Next Article in Journal / Special Issue
Catalytic Properties of the Spinel-Like CuxMn3−xO4 Copper Manganese Oxides—An Overview
Previous Article in Journal
Coupling Reactions on Secondary Allylic, Propargylic, and Alkyl Carbons Using Organoborates/Ni and RMgX/Cu Reagents
Previous Article in Special Issue
Preparing Cu2O/Al2O3 Coating via an Electrochemical Method for the Degradation of Methyl Orange in the Process of Catalytic Wet Hydrogen Peroxide Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Copper-Based Metal–Organic Frameworks (MOFs) as an Emerging Catalytic Framework for Click Chemistry

1
School of Chemical Engineering and Physical Sciences, Lovely Professional University, Phagwara 144411, India
2
Department of Chemistry and Centre of Advanced Studies in Chemistry, Panjab University, Chandigarh 160014, India
3
Department of Chemistry, Gujranwala Guru Nanak Khalsa College, Civil Lines, Ludhiana 141001, India
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 130; https://doi.org/10.3390/catal13010130
Submission received: 30 November 2022 / Revised: 27 December 2022 / Accepted: 4 January 2023 / Published: 6 January 2023
(This article belongs to the Special Issue Advances on Catalysts Based on Copper)

Abstract

:
In the extensive terrain of catalytic procedures for the synthesis of organic molecules, metal–organic frameworks (MOFs) as heterogenous catalysts have been investigated in a variety of chemical processes, including Friedel–Crafts reactions, condensation reactions, oxidations, and coupling reactions, and utilized owing to their specific properties such as high porosity, tuneability, extraordinary catalytic activity, and recyclability. The eminent copper-tailored MOF materials can be exceptionally dynamic and regioselective catalysts for click reactions (1,3-dipolar cycloaddition reaction). Considering the fact that Cu(I)-catalyzed alkyne–azide cycloaddition (CuAAC) reactions can be catalyzed by several other copper catalysts such as Cu (II)-β-cyclodextrin, Cu(OAc)2, Fe3O4@SiO2, picolinimidoamide–Cu(II) complex, and Cu(II) porphyrin graphene, the properties of sorption and reusability, as well as the high density of copper-MOFs, open an efficient and robust pathway for regimented catalysis of this reaction. This review provides a comprehensive description and analysis of the relevant literature on the utilization of Cu-MOFs as catalysts for CuAAC ‘click’ reactions published in the past decade.

Graphical Abstract

1. Introduction

Click chemistry is one of the most robust and versatile methodologies currently known to researchers, capable of synthesizing large complex compounds from relatively smaller moieties [1,2]. Due to its advantageous features and facile pathway of synthesis, the methodology has received significant interest globally. Carolyn R. Bertozzi, Morten Meldal, and K. Barry Sharpless in 2022 were awarded the Nobel Prize in Chemistry for their ground-breaking contribution, which was a big step in advancement. The advent of click chemistry has led to the discovery of nitrogen-rich azole heterocycles, which have been shown to be fascinating compounds with broad potential significantly with their pharmacological [3] and chemosensing [4] properties. ‘Click’ reactions, specifically Cu(I)-catalyzed cycloaddition involving a terminal alkyne and an organic azide to synthesize 1,4-disubstituted-1,2,3-triazole, have received a large amount of consideration in the fields of material science, polymer science, peptide chemistry [5], and synthetic organic chemistry over the past few decades. However, due to the fact that alkynes are poor 1,3-dipole acceptors, this reaction cannot proceed regioselectively without a suitable catalyst. Significantly, the presence of Cu(I), which forms bonds with terminal alkynes, greatly enhances the pace, selectivity, and overall efficiency of click reactions [6].
The catalysts tailored from transition metals have the potential to act as fabricators in the assemblage of supplementary intricate structures, yielding complex organic structures with relative ease. Several metal catalysts such as Ru, Ag, and Au with enhanced catalytic activity and selectivity have been described in the scientific literature [7,8]. In the past decade, the 1,3 dipolar cycloaddition reaction for the synthesis of 1,2,3-triazole has received considerable attention and has been investigated using a variety of different types of copper catalysts including Cu(II)-β-cyclodextrin [9], Cu(OAc)2 [10], Fe3O4@SiO2 picolinimidoamide–Cu(II) complex [11], Cu(II) porphyrin graphene [12], bromotris(triphenylphosphine)copper(I) [13], and Cu(II)-supported graphene quantum dots [14]. In spite of this, the majority of the copper catalysts that were discussed before had a number of significant drawbacks, including laborious and pricey preparation processes; the utilization of hazardous solvents and reagents; and, most critically, challenging procedures for catalyst separation [15]. Furthermore, the intensifying public concern for environmental protection has prompted scientists to look for more sustainable synthetic routes to overcome the challenges of recyclability of catalysts and switching to less hazardous solvents/chemicals. To anticipate the ease of the synthesis of 1,2,3-triazole derivatives, the area of organic synthesis has focused significantly on the creation of novel and effective Cu(I)-based MOFs catalysts (Figure 1). Over the past decade, scientists have made great strides to manufacture various MOFs and explore of their use in numerous scientific domains, including energy storages [16], medicinal technology [17], environmental pollution [18], sensing platforms, catalysis, photocatalysis [19], oxidation, hydrogenation [20], drug delivery [21,22], and bioimaging [23] due to their diverse active sites [24,25].
Metal–organic frameworks are strong hybrid emerging materials due to their amenable physical and chemical properties [26]; porous structure consisting of inorganic/supplementary building units (atoms, molecules, and ions of metals); and organic building blocks, known as linkers/bridging ligands (carboxylates, as well as anions such as phosphonate, sulfonate, and heterocyclic chemicals) [27], as shown in Figure 2. The long organic linkers are responsible for the high porosity and great storage capacity of MOFs [28]. In addition, they have the unique capacity to intentionally alter and adjust the form and function of their pores. Moreover, due to the remarkable adaptability of MOF structures, it is possible to synthesize catalytically active sites by employing defect engineering and linker modification strategies [29,30].
The coordination number, geometry, and functional assembly in MOFs play a dominant role in shaping the structure of the framework. MOFs have been found to display a wide range of supplementary binding unit (SBU) geometries, such as the octahedron, the trigonal matrix, the square paddlewheel, and the triangle, each of which has a different number of extension points [27]. As a general rule, a connecting ligand interacts with a metal ion that has several available reactive sites [31]. SBU connections and organic linkers control the ultimate topology of any given MOF such that infinitely stretched polymers or discretely closed oligomeric structures may emerge from a given system [32,33]. This review aimed to assemble copper-embedded MOFs for the synthesis of 1,4-disubstituted-1,2,3-triazoles.

2. MOFs as Heterogenous Catalysts

The transition metal complexes can be utilized as homogenous catalysts due to their extraordinary catalytic action for a prevalent variety of organic reactions, showing regio- and chemoselectivity [34,35]. However, this recompense comes with some drawbacks, namely, the homogenous catalysts can be decomposed during the reaction process, making their recovery a challenging process [36]. The various techniques for heterogenizing a homogenous catalyst involve grafting of polymers, impregnation of solid carriers, introduction to zeolite cavities via the “ship in a bottle” method [37], and the formation of organized structured or non-organized structured mesoporous organic/inorganic hybrid systems [30].
Researchers have primarily concentrated on heterogeneous catalysts, as opposed to homogeneous catalysts, due to the greater reusability and the high selectivity of the heterogenous catalyst [38]. In catalyst research, the gold standard followed currently is a combination of computational methods for predicting catalytic activity and novel synthesis methodologies for creating the appropriate organized frameworks, with constant, well-defined active sites that enable good selectivity while avoiding deactivation [39,40]. A large number of recently discovered heterogeneous catalysts are based on metal–organic frameworks, an attractive new field that uses metal ions as active sites in the generated crystal-like structure, metal complexes that were part of the organic linkers, or supported metal nanoparticles. The microporous MOF (microporous co-ordination polymer) generated as an outcome grounded on metal ion action at nodes or at the ligands emerges instantaneously as a counterpart of the homogeneous catalytic system [41]. Due to their potential use as contributing templates in the production of heterogenous catalysts, metal–organic frameworks have recently been a key area of research [42].

3. Physico-Chemical Properties of MOFs

Since MOFs are composed of a metal ion bound to an organic linker chain, they have outstanding properties in terms of their size, porosity, etc. (Table 1, Figure 3), due to which MOFs can be used in a wide range of applications including catalysis, sensing, and drug delivery, among others. In addition, the MOFs may be changed via regulated integration in conjugation with one-of-a-kind and desirable functional materials, which can then impart the required characteristics onto the MOF structures [43,44]. Organic linkers with conjugated structures facilitate electron transfer between ligands and metals. For use as smart materials, MOFs with dynamic frameworks may undergo reversible structural changes that vastly increase their intended qualities.

4. Methods for the Synthesis of MOFs

Numerous techniques for the manufacture of MOFs have been documented in published works. Some examples of these techniques, which are included in the table below, are microwave-assisted synthesis [50], solvothermal synthesis, sonochemical synthesis, electrochemical synthesis [51], among others (Table 2). The choice of solvent is often influenced by a variety of characteristics, such as the redox potential, reactivity, stability constant, and solubility of the compound [52,53].

5. MOFs as Catalysts for Innumerable Reactions

MOFs are becoming an integral part of the research society since they can be utilized as a low-cost heterogeneous catalyst in many different chemical processes [57]. This can be attributed to MOFs having large peripheral areas, size, shape, enantioselectivity, persistent porosity, and multifunctional ligands, with these being some of the interesting and modifiable features of MOFs [45]. There are several benefits of using MOFs as heterogeneous catalysts, including improved catalytic reactivity, flexibility, and facile tunability [52]. However, the use of MOFs is unfavorable for reactions that need extreme conditions because of their limited chemical and thermal stability [58]. Some of the many organic processes for which MOFs were utilized as catalysts or catalyst support include Friedel–Crafts reactions [59]; Knoevenagel condensation [60]; aldol condensation [61]; oxidation [51,62]; and coupling reactions [63,64,65].
Nguyen et al. described a solvothermal acylation process using IRMOF-8 for the Friedel–Crafts reaction, wherein 2,6-naphthalenedicarboxylic acid and zinc nitrate tetrahydrate are dissolved in dimethylformamide (DMF) for the synthesis of IRMOF-8. Moreover, a high yield of the product was achieved when the catalyst was employed in the reaction and was recovered successfully [59]. Similarly, MOF-5 was synthesized by Phan et al.—1,4-benzene dicarboxylic acid and zinc nitrate hexahydrate were combined in dimethylformamide (DMF) under solvothermal conditions to perform a Friedel–Crafts benzylation of toluene [66]. Gascon and teammates produced MOF with non-coordinated sites with an aniline-like amino group for Knoevenagel condensation with 100% selectivity. Xamena and colleagues studied the catalytic activity of IRMOF-3 and MOF-5 on the Knoevenagel condensation reaction [67]. The alkene oxidation of α-pinene and cyclohexene in a solvent-free environment was investigated by Kholdeeva et al. using MOFs (Fe-MIL-101 and Cr-MIL-10), and the allylic oxidation products were produced. Both Cr-MIL-101 and Fe-MIL-101 exhibited selectivity, with Cr-MIL-101 favoring the formation of α,β-unsaturated ketones, and Fe-MIL-101 forming 2-cyclohexen-1-ol as the desired product [51]. Under ideal circumstances, both catalysts demonstrated improved reusability. Torbina et al. determined the catalytic activity of chromium-based MOFs as a heterogenous catalyst for the oxidation reaction of propylene glycol with tert-butyl hydroperoxide [68]. In another example, Phan et al. employed the very porous metal–organic framework MOF-199 as a heterogeneous catalyst for the Ullmann reaction between phenols and aryl iodides to generate diaryl ethers. After six hours at 120 °C, the product yield was 82%. It was shown that the MOF-199 catalyst may be used several times without significantly losing its catalytic efficiency [69]. Several MOFs are used to catalyze the different reactions reported (Table 3). From this perspective, we focused on a few noteworthy instances that provide a balanced comparison to currently available copper-based MOFs over other MOFs to catalyze a wide variety of reactions.

6. Reactions Catalyzed by Cu-Tailored MOFs

Cu-MOFs have gained a great deal of popularity in the area of study as a result of their high catalytic activity, a wealth of resources, and ease of synthesis, all of which are facilitated by the excessive natural abundance of copper. Copper-containing catalysts provide a promising route for activating starting materials while maintaining reaction selectivity under tunable conditions, allowing for the generation of complex scaffolds from easily accessible building pieces [77,78]. A few of the reaction methodologies that are promoted by Cu(I) scaffolds are CuAAC, oxidation, coupling reactions, and Friedel craft acylation reactions, among others.
Maleki et al. employed the hydrothermal synthesis of Cu-MOF, presenting an effective and environmentally friendly technique of synthesizing tetrazole derivatives using Cu-MOF as a catalyst due to the high crystallinity and purity of HKUST-1 (Cu-MOF) (Figure 4). The structure’s face-centered cubic symmetry can be ascribed to the detail that three-dimeric copper wheels are covalently bonded to each BTC ligand. However, two copper atoms are octahedrally bonded to eight oxygen atoms. As a consequence of activation of the carbonyl and ethoxy groups during synthesis, Cu-MOF demonstrated good catalytic activity in three- and four-component processes involving aldehydes, amines, triethyl orthoformate, -keto-esters, and 2-cyano-guanidine [79]. Using aromatic aldehydes, malononitrile, 2-hydroxynaphthalene-1,4-dione, and cycloketone in a one-pot domino four-component reaction using a copper-based MOF as a heterogeneous catalyst and aluminum chloride as a Lewis acid, Taheri and co-workers examined the synthesis of tacrine derivatives. The reaction for the synthesis of MOF was completed at room temperature in the presence of a Cu(II) nitrate trihydrate and 1,4-benzenedioic acid in DMF. The use of Cu-MOF as the Lewis acid catalyst might facilitate the production of pyranic intermediates, which would then be followed by the Friedlander quinoline reaction being supplemented with AlCl3 in order to produce the desired products. In addition to higher yields, utilizing Cu-MOF as a heterogeneous catalyst for the production of the required products (Figure 4) also eliminates the need for a tremendous separation procedure for pyranic intermediates [80].
Ghasemzadeh et al. reported a condensation process between curcumin, aromatic aldehydes, and barbituric acid, used to efficiently synthesize novel pyrano [2,3-d] pyrimidine2,4-diones in a single step. The magnetic CoFe2O4@OCMC@Cu(BDC) showed superior catalytic activity in the generation of heterocyclic compounds (Figure 4) and had a remarkable degree of reusability through at least six cycles. The CoFe2O4@OCMC@Cu(BDC) catalyst may play a crucial role in promoting the reactions with high efficiency due to its abundance of Lewis acid sites (Co2+, Fe3+, and Cu2+) [81]. Moreover, a green three-component reaction carried between aldehydes, dimedone, and aryl amines/2-naphthol/urea to produce a wide range of substituted xanthene derivatives catalyzed by CoFe2O4@OCMC@Cu(BDC) MOF is reported by Ghasemzadeh and Ghaffarian. A metal–organic framework was shown to have a flower-like shape with surface area, total pore volume, and average pore diameters of 64,933 m2g−1, 0.186 cm3g−1, and 11.46 nm, respectively. This approach had several benefits, including good yields, clean and ecologically friendly settings, minimal catalyst loading, simple catalyst separation, and a short reaction time [82].
Three-component coupling and cyclization processes for the synthesis of propargylamines, indoles, and imidazopyridines were developed by Corma et al. via copper-containing MOF [Cu(2-pymo)2]. These crystal structures show that copper ions interact with adsorption substrates more efficiently attributable to a highly deformable, three-dimensional sodalite-like framework consisting of two Cu2+ ions attached to each heterocyclic ligand (pyrimidine, imidazole, or benzene). The MOFs possessed this characteristic, making them superior and more active than conventional N-heterocyclic copper coordination complexes. Using a three-component coupling process of amines, aldehydes, and alkynes catalyzed by [Cu(2-pymo)2], the propargylamine derivatives were synthesized with good yield and selectivity. On the other hand, the [Cu(BDC)] catalyst performed admirably in the 5-exo-dig cyclization of 2-aminopyiridine, aldehydes, and terminal alkynes to produce imidazopyridines (Figure 5) [79].
Figure 4. Different reactions catalyzed by Cu-MOFs (a) formation tetrazole from HKUST-1 MOF3 (b) three component coupling and cyclization reaction using Cu-MOF (c) single step three component coupling and cyclization reaction using CoFe2O4@OCMC@Cu(BDC) [82].
Figure 4. Different reactions catalyzed by Cu-MOFs (a) formation tetrazole from HKUST-1 MOF3 (b) three component coupling and cyclization reaction using Cu-MOF (c) single step three component coupling and cyclization reaction using CoFe2O4@OCMC@Cu(BDC) [82].
Catalysts 13 00130 g004
Figure 5. Schematic representation of the synthesis of (a) xanthene, (b) propargylamines, and (c) imidazopyridines using various copper MOFs as the catalysts [79].
Figure 5. Schematic representation of the synthesis of (a) xanthene, (b) propargylamines, and (c) imidazopyridines using various copper MOFs as the catalysts [79].
Catalysts 13 00130 g005
The use of copper-tailored metal–organic frameworks has attracted a large amount of interest in a wide variety of processes such as condensation reactions, coupling reaction, and oxidation, but especially in click reactions [83]. It can be concluded that the reactions catalyzed by Cu-MOFs are significantly more versatile and robust than other metal catalysts, as well as the fact that Cu-MOFs have high recyclability, as depicted in Table 4. Furthermore, Cu(I) may coordinate with both hard and soft donor ligands in its coordination sphere, facilitating interactions between ligands of varying compositions [84,85].

7. Reaction Mechanism for CuAAC Using MOF Catalyst

The CuAAC reaction stands out from other click reactions, owing to its remarkable process that involves the rapid devising of a super-reactive coordination complex of alkynes and azide utilizing a copper(I) catalyst [95]. However, the catalyst in CuAAC triggers the reaction and brings the reactants together even at extremely low concentrations that are physiologically relevant, while without the catalyst, the reaction rate and selectivity are poor [84]. Since the deprotonation of terminal alkynes is required in the CuAAC reaction process, to produce the Cu–alkynyl intermediate, the procedure can only be carried out using terminal alkynes. This intermediate makes a connection with azide during the subsequent step, which results in the formation of a six-membered cyclic intermediate that rapidly converts into a stable triazole.
The chemical mechanism demonstrating the synthesis of 1,4-disubstituted 1,2,3-triazole using Cu alkynyl intermediate was suggested by Sharpless and colleagues. However, in addition to increased regioselectivity, a strong yield and rapid kinetics of the reaction are also two of the most significant advances that have been documented by a variety of scientists. Cu in the +1 oxidation state can be generated in situ from CuSO4.5H2O and sodium ascorbate or utilized directly as CuI, [Cu(PPh3)3Br], CuBr, Cu2O, and other compounds [96,97,98]. A mononuclear mechanism for the reaction’s progression was initially reported by Sharpless and colleagues; however, as scientific knowledge expanded, a binuclear mechanism was discovered and reported by Finn and Folkin. This mechanism involves two copper atoms binding to the alkyne in different positions to form a complex, which is then attacked by the nucleophilic azide to form a six-membered ring. The creation of a new C-N bond then yields the desired result, as highlighted in Figure 6.
In fact, over the past several decades, A3 coupling reactions, synthesis of tetrazoles, and condensation reactions have garnered a large amount of attention. The mechanism for the A3 coupling reaction involves the activation of the C-H bond by the metal catalyst, which is a key step in the reaction process for A3 coupling. The terminal alkyne is made more acidic due to the formation of a metal complex with the alkyne proton. With a weak base, the amines in the reaction medium deprotonate the terminal alkyne to produce the alkynyl nucleophile. Metal acetylide interacts with iminium ions to make propargylamine, while simultaneously the catalyst is regenerated for use in a subsequent reaction cycle [99], as illustrated in Figure 7.
Tetrazoles are a family of synthetic organic heterocyclic compounds that are made up of a five-member ring that contains four nitrogen atoms and one carbon atom. The formal [3 + 2] cycloaddition of azides and nitriles is the simplest and most efficient way to produce tetrazoles, which are finding widespread use in pharmaceuticals as bioisosteres, information recording systems, etc. Since the nitrile group is activated in the tetrazole synthesis by its coordination with nitrogen and the triple bond, increasing the electrophilic character of the cyanide group, these intermediates speed up the 3 + 2 cycloaddition process. The complex is subsequently exposed to sodium azide, resulting in the formation of a second intermediate that is acidified to provide the final tetrazole product (Figure 8) [100].
The condensation reaction mechanism proceeds with the formation of Lewis acid sites in the vacant orbitals of Zn-MOF caused by Zn(II). This is due to the interaction of Zn(II) with the unpaired electrons on the carbonyl oxygen, suggesting that the O atom’s positive charge may promote further polarization of the carbonyl group. Meanwhile, the hydroxyl O atom connected to the Zn(II) ion acted as the central Lewis basic site, stimulating the active methylene of malononitrile to donate protons and create a carbon anion near the benzaldehyde. The carbonyl C atom of benzaldehyde was then reacted with malononitrile to generate the final product (Figure 9) [101].

8. Copper-Based MOFs as Click Catalysts for Spectacular CuAAC

The Cu(I)-catalyzed azide-alkyne cycloaddition (CuAAC) process is one of the finest procedures for the regioselective production of 1,4-disubstituted-1,2,3-triazoles (Figure 10) in either aqueous or organic media. The wide variety of uses for 1,2,3-triazoles—as pharmaceuticals, agrochemicals, and photographic materials—have attracted a large amount of attention [95]. Furthermore, the copper (II) complex is reduceable to copper(I) in situ in the presence of a reducing agent such as sodium ascorbate, which is one of the most communal and unswerving methodologies used in the click reaction that occurs during the assembly of diverse compounds. Several researchers have discovered and reported the uses of Cu-MOFs that make them one of the great catalysts, but not much work has been done on this topic thus far. The HKUST-1 MOF has been found to contain both Cu(II) and Cu(I) ions, proving the existence of a mixed valence state. More importantly, many scientists have demonstrated that the CuMOF exhibited a distribution of double pore sizes and had excellent water stability. Indeed, these researches show the way toward a better understanding of the catalytic activity of Cu-based MOFs [80].
A novel and highly efficient copper-containing metal–organic framework has been reported by Li et al. for the regioselective synthesis of 1,2,3-triazoles without the use of solvents, yielding an excellent quantity of product, with catalyst loading being comparatively low. However, little efforts were required to recover the catalyst and it was further used for up to five cycles without appreciably diminishing its reactivity (Table 5) [102]. The catalytic activity of CuBTC–PyDC by Fan et al. [103] (Table 5) was investigated in the CuAAC reaction between benzyl azide and phenylacetylene to form 1-benzyl-4-phenyl 1,2,3-triazole. Lu et al. synthesized a unique copper(I)-based metalorganic framework (MOF), exhibiting an intriguing 3D architecture using Keggin-type polyoxometalate anions and a wheel-like resorcin [4]arene-based ligand (Table 5). With an abundance of solvent-accessible empty space and possible active Cu(I) sites, Xu et al. described the Cu-MOF (Table 5) as a catalyst from a three-anion-dependent Cu(I) coordination network (Figure 11). The catalyst’s structure, recycling potential, and catalytic activity may be studied in relation to anions, allowing for fine-tuning of the catalyst’s porosity and potential active metal sites [104].
Murugan et al. reported the synthesis of a metal–organic framework (Cu-MOF) (Table 5) based on ultrasmall cubic copper nanoclusters with relative ease to catalyze the CuAAC reaction in an efficient and chemoselective manner. Selective stepwise conversion of the terminal azide afforded the highest yields in the 70–88% range for the unsymmetric bistriazoles arising from the click reaction of bifunctional azides and alkynes in the presence of Cu-MOFs (Figure 12). The bifunctional terminal contacts through thiol and sulfonate groups supplied the site-isolation-based active sites for selective catalysis, which explains how Cu-MOFs may perform selective mono-click reactions. Because Cu-MOFs may be recovered easily and reused up to five times without substantial loss of activity, they have great potential for selective organic conversions in medicinal and industrial formulations [105].
In order to accomplish the tandem click reactions with excellent yield under optimal circumstances, Arnanz et al. created a bifunctional metal–organic framework catalyst incorporating palladium and copper(II) benzene-1,3,5-tricarboxylate MOF-Cu(BTC)-[Pd] (Table 5). The development of thermally stable pyridine species grafted on the surface is attributed to the CuBTC’s inherent chelating capability, which is due to its electron-rich functional groups. A pyridine-Pd-grafted bifunctional catalyst with high activity for one-pot click reactions was prepared [107]. In a recent study, Jayaramulu and colleagues revealed a straightforward approach for stabilizing Cu2O nanoparticles of size 2–4 nm on the polar pore surface of a 2D layered metal–organic framework [Zn(Himdc)(bipy)0.5]DMF (Table 5). The 1,3,-dipolar cycloaddition reaction of terminal alkynes and aliphatic/aromatic azides to produce 1,2,3-triazoles has been catalyzed by a Cu2O@MOF composite (Figure 13), which will be recycled after use [108].
The catalytic activity of Cu(INA)2-MOF in CuAAC (copper-catalyzed azide-alkyne cycloaddition) was reported by Wiling and colleagues. The simple procedures enabled 1,2,3-triazoles to be prepared in a solvent-free environment. Even after five catalytic cycles, the Cu(INA)2-MOF crystal structure remained constant if the reaction could be carried out on a gramme scale at the optimized conditions to better understand the synthetic applicability of the catalyst [86]. On the basis of a novel premodified ligand (n-pentadecyl-attached imidazolium (IM) decorated dicarboxylic acid) (Table 5) and ZrCl4, Hu et al. reported the formation of a CuI-loaded and n-pentadecyl-attached, imidazolium-salt decorated, UiO-67-type metal–organic framework (CuI@UiO-67-IM) (Figure 14) [106].
Jiang et al. described the azide–alkyne cycloaddition reactions carried out with the help of Cu(I,II)–porphyrin networks as heterogenous catalysts that were produced and effectively implemented. It is anticipated that the efficiency of the mixed valent MOFs as catalysts would increase when they are implied directly. To our astonishment, however, analogous azide–alkyne cycloaddition studies employing porphyrin-based MOFs were investigated and were found to be the effective methodology for the synthesis of various triazole derivatives with high yield (Table 5) [109].
Arefi et al. reported the development of a core–shell magnetic MOF catalyst by the combination of Fe3O4 MNPs (magnetic nanoparticles) covered with metal–organic frameworks (MOFs) composed of copper nodes and 1,3,5-benzenetricarboxylic acid linkers (Table 5, Figure 15). Good-to-exceptional yields may be achieved when using water as the green solvent together with the Cu moieties in MOF and shell structures to catalyze the production of 1,2,3-triazole derivatives. In addition, the superparamagnetic characteristics of this catalyst indicated a high degree of reusability [76].
A mesoporous 2D Cu-MOF (Figure 16) comprising a 2D sheet with a square grid structure (Table 2) was created by Naskar et al. and supported by structural investigations. This porous grid effectively enables organic transformations including click reactions in the aqueous methanolic media by acting as a heterogeneous green catalytic agent with superficial recovery and heat resistance [60].
The Cu(II) coordination polymer synthesized by Bikas and colleagues had an azido bridge and a 1,3-oxazolidine-based ligand. The molecule was then used as a catalyst in a one-pot, three-component cycloaddition process that combined epoxide, azide, and an alkyne to generate -hydroxy-1,2,3-triazoles by employing water as the solvent in their catalytic processes because of its non-toxic nature, inexpensive cost, and negligible effect on the environment. The synthesized 1D azido-bridged Cu(II) coordination polymer (Table 6) was found to be an effective catalyst for the synthesis of hydroxy-1,2,3-triazoles [112]. Likewise, another 3D Cu-based MOF (Figure 17) was reported for the smooth synthesis of 1,4-disubstituted 1,2,3-triazole [113].
In a Huisgen 1,3-dipolar cycloaddition method, the porous MOF [Cu2(BDC)2(DABCO)] was employed as a catalyst to create a wide range of 1,2,3-triazole derivatives. This was done so that the technology could be used to its fullest potential, allowing the reaction to be completed successfully. In this work, Cu2(BDC)2(DABCO) was synthesized using 1,4-benzenedicarboxylic acid, diacetyl bis(2-aminoethyl) ketone, and copper(II) acetate as reactants. This method was executed at room temperature a in solvent-free environment. The catalyst has a large surface area (1012 m2g−1) and a small average pore size (3.9 nm), making it very effective for the synthesis of 1,2,3-triazole analogues with good regioselectivity (Table 6). By using the microdilution approach, the determination of the synthesized triazoles as useful skeletons was performed in order to demonstrate adequate antibacterial and antifungal properties against a variety of pathogenic organisms, including S. aureus, E. coli, and C. albicans [87].
Wang et al. reported a novel functionalized thiacalix [4]arenes (L), and halogen anions were used in a solvothermal assembly of CuI-based MOFs [Cu4Cl4L]CH3OH1.5H2O (1-Cl). In addition, the synthesized MOF has the potential to be a reusable heterogeneous catalyst that can catalyze click reactions in water at very high efficiency (Table 6, Figure 18). The enhanced catalytic activity may be attributed in large part to the MOF structures, particularly the exposed active CuI sites and 1D channels [118].
Yamane et al. reported the synthesis of metal nanoparticles that are supported on MOF-derived carboniferous substances using the HPHT synthesis procedure (Table 6). Cu-BTC was modified for conversion into Cu@C (copper carbon composite) at low temperatures, and the composites were pyrolysed HPHT methods. After HPHT treatment, the composite samples were analysed for their catalytic activity in cycloaddition reactions. Phenyacetylene and benzyl azide were reacted in the presence of 1,4-dioxane as a solvent and triethylamine as a base using Cu@C as the catalyst, obtaining the desired stereoselective product with good yield (Figure 19) [119].
Fu et al. created a copper MOF in reduced form (rCu-MOF). Many ‘click’ reactions were carried out with high yield using this reduced Cu-MOF (Figure 20). The rCu-MOF catalyst was shown to be more active than the traditional copper halide catalyst, with the added benefit of being easily recyclable (Table 6). rCu-MOF is appealing as a catalytic material for industrial applications due to these features as well as due to its simplicity of synthesis, absence of metallic contamination, and great mechanical constancy. This research paves the way for further exploration of rCu-MOFs in polymer chemistry [120].
The one-pot click reaction was reported by Luz et al. using [Cu(2-pymo)2] MOF as a catalyst for the 1,3-dipolar cycloaddition reaction between benzyl azide and phenylacetylene at 70 °C in ethanol [30]. The 3-propargyl-4H-pyrano[2,3-d] pyrimidine compounds were synthesized through a click synthesis involving azide (tetra-O-acetyl—D-glucopyranosyl) and the metal–organic framework Cu@MOF-5 (Table 7). The resulting 1,2,3-triazole product was tested for its ability to inhibit the protein tyrosine phosphatase B (MtbPtpB) activity of Mycobacterium tuberculosis. The most active compound’s kinetic experiments revealed competitive inhibition of the MtbPtpB enzyme. The strong inhibitory effect against MtbPtpB was believed to depend on the interaction of Arg63 amino acids with an anion-type para-hydroxyl group through a salt bridge of iminium cation, according to an extensive structure–activity relationship (SAR) in vitro and in silico investigations. The experimental evidence obtained by cross-docking showed that the chemical synthesized was more potent against the mycobacterium TB protein tyrosine phosphatase B. It was found by docking that the amino acids were inside the obligatory site and were likely hotspots for ligand binding (Figure 21) [121].
Click chemistry has developed into a highly effective toolbox for manipulating biological processes in their native settings. The copper(I)-catalyzed azide–alkyne cycloaddition (CuAAC) reaction is a common click process utilized in drug production and design. For in vivo tumor treatment, Wang et al. built a heterogeneous copper catalyst using metal–organic frameworks (Table 7) that preferentially accumulated in the mitochondria of live cells and showed significant catalytic activity, as illustrated in Figure 22. To execute the targeted drug production in mitochondria, the catalyst may be employed to activate the profluorophobic unit. There is no better way to reduce cytotoxic medications’ deleterious side effects while increasing their effectiveness than the in situ synthesis of active pharmaceuticals from inert prodrugs in subcellular organelles [123].
The reported copper metal–organic frameworks for the synthesis of 1,2,3-triazole appended probes via CuAAC are often recyclable and have a high reusability of up to four or five times. Additionally, the catalysts are also recoverable at an astounding pace [111,113,114,115,116]. Depending on the reactants and the reaction conditions, the quantity of MOF catalysts may be anywhere ranging from 0.5 mol % to 10 mol % or 0.2 mg to 30 mg. The desired products were obtained with excellent yields of more than 90%, and in some cases, an exceptional yield of 100% was reported, as highlighted in Table 5 (entries 6, 9, and 11) [105,107,109]. Cu-MOFs have relatively outstanding chemical stability, particularly in organic solvents and aqueous solutions, due to the involvement of Cu(I) sites, making the MOF a heterogeneous catalyst for the azide–alkyne cycloaddition (AAC) process [7,8].

9. Challenges in the Synthesis and Applications of Cu-Based MOFs

The challenges in the synthesis of Cu-based MOFs emerge from the need to comprehend and manage the inherent complexity of the many different topologies and compositions of MOFs. However, Cu has dynamic stability and requires more basic compounds, which often leads to the carboxylate products and waste disposal issues and separation challenges [120]. Moreover, due to the microporous structure of Cu-MOFs and MOF-based materials, mass transfer constraints occur in large molecules, which limits researchers from advancing the field of Cu-MOF synthesis. Hierarchical porosity can be introduced to resolve the problem of mass transfer. The key challenges in the application of Cu-MOFs are that they are not very mechanically stable due to their increased pore size, which sometimes hinders their use as catalysts in practical catalytic reactions [102]. This makes it necessary to focus on increasing their mechanical and chemical stability to preserve catalytic efficacy and selectivity [111]. After reviewing the relevant literature, we conclude that MOFs and MOF-derived materials are promising starting points for investigating catalytic applications, even though they provide both challenges and opportunities.

10. Conclusions and Future Aspects

There has been exponential growth in recent years in the synthesis and application of novel MOFs with exciting new architectures and functionalities. Thus, it is reasonable to anticipate that MOFs will continue to provide increasingly more efficient responses to issues in contemporary organic chemistry and modern organometallic catalysis, since the catalytic processes they mediate have grown more pervasive over the previous years. Copper-tailored metal–organic frameworks have gained a large amount of attention in the diversity of reactions, especially in CuAAC, since the MOF catalysts can be easily recovered and used multiple times, owing to their structure, durability, and porosity. This review focused on the copper-based MOFs that have been widely employed for the synthesis of 1,2,3-triazoles across a wide range of reaction conditions, with benefits including improved yield and reduced reaction time, as well as the effective recovery of the catalyst. MOF catalysts enhance the synthesis of such compounds that would otherwise be difficult to produce without the aid of a catalyst, or for which reaction improvements would be beneficial. Since MOFs are capable of catalysis, future research and industrial development stand to benefit from the synthesis of ever more creative and recyclable MOFs in the forthcoming years.

Author Contributions

Conceptualization, J.S. and R.S.; methodology, R.S. and G.S. (Gurleen Singh); validation, J.S. and G.K.; formal analysis, N.G., S.G.; investigation, G.S. (Gurjaspreet Singh), G.S. (Gurleen Singh); resources, J.S., writing—original draft preparation, R.S..; writing—review and editing, J.S.; visualization, H.S. and G.K.; supervision, J.S. and G.K.; project administration, J.S.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were generated.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MOFMetal–organic framework
ImImidazolate
BTCBenzene tricarboxylate
BDCBenzene dicarboxylate
PyDCPyridine-3,5-dicarboxylate
TPPTriphenylphosphonium
CPA4-Chloro-phenyl)-pyridin-4-ylmethylene-amine
Himdc4,5-Imidazoledicarboxylate
Bipy4,4′-Bipyridine
DABCO1,4-Diazabicyclo [2.2.2] octane
2-pymo2-Hydroxypyrimidinolate
SBUSupplementary building units
OCMCO-Carboxymethyl chitosan
IMImidazolium
HPHTHigh pressure and high temperature
MtbPtpBMycobacterium tuberculosis protein tyrosine phosphatase B

References

  1. Kolb, H.C.; Finn, M.G.; Sharpless, K.B. Click Chemistry: Diverse Chemical Function from a Few Good Reactions. Angew. Chem. Int. Ed. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  2. Hou, J.; Liu, X.; Shen, J.; Zhao, G.; Wang, P.G. The Impact of Click Chemistry in Medicinal Chemistry. Expert Opin. Drug Discov. 2012, 7, 489–501. [Google Scholar] [CrossRef] [PubMed]
  3. Kolb, H.C.; Sharpless, K.B. The Growing Impact of Click Chemistry on Drug Discovery. Drug Discov. Today 2003, 8, 1128–1137. [Google Scholar] [CrossRef]
  4. Singh, G.; George, N.; Singh, R.; Singh, G.; Sushma; Kaur, G.; Singh, H.; Singh, J. Ion Recognition by 1,2,3-Triazole Moieties Synthesized via “Click Chemistry”. Appl. Organomet. Chem. 2022, 37, e6897. [Google Scholar] [CrossRef]
  5. Ahmad Fuaad, A.A.H.; Azmi, F.; Skwarczynski, M.; Toth, I. Peptide Conjugation via CuAAC ‘Click’ Chemistry. Molecules 2013, 18, 13148–13174. [Google Scholar] [CrossRef] [Green Version]
  6. Hein, J.E.; Fokin, V.V. Copper-Catalyzed Azide–Alkyne Cycloaddition (CuAAC) and beyond: New Reactivity of Copper(I) Acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef]
  7. Kalra, P.; Kaur, R.; Singh, G.; Singh, H.; Singh, G.; Pawan; Kaur, G.; Singh, J. Metals as “Click” Catalysts for Alkyne-Azide Cycloaddition Reactions: An Overview. J. Organomet. Chem. 2021, 944, 121846. [Google Scholar] [CrossRef]
  8. Saini, P.; Sonika; Singh, G.; Kaur, G.; Singh, J.; Singh, H. Robust and Versatile Cu(I) Metal Frameworks as Potential Catalysts for Azide-Alkyne Cycloaddition Reactions: Review. Mol. Catal. 2021, 504, 111432. [Google Scholar] [CrossRef]
  9. Kaboudin, B.; Mostafalu, R.; Yokomatsu, T. Fe3O4 Nanoparticle-Supported Cu(II)-β-Cyclodextrin Complex as a Magnetically Recoverable and Reusable Catalyst for the Synthesis of Symmetrical Biaryls and 1,2,3-Triazoles from Aryl Boronic Acids. Green Chem. 2013, 15, 2266–2274. [Google Scholar] [CrossRef]
  10. Zhang, W.; He, X.; Ren, B.; Jiang, Y.; Hu, Z. Cu(OAc)2·H2O—An efficient catalyst for Huisgen-click reaction in supercritical carbon dioxide. Tetrahedron Lett. 2015, 56, 2472–2475. [Google Scholar] [CrossRef]
  11. Zirak, M.; Garegeshlagi, E.J. Picolinimidoamide-Cu(II) complex anchored on Fe3O4@SiO2 core–shell magnetic nanoparticles: An efficient reusable catalyst for click reaction. J. Coord. Chem. 2018, 71, 1168–1179. [Google Scholar] [CrossRef]
  12. Khojastehnezhad, A.; Bakavoli, M.; Javid, A.; Khakzad Siuki, M.M.; Shahidzadeh, M. Synthesis, characterization, and investigation of catalytic activity of copper(II) porphyrin graphene oxide for azide–alkyne cycloaddition. Res. Chem. Intermed. 2019, 45, 4473–4485. [Google Scholar] [CrossRef]
  13. Kaur, P.; Lal, B.; Kaur, N.; Singh, G.; Singh, A.; Kaur, G.; Singh, J. Selective Two Way Cd(II) and Co(II) Ions Detection by 1,2,3–Triazole Linked Fluorescein Derivative. J. Photochem. Photobiol. Chem. 2019, 382, 111847. [Google Scholar] [CrossRef]
  14. Deilam, R.; Moeinpour, F.; Mohseni-Shahri, F.S. Catalytic performance of Cu(II)-supported graphene quantum dots modified NiFe2O4 as a proficient nano-catalyst in the synthesis of 1,2,3-triazoles. Mon. Chem. 2020, 151, 1153–1162. [Google Scholar] [CrossRef]
  15. Jin, T.; Yan, M.; Yamamoto, Y. Click Chemistry of Alkyne-Azide Cycloaddition using Nanostructured Copper Catalysts. ChemCatChem 2012, 4, 1217–1229. [Google Scholar] [CrossRef]
  16. Chen, Y.; Zhang, Y.; Huang, Q.; Lin, X.; Zeb, A.; Wu, Y.; Xu, Z.; Xu, X. Recent Advances in Cu-Based Metal–Organic Frameworks and Their Derivatives for Battery Applications. ACS Appl. Energy Mater. 2022, 5, 7842–7873. [Google Scholar] [CrossRef]
  17. Singh, N.; Thakur, A. Applications of Copper Based Metal Organic Frameworks. Mater. Today Proc. 2022, 50, 1906–1911. [Google Scholar] [CrossRef]
  18. Li, L.; Zou, J.; Han, Y.; Liao, Z.; Lu, P.; Nezamzadeh-Ejhieh, A.; Liu, J.; Peng, Y. Recent Advances in Al(III)/In(III)-Based MOFs for the Detection of Pollutants. New J. Chem. 2022, 46, 19577–19592. [Google Scholar] [CrossRef]
  19. Dong, X.; Li, Y.; Li, D.; Liao, D.; Qin, T.; Prakash, O.; Kumar, A.; Liu, J. A New 3D 8-Connected Cd(II) MOF as a Potent Photocatalyst for Oxytetracycline Antibiotic Degradation. CrystEngComm 2022, 24, 6933–6943. [Google Scholar] [CrossRef]
  20. Guo, M.; Zhang, M.; Liu, R.; Zhang, X.; Li, G. State-of-the-Art Advancements in Photocatalytic Hydrogenation: Reaction Mechanism and Recent Progress in Metal-Organic Framework (MOF)-Based Catalysts. Adv. Sci. 2022, 9, 2103361. [Google Scholar] [CrossRef]
  21. 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]
  22. Qin, L.; Liang, F.; Li, Y.; Wu, J.; Guan, S.; Wu, M.; Xie, S.; Luo, M.; Ma, D. A 2D Porous Zinc-Organic Framework Platform for Loading of 5-Fluorouracil. Inorganics 2022, 10, 202. [Google Scholar] [CrossRef]
  23. Butler, K.S.; Pearce, C.J.; Nail, E.A.; Vincent, G.A.; Gallis, D.F.S. Antibody Targeted Metal–Organic Frameworks for Bioimaging Applications. ACS Appl. Mater. Interfaces 2020, 12, 31217–31224. [Google Scholar] [CrossRef]
  24. Song, Z.; Zhang, L.; Doyle-Davis, K.; Fu, X.; Luo, J.-L.; Sun, X. Recent Advances in MOF-Derived Single Atom Catalysts for Electrochemical Applications. Adv. Energy Mater. 2020, 10, 2001561. [Google Scholar] [CrossRef]
  25. Zhang, W.; Ye, G.; Liao, D.; Chen, X.; Lu, C.; Nezamzadeh-Ejhieh, A.; Khan, M.S.; Liu, J.; Pan, Y.; Dai, Z. Recent Advances of Silver-Based Coordination Polymers on Antibacterial Applications. Molecules 2022, 27, 7166. [Google Scholar] [CrossRef]
  26. Baumann, A.E.; Burns, D.A.; Liu, B.; Thoi, V.S. Metal-Organic Framework Functionalization and Design Strategies for Advanced Electrochemical Energy Storage Devices. Commun. Chem. 2019, 2, 86. [Google Scholar] [CrossRef] [Green Version]
  27. Sharmin, E.; Zafar, F.; Sharmin, E.; Zafar, F. Introductory Chapter: Metal-organic Frameworks (MOFs); IntechOpen: London, UK, 2016; ISBN 978-953-51-2663-8. [Google Scholar] [CrossRef] [Green Version]
  28. Kirchon, A.; Feng, L.; Drake, H.F.; Joseph, E.A.; Zhou, H.-C. From fundamentals to applications: A toolbox for robust and multifunctional MOF materials. Chem. Soc. Rev. 2018, 47, 8611. [Google Scholar] [CrossRef] [PubMed]
  29. Li, P.-Z.; Wang, X.-J.; Zhao, Y. Click chemistry as a versatile reaction for construction and modification of metal-organic frameworks. Coord. Chem. Rev. 2019, 380, 484–518. [Google Scholar] [CrossRef]
  30. Luz, I.; i Xamena, F.X.L.; Corma, A. Bridging Homogeneous and Heterogeneous Catalysis with MOFs: “Click” Reactions with Cu-MOF Catalysts. J. Catal. 2010, 276, 134–140. [Google Scholar] [CrossRef]
  31. Zhou, H.-C.; Long, J.R.; Yaghi, O.M. Introduction to Metal–Organic Frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef] [PubMed]
  32. Wei, Y.-S.; Zhang, M.; Zou, R.; Xu, Q. Metal–Organic Framework-Based Catalysts with Single Metal Sites. Chem. Rev. 2020, 120, 12089–12174. [Google Scholar] [CrossRef] [PubMed]
  33. Doonan, C.J.; Sumby, C.J. Metal–Organic Framework Catalysis. CrystEngComm 2017, 19, 4044–4048. [Google Scholar] [CrossRef]
  34. Li, S.; Gao, Y.; Li, N.; Ge, L.; Bu, X.; Feng, P. Transition Metal-Based Bimetallic MOFs and MOF-Derived Catalysts for Electrochemical Oxygen Evolution Reaction. Energy Environ. Sci. 2021, 14, 1897–1927. [Google Scholar] [CrossRef]
  35. Kumar, S.; Mohan, B.; Tao, Z.; You, H.; Ren, P. Incorporation of Homogeneous Organometallic Catalysts into Metal–Organic Frameworks for Advanced Heterogenization: A Review. Catal. Sci. Technol. 2021, 11, 5734–5771. [Google Scholar] [CrossRef]
  36. Wang, Q.; Astruc, D. State of the Art and Prospects in Metal–Organic Framework (MOF)-Based and MOF-Derived Nanocatalysis. Chem. Rev. 2020, 120, 1438–1511. [Google Scholar] [CrossRef]
  37. Ali, M.; Pervaiz, E.; Noor, T.; Rabi, O.; Zahra, R.; Yang, M. Recent advancements in MOF- based catalysts for applications in electrochemical and photoelectrochemical water splitting: A review. Int. J. Energy Res. 2020, 45, 1190–1226. [Google Scholar] [CrossRef]
  38. Chassaing, S.; Bénéteau, V.; Pale, P. When CuAAC “Click Chemistry” Goes Heterogeneous. Catal. Sci. Technol. 2016, 6, 923–957. [Google Scholar] [CrossRef]
  39. Chen, B.W.J.; Xu, L.; Mavrikakis, M. Computational Methods in Heterogeneous Catalysis. Chem. Rev. 2021, 121, 1007–1048. [Google Scholar] [CrossRef] [PubMed]
  40. Shi, H.; Zhou, T. Computational Design of Heterogeneous Catalysts and Gas Separation Materials for Advanced Chemical Processing. Front. Chem. Sci. Eng. 2021, 15, 49–59. [Google Scholar] [CrossRef]
  41. Vojvodic, A.; Nørskov, J.K. New design paradigm for heterogeneous catalysts. Natl. Sci. Rev. 2015, 2, 140–149. [Google Scholar] [CrossRef] [Green Version]
  42. Kang, Y.-S.; Lu, Y.; Chen, K.; Zhao, Y.; Wang, P.; Sun, W.-Y. Metal–organic frameworks with catalytic centers: From synthesis to catalytic application. Coord. Chem. Rev. 2019, 378, 262–280. [Google Scholar] [CrossRef]
  43. Butova, V.V.; Soldatov, M.A.; Guda, A.A.; Lomachenko, K.A.; Lamberti, C. Metal-Organic Frameworks: Structure, Properties, Methods of Synthesis and Characterization. Russ. Chem. Rev. 2016, 85, 280–307. [Google Scholar] [CrossRef]
  44. Redfern, L.R.; Farha, O.K. Mechanical Properties of Metal–Organic Frameworks. Chem. Sci. 2019, 10, 10666–10679. [Google Scholar] [CrossRef] [Green Version]
  45. Moosavi, S.M.; Nandy, A.; Jablonka, K.M.; Ongari, D.; Janet, J.P.; Boyd, P.G.; Lee, Y.; Smit, B.; Kulik, H.J. Understanding the Diversity of the Metal-Organic Framework Ecosystem. Nat. Commun. 2020, 11, 4068. [Google Scholar] [CrossRef]
  46. Kabtamu, D.M.; Wu, Y.; Li, F. Hierarchically Porous Metal–Organic Frameworks: Synthesis Strategies, Structure(s), and Emerging Applications in Decontamination. J. Hazard. Mater. 2020, 397, 122765. [Google Scholar] [CrossRef] [PubMed]
  47. Yang, X.; Yan, D. Long-Afterglow Metal–Organic Frameworks: Reversible Guest-Induced Phosphorescence Tunability. Chem. Sci. 2016, 7, 4519–4526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kuang, X.; Ma, Y.; Su, H.; Zhang, J.; Dong, Y.-B.; Tang, B. High-Performance Liquid Chromatographic Enantioseparation of Racemic Drugs Based on Homochiral Metal–Organic Framework. Anal. Chem. 2014, 86, 1277–1281. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, T.; Wang, J.; Zhang, W.; Yang, C.; Zhang, L.; Zhu, W.; Sun, J.; Li, G.; Li, T.; Wang, J. Amorphous Fe/Mn Bimetal–Organic Frameworks: Outer and Inner Structural Designs for Efficient Arsenic(III) Removal. J. Mater. Chem. A 2019, 7, 2845–2854. [Google Scholar] [CrossRef]
  50. McKinstry, C.; Cussen, E.J.; Fletcher, A.J.; Patwardhan, S.V.; Sefcik, J. Scalable Continuous Production of High Quality HKUST-1 via Conventional and Microwave Heating. Chem. Eng. J. 2017, 326, 570–577. [Google Scholar] [CrossRef]
  51. Kholdeeva, O.A.; Skobelev, I.Y.; Ivanchikova, I.D.; Kovalenko, K.A.; Fedin, V.P.; Sorokin, A.B. Hydrocarbon Oxidation over Fe- and Cr-Containing Metal-Organic Frameworks MIL-100 and MIL-101–a Comparative Study. Catal. Today 2014, 238, 54–61. [Google Scholar] [CrossRef]
  52. Chen, Y.; Xiao, J.; Lv, D.; Huang, T.; Xu, F.; Sun, X.; Xi, H.; Xia, Q.; Li, Z. Highly Efficient Mechanochemical Synthesis of an Indium Based Metal-Organic Framework with Excellent Water Stability. Chem. Eng. Sci. 2017, 158, 539–544. [Google Scholar] [CrossRef]
  53. Yang, H.; Orefuwa, S.; Goudy, A. Study of Mechanochemical Synthesis in the Formation of the Metal–Organic Framework Cu3(BTC)2 for Hydrogen Storage. Microporous Mesoporous Mater. 2011, 143, 37–45. [Google Scholar] [CrossRef]
  54. Li, Y.-P.; Zhang, L.-J.; Ji, W.-J. Synthesis, Characterization, Crystal Structure of Magnesium Compound Based 3, 3′, 5, 5′-Azobenzentetracarboxylic Acid and Application as High-Performance Heterogeneous Catalyst for Cyanosilylation. J. Mol. Struct. 2017, 1133, 607–614. [Google Scholar] [CrossRef]
  55. Du, J.; Zou, G. A Novel Microporous Zinc(II) Metal-Organic Framework with Highly Selectivity Adsorption of CO2 over CH4. Inorg. Chem. Commun. 2016, 69, 20–23. [Google Scholar] [CrossRef]
  56. Lestari, W.W.; Arvinawati, M.; Martien, R.; Kusumaningsih, T. Green and Facile Synthesis of MOF and Nano MOF Containing Zinc(II) and Benzen 1,3,5-Tri Carboxylate and Its Study in Ibuprofen Slow-Release. Mater. Chem. Phys. 2018, 204, 141–146. [Google Scholar] [CrossRef]
  57. Czaja, A.U.; Trukhan, N.; Müller, U. Industrial applications of metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1284–1293. [Google Scholar] [CrossRef]
  58. Dey, C.; Kundu, T.; Biswal, B.P.; Mallick, A.; Banerjee, R. Crystalline Metal-Organic Frameworks (MOFs): Synthesis, Structure and Function. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2014, 70, 3–10. [Google Scholar] [CrossRef] [PubMed]
  59. Nguyen, L.T.L.; Nguyen, C.V.; Dang, G.H.; Le, K.K.A.; Phan, N.T.S. Towards Applications of Metal–Organic Frameworks in Catalysis: Friedel–Crafts Acylation Reaction over IRMOF-8 as an Efficient Heterogeneous Catalyst. J. Mol. Catal. Chem. 2011, 349, 28–35. [Google Scholar] [CrossRef]
  60. Naskar, K.; Maity, S.; Maity, H.S.; Sinha, C. A Reusable Efficient Green Catalyst of 2D Cu-MOF for the Click and Knoevenagel Reaction. Molecules 2021, 26, 5296. [Google Scholar] [CrossRef]
  61. Kikhtyanin, O.; Kubička, D.; Čejka, J. Toward Understanding of the Role of Lewis Acidity in Aldol Condensation of Acetone and Furfural Using MOF and Zeolite Catalysts. Catal. Today 2015, 243, 158–162. [Google Scholar] [CrossRef]
  62. Zhao, T.; Dong, M.; Yang, L.; Liu, Y. Synthesis of Stable Hierarchical MIL-101(Cr) with Enhanced Catalytic Activity in the Oxidation of Indene. Catalysts 2018, 8, 394. [Google Scholar] [CrossRef] [Green Version]
  63. Rostamnia, S.; Alamgholiloo, H.; Liu, X. Pd-grafted Open Metal Site Copper-benzene-1,4-dicarboxylate Metal Organic Frameworks (Cu-BDC MOF’s) as Promising Interfacial Catalysts for Sustainable Suzuki Coupling. J. Colloid Interface Sci. 2016, 469, 310–317. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, L.; Tai, X.; Zhou, X. Au3+/Au0 Supported on Chromium(III) Terephthalate Metal Organic Framework (MIL-101) as an Efficient Heterogeneous Catalystfor Three-Component Coupling Synthesis of Propargylamines. Materials 2017, 10, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Hajiashrafi, T.; Karimi, M.; Heydari, A.; Tehrani, A.A. Erbium-Organic Framework as Heterogeneous Lewis Acid Catalysis for Hantzsch Coupling and Tetrahydro-4H-Chromene Synthesis. Catal. Lett. 2017, 147, 453–462. [Google Scholar] [CrossRef]
  66. Phan, N.T.S.; Le, K.K.A.; Phan, T.D. MOF-5 as an Efficient Heterogeneous Catalyst for Friedel–Crafts Alkylation Reactions. Appl. Catal. Gen. 2010, 2, 246–253. [Google Scholar] [CrossRef]
  67. Gascon, J.; Aktay, U.; Hernandez Alonso, M.; van Klink, G.; Kapteijn, F. Amino-Based Metal Organic Frameworks as Stable, Highly Active Basic Catalysts. J. Catal. 2009, 261, 75–87. [Google Scholar] [CrossRef]
  68. Torbina, V.; Salaev, M.; Vodyankina, O. Effect of Solvent Nature on Propylene Glycol Oxidation with Tert-Butyl Hydroperoxide over Metal–Organic Framework Cr-MIL-101. RSC Adv. 2019, 9, 25981–25986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Phan, N.T.S.; Nguyen, T.T.; Nguyen, C.V.; Nguyen, T.T. Ullmann-Type Coupling Reaction Using Metal-Organic Framework MOF-199 as an Efficient Recyclable Solid Catalyst. Appl. Catal. Gen. 2013, 457, 69–77. [Google Scholar] [CrossRef]
  70. Ullah, L.; Zhao, G.; Xu, Z.; He, H.; Usman, M.; Zhang, S. 12-Tungstophosphoric Acid Niched in Zr-Based Metal-Organic Framework: A Stable and Efficient Catalyst for Friedel-Crafts Acylation. Sci. China Chem. 2018, 61, 402–411. [Google Scholar] [CrossRef]
  71. Calleja, G.; Sanz, R.; Orcajo, G.; Briones, D.; Leo, P.; Martínez, F. Copper-based MOF-74 material as effective acid catalyst in Friedel–Crafts acylation of anisole. Catal. Today 2014, 227, 130–137. [Google Scholar] [CrossRef]
  72. Zhu, C.; Mao, Q.; Li, D.; Li, C.; Zhou, Y.; Wu, X.; Luo, Y.; Li, Y. A Readily Available Urea Based MOF That Act as a Highly Active Heterogeneous Catalyst for Friedel-Crafts Reaction of Indoles and Nitrostryenes. Catal. Commun. 2018, 104, 123–127. [Google Scholar] [CrossRef]
  73. Zhu, Y.; Chen, J.; Kaskel, S. Porphyrin-Based Metal-Organic Frameworks for Biomedical Applications. Angew. Chem. Int. Ed. 2020, 60, 5010–5035. [Google Scholar] [CrossRef] [Green Version]
  74. Dong, W.; Zhang, L.; Wang, C.; Feng, C.; Shang, N.; Gao, S.; Wang, C. Palladium Nanoparticle Embedded in Metal Organic Framework Derived Porous Carbon: Synthesis and Application for Efficient Suzuki-Miyaura Coupling Reaction. RSC Adv. 2016, 6, 37118–37123. [Google Scholar] [CrossRef]
  75. Hartmann, M.; Fischer, M. Amino-functionalized basic catalysts with MIL-101 structure. Microporous Mesoporous Mater. 2012, 164, 38–43. [Google Scholar] [CrossRef]
  76. Arefi, E.; Khojastehnezhad, A.; Shiri, A. A Magnetic Copper Organic Framework Material as an Efficient and Recyclable Catalyst for the Synthesis of 1,2,3-Triazole Derivatives. Sci. Rep. 2021, 11, 20514. [Google Scholar] [CrossRef]
  77. Konnerth, H.; Matsagar, B.M.; Chen, S.S.; Prechtl, M.H.G.; Shieh, F.-K.; Wu, K.C.-W. Metal-Organic Framework (MOF)-Derived Catalysts for Fine Chemical Production. Coord. Chem. Rev. 2020, 416, 213319. [Google Scholar] [CrossRef]
  78. Luz, I.; i Xamena, F.L.; Corma, A. Bridging Homogeneous and Heterogeneous Catalysis with MOFs: Cu-MOFs as Solid Catalysts for Three-Component Coupling and Cyclization Reactions for the Synthesis of Propargylamines, Indoles and Imidazopyridines. J. Catal. 2012, 285, 285–291. [Google Scholar] [CrossRef]
  79. Kal-Koshvandi, A.T.; Maleki, A.; Tarlani, A.; Soroush, M.R. Synthesis and Characterization of Ultrapure HKUST-1 MOFs as Reusable Heterogeneous Catalysts for the Green Synthesis of Tetrazole Derivatives. ChemistrySelect 2020, 5, 3164–3172. [Google Scholar] [CrossRef]
  80. Mollabagher, H.; Taheri, S.; Mojtahedi, M.M.; Seyedmousavi, S. Cu-Metal Organic Frameworks (Cu-MOF) as an Environment-Friendly and Economical Catalyst for One Pot Synthesis of Tacrine Derivatives. RSC Adv. 2020, 10, 1995–2003. [Google Scholar] [CrossRef]
  81. Ghaffarian, F.; Ghasemzadeh, M.; Aghaei, S. An efficient synthesis of some new curcumin based pyrano[2,3-d]pyrimidine-2,4(3H)-dione derivatives using CoFe2O4@OCMC@Cu(BDC) as a novel and recoverable catalyst. J. Mol. Struct. 2019, 1186, 204–211. [Google Scholar] [CrossRef]
  82. Ghasemzadeh, M.A.; Ghaffarian, F. Preparation of Core/Shell/Shell CoFe2O4/OCMC/Cu (BDC) Nanostructure as a Magnetically Heterogeneous Catalyst for the Synthesis of Substituted Xanthenes, Quinazolines and Acridines under Ultrasonic Irradiation. Appl. Organomet. Chem. 2020, 34, e5580. [Google Scholar] [CrossRef]
  83. Barea, E.; Navarro, J.A.R.; Salas, J.M.; Masciocchi, N.; Galli, S.; Sironi, A. Mineralomimetic Sodalite- and Muscovite-Type Coordination Frameworks. Dynamic Crystal-to-Crystal Interconversion Processes Sensitive to Ion Pair Recognition. J. Am. Chem. Soc. 2004, 126, 3014–3015. [Google Scholar] [CrossRef]
  84. Meldal, M.; Diness, F. Recent Fascinating Aspects of the CuAAC Click Reaction. Trends Chem. 2020, 2, 569–584. [Google Scholar] [CrossRef]
  85. Basu, P.; Bhanja, P.; Salam, N.; Dey, T.K.; Bhaumik, A.; Das, D.; Islam, S.M. Silver Nanoparticles Supported over Al2O3@Fe2O3 Core-Shell Nanoparticles as an Efficient Catalyst for One-Pot Synthesis of 1,2,3-Triazoles and Acylation of Benzyl Alcohol. Mol. Catal. 2017, 439, 31–40. [Google Scholar] [CrossRef]
  86. Willig, J.C.M.; Granetto, G.; Reginato, D.; Dutra, F.R.; Poruczinski, F.; de Oliveira, I.M.; Stefani, H.A.; de Campos, S.D.; de Campos, A.; Manarin, F.; et al. A Comparative Study between Cu(INA)2-MOF and [Cu(INA)2(H2O)4] Complex for a Click Reaction and the Biginelli Reaction under Solvent-Free Conditions. RSC Adv. 2020, 10, 3407–3415. [Google Scholar] [CrossRef] [Green Version]
  87. Tourani, H.; Naimi-Jamal, M.R.; Panahi, L.; Dekamin, M.G. Nanoporous Metal-Organic Framework Cu2(BDC)2(DABCO) as an Efficient Heterogeneous Catalyst for One-Pot Facile Synthesis of 1,2,3-Triazole Derivatives in Ethanol: Evaluating Antimicrobial Activity of the Novel Derivatives. Sci. Iran. 2019, 26, 1485–1496. [Google Scholar] [CrossRef] [Green Version]
  88. Singh, G.; Mangat, S.S.; Sharma, H.; Singh, J.; Arora, A.; Singh Pannu, A.P.; Singh, N. Design and Syntheses of Novel Fluorescent Organosilicon-Based Chemosensors through Click Silylation: Detection of Biogenic Amines. RSC Adv. 2014, 4, 36834–36844. [Google Scholar] [CrossRef]
  89. Velpuri, V.R.; Muralidharan, K. Multicomponent Click Reaction Catalyzed by Organic Surfactant-Free Copper Sulfide (Sf-CuS) Nano/Micro Flowers. J. Organomet. Chem. 2019, 884, 59–65. [Google Scholar] [CrossRef]
  90. Beghdadi, S.; Miladi, I.A.; Addis, D.; Romdhane, H.B.; Bernard, J.; Drockenmuller, E. Synthesis and Polymerization of C-Vinyl- and N-Vinyl-1,2,3-Triazoles. Polym. Chem. 2012, 3, 1680–1692. [Google Scholar] [CrossRef]
  91. Ju, C.; Meng, C.; Ma, J.; Zhang, X.; Ding, S. Construction of Sequence-Defined Polytriazoles by IrAAC and CuAAC Reactions. Chem. Commun. 2020, 56, 3955–3958. [Google Scholar] [CrossRef]
  92. Chinchilla, R.; Nájera, C. Recent Advances in Sonogashira Reactions. Chem. Soc. Rev. 2011, 40, 5084–5121. [Google Scholar] [CrossRef] [PubMed]
  93. Bozorov, K.; Zhao, J.; Aisa, H.A. 1,2,3-Triazole-Containing Hybrids as Leads in Medicinal Chemistry: A Recent Overview. Bioorg. Med. Chem. 2019, 27, 3511–3531. [Google Scholar] [CrossRef] [PubMed]
  94. Shil, A.K.; Kumar, S.; Sharma, S.; Chaudhary, A.; Das, P. Polystyrene Resin Supported Palladium(0) (Pd@PR) Nanocomposite Mediated Regioselective Synthesis of 4-Aryl-1-Alkyl/(2-Haloalkyl)-1H-1,2,3-Triazoles and Their N-Vinyl Triazole Derivatives from Terminal Alkynes. RSC Adv. 2015, 5, 11506–11514. [Google Scholar] [CrossRef]
  95. Meldal, M.; Tornøe, C.W. Cu-Catalyzed Azide−Alkyne Cycloaddition. Chem. Rev. 2008, 108, 2952–3015. [Google Scholar] [CrossRef]
  96. Guo, X.; Lin, C.; Zhang, M.; Duan, X.; Dong, X.; Sun, D.; Pan, J.; You, T. 2D/3D Copper-Based Metal-Organic Frameworks for Electrochemical Detection of Hydrogen Peroxide. Front. Chem. 2021, 9, 743637. [Google Scholar] [CrossRef] [PubMed]
  97. Benjamin, R.B.; Sandra E., D.; Harry, H. Experimental Evidence for the Involvement of Dinuclear Alkynylcopper(I) Complexes in Alkyne–Azide Chemistry. Chem. Eur. J. 2010, 16, 6278–6284. [Google Scholar] [CrossRef]
  98. Worrell, B.T.; Malik, J.A.; Fokin, V.V. Direct Evidence of a Dinuclear Copper Intermediate in Cu(I)-Catalyzed Azide-Alkyne Cycloadditions. Science 2013, 340, 457–460. [Google Scholar] [CrossRef] [Green Version]
  99. Mo, J.-N.; Su, J.; Zhao, J. The Asymmetric A3(Aldehyde–Alkyne–Amine) Coupling: Highly Enantioselective Access to Propargylamines. Molecules 2019, 24, 1216. [Google Scholar] [CrossRef] [Green Version]
  100. Swami, S.; Sahu, S.N.; Shrivastava, R. Nanomaterial Catalyzed Green Synthesis of Tetrazoles and Its Derivatives: A Review on Recent Advancements. RSC Adv. 2021, 11, 39058–39086. [Google Scholar] [CrossRef]
  101. He, Z.; Zhao, X.; Pan, X.; Li, Y.; Wang, X.; Xu, H.; Xu, Z. Ligand Geometry Controlling Zn-MOF Partial Structures for Their Catalytic Performance in Knoevenagel Condensation. RSC Adv. 2019, 9, 25170–25176. [Google Scholar] [CrossRef] [Green Version]
  102. LiLi, P.; Regati, S.; Huang, H.; Arman, H.D.; Zhao, J.C.-G.; Chen, B. A Metal–Organic Framework as a Highly Efficient and Reusable Catalyst for the Solvent-Free 1,3-Dipolar Cycloaddition of Organic Azides to Alkynes. Inorg. Chem. Front. 2015, 2, 42–46. [Google Scholar] [CrossRef]
  103. Fan, Z.; Wang, Z.; Cokoja, M.; Fischer, R.A. Defect Engineering: An Effective Tool for Enhancing the Catalytic Performance of Copper-MOFs for the Click Reaction and the A 3 Coupling. Catal. Sci. Technol. 2021, 11, 2396–2402. [Google Scholar] [CrossRef]
  104. Xu, Z.; Han, L.; Zhuang, G.; Bai, J.; Sun, D. In Situ Construction of Three Anion-Dependent Cu(I) Coordination Networks as Promising Heterogeneous Catalysts for Azide–Alkyne “Click” Reactions. Inorg. Chem. 2015, 54, 4737–4743. [Google Scholar] [CrossRef]
  105. Murugan, R.; Reddy, M.B.; Pandurangan, P.; Anandhan, R. Gold-like Thiolate-Protected Ultrasmall Cubic Copper Nanocluster-Based Metal–Organic Framework as a Selective Catalyst for Stepwise Synthesis of Unsymmetric Bistriazole by Click Reaction. ACS Appl. Mater. Interfaces 2020, 12, 56004–56016. [Google Scholar] [CrossRef]
  106. Hu, Y.-H.; Wang, J.-C.; Yang, S.; Li, Y.-A.; Dong, Y.-B. CuI@UiO-67-IM: A MOF-Based Bifunctional Composite Triphase-Transfer Catalyst for Sequential One-Pot Azide–Alkyne Cycloaddition in Water. Inorg. Chem. 2017, 56, 8341–8347. [Google Scholar] [CrossRef]
  107. Arnanz, A.; Pintado-Sierra, M.; Corma, A.; Iglesias, M.; Sanchez, F. Bifunctional Metal Organic Framework Catalysts for Multistep Reactions: MOF-Cu(BTC)-[Pd] Catalyst for One-Pot Heteroannulation of Acetylenic Compounds. Adv. Synth. Catal. 2012, 354, 1347–1355. [Google Scholar] [CrossRef]
  108. Jayaramulu, K.; Suresh, V.M.; Maji, T.K. Stabilization of Cu2O Nanoparticles on a 2D Metal–Organic Framework for Catalytic Huisgen 1,3-Dipolar Cycloaddition Reaction. Dalton Trans. 2014, 44, 83–86. [Google Scholar] [CrossRef]
  109. Jiang, W.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Porphyrin-Based Mixed-Valent Ag(I)/Ag(II) and Cu(I)/Cu(II) Networks as Efficient Heterogeneous Catalysts for the Azide–Alkyne “Click” Reaction and Promising Oxidation of Ethylbenzene. Chem. Commun. 2016, 52, 1373–1376. [Google Scholar] [CrossRef]
  110. Kim, A.; Muthuchamy, N.; Yoon, C.; Joo, S.H.; Park, K.H. MOF-Derived Cu@Cu2O Nanocatalyst for Oxygen Reduction Reaction and Cycloaddition Reaction. Nanomaterials 2018, 8, 138. [Google Scholar] [CrossRef]
  111. Thanh, N.D.; Hai, D.S.; Bich, V.T.N.; Hien, P.T.T.; Duyen, N.T.K.; Mai, N.T.; Dung, T.T.; Toan, V.N.; Van, H.T.K.; Dang, L.H.; et al. Efficient Click Chemistry towards Novel 1H-1,2,3-Triazole-Tethered 4H-Chromene−d-Glucose Conjugates: Design, Synthesis and Evaluation of in Vitro Antibacterial, MRSA and Antifungal Activities. Eur. J. Med. Chem. 2019, 167, 454–471. [Google Scholar] [CrossRef]
  112. Bikas, R.; Ajormal, F.; Noshiranzadeh, N.; Emami, M.; Kozakiewicz, A. 1D Azido Bridged Cu(II) Coordination Polymer with 1,3-oxazolidine Ligand as an Effective Catalyst for Green Click Synthesis of 1,2,3-triazoles. Appl. Organomet. Chem. 2020, 34. [Google Scholar] [CrossRef]
  113. Lu, B.-B.; Yang, J.; Che, G.-B.; Pei, W.-Y.; Ma, J.-F. Highly Stable Copper(I)-Based Metal–Organic Framework Assembled with Resorcin[4]Arene and Polyoxometalate for Efficient Heterogeneous Catalysis of Azide–Alkyne “Click” Reaction. ACS Appl. Mater. Interfaces 2018, 10, 2628–2636. [Google Scholar] [CrossRef] [PubMed]
  114. Tourani, H.; Naimi-Jamal, M.R.; Dekamin, M.G. Preparation of 5-Substituted-1H-Tetrazoles Catalyzed by MOFs via Two Strategies: Direct Condensation of Aryl Nitriles with SodiumAzide, and Tri-Component Reaction Method. ChemistrySelect 2018, 3, 8332–8337. [Google Scholar] [CrossRef]
  115. Jia, X.; Xu, G.; Du, Z.; Fu, Y. Cu(BTC)-MOF Catalyzed Multicomponent Reaction to Construct 1,4-Disubstituted-1,2,3-Triazoles. Polyhedron 2018, 151, 515–519. [Google Scholar] [CrossRef]
  116. Wang, Z.; Zhou, X.; Gong, S.; Xie, J. MOF-Derived Cu@N-C Catalyst for 1,3-Dipolar Cycloaddition Reaction. Nanomaterials 2022, 12, 1070. [Google Scholar] [CrossRef] [PubMed]
  117. Maity, T.; Saha, D.; Koner, S. Aromatic N-Arylations Catalyzed by Copper-Anchored Porous Zinc-Based Metal-Organic Framework under Heterogeneous Conditions. ChemCatChem 2014, 6, 2373–2383. [Google Scholar] [CrossRef]
  118. Wang, X.-X.; Yang, J.; Xu, X.; Ma, J.-F. Highly Stable Copper(I)–Thiacalix[4]Arene-Based Frameworks for Highly Efficient Catalysis of Click Reactions in Water. Chem.-Eur. J. 2019, 25, 16660–16667. [Google Scholar] [CrossRef]
  119. Yamane, I.; Sato, K.; Otomo, R.; Yanase, T.; Miura, A.; Nagahama, T.; Kamiya, Y.; Shimada, T. Ultrahigh-Pressure Preparation and Catalytic Activity of MOF-Derived Cu Nanoparticles. Nanomaterials 2021, 11, 1040. [Google Scholar] [CrossRef]
  120. Fu, Q.; Xie, K.; Tan, S.; Ren, J.M.; Zhao, Q.; Webley, P.A.; Qiao, G.G. The Use of Reduced Copper Metal–Organic Frameworks to Facilitate CuAAC Click Chemistry. Chem. Commun. 2016, 52, 12226–12229. [Google Scholar] [CrossRef]
  121. Thanh, N.D.; Hai, D.S.; Ha, N.T.T.; Tung, D.T.; Le, C.T.; Van, H.T.K.; Toan, V.N.; Toan, D.N.; Dang, L.H. Synthesis, Biological Evaluation and Molecular Docking Study of 1,2,3-1H-Triazoles Having 4H-Pyrano[2,3-d]Pyrimidine as Potential Mycobacterium Tuberculosis Protein Tyrosine Phosphatase B Inhibitors. Bioorg. Med. Chem. Lett. 2019, 29, 164–171. [Google Scholar] [CrossRef]
  122. Shinde, K.S.; Michael, P.; Fuhrmann, D.; Binder, W.H. A Mechanochemically Active Metal-Organic Framework (MOF) Based on Cu-Bis-NHC-Linkers: Synthesis and Mechano-Catalytic Activation. Macromol. Chem. Phys. 2022, 223, 2200207. [Google Scholar] [CrossRef]
  123. De, D.; Pal, T.K.; Neogi, S.; Senthilkumar, S.; Das, D.; Gupta, S.S.; Bharadwaj, P.K. A Versatile CuII Metal–Organic Framework Exhibiting High Gas Storage Capacity with Selectivity for CO2: Conversion of CO2 to Cyclic Carbonate and Other Catalytic Abilities. Chem.-Eur. J. 2016, 22, 3387–3396. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, F.; Zhang, Y.; Liu, Z.; Du, Z.; Zhang, L.; Ren, J.; Qu, X. A Biocompatible Heterogeneous MOF–Cu Catalyst for In Vivo Drug Synthesis in Targeted Subcellular Organelles. Angew. Chem. Int. Ed. 2019, 58, 6987–6992. [Google Scholar] [CrossRef]
  125. Zeng, F.; Pan, Y.; Luan, X.; Gao, Y.; Yang, J.; Wang, Y.; Song, Y. Copper Metal-Organic Framework Incorporated Mesoporous Silica as a Bioorthogonal Biosensor for Detection of Glutathione. Sens. Actuators B Chem. 2021, 345, 130382. [Google Scholar] [CrossRef]
Figure 1. An illustration of MOF assembly and subsequent use in click chemistry.
Figure 1. An illustration of MOF assembly and subsequent use in click chemistry.
Catalysts 13 00130 g001
Figure 2. MOF structure based on the metal nodes and bridging ligands [31].
Figure 2. MOF structure based on the metal nodes and bridging ligands [31].
Catalysts 13 00130 g002
Figure 3. Depicting physicochemical properties of MOFs.
Figure 3. Depicting physicochemical properties of MOFs.
Catalysts 13 00130 g003
Figure 6. Plausible reaction mechanism for CuAAC using Cu-MOF as a catalyst [95].
Figure 6. Plausible reaction mechanism for CuAAC using Cu-MOF as a catalyst [95].
Catalysts 13 00130 g006
Figure 7. Plausible reaction mechanism for A3 coupling reactions [99].
Figure 7. Plausible reaction mechanism for A3 coupling reactions [99].
Catalysts 13 00130 g007
Figure 8. Plausible reaction mechanism for tetrazole synthesis [100].
Figure 8. Plausible reaction mechanism for tetrazole synthesis [100].
Catalysts 13 00130 g008
Figure 9. Plausible reaction mechanism for condensation reaction. [101].
Figure 9. Plausible reaction mechanism for condensation reaction. [101].
Catalysts 13 00130 g009
Figure 10. General reaction scheme depicting the formation of 1,4-disubstituted 1,2,3-triazole.
Figure 10. General reaction scheme depicting the formation of 1,4-disubstituted 1,2,3-triazole.
Catalysts 13 00130 g010
Figure 11. Schematic representation of the simplified 3D network of Cu-MOF. Reproduced from [104] with permission from the American Chemical Society, copyright year: 2015.
Figure 11. Schematic representation of the simplified 3D network of Cu-MOF. Reproduced from [104] with permission from the American Chemical Society, copyright year: 2015.
Catalysts 13 00130 g011
Figure 12. Synthesis of a metal–organic framework (Cu-MOF) to catalyze the CuAAC reaction in an efficient and chemoselective manner. Reproduced from [105] with permission from the American Chemical Society Copyright (2020).
Figure 12. Synthesis of a metal–organic framework (Cu-MOF) to catalyze the CuAAC reaction in an efficient and chemoselective manner. Reproduced from [105] with permission from the American Chemical Society Copyright (2020).
Catalysts 13 00130 g012
Figure 13. A schematic showing the stabilization of Cu2O nanoparticles in 2D layers and the catalytic azide–alkyne cycloaddition reaction. Reproduced from [108] with permission from RSC, copyright: 2015.
Figure 13. A schematic showing the stabilization of Cu2O nanoparticles in 2D layers and the catalytic azide–alkyne cycloaddition reaction. Reproduced from [108] with permission from RSC, copyright: 2015.
Catalysts 13 00130 g013
Figure 14. Copper-based MOF (CuI@UiO-67-IM) as a catalyst for alkyne azide cycloaddition reaction. Reproduced from [106] with permission from the American Chemical Society, copyright year: 2017.
Figure 14. Copper-based MOF (CuI@UiO-67-IM) as a catalyst for alkyne azide cycloaddition reaction. Reproduced from [106] with permission from the American Chemical Society, copyright year: 2017.
Catalysts 13 00130 g014
Figure 15. Synthesis of the catalyst Fe3O4@HKUST-1. [76].
Figure 15. Synthesis of the catalyst Fe3O4@HKUST-1. [76].
Catalysts 13 00130 g015
Figure 16. Schematic representation of synthesis of Cu-MOF [60].
Figure 16. Schematic representation of synthesis of Cu-MOF [60].
Catalysts 13 00130 g016
Figure 17. Three-dimensional copper-based MOF for the synthesis of 1,2,3 triazole. Reproduced from [112] with permission from the American Chemical Society, copyright year: 2018.
Figure 17. Three-dimensional copper-based MOF for the synthesis of 1,2,3 triazole. Reproduced from [112] with permission from the American Chemical Society, copyright year: 2018.
Catalysts 13 00130 g017
Figure 18. The metal–organic framework structures with exposed active CuI sites and a depiction of the catalytic activity for the formation of 1,2,3-triazole. Reproduced from [118] with permission from Wiley, copyright year: 2019.
Figure 18. The metal–organic framework structures with exposed active CuI sites and a depiction of the catalytic activity for the formation of 1,2,3-triazole. Reproduced from [118] with permission from Wiley, copyright year: 2019.
Catalysts 13 00130 g018
Figure 19. Cycloaddition reaction depicting the catalytic performance of Cu@C [119].
Figure 19. Cycloaddition reaction depicting the catalytic performance of Cu@C [119].
Catalysts 13 00130 g019
Figure 20. A simplified depiction of the steps required to make a reduced copper metal–organic framework (rCu-MOF). (i) Benzene 1,3,5-tricarboxylate (BTC) undergoes a coordination reaction with Cu2+. and (ii) a hydroquinone-induced Cu-BTC reduction at 150 °C. Reproduced from [120] with permission from RSC, copyright year: 2016.
Figure 20. A simplified depiction of the steps required to make a reduced copper metal–organic framework (rCu-MOF). (i) Benzene 1,3,5-tricarboxylate (BTC) undergoes a coordination reaction with Cu2+. and (ii) a hydroquinone-induced Cu-BTC reduction at 150 °C. Reproduced from [120] with permission from RSC, copyright year: 2016.
Catalysts 13 00130 g020
Figure 21. Amino acids involved in the intermolecular interactions of 1H-1,2,3-triazole. Reproduced from [121] with permission from Elsevier, copyright year: 2019.
Figure 21. Amino acids involved in the intermolecular interactions of 1H-1,2,3-triazole. Reproduced from [121] with permission from Elsevier, copyright year: 2019.
Catalysts 13 00130 g021
Figure 22. Intracellular bioorthogonal reaction for in situ drug synthesis under mitochondria catalyzed by MOF-Cu-TPP. Reproduced from [123] with permission from Wiley, copyright year: 2019.
Figure 22. Intracellular bioorthogonal reaction for in situ drug synthesis under mitochondria catalyzed by MOF-Cu-TPP. Reproduced from [123] with permission from Wiley, copyright year: 2019.
Catalysts 13 00130 g022
Table 1. Physico-chemical properties of MOFs. * Reproduced from references with permission.
Table 1. Physico-chemical properties of MOFs. * Reproduced from references with permission.
EntryPropertyDescription
1.Surface area
Catalysts 13 00130 i001
  • The formation of a cage due to the binding between the metal ion with the organic linker creates a hollow structure from the inside leading to the formation of a large surface area of MOFs [45].
  • The sensing potential of surface-rich MOFs is intriguing. They may be used as a membrane because of their remarkable gas adsorption and separation capabilities.
2.Porosity
Catalysts 13 00130 i002
  • One of the properties that lends these materials their greatest value in catalysis is that the pore size may be controlled during synthesis by altering the synthesis process and chemical precursors [44].
  • Most of the MOFs described thus far, however, are microporous, which slows down mass transfer and restricts accessibility of heavier molecules to their interior surface [46]. *
3.Structure tunability and Luminescence
Catalysts 13 00130 i003
  • MOFs are highly tunable due to their porous and polymeric cage-like networks.
  • The synthetic tunability of MOFs is opportune for multifunctional catalytic systems [43,44].
  • One of the most distinguishing features of MOFs is luminescence that involves the release of light from excited molecules.
  • This is not only associated with the type of metal in the composition of MOFs but it may also be potentialized by organic ligands that provide excellent chromophores for luminescence [47]. *
4.Chirality
Catalysts 13 00130 i004
  • The synthesis of chiral metal–organic frameworks (MOFs) is highly relevant for asymmetric heterogenous catalysis. Chirality can be introduced to MOFs by using chiral linkers for synthesis [45].
  • MOFs can be regarded as molecular sieve-like materials as their applications include chiral recognition, chiral separation, and chiral catalysis [48]. *
5.Crystallinity
Catalysts 13 00130 i005
  • MOFs have a crystalline property because the synthesis of MOFs involves the key process of crystallization.
  • Highly crystalline mesoporous designs are particularly promising, owing to their unique features coming from the crystallized frameworks, having a large amount of potential in energy and environment-related areas [49]. *
Table 2. General methods for synthesis of MOFs.
Table 2. General methods for synthesis of MOFs.
EntryMethodDescriptionExamplesReferences
1.Electrochemical synthesisElectrochemical synthesis involves the addition of metal ions into a reaction mixture consisting of organic linkers and electrolytes via anodic dissolution. The significant benefit of this technique is that the anions often found with metals in salts may be eliminated, resulting in very pure compounds, since less time and energy are needed for reactions to take place under more benign circumstances.ZIF-8, HKUST-1, MIL-100 (Fe).[51]
2.Microwave-assisted synthesisThe use of microwaves to irradiate the reaction mixture results in the creation of nanoscale MOF crystals. There are several benefits to this synthesis, including high efficiency, shorter reaction times, phase selectivity, morphological control, and particle size reduction.[Cu3(btc)2(H2O)3], HKUST-1[50]
3.Mechanochemical synthesisMechanical forces are introduced to complete the chemical reaction. Mechanochemical synthesis has the benefit of not requiring the use of organic solvents, which may be carcinogenic, poisonous, and damaging to the environment. Metal oxides are often employed as precursors in this procedure rather than metal salts.InOF-1, Cu3(BTC)2-MOF[52,53]
4.Slow evaporation methodThis conventional method for MOF preparation involves the use of suitable solvents to dissolve the precursor materials and later on slow evaporation of the solvent at an adequate temperature. The synthesis of MOFs using this approach is hindered by the insolubility of the chemicals. Accordingly, a combination of solvents may be utilised to improve solubility.[Cu(2,3-pydc)(bpp)]·2.
5H2O, [Zn(2,3-pydc)(bpp)]·2.
5H2O
[50]
5.Solvothermal synthesisThe reaction between the organic linker and metal ion takes place in a suitable solvent at a temperature above the boiling point of the solvent used. The primary benefit of this technique is that the relatively greater yield can be obtained.Zn2(bpabdc)2(DMF)2(H2O)n,
[Cu(tdc) (H2O)]n.n(DMA)
[54,55]
6.Sonochemical synthesisThe solution mixture was subjected to ultrasonic radiations (20 kHz–10 MHz) to synthesize MOF with novel morphologies. The sonochemical technique allows for the rapid production of MOFs with a tiny crystal size in a very short reaction time. This approach has the benefits of being fast, cheap, repeatable, and eco-friendly.HKUST-1, TMU-7, [Zn3(btc)2][56]
Table 3. Several reactions catalyzed by MOF-based catalysts.
Table 3. Several reactions catalyzed by MOF-based catalysts.
EntryReaction CatalyzedMOF TypeMetal in MOFReference
1.Friedel–Crafts reactionHPW@Zr-BTCZirconium[70]
Cu-MOF-74Copper[71]
Urea containing 2D MOFCopper[72]
4.Suzuki couplingCu-BDC MOFCopper[63]
5.Heck couplingPd(II)-porphyrinic MOFPalladium[73]
6.Suzuki–Miyaura couplingNPC-Pd MOFPalladium[74]
7.Knoevenagel condensationAl-MIL-101-NH2Aluminium[75]
8.Aldol condensationBasolite F300Iron[61]
9.CuAAC reactionFe3O4@HKUST-1Copper[76]
Table 4. Comparison of Cu-based MOF over other catalysts used to catalyze an alkyne azide cycloaddition reaction.
Table 4. Comparison of Cu-based MOF over other catalysts used to catalyze an alkyne azide cycloaddition reaction.
EntryCatalystMetalReaction ConditionsSolventRecyclabilityReferences
1.Cu(INA)2-MOFCu(I)1.5 h, 80 °Csolvent free5 times[86]
2.[Cu(CPA)(BDC)]nCu(I)7 h, 90 °CH2O-MeOH (1:4)4 times[60]
3.Cu2(BDC)2(DABCO)Cu(I)45 min, 60 °Cethanol5 times[87]
4.CuICu(I)rtno solventNA[84]
5.CuBr(PPh3)3]Cu(I)60 °C, 5 hTHF/TEANA[88]
6.sf-CuSCu(I)rt, 30 minH2ONA[89]
7.AgN(CN)2Ag(I)rtDIPEAH
2O/ethylene glycol
NA[90]
8.ZnEt2Zn(II)rt, 18 hTHFNA[91]
9.Au(RD32)Au(III)60 °C, H2OEtN3NA[92]
10.RuH2(CO)(PPh3)3Ru(II)80 °C, 2 hPTC, H2ONA[93]
11.Pd@PRPd(0)8–12 h, 100 ℃DMFNA[94]
Table 5. Copper-based MOFs for the synthesis of 1,2,3-triazole via CuAAC reaction.
Table 5. Copper-based MOFs for the synthesis of 1,2,3-triazole via CuAAC reaction.
EntryMOFCatalytic Amount (mol %)Reaction ConditionsSolventYield of TriazoleRecovery of Catalyst %Catalytic RunsReferences
1.Cu-MOF5 mol %7 h, rtCH2Cl294875[102]
2.CuBTC–PyDC5 mol %3 h, 70 ℃Ethanol-764[103]
3.{[Cu6(bpz)6(CH3CN)3(CN)3Br]·
2OH·
14CH3CN}n
10 mol %2 h, 60 ℃DMF and DMSO solution99706[104]
4.Cu-MOF5 mol %12 h,
rt
EtOH/
H2O (2:1)
95885[105]
5.CuI@UiO-67-IM2 mol %80 °C, 2 hWater, under air atmosphere90 5[106]
6.MOF-Cu (BTC)-[Pd]2 mol %2 h, 50 °CDMF100852[107]
7.Cu2O@{[Zn(Himdc)(bipy)0.5
DMF} MOF
0.5 mol %7 h, 50 °Ct-BuOH–H2O (2:1), Et3N98-3[108]
8.Cu(INA)2-MOF1 mol %1.5 h, 80 °CSolvent free98905[86]
9.[Cu(I)6I6(Cu(II)-TPPP)].2DMF1 mol %12 h, 50 °CMethanol and water (4:1)100905[109]
10.Fe3O4@HKUST-11.8 mol %2 h, refluxH2O92875[76]
11.Cu@Cu2O2.3 mol % 5 h, 50 °C Water, t-BuOH (2:1)100--[110]
12.Cu@MOF-52 mol %30 min, 79–80 °CEthanol98--[111]
13.[Cu(CPA)(BDC)]n10 mol %7 h, 90 °CH2O-MeOH (1:4)93704[60]
Table 6. Copper-based MOFs for the synthesis of 1,2,3-triazole via the CuAAC reaction.
Table 6. Copper-based MOFs for the synthesis of 1,2,3-triazole via the CuAAC reaction.
EntryMOFCatalytic Amount (mol %)Reaction ConditionsSolventYield of TriazoleRecovery of Catalyst %Catalytic RunsReferences
1.[CuI4(SiW12O40)(L)]·6H2
2DMF
30 mg12 h, 80 ℃Methanol99995[114]
2.[Cu4Cl4L]·CH3OH·1.5H2O10 mg 8 h, 60 °C Water 99-5[113]
3.Cu(BTC)-MOF10 mg16 h, RTCH3OH91873[115]
4.Cu2(BDC)2(DABCO)20 mg45 min, 60 °CEthanol98925[87]
5.Cu@N-C(600)5 mg12 h, 50 °Ct-BuOH/H2O (3/1)98934[116]
6.[Cu(H3L)(μ1,3-N3)(N3)]n4 mg3.5 h, 40 °CWater93--[112]
7.IRMOF-3-PI-Cu2 mg20 h, 90°CDMSO90805[117]
8.[Cu4Cl4L]·CH3OH·1.5H2O10 mg8 h, 60 °CWater99975[118]
9.Cu-BTC10 mg1 h, 60 °C1,4-Dioxane, TEA---[119]
10.rCu-MOF0.2 mg4 h, 60 °CH2O, N2 for 20 min--3[120]
Table 7. Copper-based MOFs for the synthesis of 1,2,3-triazole via the CuAAC reaction.
Table 7. Copper-based MOFs for the synthesis of 1,2,3-triazole via the CuAAC reaction.
EntryMOFCatalytic Amount (mol %)Reaction ConditionsSolventYield of Triazole %Catalytic RunsReferences
1.Cu-Bis-NHC- MOF10 wt %-THF/MeOH--[122]
2. {[Cu2(L)(H2O)2]·(5 DMF)(4 H2O)}n 5 wt %4 h, 50 °CCH2Cl296-[123]
3.[Cu(2-pymo)2]-4 h, 70 °CEthanol-6[30]
4.MOF-Cu-TPP----10[124]
5.MSNs@Cu- MOF-----[125]
6.Cu@MOF-5-4–5 h, 79−80 °Cabs. EtOH94-[121]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Singh, R.; Singh, G.; George, N.; Singh, G.; Gupta, S.; Singh, H.; Kaur, G.; Singh, J. Copper-Based Metal–Organic Frameworks (MOFs) as an Emerging Catalytic Framework for Click Chemistry. Catalysts 2023, 13, 130. https://doi.org/10.3390/catal13010130

AMA Style

Singh R, Singh G, George N, Singh G, Gupta S, Singh H, Kaur G, Singh J. Copper-Based Metal–Organic Frameworks (MOFs) as an Emerging Catalytic Framework for Click Chemistry. Catalysts. 2023; 13(1):130. https://doi.org/10.3390/catal13010130

Chicago/Turabian Style

Singh, Riddima, Gurleen Singh, Nancy George, Gurjaspreet Singh, Sofia Gupta, Harminder Singh, Gurpreet Kaur, and Jandeep Singh. 2023. "Copper-Based Metal–Organic Frameworks (MOFs) as an Emerging Catalytic Framework for Click Chemistry" Catalysts 13, no. 1: 130. https://doi.org/10.3390/catal13010130

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

Article Metrics

Back to TopTop