875–891. [Google Scholar] [CrossRef]
  58. Cavalcante, A.L.G.; Cavalcante, C.G.; Colares, R.P.; Ferreira, D.A.; da Silva, F.F.M.; de Sousa, E.Y.A.; da Silva Souza, J.E.; de Castro Monteiro, R.R.; de Oliveira, A.L.B.; dos Santos, J.C.S.; et al. Preparation, Characterization, and Enantioselectivity of Polyacrylate Microcapsules Entrapping Ananas Comosus Extract. Rev. Virtual Quim. 2021, 13, 1319–1329. [Google Scholar] [CrossRef]
  59. Carneiro, E.; Bastos, A.; De Oliveira, U.; De Matos, L.; Adriano, W.; Monteiro, R.; Dos Santos, J.; Gonçalves, L. Improving the Catalytic Features of the Lipase from Rhizomucor Miehei Immobilized on Chitosan-Based Hybrid Matrices by Altering the Chemical Activation Conditions. Quim. Nova 2020, 43, 1234–1239. [Google Scholar] [CrossRef]
  60. Rios, N.S.; Morais, E.G.; dos Santos Galvão, W.; Andrade Neto, D.M.; dos Santos, J.C.S.; Bohn, F.; Correa, M.A.; Fechine, P.B.A.; Fernandez-Lafuente, R.; Gonçalves, L.R.B. Further Stabilization of Lipase from Pseudomonas Fluorescens Immobilized on Octyl Coated Nanoparticles via Chemical Modification with Bifunctional Agents. Int. J. Biol. Macromol. 2019, 141, 313–324. [Google Scholar] [CrossRef]
  61. Talekar, S.; Joshi, A.; Kambale, S.; Jadhav, S.; Nadar, S.; Ladole, M. A Tri-Enzyme Magnetic Nanobiocatalyst with One Pot Starch Hydrolytic Activity. Chem. Eng. J. 2017, 325, 80–90. [Google Scholar] [CrossRef]
  62. Cui, J.; Feng, Y.; Yue, S.; Zhao, Y.; Li, L.; Liu, R.; Lin, T. Magnetic Mesoporous Enzyme-Silica Composites with High Activity and Enhanced Stability. J. Chem. Technol. Biotechnol. 2016, 91, 1905–1913. [Google Scholar] [CrossRef]
  63. Cui, J.; Feng, Y.; Jia, S. Silica Encapsulated Catalase@metal-Organic Framework Composite: A Highly Stable and Recyclable Biocatalyst. Chem. Eng. J. 2018, 351, 506–514. [Google Scholar] [CrossRef]
  64. Benucci, I.; Lombardelli, C.; Cacciotti, I.; Liburdi, K.; Nanni, F.; Esti, M. Chitosan Beads from Microbial and Animal Sources as Enzyme Supports for Wine Application. Food Hydrocoll. 2016, 61, 191–200. [Google Scholar] [CrossRef]
  65. Lee, K.Y.; Yuk, S.H. Polymeric Protein Delivery Systems. Prog. Polym. Sci. 2007, 32, 669–697. [Google Scholar] [CrossRef]
  66. Lee, E.S.; Kwon, M.J.; Lee, H.; Kim, J.J. Stabilization of Protein Encapsulated in Poly(Lactide-Co-Glycolide) Microspheres by Novel Viscous S/W/O/W Method. Int. J. Pharm. 2007, 331, 27–37. [Google Scholar] [CrossRef] [PubMed]
  67. Khoshnevisan, K.; Vakhshiteh, F.; Barkhi, M.; Baharifar, H.; Poor-Akbar, E.; Zari, N.; Stamatis, H.; Bordbar, A.K. Immobilization of Cellulase Enzyme onto Magnetic Nanoparticles: Applications and Recent Advances. Mol. Catal. 2017, 442, 66–73. [Google Scholar] [CrossRef]
  68. Hong, S.G.; Kim, B.C.; Na, H.B.; Lee, J.; Youn, J.; Chung, S.W.; Lee, C.W.; Lee, B.; Kim, H.S.; Hsiao, E.; et al. Single Enzyme Nanoparticles Armored by a Thin Silicate Network: Single Enzyme Caged Nanoparticles. Chem. Eng. J. 2017, 322, 510–515. [Google Scholar] [CrossRef]
  69. Fried, D.I.; Brieler, F.J.; Fröba, M. Designing Inorganic Porous Materials for Enzyme Adsorption and Applications in Biocatalysis. ChemCatChem 2013, 5, 862–884. [Google Scholar] [CrossRef]
  70. Kato, K.; Kawachi, Y.; Nakamura, H. Silica-Enzyme-Ionic Liquid Composites for Improved Enzymatic Activity. J. Asian Ceram. Soc. 2014, 2, 33–40. [Google Scholar] [CrossRef] [Green Version]
  71. Iyer, P.V.; Ananthanarayan, L. Enzyme Stability and Stabilization-Aqueous and Non-Aqueous Environment. Process Biochem. 2008, 43, 1019–1032. [Google Scholar] [CrossRef]
  72. Hudson, S.; Cooney, J.; Magner, E. Proteins in Mesoporous Silicates. Angew. Chem. Int. Ed. 2008, 47, 8582–8594. [Google Scholar] [CrossRef] [PubMed]
  73. Garcia-Galan, C.; Barbosa, O.; Hernandez, K.; Dos Santos, J.C.S.; Rodrigues, R.C.; Fernandez-Lafuente, R. Evaluation of Styrene-Divinylbenzene Beads as a Support to Immobilize Lipases. Molecules 2014, 19, 7629–7645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Moreira, K.d.S.; de Oliveira, A.L.B.; Júnior, L.S.d.M.; Monteiro, R.R.C.; da Rocha, T.N.; Menezes, F.L.; Fechine, L.M.U.D.; Denardin, J.C.; Michea, S.; Freire, R.M.; et al. Lipase From Rhizomucor Miehei Immobilized on Magnetic Nanoparticles: Performance in Fatty Acid Ethyl Ester (FAEE) Optimized Production by the Taguchi Method. Front. Bioeng. Biotechnol. 2020, 8, 1–17. [Google Scholar] [CrossRef] [PubMed]
  75. Monteiro, R.R.C.; Lima, P.J.M.; Pinheiro, B.B.; Freire, T.M.; Dutra, L.M.U.; Fechine, P.B.A.; Gonçalves, L.R.B.; de Souza, M.C.M.; Dos Santos, J.C.S.; Fernandez-Lafuente, R. Immobilization of Lipase a from Candida antarctica onto Chitosan-Coated Magnetic Nanoparticles. Int. J. Mol. Sci. 2019, 20, 4018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Feng, Y.; Zhong, L.; Jia, S.; Cui, J. Acid-Resistant Enzyme@MOF Nanocomposites with Mesoporous Silica Shells for Enzymatic Applications in Acidic Environments. J. Biotechnol. 2019, 306, 54–61. [Google Scholar] [CrossRef]
  77. Sun, Q.; Pan, Y.; Wang, X.; Li, H.; Farmakes, J.; Aguila, B.; Yang, Z.; Ma, S. Mapping out the Degree of Freedom of Hosted Enzymes in Confined Spatial Environments. Chem 2019, 5, 3184–3195. [Google Scholar] [CrossRef]
  78. Nunes, Y.L.; de Menezes, F.L.; de Sousa, I.G.; Cavalcante, A.L.G.; Cavalcante, F.T.T.; da Silva Moreira, K.; de Oliveira, A.L.B.; Mota, G.F.; da Silva Souza, J.E.; de Aguiar Falcão, I.R.; et al. Chemical and Physical Chitosan Modification for Designing Enzymatic Industrial Biocatalysts: How to Choose the Best Strategy? Int. J. Biol. Macromol. 2021, 181, 1124–1170. [Google Scholar] [CrossRef]
  79. Srbová, J.; Slováková, M.; Křípalová, Z.; Žárská, M.; Špačková, M.; Stránská, D.; Bílková, Z. Covalent Biofunctionalization of Chitosan Nanofibers with Trypsin for High Enzyme Stability. React. Funct. Polym. 2016, 104, 38–44. [Google Scholar] [CrossRef] [Green Version]
  80. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [Green Version]
  81. Dhaka, S.; Kumar, R.; Deep, A.; Kurade, M.B.; Ji, S.W.; Jeon, B.H. Metal–Organic Frameworks (MOFs) for the Removal of Emerging Contaminants from Aquatic Environments. Coord. Chem. Rev. 2019, 380, 330–352. [Google Scholar] [CrossRef]
  82. Liang, J.; Huang, Y.B.; Cao, R. Metal–Organic Frameworks and Porous Organic Polymers for Sustainable Fixation of Carbon Dioxide into Cyclic Carbonates. Coord. Chem. Rev. 2019, 378, 32–65. [Google Scholar] [CrossRef]
  83. Marsh, C.; Shearer, G.C.; Knight, B.T.; Paul-Taylor, J.; Burrows, A.D. Supramolecular Aspects of Biomolecule Interactions in Metal–Organic Frameworks. Coord. Chem. Rev. 2021, 439, 213928. [Google Scholar] [CrossRef]
  84. Kitagawa, S.; Kitaura, R.; Noro, S.I. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, Y.; Yan, B. A Point-of-Care Diagnostics Logic Detector Based on Glucose Oxidase Immobilized Lanthanide Functionalized Metal-Organic Frameworks. Nanoscale 2019, 11, 22946–22953. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, J.; Liang, J.; Xue, J.; Liang, K. Metal–Organic Frameworks as a Versatile Materials Platform for Unlocking New Potentials in Biocatalysis. Small 2021, 17, 1–21. [Google Scholar] [CrossRef]
  87. Birhanlı, E.; Noma, S.A.A.; Boran, F.; Ulu, A.; Yeşilada, Ö.; Ateş, B. Design of Laccase–Metal–Organic Framework Hybrid Constructs for Biocatalytic Removal of Textile Dyes. Chemosphere 2022, 292, 133382. [Google Scholar] [CrossRef] [PubMed]
  88. Drout, R.J.; Robison, L.; Farha, O.K. Catalytic Applications of Enzymes Encapsulated in Metal–Organic Frameworks. Coord. Chem. Rev. 2019, 381, 151–160. [Google Scholar] [CrossRef]
  89. Pérez Gascón, V.; Sánchez-Sánchez, M. Environmentally friendly enzyme immobilization on MOF materials. In Immobilization of Enzymes and Cells: Methods and Protocols; Guisan, J.M., Bolivar, J.M., López-Gallego, F., Rocha-Martín, J., Eds.; Humana: New York, NY, USA, 2020; pp. 271–296. ISBN 978-1-0716-0215-7. [Google Scholar]
  90. Sun, H.; Li, Y.; Yu, S.; Liu, J. Metal-Organic Frameworks (MOFs) for Biopreservation: From Biomacromolecules, Living Organisms to Biological Devices. Nano Today 2020, 35, 100985. [Google Scholar] [CrossRef]
  91. Huang, S.; Kou, X.; Shen, J.; Chen, G.; Ouyang, G. “Armor-Plating” Enzymes with Metal–Organic Frameworks (MOFs). Angew. Chem. Int. Ed. 2020, 59, 8786–8798. [Google Scholar] [CrossRef]
  92. Cong, W.J.; Nanda, S.; Li, H.; Fang, Z.; Dalai, A.K.; Kozinski, J.A. Metal-Organic Framework-Based Functional Catalytic Materials for Biodiesel Production: A Review. Green Chem. 2021, 23, 2595–2618. [Google Scholar] [CrossRef]
  93. Xia, H.; Li, N.; Zhong, X.; Jiang, Y. Metal-Organic Frameworks: A Potential Platform for Enzyme Immobilization and Related Applications. Front. Bioeng. Biotechnol. 2020, 8, 1–16. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, H.S. Metal–Organic Frameworks for Biosensing and Bioimaging Applications. Coord. Chem. Rev. 2017, 349, 139–155. [Google Scholar] [CrossRef]
  95. Liang, S.; Wu, X.L.; Xiong, J.; Zong, M.H.; Lou, W.Y. Metal-Organic Frameworks as Novel Matrices for Efficient Enzyme Immobilization: An Update Review. Coord. Chem. Rev. 2020, 406, 213149. [Google Scholar] [CrossRef]
  96. Gkaniatsou, E.; Sicard, C.; Ricoux, R.; Mahy, J.P.; Steunou, N.; Serre, C. Metal-Organic Frameworks: A Novel Host Platform for Enzymatic Catalysis and Detection. Mater. Horiz. 2017, 4, 55–63. [Google Scholar] [CrossRef]
  97. Feng, Y.; Xu, Y.; Liu, S.; Wu, D.; Su, Z.; Chen, G.; Liu, J.; Li, G. Recent Advances in Enzyme Immobilization Based on Novel Porous Framework Materials and Its Applications in Biosensing. Coord. Chem. Rev. 2022, 459, 214414. [Google Scholar] [CrossRef]
  98. Du, Y.; Jia, X.; Zhong, L.; Jiao, Y.; Zhang, Z.; Wang, Z.; Feng, Y.; Bilal, M.; Cui, J.; Jia, S. Metal-Organic Frameworks with Different Dimensionalities: An Ideal Host Platform for Enzyme@MOF Composites. Coord. Chem. Rev. 2022, 454, 214327. [Google Scholar] [CrossRef]
  99. Lian, X.; Fang, Y.; Joseph, E.; Wang, Q.; Li, J.; Banerjee, S.; Lollar, C.; Wang, X.; Zhou, H.C. Enzyme-MOF (Metal-Organic Framework) Composites. Chem. Soc. Rev. 2017, 46, 3386–3401. [Google Scholar] [CrossRef]
  100. Batten, S.R.; Champness, N.R.; Chen, X.M.; Garcia-Martinez, J.; Kitagawa, S.; Öhrström, L.; O’Keeffe, M.; Suh, M.P.; Reedijk, J. Coordination Polymers, Metal-Organic Frameworks and the Need for Terminology Guidelines. CrystEngComm 2012, 14, 3001–3004. [Google Scholar] [CrossRef] [Green Version]
  101. Thorarinsdottir, A.E.; Harris, T.D. Metal-Organic Framework Magnets. Chem. Rev. 2020, 120, 8716–8789. [Google Scholar] [CrossRef]
  102. Wang, S.; McGuirk, C.M.; d’Aquino, A.; Mason, J.A.; Mirkin, C.A. Metal–Organic Framework Nanoparticles. Adv. Mater. 2018, 30, 1–14. [Google Scholar] [CrossRef]
  103. Kumar, P.; Vellingiri, K.; Kim, K.H.; Brown, R.J.C.; Manos, M.J. Modern Progress in Metal-Organic Frameworks and Their Composites for Diverse Applications. Microporous Mesoporous Mater. 2017, 253, 251–265. [Google Scholar] [CrossRef]
  104. Zhu, Q.L.; Xu, Q. Metal-Organic Framework Composites. Chem. Soc. Rev. 2014, 43, 5468–5512. [Google Scholar] [CrossRef]
  105. Huskić, I.; Pekov, I.V.; Krivovichev, S.V.; Friščić, T. Minerals with Metal-Organic Framework Structures. Sci. Adv. 2016, 2, 1–8. [Google Scholar] [CrossRef] [Green Version]
  106. Doherty, C.M.; Buso, D.; Hill, A.J.; Furukawa, S.; Kitagawa, S.; Falcaro, P. Using Functional Nano- and Microparticles for the Preparation of Metal-Organic Framework Composites with Novel Properties. Acc. Chem. Res. 2014, 47, 396–405. [Google Scholar] [CrossRef]
  107. Coudert, F.X.; Fuchs, A.H. Computational Characterization and Prediction of Metal-Organic Framework Properties. Coord. Chem. Rev. 2016, 307, 211–236. [Google Scholar] [CrossRef] [Green Version]
  108. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular Synthesis and the Design of New Materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef]
  109. Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.; Férey, G. Very Large Breathing Effect in the First Nanoporous Chromium(III)-Based Solids: MIL-53 or CrIII(OH)·{O2C-C6H4- CO2}·{HO2C-C6H4 -CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124, 13519–13526. [Google Scholar] [CrossRef]
  110. 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 2015, 3, 732–738. [Google Scholar] [CrossRef]
  111. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastré, J. Metal-Organic Frameworks—Prospective Industrial Applications. J. Mater. Chem. 2006, 16, 626–636. [Google Scholar] [CrossRef]
  112. Bakhtiari, N.; Azizian, S. Nanoporous Carbon Derived from MOF-5: A Superadsorbent for Copper Ions. ACS Omega 2018, 3, 16954–16959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Yoo, J.; Ryu, U.J.; Kwon, W.; Choi, K.M. A Multi-Dye Containing MOF for the Ratiometric Detection and Simultaneous Removal of Cr2O72− in the Presence of Interfering Ions. Sens. Actuators B Chem. 2019, 283, 426–433. [Google Scholar] [CrossRef]
  114. Oveisi, M.; Alinia Asli, M.; Mahmoodi, N.M. Carbon Nanotube Based Metal-Organic Framework Nanocomposites: Synthesis and Their Photocatalytic Activity for Decolorization of Colored Wastewater. Inorg. Chim. Acta 2019, 487, 169–176. [Google Scholar] [CrossRef]
  115. Van Assche, T.R.C.; Desmet, G.; Ameloot, R.; De Vos, D.E.; Terryn, H.; Denayer, J.F.M. Electrochemical Synthesis of Thin HKUST-1 Layers on Copper Mesh. Microporous Mesoporous Mater. 2012, 158, 209–213. [Google Scholar] [CrossRef]
  116. Campagnol, N.; Rezende Souza, E.; De Vos, D.E.; Binnemans, K.; Fransaer, J. Luminescent Terbium-Containing Metal–Organic Framework Films: New Approaches for the Electrochemical Synthesis and Application as Detectors for Explosives. Chem. Commun. 2014, 50, 12680–12683. [Google Scholar] [CrossRef] [Green Version]
  117. Jhung, S.H.; Chang, J.S.; Hwang, J.S.; Park, S.E. Selective Formation of SAPO-5 and SAPO-34 Molecular Sieves with Microwave Irradiation and Hydrothermal Heating. Microporous Mesoporous Mater. 2003, 64, 33–39. [Google Scholar] [CrossRef]
  118. Jhung, S.H.; Yoon, J.W.; Hwang, J.S.; Cheetham, A.K.; Chang, J.S. Facile Synthesis of Nanoporous Nickel Phosphates without Organic Templates under Microwave Irradiation. Chem. Mater. 2005, 17, 4455–4460. [Google Scholar] [CrossRef]
  119. Ratera, I.; Veciana, J. Playing with Organic Radicals as Building Blocks for Functional Molecular Materials. Chem. Soc. Rev. 2012, 41, 303–349. [Google Scholar] [CrossRef]
  120. Pichon, A.; Lazuen-Garay, A.; James, S.L. Solvent-Free Synthesis of a Microporous Metal-Organic Framework. CrystEngComm 2006, 8, 211–214. [Google Scholar] [CrossRef]
  121. Yaghi, O.M.; Li, H.; Davis, C.; Richardson, D.; Groy, T.L. Synthetic Strategies, Structure Patterns, and Emerging Properties in the Chemistry of Modular Porous Solids. Acc. Chem. Res. 1998, 31, 474–484. [Google Scholar] [CrossRef]
  122. Eddaoudi, M.; Moler, D.B.; Li, H.; Chen, B.; Reineke, T.M.; O’Keeffe, M.; Yaghi, O.M. Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal-Organic Carboxylate Frameworks. Acc. Chem. Res. 2001, 34, 319–330. [Google Scholar] [CrossRef]
  123. MICHL, J. The “Molecular Tinkertoy” approach to materials. In Applications of Organometallic Chemistry in the Preparation and Processing of Advanced Materials; Springer: Dordrecht, The Netherlands, 1995; ISBN 9789401041492. [Google Scholar]
  124. Jarrah, A.; Farhadi, S. Dawson-Type Polyoxometalate Incorporated into Nanoporous MIL-101(Cr): Preparation, Characterization and Application for Ultrafast Removal of Organic Dyes. Acta Chim. Slov. 2019, 66, 85–102. [Google Scholar] [CrossRef] [PubMed]
  125. Safaei, M.; Foroughi, M.M.; Ebrahimpoor, N.; Jahani, S.; Omidi, A.; Khatami, M. A Review on Metal-Organic Frameworks: Synthesis and Applications. TrAC Trends Anal. Chem. 2019, 118, 401–425. [Google Scholar] [CrossRef]
  126. Yu, F.; Xiong, X.; Zhou, L.Y.; Li, J.L.; Liang, J.Y.; Hu, S.Q.; Lu, W.T.; Li, B.; Zhou, H.C. Hierarchical Nickel/Phosphorus/Nitrogen/Carbon Composites Templated by One Metal-Organic Framework as Highly Efficient Supercapacitor Electrode Materials. J. Mater. Chem. A 2019, 7, 2875–2883. [Google Scholar] [CrossRef]
  127. Gao, T.; Dong, B.X.; Pan, Y.M.; Liu, W.L.; Teng, Y.L. Highly Sensitive and Recyclable Sensing of Fe3+ Ions Based on a Luminescent Anionic [Cd(DMIPA)]2- Framework with Exposed Thioether Group in the Snowflake-like Channels. J. Solid State Chem. 2019, 270, 493–499. [Google Scholar] [CrossRef]
  128. Pötter, M.; Dehne, H.; Reinke, H.; Dobbertin, J.; Schick, C. Glass-Forming Terephthalic Esters with Lateral Phenylthio Groups and Their Relaxation Behavior. Mol. Cryst. Liq. Cryst. Sci. Technol. Sect. A Mol. Cryst. Liq. Cryst. 1998, 312, 55–68. [Google Scholar] [CrossRef]
  129. Yang, X.; Zhu, P.; Ren, J.; Chen, Y.; Li, X.; Sha, J.; Jiang, J. Surfactant-Assisted Synthesis and Electrochemical Properties of an Unprecedented Polyoxometalate-Based Metal-Organic Nanocaged Framework. Chem. Commun. 2019, 55, 1201–1204. [Google Scholar] [CrossRef]
  130. Wu, Y.; Kobayashi, A.; Halder, G.J.; Peterson, V.K.; Chapman, K.W.; Lock, N.; Southon, P.D.; Kepert, C.J. Negative Thermal Expansion in the Metal-Organic Framework Material Cu 3(1,3,5-Benzenetricarboxylate)2. Angew. Chem. Int. Ed. 2008, 47, 8929–8932. [Google Scholar] [CrossRef]
  131. Ameloot, R.; Stappers, L.; Fransaer, J.; Alaerts, L.; Sels, B.F.; De Vos, D.E. Patterned Growth of Metal-Organic Framework Coatings by Electrochemical Synthesis. Chem. Mater. 2009, 21, 2580–2582. [Google Scholar] [CrossRef]
  132. Yang, H.M.; Liu, X.; Song, X.L.; Yang, T.L.; Liang, Z.H.; Fan, C.M. In Situ Electrochemical Synthesis of MOF-5 and Its Application in Improving Photocatalytic Activity of BiOBr. Trans. Nonferrous Met. Soc. China 2015, 25, 3987–3994. [Google Scholar] [CrossRef]
  133. Martinez Joaristi, A.; 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. [Google Scholar] [CrossRef]
  134. Tompsett, G.A.; Conner, W.C.; Yngvesson, K.S. Microwave Synthesis of Nanoporous Materials. ChemPhysChem 2006, 7, 296–319. [Google Scholar] [CrossRef]
  135. Chen, C.; Feng, X.; Zhu, Q.; Dong, R.; Yang, R.; Cheng, Y.; He, C. Microwave-Assisted Rapid Synthesis of Well-Shaped MOF-74 (Ni) for CO2 Efficient Capture. Inorg. Chem. 2019, 58, 2717–2728. [Google Scholar] [CrossRef]
  136. Ni, Z.; Masel, R.I. Rapid Production of Metal-Organic Frameworks via Microwave-Assisted Solvothermal Synthesis. J. Am. Chem. Soc. 2006, 128, 12394–12395. [Google Scholar] [CrossRef]
  137. Zhang, W.X.; Yang, Y.Y.; Zai, S.B.; Seik, W.N.; Chen, X.M. Syntheses, Structures and Magnetic Properties of Dinuclear Copper(II)-Lanthanide(III) Complexes Bridged by 2-Hydroxymethyl-1- Methylimidazole. Eur. J. Inorg. Chem. 2008, 2, 679–685. [Google Scholar] [CrossRef]
  138. Beldon, P.J.; Fábián, L.; Stein, R.S.; Thirumurugan, A.; Cheetham, A.K.; Friščić, T. Rapid Room-Temperature Synthesis of Zeolitic Imidazolate Frameworks by Using Mechanochemistry. Angew. Chem. 2010, 122, 9834–9837. [Google Scholar] [CrossRef]
  139. Garay, A.L.; Pichon, A.; James, S.L. Solvent-Free Synthesis of Metal Complexes. Chem. Soc. Rev. 2007, 36, 846–855. [Google Scholar] [CrossRef]
  140. Kaupp, G. Mechanochemistry: The Varied Applications of Mechanical Bond-Breaking. CrystEngComm 2009, 11, 388–403. [Google Scholar] [CrossRef]
  141. Bang, J.H.; Suslick, K.S. Applications of Ultrasound to the Synthesis of Nanostructured Materials. Adv. Mater. 2010, 22, 1039–1059. [Google Scholar] [CrossRef]
  142. Mason, T.J. Ultrasound in Synthetic Organic Chemistry. Chem. Soc. Rev. 1997, 26, 443–451. [Google Scholar] [CrossRef]
  143. Yu, K.; Lee, Y.R.; Seo, J.Y.; Baek, K.Y.; Chung, Y.M.; Ahn, W.S. Sonochemical Synthesis of Zr-Based Porphyrinic MOF-525 and MOF-545: Enhancement in Catalytic and Adsorption Properties. Microporous Mesoporous Mater. 2021, 316, 110985. [Google Scholar] [CrossRef]
  144. Hamedi, A.; Zarandi, M.B.; Nateghi, M.R. Highly Efficient Removal of Dye Pollutants by MIL-101(Fe) Metal-Organic Framework Loaded Magnetic Particles Mediated by Poly L-Dopa. J. Environ. Chem. Eng. 2019, 7, 102882. [Google Scholar] [CrossRef]
  145. Tian, M.; Pei, F.; Yao, M.; Fu, Z.; Lin, L.; Wu, G.; Xu, G.; Kitagawa, H.; Fang, X. Ultrathin MOF Nanosheet Assembled Highly Oriented Microporous Membrane as an Interlayer for Lithium-Sulfur Batteries. Energy Storage Mater. 2019, 21, 14–21. [Google Scholar] [CrossRef]
  146. Li, H.; Lv, N.; Li, X.; Liu, B.; Feng, J.; Ren, X.; Guo, T.; Chen, D.; Fraser Stoddart, J.; Gref, R.; et al. Composite CD-MOF Nanocrystals-Containing Microspheres for Sustained Drug Delivery. Nanoscale 2017, 9, 7454–7463. [Google Scholar] [CrossRef] [Green Version]
  147. De Lima Neto, O.J.; Frós, A.C.d.O.; Barros, B.S.; De Farias Monteiro, A.F.; Kulesza, J. Rapid and Efficient Electrochemical Synthesis of a Zinc-Based Nano-MOF for Ibuprofen Adsorption. New J. Chem. 2019, 43, 5518–5524. [Google Scholar] [CrossRef]
  148. Wu, X.; Bao, Z.; Yuan, B.; Wang, J.; Sun, Y.; Luo, H.; Deng, S. Microwave Synthesis and Characterization of MOF-74 (M = Ni, Mg) for Gas Separation. Microporous Mesoporous Mater. 2013, 180, 114–122. [Google Scholar] [CrossRef]
  149. Lu, C.M.; Liu, J.; Xiao, K.; Harris, A.T. Microwave Enhanced Synthesis of MOF-5 and Its CO2 Capture Ability at Moderate Temperatures across Multiple Capture and Release Cycles. Chem. Eng. J. 2010, 156, 465–470. [Google Scholar] [CrossRef]
  150. Javanbakht, S.; Nezhad-Mokhtari, P.; Shaabani, A.; Arsalani, N.; Ghorbani, M. Incorporating Cu-Based Metal-Organic Framework/Drug Nanohybrids into Gelatin Microsphere for Ibuprofen Oral Delivery. Mater. Sci. Eng. C 2019, 96, 302–309. [Google Scholar] [CrossRef]
  151. Azizi Vahed, T.; Naimi-Jamal, M.R.; Panahi, L. Alginate-Coated ZIF-8 Metal-Organic Framework as a Green and Bioactive Platform for Controlled Drug Release. J. Drug Deliv. Sci. Technol. 2019, 49, 570–576. [Google Scholar] [CrossRef]
  152. Abazari, R.; Mahjoub, A.R.; Shariati, J. Synthesis of a Nanostructured Pillar MOF with High Adsorption Capacity towards Antibiotics Pollutants from Aqueous Solution. J. Hazard Mater. 2019, 366, 439–451. [Google Scholar] [CrossRef]
  153. Kumar, P.; Deep, A.; Kim, K.H. Metal Organic Frameworks for Sensing Applications. TrAC Trends Anal. Chem. 2015, 73, 39–53. [Google Scholar] [CrossRef]
  154. Falcaro, P.; Ricco, R.; Doherty, C.M.; Liang, K.; Hill, A.J.; Styles, M.J. MOF Positioning Technology and Device Fabrication. Chem. Soc. Rev. 2014, 43, 5513–5560. [Google Scholar] [CrossRef] [Green Version]
  155. Yap, M.H.; Fow, K.L.; Chen, G.Z. Synthesis and Applications of MOF-Derived Porous Nanostructures. Green Energy Environ. 2017, 2, 218–245. [Google Scholar] [CrossRef]
  156. Janiak, C.; Vieth, J.K. MOFs, MILs and More: Concepts, Properties and Applications for Porous Coordination Networks (PCNs). New J. Chem. 2010, 34, 2366–2388. [Google Scholar] [CrossRef]
  157. Usman, M.; Mendiratta, S.; Lu, K.L. Semiconductor Metal–Organic Frameworks: Future Low-Bandgap Materials. Adv. Mater. 2017, 29, 1–5. [Google Scholar] [CrossRef]
  158. Lv, K.; Fichter, S.; Gu, M.; März, J.; Schmidt, M. An Updated Status and Trends in Actinide Metal-Organic Frameworks (An-MOFs): From Synthesis to Application. Coord. Chem. Rev. 2021, 446, 214011. [Google Scholar] [CrossRef]
  159. Wu, D.; Zhou, K.; Tian, J.; Liu, C.; Tian, J.; Jiang, F.; Yuan, D.; Zhang, J.; Chen, Q.; Hong, M. Induction of Chirality in a Metal–Organic Framework Built from Achiral Precursors. Angew. Chem. 2021, 133, 3124–3131. [Google Scholar] [CrossRef]
  160. Cheng, R.; Li, W.; Wei, W.; Huang, J.; Li, S. Molecular Insights into the Correlation between Microstructure and Thermal Conductivity of Zeolitic Imidazolate Frameworks. ACS Appl. Mater. Interfaces 2021, 13, 14141–14149. [Google Scholar] [CrossRef]
  161. Kolobov, N.; Goesten, M.G.; Gascon, J. Metal–Organic Frameworks: Molecules or Semiconductors in Photocatalysis? Angew. Chem. Int. Ed. 2021, 60, 26038–26052. [Google Scholar] [CrossRef]
  162. Yu, F.; Bai, X.; Liang, M.; Ma, J. Recent Progress on Metal-Organic Framework-Derived Porous Carbon and Its Composite for Pollutant Adsorption from Liquid Phase. Chem. Eng. J. 2021, 405, 126960. [Google Scholar] [CrossRef]
  163. Younis, S.A.; Bhardwaj, N.; Bhardwaj, S.K.; Kim, K.H.; Deep, A. Rare Earth Metal–Organic Frameworks (RE-MOFs): Synthesis, Properties, and Biomedical Applications. Coord. Chem. Rev. 2021, 429, 213620. [Google Scholar] [CrossRef]
  164. Zhao, Y.; Yu, M.; Liu, C.; Li, S.; Li, Z.; Jiang, F.; Chen, L.; Hong, M. Tunable Dual-Emission Luminescence from Cu(i)-Cluster-Based MOFs for Multi-Stimuli Responsive Materials. J. Mater. Chem. C 2021, 9, 2890–2897. [Google Scholar] [CrossRef]
  165. Ma, L.L.; Yang, G.P.; Li, G.P.; Zhang, P.F.; Jin, J.; Wang, Y.; Wang, J.M.; Wang, Y.Y. Luminescence Modulation, near White Light Emission, Selective Luminescence Sensing, and Anticounterfeiting: Via a Series of Ln-MOFs with a p-Conjugated and Uncoordinated Lewis Basic Triazolyl Ligand. Inorg. Chem. Front. 2021, 8, 329–338. [Google Scholar] [CrossRef]
  166. Pan, S.; Chen, X.; Li, X.; Jin, M. Nonderivatization Method for Determination of Glyphosate, Glufosinate, Bialaphos, and Their Main Metabolites in Environmental Waters Based on Magnetic Metal-Organic Framework Pretreatment. J. Sep. Sci. 2019, 42, 1045–1050. [Google Scholar] [CrossRef]
  167. Ghanem, A.; Bados, P.; Kerhoas, L.; Dubroca, J.; Einhorn, J. Glyphosate and AMPA Analysis in Sewage Sludge by LC-ESI-MS/MS after FMOC Derivatization on Strong Anion-Exchange Resin as Solid Support. Anal. Chem. 2007, 79, 3794–3801. [Google Scholar] [CrossRef]
  168. Canivet, J.; Fateeva, A.; Guo, Y.; Coasne, B.; Farrusseng, D. Water Adsorption in MOFs: Fundamentals and Applications. Chem. Soc. Rev. 2014, 43, 5594–5617. [Google Scholar] [CrossRef] [Green Version]
  169. Wan Ngah, W.S.; Hanafiah, M.A.K.M. Removal of Heavy Metal Ions from Wastewater by Chemically Modified Plant Wastes as Adsorbents: A Review. Bioresour. Technol. 2008, 99, 3935–3948. [Google Scholar] [CrossRef]
  170. Khan, N.A.; Hasan, Z.; Jhung, S.H. Adsorptive Removal of Hazardous Materials Using Metal-Organic Frameworks (MOFs): A Review. J. Hazard Mater. 2013, 244–245, 444–456. [Google Scholar] [CrossRef]
  171. Abd, A.A.; Naji, S.Z.; Hashim, A.S.; Othman, M.R. Carbon Dioxide Removal through Physical Adsorption Using Carbonaceous and Non-Carbonaceous Adsorbents: A Review. J. Environ. Chem. Eng. 2020, 8, 104142. [Google Scholar] [CrossRef]
  172. Wang, H.; Lou, X.; Hu, Q.; Sun, T. Adsorption of Antibiotics from Water by Using Chinese Herbal Medicine Residues Derived Biochar: Preparation and Properties Studies. J. Mol. Liq. 2021, 325, 2–10. [Google Scholar] [CrossRef]
  173. Majewski, M.B.; Howarth, A.J.; Li, P.; Wasielewski, M.R.; Hupp, J.T.; Farha, O.K. Enzyme Encapsulation in Metal-Organic Frameworks for Applications in Catalysis. CrystEngComm 2017, 19, 4082–4091. [Google Scholar] [CrossRef]
  174. Li, N.; Zhou, L.; Jin, X.; Owens, G.; Chen, Z. Simultaneous Removal of Tetracycline and Oxytetracycline Antibiotics from Wastewater Using a ZIF-8 Metal Organic-Framework. J. Hazard Mater. 2019, 366, 563–572. [Google Scholar] [CrossRef]
  175. Joseph, L.; Jun, B.M.; Jang, M.; Park, C.M.; Muñoz-Senmache, J.C.; Hernández-Maldonado, A.J.; Heyden, A.; Yu, M.; Yoon, Y. Removal of Contaminants of Emerging Concern by Metal-Organic Framework Nanoadsorbents: A Review. Chem. Eng. J. 2019, 369, 928–946. [Google Scholar] [CrossRef]
  176. Dhankhar, S.S.; Nagaraja, C.M. Construction of a 3D Porous Co(Ii) Metal-Organic Framework (MOF) with Lewis Acidic Metal Sites Exhibiting Selective CO2 Capture and Conversion under Mild Conditions. New J. Chem. 2019, 43, 2163–2170. [Google Scholar] [CrossRef]
  177. Huelsenbeck, L.; Westendorff, K.S.; Gu, Y.; Marino, S.; Jung, S.; Epling, W.S.; Giri, G. Modulating and Orienting an Anisotropic Zn-Based Metal Organic Framework for Selective CH4/CO2 Gas Separation. Crystals 2019, 9, 20. [Google Scholar] [CrossRef] [Green Version]
  178. Ikreedeegh, R.R.; Tahir, M. A Critical Review in Recent Developments of Metal-Organic-Frameworks (MOFs) with Band Engineering Alteration for Photocatalytic CO2 reduction to Solar Fuels. J. CO2 Util. 2021, 43, 101381. [Google Scholar] [CrossRef]
  179. Wang, J.; Cherevan, A.S.; Hannecart, C.; Naghdi, S.; Nandan, S.P.; Gupta, T.; Eder, D. Ti-Based MOFs: New Insights on the Impact of Ligand Composition and Hole Scavengers on Stability, Charge Separation and Photocatalytic Hydrogen Evolution. Appl. Catal. B Environ. 2021, 283, 119626. [Google Scholar] [CrossRef]
  180. Gkaniatsou, E.; Ricoux, R.; Kariyawasam, K.; Stenger, I.; Fan, B.; Ayoub, N.; Salas, S.; Patriarche, G.; Serre, C.; Mahy, J.P.; et al. Encapsulation of Microperoxidase-8 in MIL-101(Cr)-X Nanoparticles: Influence of Metal-Organic Framework Functionalization on Enzymatic Immobilization and Catalytic Activity. ACS Appl. Nano Mater. 2020, 3, 3233–3243. [Google Scholar] [CrossRef]
  181. Phipps, J.; Chen, H.; Donovan, C.; Dominguez, D.; Morgan, S.; Weidman, B.; Fan, C.; Fan, C.; Beyzavi, M.H.; Beyzavi, M.H. Catalytic Activity, Stability, and Loading Trends of Alcohol Dehydrogenase Enzyme Encapsulated in a Metal-Organic Framework. ACS Appl. Mater. Interfaces 2020, 12, 26084–26094. [Google Scholar] [CrossRef]
  182. An, H.; Li, M.; Gao, J.; Zhang, Z.; Ma, S.; Chen, Y. Incorporation of Biomolecules in Metal-Organic Frameworks for Advanced Applications. Coord. Chem. Rev. 2019, 384, 90–106. [Google Scholar] [CrossRef]
  183. Li, B.; Suo, T.; Xie, S.; Xia, A.; Ma, Y.J.; Huang, H.; Zhang, X.; Hu, Q. Rational Design, Synthesis, and Applications of Carbon Dots@metal–Organic Frameworks (CD@MOF) Based Sensors. TrAC Trends Anal. Chem. 2021, 135, 116163. [Google Scholar] [CrossRef]
  184. 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. [Google Scholar] [CrossRef]
  185. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal-Organic Framework Materials as Chemical Sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef]
  186. Wang, Q.; Zhang, X.; Huang, L.; Zhang, Z.; Dong, S. GOx@ZIF-8(NiPd) Nanoflower: An Artificial Enzyme System for Tandem Catalysis. Angew. Chem. 2017, 129, 16298–16301. [Google Scholar] [CrossRef]
  187. Kornienko, N. Operando Spectroscopy of Nanoscopic Metal/Covalent Organic Framework Electrocatalysts. Nanoscale 2021, 13, 1507–1514. [Google Scholar] [CrossRef]
  188. Chen, B.; Xiang, S.; Qian, G. Metal-Organic Frameworks with Functional Pores for Recognition of Small Molecules. Acc. Chem. Res. 2010, 43, 1115–1124. [Google Scholar] [CrossRef]
  189. Wang, Z.; Gui, M.; Asif, M.; Yu, Y.; Dong, S.; Wang, H.; Wang, W.; Wang, F.; Xiao, F.; Liu, H. A Facile Modular Approach to the 2D Oriented Assembly MOF Electrode for Non-Enzymatic Sweat Biosensors. Nanoscale 2018, 10, 6629–6638. [Google Scholar] [CrossRef]
  190. Qiu, Q.; Chen, H.; Wang, Y.; Ying, Y. Recent Advances in the Rational Synthesis and Sensing Applications of Metal-Organic Framework Biocomposites. Coord. Chem. Rev. 2019, 387, 60–78. [Google Scholar] [CrossRef]
  191. Du, L.; Chen, W.; Zhu, P.; Tian, Y.; Chen, Y.; Wu, C. Applications of Functional Metal-Organic Frameworks in Biosensors. Biotechnol. J. 2021, 16, 1–12. [Google Scholar] [CrossRef]
  192. Yang, M.; Sun, Z.; Jin, H.; Gui, R. Urate Oxidase-Loaded MOF Electrodeposited on Boron Nanosheet-Doxorubicin Complex as Multifunctional Nano-Enzyme Platform for Enzymatic and Ratiometric Electrochemical Biosensing. Talanta 2022, 243, 123359. [Google Scholar] [CrossRef]
  193. Lu, X.; Zhang, F.; Sun, Y.; Yu, K.; Guo, W.; Qu, F. A 2D/2D NiCo-MOF/Ti3C2 heterostructure for the Simultaneous Detection of Acetaminophen, Dopamine and Uric Acid by Differential Pulse Voltammetry. Dalt. Trans. 2021, 50, 16593–16600. [Google Scholar] [CrossRef]
  194. Jin, X.; Li, G.; Xu, T.; Su, L.; Yan, D.; Zhang, X. Ruthenium-Based Conjugated Polymer and Metal-Organic Framework Nanocomposites for Glucose Sensing. Electroanalysis 2021, 33, 1902–1910. [Google Scholar] [CrossRef]
  195. Bao, T.; Fu, R.; Jiang, Y.; Wen, W.; Zhang, X.; Wang, S. Metal-Mediated Polydopamine Nanoparticles-DNA Nanomachine Coupling Electrochemical Conversion of Metal-Organic Frameworks for Ultrasensitive MicroRNA Sensing. Anal. Chem. 2021, 93, 13475–13484. [Google Scholar] [CrossRef]
  196. Cui, H.; Cui, S.; Zhang, S.; Tian, Q.; Liu, Y.; Zhang, P.; Wang, M.; Zhang, J.; Li, X. Cu-MOF/Hemin: A Bionic Enzyme with Excellent Dispersity for the Determination of Hydrogen Peroxide Released from Living Cells. Analyst 2021, 146, 5951–5961. [Google Scholar] [CrossRef]
  197. Liu, X.; Zhao, Y.; Li, F. Nucleic Acid-Functionalized Metal-Organic Framework for Ultrasensitive Immobilization-Free Photoelectrochemical Biosensing. Biosens. Bioelectron. 2021, 173, 112832. [Google Scholar] [CrossRef]
  198. Yang, J.; Yang, Y.W. Metal–Organic Frameworks for Biomedical Applications. Small 2020, 16, 1–24. [Google Scholar] [CrossRef]
  199. Lawson, H.D.; Walton, S.P.; Chan, C. Metal-Organic Frameworks for Drug Delivery: A Design Perspective. ACS Appl. Mater. Interfaces 2021, 13, 7004–7020. [Google Scholar] [CrossRef]
  200. Ji, X.; Mo, Y.; Li, H.; Zhao, W.; Zhong, A.; Li, S.; Wang, Q.; Duan, X.; Xiao, J. Gender-Dependent Reproductive Toxicity of Copper Metal-Organic Frameworks and Attenuation by Surface Modification. Nanoscale 2021, 13, 7389–7402. [Google Scholar] [CrossRef]
  201. Wang, Y.; Tu, R.; Hou, C.; Wang, Z. Zn–Porphyrin Metal–Organic Framework–Based Photoelectrochemical Enzymatic Biosensor for Hypoxanthine. J. Solid State Electrochem. 2022, 26, 565–572. [Google Scholar] [CrossRef]
  202. Wang, Q.; Chen, M.; Xiong, C.; Zhu, X.; Chen, C.; Zhou, F.; Dong, Y.; Wang, Y.; Xu, J.; Li, Y.; et al. Dual Confinement of High–Loading Enzymes within Metal–Organic Frameworks for Glucose Sensor with Enhanced Cascade Biocatalysis. Biosens. Bioelectron. 2022, 196, 113695. [Google Scholar] [CrossRef]
  203. Guo, J.; Yang, L.; Gao, Z.; Zhao, C.; Mei, Y.; Song, Y.Y. Insight of MOF Environment-Dependent Enzyme Activity via MOFs-in-Nanochannels Configuration. ACS Catal. 2020, 10, 5949–5958. [Google Scholar] [CrossRef]
  204. Kustov, L.M.; Isaeva, V.I.; Přech, J.; Bisht, K.K. Metal-Organic Frameworks as Materials for Applications in Sensors. Mendeleev Commun. 2019, 29, 361–368. [Google Scholar] [CrossRef]
  205. Zhou, H.C.J.; Kitagawa, S. Metal–Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef] [Green Version]
  206. Ghorbanpour, M.; Bhargava, P.; Varma, A.; Choudhary, D.K. Biogenic Nano-Particles and Their Use in Agro-Ecosystems; Springer: Singapore, 2020; ISBN 9789811529856. [Google Scholar]
  207. Shet, S.P.; Shanmuga Priya, S.; Sudhakar, K.; Tahir, M. A Review on Current Trends in Potential Use of Metal-Organic Framework for Hydrogen Storage. Int. J. Hydrogen Energy 2021, 46, 11782–11803. [Google Scholar] [CrossRef]
  208. Langmi, H.W.; Ren, J.; North, B.; Mathe, M.; Bessarabov, D. Hydrogen Storage in Metal-Organic Frameworks: A Review. Electrochim. Acta 2014, 128, 368–392. [Google Scholar] [CrossRef]
  209. Suresh, K.; Aulakh, D.; Purewal, J.; Siegel, D.J.; Veenstra, M.; Matzger, A.J. Optimizing Hydrogen Storage in MOFs through Engineering of Crystal Morphology and Control of Crystal Size. J. Am. Chem. Soc. 2021, 143, 10727–10734. [Google Scholar] [CrossRef]
  210. Ma, L.; He, Y.; Wang, Y.; Wang, Y.; Li, R.; Huang, Z.; Jiang, Y.; Gao, J. Nanocomposites of Pt Nanoparticles Anchored on UiO66-NH2 as Carriers to Construct Acetylcholinesterase Biosensors for Organophosphorus Pesticide Detection. Electrochim. Acta 2019, 318, 525–533. [Google Scholar] [CrossRef]
  211. Barea, E.; Montoro, C.; Navarro, J.A.R. Toxic Gas Removal-Metal-Organic Frameworks for the Capture and Degradation of Toxic Gases and Vapours. Chem. Soc. Rev. 2014, 43, 5419–5430. [Google Scholar] [CrossRef] [Green Version]
  212. Yilmaz, G.; Peh, S.B.; Zhao, D.; Ho, G.W. Atomic- and Molecular-Level Design of Functional Metal–Organic Frameworks (MOFs) and Derivatives for Energy and Environmental Applications. Adv. Sci. 2019, 6, 1901129. [Google Scholar] [CrossRef] [Green Version]
  213. Cortés-Súarez, J.; Celis-Arias, V.; Beltrán, H.I.; Tejeda-Cruz, A.; Ibarra, I.A.; Romero-Ibarra, J.E.; Sánchez-González, E.; Loera-Serna, S. Synthesis and Characterization of an SWCNT@HKUST-1 Composite: Enhancing the CO2 Adsorption Properties of HKUST-1. ACS Omega 2019, 4, 5275–5282. [Google Scholar] [CrossRef] [Green Version]
  214. Mehta, J.; Dhaka, S.; Paul, A.K.; Dayananda, S.; Deep, A. Organophosphate Hydrolase Conjugated UiO-66-NH2 MOF Based Highly Sensitive Optical Detection of Methyl Parathion. Environ. Res. 2019, 174, 46–53. [Google Scholar] [CrossRef] [PubMed]
  215. Ma, X.; Chai, Y.; Li, P.; Wang, B. Metal-Organic Framework Films and Their Potential Applications in Environmental Pollution Control. Acc. Chem. Res. 2019, 52, 1461–1470. [Google Scholar] [CrossRef] [PubMed]
  216. Yang, J.M.; Ying, R.J.; Han, C.X.; Hu, Q.T.; Xu, H.M.; Li, J.H.; Wang, Q.; Zhang, W. Adsorptive Removal of Organic Dyes from Aqueous Solution by a Zr-Based Metal-Organic Framework: Effects of Ce(III) Doping. Dalt. Trans. 2018, 47, 3913–3920. [Google Scholar] [CrossRef]
  217. Yu, L.L.; Luo, Z.F.; Zhang, Y.Y.; Wu, S.C.; Yang, C.; Cheng, J. hua Contrastive Removal of Oxytetracycline and Chlortetracycline from Aqueous Solution on Al-MOF/GO Granules. Environ. Sci. Pollut. Res. 2019, 26, 3685–3696. [Google Scholar] [CrossRef]
  218. Zhang, M.; Feng, G.; Song, Z.; Zhou, Y.P.; Chao, H.Y.; Yuan, D.; Tan, T.T.Y.; Guo, Z.; Hu, Z.; Tang, B.Z.; et al. Two-Dimensional Metal-Organic Framework with Wide Channels and Responsive Turn-on Fluorescence for the Chemical Sensing of Volatile Organic Compounds. J. Am. Chem. Soc. 2014, 136, 7241–7244. [Google Scholar] [CrossRef] [PubMed]
  219. Xu, D.; Qi, S.; Chen, Y.; Yin, M.; Zhang, L.; Ge, K.; Wei, X.; Tian, X.; Wang, P.; Li, M.; et al. Hierarchical Mesoporous Hollow Ce-MOF Nanosphere as Oxidase Mimic for Highly Sensitive Colorimetric Detection of Ascorbic Acid. Chem. Phys. Lett. 2021, 777, 138749. [Google Scholar] [CrossRef]
  220. Abdelhamid, H.N.; Sharmoukh, W. Intrinsic Catalase-Mimicking MOFzyme for Sensitive Detection of Hydrogen Peroxide and Ferric Ions. Microchem. J. 2021, 163, 105873. [Google Scholar] [CrossRef]
  221. Zhang, C.; Hong, S.; Liu, M.D.; Yu, W.Y.; Zhang, M.K.; Zhang, L.; Zeng, X.; Zhang, X.Z. PH-Sensitive MOF Integrated with Glucose Oxidase for Glucose-Responsive Insulin Delivery. J. Control. Release 2020, 320, 159–167. [Google Scholar] [CrossRef]
  222. Zhang, Y.; Yang, X.; Zhou, H.C. Direct Synthesis of Functionalized PCN-333 via Linker Design for Fe3+ Detection in Aqueous Media. Dalt. Trans. 2018, 47, 11806–11811. [Google Scholar] [CrossRef]
  223. Li, Z.; Zhang, H.; Zha, Q.; Zhai, C.; Li, W.; Zeng, L.; Zhu, M. Photo-Electrochemical Detection of Dopamine in Human Urine and Calf Serum Based on MIL-101 (Cr)/Carbon Black. Microchim. Acta 2020, 187, 526. [Google Scholar] [CrossRef]
  224. Haghighi, E.; Zeinali, S. Nanoporous MIL-101(Cr) as a Sensing Layer Coated on a Quartz Crystal Microbalance (QCM) Nanosensor to Detect Volatile Organic Compounds (VOCs). RSC Adv. 2019, 9, 24460–24470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Fang, B.; Guo, P.; Yang, M.; Ma, Y.; Yan, X.; Jia, Z.; Gao, W.; Ahmad, S.; Xu, C.; Liu, C.; et al. A Novel Fluorescent Enhancing Platform Based on DNA-Scaffolded Silver Nanoclusters for Potential Inflammatory Bowel Disease-Associated MicroRNA Detection. Talanta 2020, 218, 121122. [Google Scholar] [CrossRef] [PubMed]
  226. Hassanzadeh, J.; Khataee, A.; Eskandari, H. Encapsulated Cholesterol Oxidase in Metal-Organic Framework and Biomimetic Ag Nanocluster Decorated MoS2 Nanosheets for Sensitive Detection of Cholesterol. Sens. Actuators B Chem. 2018, 259, 402–410. [Google Scholar] [CrossRef]
  227. Tu, X.; Gao, F.; Ma, X.; Zou, J.; Yu, Y.; Li, M.; Qu, F.; Huang, X.; Lu, L. Mxene/Carbon Nanohorn/β-Cyclodextrin-Metal-Organic Frameworks as High-Performance Electrochemical Sensing Platform for Sensitive Detection of Carbendazim Pesticide. J. Hazard Mater. 2020, 396, 122776. [Google Scholar] [CrossRef] [PubMed]
  228. Chen, J.; Jiang, S.; Wang, M.; Xie, X.; Su, X. Self-Assembled Dual-Emissive Nanoprobe with Metal−organic Frameworks as Scaff; Olds for Enhanced Ascorbic Acid and Ascorbate Oxidase Sensing. Sens. Actuators B Chem. 2021, 339, 129910. [Google Scholar] [CrossRef]
  229. Yu, J.; Wei, Z.; Li, Q.; Wan, F.; Chao, Z.; Zhang, X.; Lin, L.; Meng, H.; Tian, L. Advanced Cancer Starvation Therapy by Simultaneous Deprivation of Lactate and Glucose Using a MOF Nanoplatform. Adv. Sci. 2021, 8, 1–13. [Google Scholar] [CrossRef]
  230. Moghaddam, Z.S.; Kaykhaii, M.; Khajeh, M.; Oveisi, A.R. Synthesis of UiO-66-OH Zirconium Metal-Organic Framework and Its Application for Selective Extraction and Trace Determination of Thorium in Water Samples by Spectrophotometry. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 194, 76–82. [Google Scholar] [CrossRef]
  231. Nadar, S.S.; Rathod, V.K. Encapsulation of Lipase within Metal-Organic Framework (MOF) with Enhanced Activity Intensified under Ultrasound. Enzym. Microb. Technol. 2018, 108, 11–20. [Google Scholar] [CrossRef]
  232. Yang, J.; Li, K.; Li, C.; Gu, J. In Situ Coupling of Catalytic Centers into Artificial Substrate Mesochannels as Super-Active Metalloenzyme Mimics. Small 2021, 17, 1–11. [Google Scholar] [CrossRef]
  233. Zhou, J.; Long, Z.; Tian, Y.; Ding, X.; Wu, L.; Hou, X. A Chemiluminescence Metalloimmunoassay for Sensitive Detection of Alpha-Fetoprotein in Human Serum Using Fe-MIL-88B-NH2 as a Label. Appl. Spectrosc. Rev. 2016, 51, 517–526. [Google Scholar] [CrossRef]
  234. Lykourinou, V.; Chen, Y.; Wang, X.; Meng, L.; Hoang, T.; Ming, L.; Musselman, R.L.; Ma, S. Immobilization of MP-11 into a Mesoporous Metal?Organic Framework, MP-11@mesoMOF: A New Platform for Enzymatic Catalysis. J. Am. Chem. Soc. 2011, 133, 10382–10385. [Google Scholar] [CrossRef] [PubMed]
  235. Zhu, G.; Cheng, L.; Qi, R.; Zhang, M.; Zhao, J.; Zhu, L.; Dong, M. A Metal-Organic Zeolitic Framework with Immobilized Urease for Use in a Tapered Optical Fiber Urea Biosensor. Microchim. Acta 2020, 187, 72. [Google Scholar] [CrossRef]
  236. Liu, W.L.; Lo, S.H.; Singco, B.; Yang, C.C.; Huang, H.Y.; Lin, C.H. Novel Trypsin-FITC@MOF Bioreactor Efficiently Catalyzes Protein Digestion. J. Mater. Chem. B 2013, 1, 928–932. [Google Scholar] [CrossRef] [PubMed]
  237. Sun, X.; Yu, T.; Du, Y.; Ding, W.; Chen, C.; Ma, X. Correction to: Metal Organic Framework HKUST-1 Modified with Carboxymethyl-β-Cyclodextrin for Use in Improved Open Tubular Capillary Electrochromatographic Enantioseparation of Five Basic Drugs (Microchimica Acta, (2019), 186, 7, (462), 10.1007/S00604-019. Microchim. Acta 2019, 186, 1–8. [Google Scholar] [CrossRef] [Green Version]
  238. Chen, Y.; Chen, F.; Zhang, S.; Cai, Y.; Cao, S.; Li, S.; Zhao, W.; Yuan, S.; Feng, X.; Cao, A.; et al. Facile Fabrication of Multifunctional Metal-Organic Framework Hollow Tubes to Trap Pollutants. J. Am. Chem. Soc. 2017, 139, 16482–16485. [Google Scholar] [CrossRef]
  239. Chen, S.; Wen, L.; Svec, F.; Tan, T.; Lv, Y. Magnetic Metal-Organic Frameworks as Scaffolds for Spatial Co-Location and Positional Assembly of Multi-Enzyme Systems Enabling Enhanced Cascade Biocatalysis. RSC Adv. 2017, 7, 21205–21213. [Google Scholar] [CrossRef] [Green Version]
  240. Aguilera-Sigalat, J.; Bradshaw, D. Synthesis and Applications of Metal-Organic Framework-Quantum Dot (QD at MOF) Composites. Coord. Chem. Rev. 2016, 307, 267–291. [Google Scholar] [CrossRef] [Green Version]
  241. Wang, K.; Li, N.; Zhang, J.; Zhang, Z.; Dang, F. Size-Selective QD@MOF Core-Shell Nanocomposites for the Highly Sensitive Monitoring of Oxidase Activities. Biosens. Bioelectron. 2017, 87, 339–344. [Google Scholar] [CrossRef]
  242. Wu, T.; Liu, X.; Liu, Y.; Cheng, M.; Liu, Z.; Zeng, G.; Shao, B.; Liang, Q.; Zhang, W.; He, Q. Application of QD-MOF Composites for Photocatalysis: Energy Production and Environmental Remediation. Coord. Chem. Rev. 2020, 403, 213097. [Google Scholar] [CrossRef]
  243. Yang, G.L.; Jiang, X.L.; Xu, H.; Zhao, B. Applications of MOFs as Luminescent Sensors for Environmental Pollutants. Small 2021, 17, 1–19. [Google Scholar] [CrossRef]
  244. Meng, W.; Wen, Y.; Dai, L.; He, Z.; Wang, L. A Novel Electrochemical Sensor for Glucose Detection Based on Ag@ZIF-67 Nanocomposite. Sens. Actuators B Chem. 2018, 260, 852–860. [Google Scholar] [CrossRef]
  245. Aguado, S.; Quirós, J.; Canivet, J.; Farrusseng, D.; Boltes, K.; Rosal, R. Antimicrobial Activity of Cobalt Imidazolate Metal-Organic Frameworks. Chemosphere 2014, 113, 188–192. [Google Scholar] [CrossRef] [PubMed]
  246. Bagheri, N.; Khataee, A.; Hassanzadeh, J.; Habibi, B. Sensitive Biosensing of Organophosphate Pesticides Using Enzyme Mimics of Magnetic ZIF-8. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 209, 118–125. [Google Scholar] [CrossRef] [PubMed]
  247. Gascón, V.; Jiménez, M.B.; Blanco, R.M.; Sanchez-Sanchez, M. Semi-Crystalline Fe-BTC MOF Material as an Efficient Support for Enzyme Immobilization. Catal. Today 2018, 304, 119–126. [Google Scholar] [CrossRef] [Green Version]
  248. Girão Neto, C.A.C.; Silva, N.C.G.E.; de Oliveira Costa, T.; de Albuquerque, T.L.; Gonçalves, L.R.B.; Fernandez-Lafuente, R.; Rocha, M.V.P. The β-Galactosidase Immobilization Protocol Determines Its Performance as Catalysts in the Kinetically Controlled Synthesis of Lactulose. Int. J. Biol. Macromol. 2021, 176, 468–478. [Google Scholar] [CrossRef]
  249. Nadar, S.S.; Rathod, V.K. Magnetic-Metal Organic Framework (Magnetic-MOF): A Novel Platform for Enzyme Immobilization and Nanozyme Applications. Int. J. Biol. Macromol. 2018, 120, 2293–2302. [Google Scholar] [CrossRef]
  250. Nadar, S.S.; Vaidya, L.; Rathod, V.K. Enzyme Embedded Metal Organic Framework (Enzyme–MOF): De Novo Approaches for Immobilization. Int. J. Biol. Macromol. 2020, 149, 861–876. [Google Scholar] [CrossRef]
  251. Wang, J.; Zhao, G.; Yu, F. Facile Preparation of Fe3O4@MOF Core-Shell Microspheres for Lipase Immobilization. J. Taiwan Inst. Chem. Eng. 2016, 69, 139–145. [Google Scholar] [CrossRef]
  252. Wu, X.; Yang, C.; Ge, J. Green Synthesis of Enzyme/Metal-Organic Framework Composites with High Stability in Protein Denaturing Solvents. Bioresour. Bioprocess 2017, 4, 1–8. [Google Scholar] [CrossRef]
  253. Gascón, V.; Castro-Miguel, E.; Díaz-García, M.; Blanco, R.M.; Sanchez-Sanchez, M. In Situ and Post-Synthesis Immobilization of Enzymes on Nanocrystalline MOF Platforms to Yield Active Biocatalysts. J. Chem. Technol. Biotechnol. 2017, 92, 2583–2593. [Google Scholar] [CrossRef]
  254. Cui, J.; Ren, S.; Sun, B.; Jia, S. Optimization Protocols and Improved Strategies for Metal-Organic Frameworks for Immobilizing Enzymes: Current Development and Future Challenges. Coord. Chem. Rev. 2018, 370, 22–41. [Google Scholar] [CrossRef]
  255. Cao, S.L.; Yue, D.M.; Li, X.H.; Smith, T.J.; Li, N.; Zong, M.H.; Wu, H.; Ma, Y.Z.; Lou, W.Y. Novel Nano-/Micro-Biocatalyst: Soybean Epoxide Hydrolase Immobilized on UiO-66-NH2 MOF for Efficient Biosynthesis of Enantiopure (R)-1, 2-Octanediol in Deep Eutectic Solvents. ACS Sustain. Chem. Eng. 2016, 4, 3586–3595. [Google Scholar] [CrossRef]
  256. Pang, S.; Wu, Y.; Zhang, X.; Li, B.; Ouyang, J.; Ding, M. Immobilization of Laccase via Adsorption onto Bimodal Mesoporous Zr-MOF. Process Biochem. 2016, 51, 229–239. [Google Scholar] [CrossRef]
  257. Cao, Y.; Wu, Z.; Wang, T.; Xiao, Y.; Huo, Q.; Liu, Y. Immobilization of: Bacillus Subtilis Lipase on a Cu-BTC Based Hierarchically Porous Metal-Organic Framework Material: A Biocatalyst for Esterification. Dalt. Trans. 2016, 45, 6998–7003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Li, P.; Moon, S.Y.; Guelta, M.A.; Harvey, S.P.; Hupp, J.T.; Farha, O.K. Encapsulation of a Nerve Agent Detoxifying Enzyme by a Mesoporous Zirconium Metal-Organic Framework Engenders Thermal and Long-Term Stability. J. Am. Chem. Soc. 2016, 138, 8052–8055. [Google Scholar] [CrossRef]
  259. Lian, X.; Chen, Y.P.; Liu, T.F.; Zhou, H.C. Coupling Two Enzymes into a Tandem Nanoreactor Utilizing a Hierarchically Structured MOF. Chem. Sci. 2016, 7, 6969–6973. [Google Scholar] [CrossRef] [Green Version]
  260. Zhang, X.; Tu, R.; Lu, Z.; Peng, J.; Hou, C.; Wang, Z. Hierarchical Mesoporous Metal–Organic Frameworks Encapsulated Enzymes: Progress and Perspective. Coord. Chem. Rev. 2021, 443, 214032. [Google Scholar] [CrossRef]
  261. Sokolov, A.V.; Limareva, L.V.; Iliasov, P.V.; Gribkova, O.V.; Sustretov, A.S. Methods of Encapsulation of Biomacromolecules and Living Cells. Prospects of Using Metal–Organic Frameworks. Russ. J. Org. Chem. 2021, 57, 491–505. [Google Scholar] [CrossRef]
  262. Xu, W.; Jiao, L.; Wu, Y.; Hu, L.; Gu, W.; Zhu, C. Metal–Organic Frameworks Enhance Biomimetic Cascade Catalysis for Biosensing. Adv. Mater. 2021, 33, 1–18. [Google Scholar] [CrossRef]
  263. Xie, J.; Zhang, Y.; Simpson, B. Food Enzymes Immobilization: Novel Carriers, Techniques and Applications. Curr. Opin. Food Sci. 2022, 43, 27–35. [Google Scholar] [CrossRef]
Figure 1. Growth in the number of published articles retrieved on Scopus over 10 years using the following keywords: (1) “metal-organic frameworks” and “MOF”; and (2) “metal-organic frameworks”, “MOF”, “enzyme”, and “immobilization”. The search was carried out on 7 January 2022, and returned (1) 45,925 and (2) 371 documents.
Figure 1. Growth in the number of published articles retrieved on Scopus over 10 years using the following keywords: (1) “metal-organic frameworks” and “MOF”; and (2) “metal-organic frameworks”, “MOF”, “enzyme”, and “immobilization”. The search was carried out on 7 January 2022, and returned (1) 45,925 and (2) 371 documents.
Molecules 27 04529 g001
Figure 2. Schematic representation of strategies of (a) hydrothermal–solvothermal synthesis, (b) diffusion synthesis, (c) electrochemical synthesis, (d) microwave-assisted synthesis, (e) mechanochemical synthesis, and (f) sonochemistry synthesis.
Figure 2. Schematic representation of strategies of (a) hydrothermal–solvothermal synthesis, (b) diffusion synthesis, (c) electrochemical synthesis, (d) microwave-assisted synthesis, (e) mechanochemical synthesis, and (f) sonochemistry synthesis.
Molecules 27 04529 g002
Figure 3. Representation of the formation of a crystalline structure of metal–organic structures (MOFs) based on organic ligands being coupled to a metallic center.
Figure 3. Representation of the formation of a crystalline structure of metal–organic structures (MOFs) based on organic ligands being coupled to a metallic center.
Molecules 27 04529 g003
Figure 4. Representation scheme of composite properties based on metal–organic framework (MOF) structures, highlighting their thermal capacity, chirality, high stability, semiconductivity, luminescence, magnetism, catalytic power, and porosity.
Figure 4. Representation scheme of composite properties based on metal–organic framework (MOF) structures, highlighting their thermal capacity, chirality, high stability, semiconductivity, luminescence, magnetism, catalytic power, and porosity.
Molecules 27 04529 g004
Figure 5. Schematic diagram showing pollutant adsorption on the surface of metal–organic frameworks (MOFs), where the contaminant particles can bind to the material, leaving a pollutant-free aqueous medium.
Figure 5. Schematic diagram showing pollutant adsorption on the surface of metal–organic frameworks (MOFs), where the contaminant particles can bind to the material, leaving a pollutant-free aqueous medium.
Molecules 27 04529 g005
Figure 6. A simplified representation of the separation of metal–organic structures (MOFs) from their reaction media by their magnetic properties, which enables their simplified removal—an excellent characteristic for catalysts.
Figure 6. A simplified representation of the separation of metal–organic structures (MOFs) from their reaction media by their magnetic properties, which enables their simplified removal—an excellent characteristic for catalysts.
Molecules 27 04529 g006
Figure 7. Representation of a metal–organic framework (MOF) as a drug-administrating carrier in tumor cells. They can be used as identifiers of the regions of inflammation and, due to their luminescence, can make it easy to detect the exact region of drug action.
Figure 7. Representation of a metal–organic framework (MOF) as a drug-administrating carrier in tumor cells. They can be used as identifiers of the regions of inflammation and, due to their luminescence, can make it easy to detect the exact region of drug action.
Molecules 27 04529 g007
Figure 8. Illustration of the mapping of overexpressed macromolecules in tumor cell lines through luminescent metal–organic framework (MOF) composites. This is a widely explored property, which was enabled by their metallic centers and mapped by confocal microscopy.
Figure 8. Illustration of the mapping of overexpressed macromolecules in tumor cell lines through luminescent metal–organic framework (MOF) composites. This is a widely explored property, which was enabled by their metallic centers and mapped by confocal microscopy.
Molecules 27 04529 g008
Figure 9. Scheme of metal–organic frameworks (MOFs) capturing hydrogen via their adsorbent properties, a function that can be used for hydrogen storage.
Figure 9. Scheme of metal–organic frameworks (MOFs) capturing hydrogen via their adsorbent properties, a function that can be used for hydrogen storage.
Molecules 27 04529 g009
Figure 10. Representation of different techniques of enzyme immobilization onto MOFs.
Figure 10. Representation of different techniques of enzyme immobilization onto MOFs.
Molecules 27 04529 g010
Table 1. Metal–organic frameworks: a summary of different synthesis strategies and their applications.
Table 1. Metal–organic frameworks: a summary of different synthesis strategies and their applications.
Synthesis StrategyMain FeaturesApplicationsMaterialRef.
Hydrothermal (Solvothermal)Generally, processes that use water as a solvent are termed hydrothermal processes, while those that use other solvents are classified as solvothermal processes [124,126]Dye removalMIL-101(Fe)@PDopa@Fe3O4[144]
Lithium–sulfur batteryCu2(CuTCPP)[145]
DiffusionDiffusion MOF synthesis methods can be subdivided into two strategies: diffusion between two liquids with different densities (no physical barrier) and gradual diffusion that occurs through physical barriers [111]Drug deliveryCD-MOF[146]
Adsorption of copper ionsMOF-5[112]
ElectrochemicalElectrochemical synthesis is widely used at an industrial scale to produce MOFs [129]Lithium-ion batteriesZn-POMCF[129]
Ibuprofen adsorption[Zn(1,3-bdc)0.5(bzim)][147]
Microwave-assistedThe microwave technique is widely used in synthesizing small particles of oxides and metals [137]Gas separationMOF-74[148]
CO2 captureMOF-5[149]
MechanochemicalIn this type of synthesis, chemical transformation is preceded by the mechanical rupture of intermolecular bonds [139,140]Drug deliveryCu-MOF/IBU@GM[150]
Drug deliveryZIF-8@alginate NPs[151]
SonochemistryThis methodology uses frequencies between 10 MHz and 20 KHz, which are higher than those detectable by the human ear, for dispersion and agglomeration purposes [102]Adsorption of antibiotics[Zn6(IDC)4(OH)2(Hprz)2]n[152]
DMNP hydrolysis and BPA adsorptionMOF-525 and MOF-545[143]
Table 2. General applications of composite metal–organic frameworks (MOFs) reported in the scientific literature, and their main areas of interest, such as environmental and biomedical industries, among others.
Table 2. General applications of composite metal–organic frameworks (MOFs) reported in the scientific literature, and their main areas of interest, such as environmental and biomedical industries, among others.
N0MOFsEnzymeApplicationsRef.
1ZIF-90/Ce-MOFCatalaseSensitive detection and degradation of hydrogen peroxide[219,220]
2L-MOFsGlucose oxidaseInsulin delivery[85,221]
3PCN-333(Fe)Alcohol DehydrogenaseCatalysis of the conversion of toxic levels of alcohols to aldehydes in cells[181,222]
4MIL-101(Cr)Microperoxidase 8Dual catalytic activity in the selective oxidation of organic molecules[180,223,224]
5AgNC/Mo(II)-NSCholesterol oxidaseDetection and concentration in blood vessels or other body tissues[225,226]
6QDs/CDs@MOFsAscorbate oxidaseImproved ascorbic acid detection[227,228]
7ZIF-8Lactate/glucose oxidaseTumor cell mapping and energy reduction for tumor cycles[229]
8UiO-66LipaseDrug synthesis against venous thromboembolism[230,231]
9OMUiO-66 (Ce)Glutamate oxidaseScreening of specific chiral amino acids in complex biological samples[198,232]
10ZIF-8Glucose oxidaseElectrochemical glucose detection[186]
11MIL-88B-NH2(Cr)TrypsinProtein degradation by enzymatic hydrolysis[99,233]
12ZIF-8Glucose oxidaseElectrochemical glucose detection[99]
13Tb-mesoMOFMbOxidation of ABTS and THB[99,234]
14ZIF-8UreaseSensitive biosensor for urea detection[235]
15CYCU-4TrypsinProtein digestion[99,236]
16HKUST-1PeroxidaseCO2 adsorption[99,213,237]
17UIO66-NH2AcetylcholinesteraseBiosensors for organophosphorus pesticide detection[166,210]
18MOF-199LaccaseRemoval of heavy metals from fluids and aquatic environments[238,239]
19QD-MOFOxidaseDegradation of organic dyes in industrial wastewaters[240,241,242]
20L-MOFsLipaseLuminescent sensors for environmental pollutants[125,243]
21ZIF-90CatalaseEffluent treatment in wastewater[214,216,244]
22ZIF-67Glucose oxidaseAntimicrobial agent[244,245]
23Ce (III)/UiO-66HydrolasesAdsorptive removal of organic dyes from aqueous solutions[214,216]
24ZIF-8Choline oxidaseDetection and removal of water pollutants[215,246]
Table 3. Advantages and disadvantages of different enzyme immobilization strategies in/onto MOF.
Table 3. Advantages and disadvantages of different enzyme immobilization strategies in/onto MOF.
StrategyAdvantagesDisadvantagesRef.
In situ synthesisEasily conducted; requires only mild reaction conditionsNot all MOFs are ideal candidates to the process[252]
Covalent bondingThe enzyme is strongly attached to the surface of the support; several MOFs can be usedIt can change the morphology of the enzyme, altering its activity or even inactivating it[93,251]
Surface immobilization (adsorption)Relative low cost and simple methodologyEnzymes can be easily leached from supports due to variations in pH and temperature[93,254]
EntrapmentGives proteins greater stability against denaturation caused by organic solvents, high temperatures, or sudden changes in pHMass transfer limitations may occur; difficult for substrates to reach the active site of enzymes[93,173]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Silva, A.R.M.; Alexandre, J.Y.N.H.; Souza, J.E.S.; Neto, J.G.L.; de Sousa Júnior, P.G.; Rocha, M.V.P.; dos Santos, J.C.S. The Chemistry and Applications of Metal–Organic Frameworks (MOFs) as Industrial Enzyme Immobilization Systems. Molecules 2022, 27, 4529. https://doi.org/10.3390/molecules27144529

AMA Style

Silva ARM, Alexandre JYNH, Souza JES, Neto JGL, de Sousa Júnior PG, Rocha MVP, dos Santos JCS. The Chemistry and Applications of Metal–Organic Frameworks (MOFs) as Industrial Enzyme Immobilization Systems. Molecules. 2022; 27(14):4529. https://doi.org/10.3390/molecules27144529

Chicago/Turabian Style

Silva, Allison R. M., Jeferson Y. N. H. Alexandre, José E. S. Souza, José G. Lima Neto, Paulo G. de Sousa Júnior, Maria V. P. Rocha, and José C. S. dos Santos. 2022. "The Chemistry and Applications of Metal–Organic Frameworks (MOFs) as Industrial Enzyme Immobilization Systems" Molecules 27, no. 14: 4529. https://doi.org/10.3390/molecules27144529

Article Metrics

Back to TopTop