Next Article in Journal
The bis(Biphenyl)phosphorus Fragment in Trivalent and Tetravalent P-Environments
Next Article in Special Issue
Functionalized Tris(anilido)triazacyclononanes as Hexadentate Ligands for the Encapsulation of U(III), U(IV) and La(III) Cations
Previous Article in Journal
Density Functional Theory Study on the Adsorption Mechanism of Sulphide Gas Molecules on α-Fe2O3(001) Surface
Previous Article in Special Issue
Pyridinesilver Tetraoxometallate Complexes: Overview of the Synthesis, Structure, and Properties of Pyridine Complexed AgXO4 (X = Cl, Mn, Re) Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal Organic Frameworks as Heterogeneous Catalysts in Olefin Epoxidation and Carbon Dioxide Cycloaddition

by
Alessia Tombesi
1 and
Claudio Pettinari
2,*
1
Chemistry Section, School of Science and Technology, University of Camerino, Via S. Agostino 1, 62032 Camerino, Italy
2
Chemistry Section, School of Pharmacy, University of Camerino, Via S. Agostino 1, 62032 Camerino, Italy
*
Author to whom correspondence should be addressed.
Submission received: 28 September 2021 / Revised: 5 November 2021 / Accepted: 9 November 2021 / Published: 15 November 2021
(This article belongs to the Special Issue Cornerstones in Contemporary Inorganic Chemistry)

Abstract

:
Metal–organic frameworks (MOFs) are a family of porous crystalline materials that serve in some cases as versatile platforms for catalysis. In this review, we overview the recent developments about the use of these species as heterogeneous catalysts in olefin epoxidation and carbon dioxide cycloaddition. We report the most important results obtained in this field relating them to the presence of specific organic linkers, metal nodes or clusters and mixed-metal species. Recent advances obtained with MOF nanocomposites were also described. Finally we compare the results and summarize the major insights in specific Tables, outlining the major challenges for this emerging field. This work could promote new research aimed at producing coordination polymers and MOFs able to catalyse a broader range of CO2 consuming reactions.

Graphical Abstract

1. Introduction

International Union of Pure and Applied Chemistry (IUPAC) defines MOFs as a coordination network with an open framework containing potential voids [1]. This emerging class of porous coordination polymers are formed by metal ion or cluster nodes and functional organic ligands, all connected through coordination bonds to form 1D, 2D o 3D networks (Figure 1) [2,3,4,5,6]. MOFs can be easily obtained by several different synthetic methods, such as electrochemical [7], solvothermal [8] and mechanochemical [9], slow diffusion [10], and more recently also by microwave-assisted heating [11].
The crystal structures of MOFs can be customized depending on the metal and ligand choice as also on the solvents and reaction conditions employed. [12] Due to the high surface areas [13] and ultrahigh porosity they are attractive for CH4, CO2, and H2 sorption and storage. Most MOFs have higher volumetric H2 and CH4 storage capacities concerning traditional porous materials.
In recent years nanoscale MOFs have been also investigated for their potential applications in biomedicine, for example for drug delivery [14] and biological imaging [15], mainly for the possibility to use biocompatible building blocks. MOFs were employed as electrode materials for supercapacitors using Co-based coordination polymers [16], for magnetic and electronic devices [17], for water harvesting where H2O is extracted from the air by solar energy [18], and finally also for non-linear optics [19].
The use of MOFs as a catalyst has been widely explored and several applications have been developed, for example in the production of fine chemicals [20], or the definition of possible new green protocols replacing non-eco-friendly catalysts [21]. Differences in activity and selectivity toward specific organic reactions are significantly dependent on the MOFs structure [22]. The main MOFs advantage, when we consider their use in catalysis, is in the possibility to design and predict the structural properties based on of linker features, coordination number and geometry of the metal.
The presence of coordinatively unsaturated metal sites, the variety of basic linkers available, the stability to solvents and to reaction conditions, the possibility to host guest molecules within the pores makes MOFs perspective materials for heterogenous catalysis. They have also a lot of advantages concerning other inorganic systems as zeolites and aluminophosphates, i.e., they can be modified using organic synthesis, being possible to decorate their pores with catalytic sites. MOFs can be tailored by a simple change in the initial synthetic conditions or by using post-synthetic reactions. These modifications make MOFs excellent candidates for designing functional materials to allow the attachment of different catalysts [23].
While the characterization of deposited species upon conventional catalyst supports, such as metal oxides, tends to be challenging due to the non-uniform surface and pore structures of the support, the crystalline nature of MOFs enables visualization of the catalytically active species within the framework, which leads to a detailed characterization of active catalytic sites and provides insight into structure−activity relationships.
In this review we want to focus on the most recent progress in two reactions MOFs-catalyzed, i.e., the olefin epoxidation and the cycloaddition of CO2 to epoxides to yield organic carbonates as a final product, by performing a rigorous analysis of the best MOFs in terms of conversion and selectivity. Specifically, we examined MOF-based catalytic materials producing epoxide and cyclic carbonates with percentages of conversion and selectivity exceeding 70, in the 2015 to 2021 period. Moreover, few relevant papers on the heterogeneous MOFs catalysts published before 2015, for a useful comparison have been considered.
Epoxides are important species and intermediates in the production of pharmaceuticals, agrochemicals, and relevant industrial chemicals. In the global market, the production of propylene oxide achieves 8 million tons per year with an expected annual increase of 5% [24]. Due to the industrial relevance of catalytic oxidation of olefins to fine chemicals, numerous studies have been devoted to the development of efficient homogeneous [24] and heterogeneous catalysts [24]. However, high selectivity and enantioselectivity in epoxidation reactions remain a challenge. While recovery and product separation are the main drawbacks for homogenous catalysts, MOFs used as heterogeneous catalysts in the oxidation of olefins have attracted significant attention (Figure 2) [25].
CO2 is the primary greenhouse gas in the atmosphere, and it is the cause of environmental and energy-related problems in the world. Nowadays, the development of new methods is fundamental to capture and convert CO2 into useful chemical products to improve the environment and promote sustainable development. Several studies have been carried out on MOF’s efficiency to capture CO2. The linkers that connect the MOFs metal nodes are the major sites for CO2 binding. The linkers that connect the MOFs metal nodes are the major sites for CO2 binding, and they can be chemically modified with functional groups to increase their interaction with CO2. Moreover, unsaturated metals ions can be introduced in the MOFs structure. A significantly benefit generated from the possibility to have adequate quantities of CO2 in concentrated form within a MOF is the possible use of CO2 as a chemical reagent (Figure 3).
A significant number of MOFs has been recently reported to catalyse the CO2 cycloaddition reaction to epoxides to give cyclic organic carbonates (OCs) and several papers describe the potential and effectiveness of MOFs in this important process, so it is necessary to identify better strategies to build new advanced materials as MOFs or MOF-based species to grow selectivity, capacity, and conversion of this catalytic reaction.

2. Olefin Epoxidation

C=C bond epoxidation is an attractive reaction for industrial process to obtain raw materials for epoxy resin, polymers, and pharmaceutical intermediates. Although homogenous catalysts in the epoxidation of alkene have been largely studied in the past few decades, the separation from the reaction mixture and its subsequent reusability remain open challenges [26,27].
Very recently, taking advantage of the tunability of MOFs, several transition metal-based epoxidation catalysts have been developed using MOFs synthesis in combination with post-synthetic modification. Several literature reports require utilization of expensive transition metals, but in the last period also metals like Cu, a classic non-noble transition metal, abundant, inexpensive, and non-toxic, become appealing catalyst sources.

2.1. Metal Nodes/Clusters as Catalytically Active Sites

MOFs can be applied as ideal platforms for heterogeneous catalysis towards olefin epoxidation thanks to several structural features with intrinsic catalytic activity such as the coordinatively unsaturated sites on MOFs nodes, defects, and catalytically active organic linkers.
Three pillared-layered Co6-MOFs were utilized as heterogeneous catalysts for the selective oxidation of styrene using air, and benzyl alcohol with oxygen. The hexaprismatic [Co63−OH)6] cluster with different variable valences activate the oxygen molecule for aerobic epoxidation of alkenes [28]. In Co6-MOF-3, large pores facilitated the mass transfer giving the fastest reaction rate with high conversion and good selectivity for oxidation of both styrene and benzyl alcohol [29].
The 2D-cobalt (II)-based coordination polymer, {(Co(L2)H2O))2·H2O)}n, have been obtained by hydrothermal synthesis using the histidine derivative 4-((1-carboxy-2-(1H-imidazol-4-yl)ethylamino)methyl)benzoic acid (H2L2) as ligand. It has been investigated as heterogeneous catalysts on the allylic oxidation of cyclohexene (Appendix A).
The presence of a Co(II) open site on the surface maximizes the catalytic productivity, giving 82.56% of conversion and 71% of ter-butyl-2-cyclohexenyl-1-peroxide. Moreover, a Co(II)-based catalyst exhibits similar activity over five cycles without metal leaching [30] (Table 1).
The static and rotary hydrothermally synthetic method could affect significantly both the process of crystallization and heterogeneous catalytic activity of MOFs in the epoxidation reaction. For example, Co-MOF-150-2, hydrothermally synthesized by rotary crystallization at 150 rpm for 2 h, has reached 95.7% yield of 2,3-epoxypinane from α-pinene in aerobic conditions. The high catalytic activity of Co-MOF-150-2 is due to the better exposure of the metal active in the high crystalline structure, where the lamellar layer was more homogenous. The thinner Co-MOF-150-2 was also investigated in the epoxidation of the other olefins. Additionally, the catalytic activity was relevant for cyclic olefins like cyclooctene (78.5% of conversion after 5h of reaction) and for linear olefins (after 12 h 87.2% of 1-decene was transformed into the epoxide) [31].
A Co-MOF has been prepared under surfactant-thermal condition: NTUZ30 has been obtained with two different secondary building units (SBU), i.e., the unusual trinuclear [Co33-OH)(COO)7] and [Co(COO)4]. The cobalt sites onto the surface can convert trans-stilbene into the corresponding epoxide with excellent selectivity and high conversion [32].
A new 8-connected cobalt network (NH4)2[Co3(Ina)(BDC)3(HCOO)] has been generated from dicarboxylate (BDC = 1,4-benzenedicarboxylate) and pillar isonicotinate (Ina = isonicotinate) ligands with unusual Co3 paddle-wheel cluster. The high number of unsaturated Co active sites available and the regular crystalline structure gave a great catalytic performance for cyclooctene epoxidation with a satisfactory TOF (turnover frequency) of 1370 [33].
Cu-containing MOFs look like promising catalysts for selective oxidation reactions. Generally, the selective oxidation of alkene goes through a radical reaction pathway in which molecular oxygen, or an oxidizing agent is present. Upon coordination of the oxidizing agent to Cu(II) (Figure 4), peroxyl radicals are formed which then react with olefin to form the oxidized products [34,35].
By using Cu4O(OH)2(Me2trz-pba)4 (Me2trz-pba = 4-(3,5-dimethyl-4-H-1,2,4-triazol-4-yl)benzoate) and Cu(Me-4py-trz-ia) (Me-4py-trz-ia = 5-(3methyl-5-(pyridine-4-yl)4H-1,2,4-triazol-4-yl)isophthalate) a significantly higher catalytic activity in the epoxidation of cyclooctane with respect to Cu3(BTC)2 (HKUST-1) has been found, due to the different coordination environment at the catalytically active Cu sites [36]. The catalytic performances of Cu-MOF nanosheets for cyclooctene and 1-hexene epoxidation were nearly twice higher than that of bulk Cu3(BDC)2 crystals. This behaviour is attributed to the better exposure of a greater number of active sites on the surface of the Cu(BDC) nanosheets, which become more available during the reaction. The synthetic procedure must regulates the nanosheet thickness by controlling the dissolution rate of Cu2+ from Cu(OH)2 precursor and tuning the solvent composition. Moreover, the epoxide yield, after 5 cycles with CuMOF nanosheets, remains 96% [37] (Table 1).
In Cu4O(OH)2(Me2trz-pba)4, the Cu44-O)(µ2-OH) tetrahedral node possesses two Cu2+ ions bridged by hydroxyl group, which take part in the activation of oxidating agent TBHP, promoting a quicker formation of tert-butoxyl and tert-butylperoxyl radicals, whereas in Cu(Me-4py-trz-ia) the asymmetric unit contains two crystallographically independent Cu2+ ions. One of them possesses two unsaturated sites that could cause a change in Lewis acidity and generate different redox properties [36].
Oxidation of nonterminal olefin, such as cis-stilbene and cyclooctene, occurs with 94% and 98% conversion, when the activated {[Cu(L3-H)(DMA)]·DMA·2H2O} MOF (H3L3 = tris(4′-carboxybiphenyl)amine; DMA = N,N-dimethylacetamide) was used. The activation was carried out at 200° C for 8 h under vacuum to remove DMA coordinated molecules and produce unsaturated Cu sites that act as Lewis acid [38].
Nbo-type Cu-MOFs, synthesized from the meta-substituted ligand 2,2′,6,6′-tetramethoxy-4,4′-biphenyldicarboxylic acid (H2L4) and copper nitrate [Cu3(L4)3(H2O)2(DMF)]n, possess a high density of catalytic sites in optimal position within the channels, in which oxidation of nonterminal olefins (e.g., norbornene, trans-β-methylstyrene, cis-β-methylstyrene, and trans-stilbene) occurs with 99% conversion and 99% selectivity [39]. Moreover, the less reactive aliphatic alkenes such as 1-octene and trans-4-octene showed moderate conversions with good selectivity.
Epoxidation of cyclohexene achieves the 100% of conversion in presence of H2O2 after 8h when 2D metal carboxylate framework {2(Him)·[Cu(pdc)2]}n has been involved as a heterogeneous catalyst. {2(Him)·[Cu(pdc)2]}n (H2pdc = pyridine-2,5-dicarboxylic acid, Him = imidazole) was obtained through structural inter-conversions starting from {[Mg(H2O)6][Cu(pdc)2]·2H2O}n increasing the imidazole concentration by hydrothermal treatment. The structure of {2(Him)·[Cu(pdc)2]}n derives from the connection of {[Cu(pdc)2]}n- ribbon-like 1D chains by intermolecular H-bonding between hydrogen in the imidazolium ion and the free carboxylate oxygens of pdc2−, this 2D supramolecular structure being crucial to ensure the reaction heterogeneity. Likewise, 1-hexene showed almost complete conversion but increasing the chain length of alkene, the double bond becomes sterically hindered limiting the approach to the active site, and the catalytic activity decreases [40] (Table 1).
High stable zirconium-based MOFs are largely used as active and recyclable catalysts for a variety of catalytic transformations. The catalytic activity of UiO-66 and other Zr-MOFs can be greatly attributed to the presence of random defects in their crystalline structure [41,42,43]. These accessible Lewis acid centers, sometimes in conjunction with Lewis basic sites (e.g., amine groups) in functionalized linker, lead to a significant increase in the catalytic activity [44,45].
Recently, the reaction mechanism underlying both thioether oxidation in nonprotic solvents and epoxidation of electron-deficient C=C bonds in α,β-unsaturated ketones, catalysed by UiO-66 and UiO-67 has been exhaustively investigated [46]. This study suggests the formation of hydroperoxo zirconium species as an oxidant. As already known, the oxidation of less-reactive α,β-unsaturated carbonyl compounds was accompanied by oxidation of MeCN solvent and H2O2 under basic conditions [47], but this nucleophilic peroxo species derived from H2O2 and Zr-MOF can contribute to the epoxidation of the electron-deficient C=C bonds because the reaction readily proceeds even in ethyl acetate.
Table 1. MOFs with metal Nodes/clusters active in olefin epoxidation.
Table 1. MOFs with metal Nodes/clusters active in olefin epoxidation.
MOFSubstrateReaction Data
T (°C) P (atm) Time (h)
Oxidant/Cocatalyst/
Solvent a
Conversion
%
Epoxide
Selectivity%
Ref.
Co6-MOF-3Styrene100114Air/-/DMF9990[29]
Co-MOF-150-2α-Pinene9015Air/CHP/- 99.596.2[31]
Cyclooctene9015Air/CHP/- 78.5-[31]
1-Decene9015Air/CHP/-87.2-[31]
NTUZ30trans-Stilbene10071O2/-/-98.295.6[32]
(NH4)2[Co3(Ina)(BDC)3(HCOO)]Cyclooctene351.51IBA/-/CH3CN9892[33]
Cu3(BTC)2Cyclooctene75124Air/TBHP/Toluene20-[36]
1-Hexene251 Air/TBHP/Toluene30.5-[36]
Cu-MOF nanosheetsCyclooctene 25112O2/-/CH3CN100-[37]
1-Hexene251 O2/-/CH3CN67.2-[37]
Cu4O(OH)2(Me2trz-pba)4Cyclooctene75124Air/TBHP/Toluene9080[36]
Cu(Me-4py-trz-ia)Cyclooctene75124Air/TBHP/Toluene3860[36]
{[Cu(L3-H)(DMA)]·DMA·2H2O}cis-Stilbene 60124t-BuOOH/-/CH3CN94.4-[38]
Cyclooctene60124t-BuOOH/-/CH3CN98-[38]
{2(Him)·[Cu(pdc)2]}nCyclohexene6018Air/H2O2/EtOH100-[38]
Cyclooctene6018Air/H2O2/EtOH100-[38]
[Cu3(L4)3(H2O)2(DMF)]nStyrene 4016O2/TMA/CH3CN9088[39]
Cyclooctene4016O2/TMA/CH3CN9999[39]
{(Co(L2)H2O))2·H2O)}nCyclohexene6016t-BuOOH/-/-82.5671.93[30]
UiO-662-Cyclohexen-1-one7011H2O2/-/CH3CN2060[46]
2-Cyclohexen-1-one7012H2O2/-/EtOAc1845[46]
Chalcone7010.5H2O2/-/EtOAc3050[46]
UiO-672-Cyclohexen-1-one7011H2O2/-/CH3CN2055[46]
2-Cyclohexen-1-one7012H2O2/-/EtOAC2055[46]
Chalcone7010.5H2O2/-/CH3CN 4040[46]
a tBuOOH = tert-butyl hydroperoxide; CHP = cumene hydroperoxide; TBHP = tert-butylhydroperoxide; IBA = isobutyraldehyde; TMA = trimethylacetaldehyde.

2.2. Mixed-Metal Species

Many efforts have been made to improve the catalytic performance of MOFs, and one possible way is the construction of bimetallic clusters by functionalization of metal nodes/clusters with active transition metals to afford MOF-based catalysts with high performance.
A hydrothermal reaction has been used to synthesise Cux-Coy-MOF, where Co(NO3)2∙6H2O and Cu(NO3)2·3H2O inorganic metal salts have been one-pot added to a ligand solution in different molar ratios. In addition to the high catalytic activity by doping Cu-MOF with Co, a better selectivity to produce styrene oxide is achieved. At the optimal reaction conditions, the conversion and the selectivity of styrene to styrene oxide increased to 97.81% and 83.04%, respectively, by using Cu0.25-Co0.75-MOF, the catalyst of this series with higher content of Co2+ [48]. Another study showed how the conversion of styrene-to-styrene oxide increased rapidly when Mn ions were introduced into a Cu-MOF with the two ligands 2,5-dihydroxyterephthalic acid (H4DHTA) and 2-picolinic acid (PCA). Mn0.1Cu0.9-MOF exhibits interesting catalytic activity for the epoxidations of various aromatic and cyclic olefins and a weak activity on decomposition of H2O2. Styrene can be oxidized by H2O2, through peroxybicarbonate-assisted catalysis, the styrene oxide yield achieving 85% in the presence of Mn0.1Cu0.9-MOF at 0 °C for 6 h [49].
To increase conversion and selectivity in the solvent-free aerobic oxidation of olefins, MOF catalysts based on 3d metal copper (II), cobalt (II) and H2ODA (oxydiacetic acid) containing lanthanum (III) as 4f ions {[La2Cu3(µ-H2O)(ODA)6(H2O)3]∙3H2O}n (LaCuODA) and {[La2Co3(ODA)6(H2O)6]·12H2O}n (LaCoODA) were employed. Catalytic studies pointed out the difference in aerobic oxidation of cyclohexene performances due to different physicochemical properties, surface area and redox properties of the metals (Table 2). [50] LaCoODA, based on Co(II), showed better conversion and selectivity for 2-cyclohexen-1-one. This is due to the structural differences between the square planar LaCuODA and the octahedral LaCoODA, in the latter case the water molecules could easily leave the channels to foster interaction between the active sites and the oxidant/catalyst. Moreover, the acid properties of the copper(II) ions are less effective than the redox properties of cobalt(II) ones, as far as the catalytic performances [34,51].
In NU-1000 single-ion-based iron(III) species have been incorporated using solution-phase post-synthetic metalation with two different iron(III) precursors. The resulting NU-1000-Fe-NO3 and NU-1000-Fe-Cl frameworks show two crystallographically independent Fe sites (Fe1 resides in the c-pore and Fe2 in the hexagonal mesopore), coordinated to the bridging and terminal oxygens of the Zr6 node, with Fe−O distances in NU-1000-Fe-Cl being much longer than those of NU-1000-Fe-NO3 (Figure 5) [52]. Epoxidation of cyclohexene in vapour H2O2 with NU-1000-Fe-NO3 as catalysts initially yields cyclohexene epoxide derived from heterolytic activation of H2O2, which in turn hydrolyzes rapidly to trans-cyclohexanediol. Otherwise, NU-1000-Fe-Cl yields a mixture of products and by-products, derived from the radical oxidation products due to homolytic activation of H2O2 [53,54]. This behaviour is probably due to the difference in the metal−node distance between the frameworks, the active site rearranging differently.
One-step template-free synthesis of ultrathin (∼5 nm) mixed-valence {V16} clusters-based MOF nanosheets [Ni(4,4′-bpy)2]2 [V7IVV9VO38Cl]·(4,4′-bpy)·6H2O (NENU-MV-1) has been also reported. A large number of vanadium catalytically active sites in the NENU-MV-1 nanosheet allowed excellent cyclohexene oxidation under air exhibiting a conversion of 95%. Moreover, the nanometer scale of the catalyst increased the catalytic activity 2.7 times compared to the bulk crystal (0.25 mm) for olefin epoxidation. Excellent catalytic performances have been shown for different olefin substrates [55].
Table 2. Mixed Metal MOFs in Olefin Epoxidation.
Table 2. Mixed Metal MOFs in Olefin Epoxidation.
MOFSubstrateReaction Data
T (°C) P (atm) Time (h)
Oxidant/Cocatalyst/Solvent aConversion
%
Epoxide Selectivity%Ref.
Mn0.1Cu0.9-MOFStyrene016H2O2/-/DMF90.294.3[49]
Cu0.25-Co0.75-MOFStyrene8018Air/TBHP/t-BuOH/H2O297.8183.4[48]
Zn1Co1-ZIFStyrene 100124TBHP/-/DMF9971.31[56]
LaCoODACyclohexene75124O2 flow/-/-8575[50]
LaCuODA Cyclohexene75124O2 flow/-/-6755[50]
NENU-MV-1Cyclohexene3514Air/IBA/CH3CN9586[55]
NU-1000-Fe-ClCyclohexene1200.033H2O2/-/--70[52]
NU-1000-Fe-ClCyclohexene1200.033H2O2/-/--70[52]
a tBuOH = tert-butyl alcohol; TBHP = tert-butylhydroperoxide; IBA = isobutyraldehyde.

2.3. Organic Linkers with Functional Catalytically Active Sites

Functional groups such as amino, pyridyl, amide, sulfonic acid, etc. present in organic linkers serve as active sites for catalysis and strongly influence the intrinsic catalytic activity of the MOFs through inductive effects. In addition, organic linkers can be catalytically active when organic functional groups and/or functional molecular catalysts (e.g., metalloporphyrins, salen and related ligands, chiral molecules, Schiff-base complexes, etc.) are introduced by post-synthetic ways. Alternatively, the same functional molecular catalysts can also be used as building units to fabricate MOFs.
Molybdenum complexes have been widely applied as homogeneous catalysts for the epoxidation of alkenes by H2O2 and organic hydroperoxide, a complete conversion and selectivity being reported. To overcome the recoverability and reusability issues correlated to the use of homogeneous molybdenum catalysts, molybdenylacetylacetonate has been supported on TMU-16-NH2 [Zn2(NH2-BDC)2(4-bpdh)]·3DMF, an amine-functionalized two-fold interpenetrated MOF via dative and combined covalent and dative post-synthetic modification [57].
A high porous NU-1000 MOF has been post-modified with the chiral L-tartaric acid, by SALI (solvent-assisted ligand incorporation) to build a chiral Zr-based MOF [C-NU-1000] [58]. Moreover, another active catalytic site, molybdenyl acetylacetonate, MoO2(acac)2, was incorporated on chiral NU-1000 to explore catalytic performance in the asymmetric epoxidation of olefins (Figure 6a) [59]. When olefins approach by pro-S- or R-face to the catalytic active center, they interact with the OH group of the tartrate through H-bond which induces chirality generating two chiral intermediates. The [C-NU-1000-Mo] catalyst, used in the epoxidation of styrene and 1-decene, can discriminate the S configuration in epoxides (Figure 6b).
The free amine group available on UiO-66-NH2 has been post-synthetically modified with salicylaldehyde (SA) or thiophene-2-carbaldehyde (TC) to graft a Schiff base in which MoO2(acac)2 could be immobilized. Efficient olefin epoxidation catalysed by UiO-66-NH2-SA-Mo and UiO-66-NH2-TC-Mo has been described and no Mo active site leaching was detected [60]. In the same way, MoO(O2)2·2DMF was immobilized onto UiO-66(NH2) functionalized with salicylaldehyde (Sal) (UiO-66-sal-MoD), pyridine-2-aldehyde (PI) (UiO-66-PI-MoD) and 2-pyridine chloride (PC) (UiO-66-PC-MoD). All of them allowed a high dispersion of Mo catalyst, the large pores of MOFs guarantee adequate contact between the substrate and the catalytic active center, thus improving the efficiency of cyclic olefins epoxidation [61].
Molybdenum(VI) oxide was deposited on the eight-connected Zr63-O)43-OH)4(H2O)4(OH)4 nodes connected by 1,3,5,8-(p-benzoate) pyrene linkers (TBAPy4−) of the mesoporous NU-1000, via condensation phase through solvothermal deposition in MOF (SIM) [62]. The stable Mo-SIM system exhibits a high conversion for cyclohexene epoxidation without leaching of molybdenum catalyst compared to Mo supported on bulk zirconia (Mo-ZrO2), in which significant leaching of the catalytic species has been observed [60].
Molybdenum tricarbonyl complexes are known to be effective catalysts for the epoxidation of olefins. They form an oxomolybdenum (VI) species in the presence of tert-butyl hydroperoxide (TBHP) as an oxidant which acts as highly active catalytic sites for the epoxidation of olefin. M(CO)6 was deposited on UiO-66 and UiO-67 by chemical vapor deposition (CVD) treatment, UiO-66-Mo(CO)3, and UiO-67-Mo(CO)3 heterogeneous catalysts being fabricated. Herein, the larger tetrahedral and octahedral cavities of UiO-67 enable more accessibility of cyclooctene to catalytically active sites showing higher catalytic activity for the cyclooctene conversion than UiO-66-M(CO)3 [33].
Several attempts have been made to immobilize oxovanadium(IV) complexes on different solid materials and create heterogeneous catalytic systems for the epoxidation of allylic alcohols [63,64,65,66]. A catalyst has been designed by immobilizing oxovanadium(IV) species on UiO-66(NH2) via post-synthetic modification and by using two different pathways. At first, the amino-functionalized UiO-66(NH2) was modified with salicylaldehyde to produce salicylideneimine modified UiO-66 (UiO-66-SI), subsequently [VO(acac)2] was reacted with UiO-66-SI to give UiO-66-SI/VO(acac). In another pathway, UiO-66(NH2) directly reacted with [VO(acac)2] to produce UiO-66-N/VO(acac)2 [67]. Excellent catalytic activity in the regioselective epoxidation of geraniol was obtained when UiO-66-SI/VO(acac) and UiO-66-N/VO(acac)2 systems were employed by using reaction times of 60 and 120 min, respectively. The proper pore size of and the high dispersion of the catalytic sites on UiO-66-SI/VO(acac) and UiO-66-N/VO(acac)2 guarantee good access of the substrate to the active sites.
Metallosalen-based crystalline porous materials have been realized for heterogeneous catalytic applications towards cyclopropanation, alkene epoxidation and hydrolytic kinetic resolution of epoxides with interesting enantioselectivities [5]. UiO-68-Me has been modified via post-synthetic exchange (PSE) with single- and mixed-M(salen) linker (M = Cu, Fe, Cr, V, Mn) to fabricate UiO-66(NH2) attractive species for heterogeneous asymmetric catalysis, useful to overcome the problem of metal leaching. It was found that the single-M(salen) chiral MOFs (R)-UiO-68-Mn and (R)-UiO-68-Fe catalyse the epoxidation of alkenes to epoxides with up to a 98% ee of epoxide and 97% ee, respectively. The different catalytic metal centers in the mixed-(M)salen species UiO-68-Mn-Cr gave consecutive reactions starting from the epoxidation of alkene followed by ring–opening reaction of epoxide to produce the desired amino alcohol in 80−85% yields with 80−99.5% ee. Catalytic activity and enantioselectivity of all chiral UiO-68 catalysts remain unchanged for 10 cycles [68].
Encapsulation of Cu- or Ni-salen species in NH2-MIL-101(Cr) through one-pot method gave a series of effective heterogeneous catalysts in the styrene oxidation under mild conditions. Specifically, the styrene conversion obtained using TBHP was 98.78%. The concentrated electronic density around Cu(II) in the Cu salen@NH2-MIL-101(Cr) catalyst promoted the formation of tBuOOCu(III)-salen enhancing the selectivity to epoxide [69].
A recent synthetic strategy resides in the incorporation of different functionality in one single framework to generate a multivariate MOF (MTV-MOFs). On this basis, a chiral MOF based on multiple metallosalen bridging ligands has been synthesised. Firstly, M(salen)-derived dicarboxylate ligands H2L5M [M = Cu, VO, CrCl, MnCl, Fe(OAc), and Co(OAc)] were synthesized by reactions of N,N′-bis(3-tert-butyl-5-(carboxyl)salicylide (H4L5) and the corresponding metal salts in MeOH at room temperature. Secondly, the crystals of binary or ternary MTV-MOFs (CuV, CuMn, CuCr, CuFe, and CuCo) were obtained by heating a 1:1 or 1:1:1 mixture of H2LM with Zn(NO3)2·6H2O at 80 °C in DMF, [Zn4O(L5M,M′)3] species is obtained. Both [Zn4O(L5Cu,Mn)3] and [Zn4O(L5Cu,Fe)3] showed efficient catalytic performances for asymmetric epoxidation of alkenes, affording up to 93% and 90% ee of the epoxides, respectively. Moreover, in the ternary heterogeneous catalyst [Zn4O(L5Cu,Mn,Co)3], the combination of Mn3+ and Co3+ promotes the epoxidation of alkene followed by enantioselective hydrolysis of epoxide to afford diols [70].
The achiral Zr-MOF, [PCN-224(Mn(Cl)], based on tetratopic ligand [manganese (chloride) tetrakis(4-carboxyphenyl)porphyrin [Mn(Cl)-TCPP], has been post-synthetically modified with tartrate anion, as a chiral auxiliary. The final chiral PCN-224-Mn(tart) contains two active metal sites (Zr and Mn) as Lewis acid centers and the chiral tartrate counterion, as Brønsted acid sites (OH functional group) has been investigated as chiral nucleophile catalyst towards both asymmetric epoxidation and CO2 fixation. Asymmetric epoxidation of several aromatic and aliphatic olefins like styrene, trans-stilbene, 4-methylstyrene, α-methylstyrene, 1-phenyl-1-cyclohexene, 1-decene, and 1-octene has been tested by using PCN-224-Mn(tart) with aldehyde as co-catalyst, CH3CN, and O2. The all-reaction conversions were completed with an optimum range of epoxide selectivity 83–100% and high ee (84−100%). Several factors allow high enantioselectivity in the formation of the epoxide: the framework porosity, the active Mn center, a preferred face of the olefin (pro-S or -R face) close to produce the more stable configuration, and the noncovalent interactions between H atom of the olefinic double bond of the preferred face and chiral centers (Table 3) [71].

3. Epoxidation with MOF-Based Composites

One possible way to improve the chemical and mechanical stability of MOFs as potentially heterogeneous catalysts is their immobilization onto/into supports. In this contest, solid polymer, graphene, and inorganic particles [72] or inorganic polymers [73] are largely employed as supports.
To overcome the poor hydrostability of [Cu3-BTC2] [74], a porous dendrimer-like porous silica nanoparticles (DPSNs) has been utilized as a carrier to support Cu-BTC Nps. The nanocomposites DPSNs@Cu-BTC were prepared by growing Cu2O NPs in the center-radial porous channels of DPSNs. After that, Cu2O NPs were dissolved in the presence of acid, oxidant and 1,3,5-benzenetricarboxylic acid (H3BTC) [75]. The obtained Cu-BTC NPs have shown limited growth and a uniform distribution without agglomeration. The small size of Cu-BTC NPs (40 ± 25 nm) is useful in the aerobic epoxidation of various cyclic olefins achieving high catalytic activity without by-products. Good yield and selectivity were detected with inert terminal linear alkenes. Otherwise, epoxidation of styrene only achieved 65% of conversion due to the kinetic instability of styrene oxide (Table 4) [76].
The amphiphilic MIL-101-GH, a porous hierarchical material, has been explored as catalyst for the biphasic epoxidation reaction of 1-octene with H2O2. MIL-101-GH hydrogel was obtained by dispersing MIL-101 nanoparticles homogeneously in aqueous graphene oxide (GO) solutions. The TS-1 catalyst, commercially used in this biphasic reaction, was then introduced in MIL-101-GH. The resulting system, MIL-101-GH-TS-1, overcame the lower activity toward olefin epoxidation of TS-1, and the amphiphilic MIL-101-GH increased the contact areas of TS-1 with both H2O2 and 1-octene. The catalytic performance of MIL-101-GH-TS-1 has been much higher than that of single TS-1 and the 1,2-epoxyoctane was obtained without other by-products [77].
Polyoxometalate-based (POMs) heterogeneous catalysts are attractive species in the catalytic epoxidation of olefin. They have got great catalytic activity, selectivity, and easy separation but their leaching mainly due to the strong complexing capability of solvent and H2O2 oxidants, represents the major obstacle in the possible applications [78,79]. To overcome the stability issue of POMs, the polyoxomolybdic cobalt (CoPMA) and polyoxomolybdic acid (PMA) species were incorporated into UiO-bpy, a Zr-based MOFs, through self-assembly process under solvothermal condition [80]. CoPMA@UiO-bpy showed the highest catalytic activity for cyclooctene oxidation with H2O2 and also for the oxidation of styrene and 1-octene with O2 as oxidant and tert-butyl hydroperoxide (t-BuOOH) as initiator. This is due to the uniform distribution and better immobilization of POM clusters within the size-matched cages of Zr-MOFs owing to the presence of bipyridine groups in the UiO-bpy framework. It is noteworthy that CoPMA@UiO-bpy shows excellent recyclability and stability against the leaching of active POM species.
Composite material has been obtained by encapsulating H5-PMo10V2O40 polyoxometalates (POMs) and 1-octyl-3-methylimidazolium bromide, ionic liquids (ILs), in the mesoporous cages and large surface area of MIL-100 (Fe). The synergic effect of ILs, Lewis and Brønsted acid sites in both PMo10V2 species and MOF created a PMo10V2-ILs@MIL-100(Fe) hybrid with significant catalytic properties in cycloolefins epoxidation. Indeed, the PMo10V2 was activated by the imidazolium cations originated from ILs and the incorporation on MIL-100(Fe) prevented the leaching of POMs [81]. This composite is easily regenerated for 12 cycles without loss catalytic performance [82].
MIL-100(Fe) combined with the polyoxometalate (C16H36N)6K2[γ-SiW10O36] has been reported to catalyse epoxidation of 3Z,6Z,9Z-octadecatriene to the corresponding 6,7-epoxide with high site selectivity (82.35%). The conversion catalysed by POM/MIL-100(Fe) exhibits a greater performance when the MOF contains unsaturated Lewis acid iron ions [83]. The main product of this epoxidation is a sex pheromone of E. obliqua Prout and can be potentially used in pest insect control with environmental friendliness.
Two POMs-based MOFs, [Cu6(bip)12(PMoVI12O40)2(PMoVMoVI11O40O2)]·8H2O and [Co3IICo2III(H2bib)2(Hbib)2(PW9O34)2(H2O)6]·6H2O (H2bip = 1,3-bis(imidazolyl)propane; bib = 1,4-bis(imidazol)butane)), have been fabricated using a flexible N-containing bidentate ligands via hydrothermal condition. They have been employed in the catalytic processes for selective alkene epoxidation and recycled four times without loss of quality (Figure 7) [84].
Metal nanoparticles can grow without agglomeration in a porous matrix to produce a stable and active heterogeneous catalyst. Pd NPs have been loaded on the pre-synthesized UiO-66-NH2 using a simple solution impregnation method and NaBH4 reduction. The amino groups in the linkers allow a strong interaction with Pd (II) ions which is essential to yielding well-dispersed Pd/UiO-66-NH2 catalyst. The experiments suggest that the best catalytic activity for styrene epoxidation has been found under Pd NPs loadings of 3.69 wt% [85].
A dually functionalized catalytic system for the tandem H2O2-generation/alkene-oxidation reaction has been realized. A microcrystal of UiO-66-NH2 has been used as a platform to encapsulate Au and Pd metal NPs and later Pd/Au@UiO-66-NH2 surfaces have been post-synthetically modified with a (sal)MoVI (sal = salicylaldimine) molecular epoxidation catalyst. The porosity of Pd@UiO-66-sal(Mo) allows H2 and O2 gases to come into contact with the encapsulated NPs to generate H2O2. The synergic effect of the generated H2O2 and (sal)MoVI in a MOF enhanced epoxide productivity reducing alkene hydrogenation side reaction. This study showed that (sal)Mo moieties in Pd@UiO-66-NH2 epoxidize cis-cyclooctene substrate faster, leading to the more effective usage of the H2O2 oxidant [86].
Systems composed of a magnetic uniform Fe3O4(PAA) microspheres core and of a copper-doped MOF shell demonstrated an easily catalyst recovery approach improving turnover number and turnover frequency. In addition, these magnetic core–shell heterogeneous catalysts improve both stability of the metal active site and dispersity of catalyst materials reducing the metal leaching. Two interesting magnetic core-shell copper-doped catalysts, Fe3O4@P4VP@ZIF-8 and Fe3O4/Cu3(BTC)2 have been prepared by combining the solvothermal method with layer-by-layer assembly. Initially, monodispersed PAA-modified Fe3O4 particles were synthesized by solvothermal methods [87]. In the case of Fe3O4/Cu3(BTC)2, Fe3O4 particles were alternately immersed in solutions containing Cu(CH3COO)2·H2O and H3BTC such that Cu3(BTC)2 nanocrystals grow layer-by-layer on the surface of PAA- modified Fe3O4 particles. This nanosized porous structure increases the contact between the Cu(II) active sites present in the Cu3(BTC)2 shell and the catalytic substrates [88]. In Fe3O4@P4VP@ZIF-8 catalyst, on the other hand, the Fe3O4(PAA) core has been coated with P4VP middle layer to adsorb a large number of Zn2+ for the growth of the ZIF-8 shell thickness on the surface of the core–shell Fe3O4(PAA)@P4VP. Then, the Zn2+ ions were partially substituted by Cu2+ ions in the ZIF-8 shell framework. The ions exchange allowed a well-dispersed copper active site in the resulting copper-doped ZIF-8 structure, avoiding their leaching [89].
Aerobic epoxidation of cyclic olefins (e.g., cyclohexene, norbornene) using both magnetic core–shell copper-doped Fe3O4@P4VP@ZIF-8 and Fe3O4/Cu3(BTC)2 as heterogeneous catalyst achieved high conversion and selectivity (99%) in the formation of the epoxide under mild reaction conditions. Epoxidation of styrene by using Fe3O4@P4VP@ZIF-8 as a catalyst has brought only 54% selectivity of the desired epoxide owing to the kinetic instability of styrene oxide and its oxidation into benzaldehyde [90].
A series of Zr-based core-shell MOF composites with mesoporous cores and microporous shells have been synthesized by solvothermal under kinetic control. PCN-222(Fe) crystals have been synthesized and used as seed crystals to grow the Zr-BPDC(UiO-67) crystals. Meso- and micro-porosity inside of PCN-222(Fe)@Zr-BPDC(UiO-67) drives the catalytic performances for olefin epoxidation reaction [91]. Indeed, the core MOF with Fe-porphyrin moieties represents the catalytic center, while the shell controls the selectivity of the substrate through tuneable pore size. This size-selective catalyst showed almost complete conversions for small olefins.
Table 4. MOF-based composites for epoxidation reaction.
Table 4. MOF-based composites for epoxidation reaction.
MOFSubstrateReaction Data
T (°C) P (atm) Time (h)
Oxidant/Cocatalyst/Solvent aConversion
%
Epoxide
Selectivity%
Ref.
DPSNs@Cu-BTCCyclooctene4014O2/TMA/CH3CN9999[75]
Styrene4016O2/TMA/CH3CN6265[75]
Fe3O4@P4VP@ZIF-8Cyclohexene
Cyclooctene
Norbornene
60112O2/TMA/CH3CN9999[90]
Fe3O4/Cu3(BTC)2Cyclohexene
Cyclooctene
Norbornene
4016–8O2/IBA/CH3CN9999[88]
Styrene4016–8O2/IBA/CH3CN9984[88]
PCN-222(Fe)@Zr-BPDC(UiO-67)1-Hexener.t112PhIO/-/CH3CN99-[91]
Cyclopentener.t112PhIO/-/CH3CN99-[91]
Cyclohexener.t112PhIO/-/CH3CN99-[91]
CoPMA@UiO-bpyCyclooctene7016H2O2/-/CH3CN9199[80]
Styrene8016O2/t-BuOOH/- 8056[80]
PMo10V2-ILs@MIL-100(Fe)Cyclohexene6014H2O2/-/CH3CN9293[91]
[Cu6(bip)12(PMoVI12O40)2(PMoVMoVI11O40O2)]·8H2OCyclooctene2014H2O2/tBuOH/CH3CN>9974.1[84]
1−Hexene2014H2O2/tBuOH/CH3CN>9991.9[84]
1−Octene2014H2O2/tBuOH/CH3CN>9971.5[84]
Pd/UiO-66-NH2Styrene80112N2/TBHP/CH3CN90.896.5[85]
[Co3IICo2III(H2bib)2(Hbib)2(PW9O34)2(H2O)6] ·6H2OCyclohexene2014H2O2/tBuOH/CH3CN72.995.3[84]
1−Hexene2014H2O2/tBuOH/CH3CN>9985.9[84]
1−Octene2014H2O2/tBuOH/CH3CN95.570.1[84]
POM/MIL-100(Fe)3Z,6Z,9Z-Octadecatriene40124H2O2/-/CH3CN3082[83]
MIL-101-GH-TS-1Octane40112H2O2 (30%)/-/-15-[77]
Pd@UiO-66-sal(Mo)cis-Cyclooctener.t16H2O2/CH3OH/H2O--[86]
a tBuOH = tert-butyl alcohol; TMA = trimethylacetaldehyde; IBA = isobutyraldehyde.

4. CO2 Epoxide Cycloaddition to Cyclic Carbonates

Cycloaddition reaction of CO2 with epoxides represents one of the most economically efficient approaches in the production of cyclic organic carbonates with relevant applications ranging from raw materials in the pharmaceuticals industry, polar aprotic solvents, electrolytes in lithium batteries, lubricants, precursors for polycarbonate materials, and other fine chemicals.
CO2, being a C1 feedstock has, in fact, a high potential from the chemical point of view [92]. CO2 can be employed in the highly atom-economical acid-catalysed epoxides cycloaddition to give cyclic organic carbonates, relevant species for industrial applications [93]. The cyclic carbonates (OCs) have been also used as intermediates for engineered polymers, as a lubricant (in 1987 Agip Petroli added dialkylcarbonates as lubricant in a formulation of semisynthetic gasoline engine oil components), and more recently found application in varnish production, green solvents or electrolytes in lithium-ion batteries.
The CO2 cycloaddition mechanism involves an acid catalyst (Lewis or Brønsted acid) that coordinates to the epoxide substrate activating it toward nucleophilic attack by the co-catalyst (e.g., typically a tetraalkylammonium halide). The resulting halo-alkoxide intermediate reacts with carbon dioxide to generate the cyclic carbonate and subsequently regeneration of both catalyst and co-catalyst [93].
The CO2 fixation reaction catalysed by homogeneous or heterogeneous catalysts has been extensively investigated, however some drawbacks remain. Differently from the homogeneous, heterogeneous catalysts (e.g., ionic liquid-supported solids [94,95,96], polymers [96,97], and porous organic frameworks [98]) have the advantages of easy separation and regeneration of the catalyst, but they often required rough conditions (high temperature, pressure, and time) due to a lack of accessible surface area for accelerating interactions of CO2 and reagents with active sites. Therefore, the high surface area, tunability, and CO2 sorption capacity of MOFs can be beneficial for improving the efficiency of the CO2 cycloaddition reaction. Lewis acid metal centers and Brønsted acid groups in MOFs can promote the activation of the epoxide ring, while the functional groups in the ligands can act as Lewis/basic sites improving not only the CO2 affinity inside the pore but also can fulfil the role of co-catalyst (Figure 8) [99].
The Hf-cluster-based NU-1000 (Hf-NU-1000) demonstrated excellent catalytic activity, greater than the Zr-cluster-based NU-1000 under the same mild reaction conditions [100]. Indeed, the presence of high density stronger acidic Brønsted sites, due to stronger M−O bonds, gave a complete and quantitative conversion of styrene oxide and propylene oxide to form cyclic carbonates. Moreover, high yields have been detected for the cycloaddition reaction of CO2 with industrially important epoxide divinylbenzene dioxide (DVBDO) [101].
Large pores in the MOFs, easily functionalized by polar groups, can promote CO2 fixation in a short reaction time under ambient CO2 pressure and moderate temperature without the use of solvent. Within the mesoporous M-MOF-184 series (M = Co, Ni, Mg, Zn), Zn-MOF-184 achieved efficient catalysis performances to convert CO2 to cyclic carbonates under ambient conditions for several epoxy substrate, due to the presence of high concentration of accessibly acidic metals, basic 2-oxidobenzoate anion sites and to the high polarity induced by C C bonds and π systems from the phenyl rings in the linkers. Low conversion has been detected for larger epoxides due to limit diffusion into the MOF pores of reactants toward the active sites [102]. The hydrothermally synthesized flexible Zn-based {[Zn2(TBIB)2(HTCPB)2]·9DMF·19H2O}n, has been synthesized employing two types of large linkers 1,3,5-tri(1H-benzo[d]imidazol-1-yl)benzene (TBIB) and 1,3,5-tris(4′-carboxyphenyl-)benzene (H3TCPB). A porous structure with 1D channels was generated via noncovalent supramolecular interactions between the layers. The presence of free protonated carboxylic acid groups(−COOH), carbonyl groups (−C=O), and the presence of Lewis basic sites from the rich N-containing TBIB on the surface pores enhance the selectivity toward CO2. Moreover, the COOH group helps in catalysing the CO2 cycloaddition reaction efficiently through noncovalent interaction with the epoxide substrate, followed by ring-opening upon nucleophilic attack of co-catalyst [103].
Excellent conversions of epichlorohydrin and 2-vinyloxirane have been obtained using as heterogeneous catalyst [Zn4OL43]n based on the meta-substituted 2,2′,6,6′-tetramethoxy-4,4′-biphenyldicarboxyate ligand [39].
Zeolitic imidazolate frameworks are known for their high CO2 solubility and capture ability [104], especially the chloro-functionalized ZIF-95 [105]. The CO2 cycloaddition to propylene oxide by using ZIF-95 and a quaternary ammonium salt as cocatalyst procured over 99% selectivity to the desired propylene carbonate product under moderate conditions [106]. Also the imidazolate-containing species Im-UiO66(Zr)MOF reacts with methyl iodine to produce (I)MeIm-UiO-66 that demonstrate efficiency in the CO2 cycloaddition reaction toward a broad range of substrates, in this case without the addition of co-catalyst [107].
Conversely, imidazolium-based IL units were grafted and immobilized into UiO-67 via direct ligand functionalization that, considering the post synthetic approach, is a quantitative method. The obtained species show a high density of IL sites. UiO-67-IL converts epichlorohydrin substrate in 95% yield under co-catalyst and solvent-free conditions. The yield increases to 99% in a shorter time when TBAB was employed (TBAB = tetrabutylammonium bromide) [108] (Table 5).
UiO-66-NH2 pores were modified with ILs such as methylimidazolium bromide and methylbenzimidazolium bromide by coupling reactions, to generate ILA@U6N and ILB@U6N MOFs. The Lewis acid sites (for activation of the epoxide) and the IL functional sites (for epoxide ring-opening) efficiently catalyse the epichlorohydrin conversion under mild conditions [109].
A linear ionic polymer was inserted inside the MIL-101(Cr) via in situ polymerization to form polyILs@MIL-101(Cr) stable heterogeneous composites. This polyILs@MIL-101 is able to catalyse the CO2 cycloaddition reaction with various epoxides with good to excellent conversions, including terminal epoxides with both electron-withdrawing and electron-donating substituents without the need of co-catalyst [110].
A new multimodal catalytic system has been designed via two steps post-synthetic modification of the metal nodes in the NU-1000 framework. A tandem functionalization was performed starting from the incorporation of ortho-, meta-, and para-pyridinecarboxylic acids into the framework of NU-1000(M), then the pyridine moieties were alkylated with various haloalkanes (CH3I, C4H9I, C4H9Br, and C6H4F9I) to introduce co-catalyst moieties near to the inorganic node [111]. Among catalysts, NU-1000(Zr) functionalized with 4-PyCOOH and CH3I, i.e., SALI-4-Py-I-(Zr), showed the highest styrene carbonate yield without co-catalyst, the epoxy ring being activated upon coordination to Zr4+ center (Lewis acid site) and the halogen anion opening the epoxy ring by nucleophilic attack on the less sterically hindered carbon atom [111].
Two 3D metal-cyclam-based zirconium MOFs [Zr63-OH)8(OH)8(M-L)4] (where M = Cu(II) or Ni(II), L1 = 6,13-dicarboxy-1,4,8,11-tetraazacyclotetradecane) were prepared, namely VPI-100 (Cu) and VPI-100 (Ni) (VPI = Virginia Polytechnic Institute), respectively. A two-step solvothermal synthesis has been necessary to build the MOFs. Initially, a zirconium-oxo cluster was assembled, then cyclam was added. The presence of accessible Cu2+/Ni2+ metal active sites in the metallocyclams and of the coordinatively unsaturated Zr4+ sites in the equatorial plane of the Zr6 cluster in VPI-100 improved their catalytic activity toward CO2 cycloaddition to various organic epoxides [112].
Another strategy developed to increase the catalytic performances is based on the incorporation of an amine group in MOFs. Essentially, the amino group has the dual advantage of acting as an electron donor (Lewis base) toward CO2 and increasing the local concentration of CO2 near catalytic centres through a high CO2 adsorption [113,114].
The amine-functionalized NH2-MIL-101(Al) has been synthesized using a solvothermal or microwave method and its catalytic activity in the solvent-free cycloaddition of CO2 to styrene oxide achieved nearly total conversion and selectivity in 96% yield, with a TOF of 23.5 h−1 [115]. The coordinatively unsaturated aluminium centers present in the SBUs (Lewis acidic sites) bind the epoxides and activate them toward ring-opening, this step is immediately followed by the attack of the bulky bromide ions of TBAB. The pendant amino groups polarize the CO2 molecules, through the nucleophilic attack at the carbon atom, and facilitate CO2 insertion and cycloaddition (Figure 9). During the catalytic reaction, the micro and mesoporous of the framework facilitate the diffusion of substrates and reactants to enhance their interactions [116].
Recently, the acid-base pair UiO-66-NH2 has been used to synthesize bio-based five-membered cyclic carbonate from vegetable oil methyl ester by CO2 fixation. At first, 95% of double bonds in the O-acetyl methyl ricinoleate starting material were converted to epoxide through an enzyme-catalyzed process. Then, the cycloaddition of epoxy fatty acid methyl esters was performed in the presence of UiO-66-NH2 as catalyst and TBAB as co-catalyst for CO2 fixation. At 120 °C under 3 MPa CO2 pressure for 12 h, the reaction conversion reached 94.4% [117] (Table 5).
A series of diamino-tagged zinc bipyrazolate MOFs have been investigated as heterogeneous catalyst in the reaction of CO2 with the epoxides epichlorohydrin and epibromohydrin to give the corresponding cyclic carbonates at 393 K and pCO2 5 bar under relatively mild conditions (solvent and co-catalyst-free) [118]. The presence of amino group in the MOFs pores increased the CO2 storage capacity as well as the catalytic performances. The epoxide has been activated through halogen-amine interaction which was observed in structure of the [epibromohydrin@Zn(3,3′-(NH2)2BPZ)] adduct. The isomeric Lewis basic site (NH2) in Zn(3,5 NH2-Bpz) (64% yield) improves more than twice the catalytic transformation of epichlorohydrin compared to its mono(amino) parent Zn(BPZNH2) (32% yield) [118].
Post-synthetic metalation of organic linkers is employed strategically to tailor the MOFs’ properties. In Hf-Bipy-UiO-67, the 2,2-bipyridine-5,5-dicarboxylate ligand was grafted with Mn(OAc)2 and the resulting Hf-Bipy-UiO-67(Mn(OAc)2 showed that synergy of the binary Lewis acid function significantly enhances the CO2 uptake capacity and catalytic performance of the cycloaddition reaction under mild conditions [119].
Vanadium chlorides have been used to produce the post-metalated Zr-based MOF-VCl3 and MOF-VCl4, with biphenyl-4,4′-dicarboxylic acid, and 2,2′-bipyridine- 5,5′-dicarboxylic acid, respectively, which provide Lewis basic sites. Their high catalytic activity in the CO2 cycloaddition to various organic epoxides was attributed to the accessible Cu2+/Ni2+ metal active sites in the metallocyclams and the presence of coordinatively unsaturated Zr4+ sites in the equatorial plane of the Zr6 cluster in VPI-100 MOFs [112].
UiO-type MOFs become susceptible to water and alkaline solution when the length of the carboxylic linker increase. A series of UiO-type MOF named ZSF, incorporating chiral metallosalen as linker has been produced. The chemically stable ZSF-1 MOF, synthesized by dissolving a mixture of ZrCl4, Cy-salen-Ni, and modulators (trifluoroacetic acid), showed excellent catalytic performance for the conversion of CO2 with epoxides into cyclic carbonates. The tetrahedral cages of ZSF-1 decorated with salen-Ni moieties entrap efficiently CO2 and activate the substrate. ZSF-1 catalyses efficiently the asymmetric cycloaddition of CO2 with styrene oxide giving 94% yield of the resulting cyclic carbonate [120]. With other epoxides, specifically epichlorohydrin, the catalytic activity of ZSF-1 increases until to 99% of conversion thanks to the presence of electron-withdrawing Cl group, which promotes the nucleophilic attack of Br during the ring-opening process.
The chiral PCN-224-Mn(tart) (see Section 2.3) has been used in asymmetric CO2 cycloaddition to styrene epoxide, its derivative showing conversions of 96% and 87%, respectively. The missing-linker defects in the Zr cluster and in the Mn center are Lewis acids inducing catalytic ability into the framework for CO2 chemical fixation. In addition, the auxiliary chiral tartrate anions, and the co-catalyst (Bu4NBr) act as nucleophiles generating a chiral epoxide, semi-intermediate, starting from prochiral styrene substrate. The CO2 addition leads asymmetrically to cyclic carbonate with a high ee, and it is related to the interaction of the chiral centers and substrate pro R/S face. Moreover, catalytic reactions with PCN-224-Mn(tart) were performed at low energy and ambient pressure and temperature [71].
Table 5. MOF-based composites for cycloaddition reaction.
Table 5. MOF-based composites for cycloaddition reaction.
MOFSubstrateReaction Data
T (°C) CO2 P (atm) Time (h)
Cocatalys aConversion
%
Cyclic Carbonate
Selectivity%
Ref.
Hf-NU-1000Styrene epoxider.t.156TBAB100100[101]
Propylene oxider.t.126TBAB100100[101]
Epoxide divinylbenzene dioxider.t.119TBAB100100[101]
PCN-224-Mn(tart)Styrene epoxide60115TBAB96
94 ee (S)
100[71]
(2,3-Epoxypropyl)benzene60115TBAB87
90 ee (S)
100[71]
Propylene oxide60115TBAB99
98 ee (S)
100[71]
1,2-Epoxybutane60115TBAB91
97 ee (S)
100[71]
1,2-Epoxyoctane60115TBAB78
96 ee
100[71]
polyILs@MIL-101(Cr)1-Butene oxide45148-94100[110]
1,2-Epoxyhexane70124-89100[110]
3-Hydroxy-1,2-epoxypropane70124->99100[110]
1,2-Epoxy-3-phenoxypropane70124-95100[110]
NH2-MIL-101(Al)Styrene oxide120186TBAB93.699[115]
UiO-66-NH2Epoxy fatty acid methyl ester1203012TBAB9480[117]
ILB@U6NEpichlorohydrin801184-9499[109]
(I)Meim-UiO-66Epichlorohydrin120124-10093[107]
VPI-100 (Ni)Epichlorohydrin90106TBAB96-[112]
VPI-100 (Cu) Epichlorohydrin90106TBAB94-[112]
[Zn4OL43]nEpichlorohydrin 5014-9699[107]
2-Vinyloxirane5014-9981[107]
Hf-Bipy-UiO-67(Mn(OAc)2Epichlorohydrin25112TBAB83.299[119]
Zn-MOF-184Styrene oxide8016TBAB9685[102]
Propylene oxide8016TBAB10075[102]
Epichlorohydrin8016TBAB10070[102]
Cyclohexene oxide8016TBAB6985[102]
SALI-4-Py-I-(Zr),Styrene oxide 8044-99 98[111]
ILA@U6N Epichlorohydrin801184-6599[109]
UiO-67-ILEpichlorohydrin9013TBAB99100[108]
Epichlorohydrin9013-9996[108]
{[Zn2(TBIB)2(HTCPB)2]·9DMF·19H2O}nEpichlorohydrinr.t124TBAB99100[105]
ZIF-95Propylene oxide12011824TBAB9199[106]
ZSF-1Styrene oxide100120TBAB93-[120]
Epichlorohydrin100120TBAB99-[120]
Zn(3,5-NH2-Bpz)Epichlorohydrin120524 -9850[118]
Zn(BPZNH2)Epichlorohydrin120524-9633[118]
a TBAB = tetrabutylammonium bromide.

5. Conclusions

MOF-based catalysts are now a very promising class of compounds as they merge relevant characteristics of both homogeneous and heterogeneous catalysts. They can be easily modified by changing linkers substituents to increase affinity for reactants, or by growing the number of active catalytic sites.
In this review, we have explored the ability of MOFs, MOF nanocomposites and mixed metal species toward olefin epoxidation and carbon dioxide cycloaddition.
We have observed that the olefin conversion and the epoxide selectivity are strongly dependent on the metal nodes/clusters, Co and Cu species being the most efficient, in some cases as for the epoxidation of a-pinene by Co-MOF-150-2 a conversion and an epoxide selectivity close to 100% being found.
Mixed metal MOFs can be also successfully employed in styrene and cyclohexene epoxidation, the best results being obtained with Cu/Co, Mn/Cu, and Ni/V species.
Selected functional groups introduced in organic linkers can also act as catalytically active sites. Amino, pyridyl, amide and sulfonic acid groups, but also metalloporphyrins, vanadium and molybdenum acetylacetonate, tartaric acid, salen and analogous molecules can be inserted or deposited to obtain also greater selectivity. UiO-66, UiO67, and PCN-224, appropriately functionalized can induce a complete conversion and selectivity as in the case of the geraniol epoxidation.
MOF-based composites are often employed to increase the hydrostability of selected MOFs or to perform epoxidation also of specific substrates as norbornene or octadecatriene. Specifically, a porous dendrimer-like porous silica nanoparticles (DPSNs) used as a carrier to support Cu-BTC NPs overcame the poor hydrostability of [Cu3-BTC2] MOF achieving high catalytic activity without by-products under mild reaction conditions.
Finally, MOFs and MOF-based composites show a great efficiency toward CO2 cycloaddition to epoxides, conversion being generally in the range 70–100% and selectivity close to 100%. The use of chiral ligands and amine-functionalized ligands seems to be very promising. The CO2 binding mode can in fact open new strategies for activation of CO2 and its transformation.
However, the low reactivity and inert nature of CO2 make its incorporation and activation into organic substrates still a challenge. Currently, the heterogeneous MOFs-based catalysts, as well as the technical system, remain at the laboratory scale and that makes the costs of productions of these materials extremely pricey. It is desirable that the improvement of MOFs-based catalysts might lead to technically viable efficiencies to industrial production to allow their large-scale application, in the next future. This review clearly shows that MOFs are now perspective materials and valid candidates for catalytic epoxidation and CO2 cycloaddition reactions.

Author Contributions

Conceptualization, A.T. and C.P.; methodology, A.T. and C.P.; software, A.T. and C.P.; validation, A.T. and C.P.; data curation, A.T. and C.P.; writing—original draft preparation, A.T. and C.P.; writing—review and editing, A.T. and C.P.; funding acquisition, C.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by University of Camerino.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MOFs
Co6-MOF-3[(Co6(OH)6(TCA)2(BPB)3]n
Co-MOF-150-2[Co(BDC)]n
Cu0.25-Co0.75-MOF[(Cu0.25-Co0.75)3(BTC)2]n
HKUST-1[Cu3(BTC)2]n
LaCoODa{[La2Co3(ODA)6(H2O)6]·12H2O}n
LaCuODA{[La2Cu3(µ-H2O)(ODA)6(H2O)3]∙3H2O}n
MIL-100(Fe)[Fe3O(OH)(H2O)2(BDC)3]n
MIL-101[Cr3O(H2O)2F(BDC)]n
Mn0.1Cu0.9-MOF[(Mn0.1-Cu0.9)3(BTC)2]n
NENU-MV-1{[Ni(4,4′-bpy)2]2[V7IVV9VO38Cl]·(4,4′-bpy)·6H2O}n
NH2-MIL-101(Al)[Al3O(OH)(H2O)2(BDCNH2)3]n
NH2-MIL-101(Cr)Cr3O(H2O)2F(NH2-BDC)
NTUZ30{[Co33-OH)(HBTC)(BTC)2Co(HBTC)]·(HTEA)3·H2O}n
PCN-222[Zr63-OH)8(OH)8-(TCPP)2]n
PCN-224[Zr63-OH)12(OH)16-(TCPP)1.5]n
TMU-16-NH2{[Zn2(NH2-BDC)2(4-bpdh)]·3DMF}n
UiO-66[Zr6O4(OH)4(BDC)6]n
UiO-66-NH2[Zr6O4(OH)4(NH2-BDC)6]n
UiO-67[Zr63-O)43-OH)4(BPDC)6]n
UiO-68[Zr63-O)43-OH)4(TPDC)6]n
UiO-bpy[Zr6O4(OH)4(bpy)6]n
VPI-100(Cu)[Zr63-OH)8(OH)8(Cu-L1)4]n
VPI-100(Ni)[Zr63-OH)8(OH)8(Ni-L1)4]n
ZIF-67[Co(MeIm)2]n
ZIF-8[Zn(MeIm)2]n
ZIF-95[Zn(cbIm)2]n
Zn-MOF-184[Zn2(EDOB)]n
Zr-NU-1000([Zr63-O)43-OH)4(OH)4(H2O)4(TBAPy)2]n
ZSF-1[Zr6O4(OH)4(metallosalen)6]n
Hf-NU-1000[(Hf63-O)43-OH)4(OH)4(OH2)4(TBAPy)2]n

Appendix A. Chart of the MOF Linkers Present in This Review and Their Relative Abbreviations

Structural FormulaNameAbbreviation
Inorganics 09 00081 i0011-(4-Cyanobenzyl)-5-methyl-1H-imidazolecbIm
Inorganics 09 00081 i0021,3,5-tri(1H-Benzo[d]imidazol-1-yl)benzeneTBIB
Inorganics 09 00081 i0031,3,5-tris(4′-Carboxy-phenyl-)benzeneH3TCPB
Inorganics 09 00081 i0041,3,6,8-(p-Benzoate)pyreneH4TBAPy
Inorganics 09 00081 i0051,3-bis(Imidazolyl)propaneH2bip
Inorganics 09 00081 i0061,4-bis(Imidazol)butanebib
Inorganics 09 00081 i0072,2′,6,6′-Tetramethoxy-4,4′-biphenyldicarboxylic acidH2L4
Inorganics 09 00081 i0082,2-Bipyridine-4,4′-dicarboxylic acidH2Bpy
Inorganics 09 00081 i0092,5-bis(4-Pyridyl)-3,4-diaza-2,4-hexadiene4-bpdh
Inorganics 09 00081 i0102-Aminoterephthalic acidH2BDCNH2
Inorganics 09 00081 i0112-MethylimidazoleHmeIm
Inorganics 09 00081 i0122-Picolinic acidPCA
Inorganics 09 00081 i0133,5-Diamino-4,4′-bipyrazoleH2-NH2-Bpz
Inorganics 09 00081 i0144-((1-Carboxy-2-(1H-imidazol-4-yl)ethylamino)methyl)benzoic acidH2L2
Inorganics 09 00081 i0154-(3,5-Dimethyl-4-H-1,2,4-triazol-4-yl) benzoateMe2trz-pba
Inorganics 09 00081 i0164,4′-Bipyridine4,4′-bipy
Inorganics 09 00081 i0174,4′-(Ethyne-1,2-diyl)bis(2-hydroxybenzoic acid)H4EDOB
Inorganics 09 00081 i0184,4′,4′′-TricarboxyltriphenylamineH3TCA
Inorganics 09 00081 i0195-(3-Methyl-5-(pyridine-4-yl)-4H-1,2,4-triazol-4-yl) isophthalateMe-4py-trz-ia
Inorganics 09 00081 i0205,10,15,20-Tetrakis(4-carboxyphenyl)porphyrinH6TCPP
Inorganics 09 00081 i0215-Dihydroxyterephthalic acidH4DHTA
Inorganics 09 00081 i0226,13-Dicarboxy-1,4,8,11-tetraazacyclotetradecaneL1
Inorganics 09 00081 i023Benzene-1,3-5 tricarboxylic acidH3BTC
Inorganics 09 00081 i024Biphenyl-4,4′-dicarboxylic acidH2BPDC
Inorganics 09 00081 i025ImidazoleHIm
Inorganics 09 00081 i026IsonicotinateIna
Inorganics 09 00081 i027N,N′-bis(3-tert-Butyl-5-(carboxy)salicylideH4L5
Inorganics 09 00081 i028Oxydiacetic acidH2ODA
Inorganics 09 00081 i029Pyridine-2,5-dicarboxylic acidH2pdc
Inorganics 09 00081 i030Pyridine-2-aldehydePI
Inorganics 09 00081 i031SalicylaldehydeSA
Inorganics 09 00081 i032Thiophene-2-carbaldehydeTC
Inorganics 09 00081 i033TriethylamineTEA
Inorganics 09 00081 i034Tris(4′-carboxybiphenyl)amineH3L3

References

  1. Batten, S.R.; Champness, N.R.; Chen, X.-M.; Garcia-Martinez, J.; Kitagawa, S.; Öhrström, L.; O’Keeffe, M.; Suh, M.P.; Reedijk, J. Terminology of metal–organic frameworks and coordination polymers (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85, 1715–1724. [Google Scholar] [CrossRef] [Green Version]
  2. 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]
  3. Pettinari, C.; Marchetti, F.; Mosca, N.; Drozdov, A. Application of metal-organic frameworks. Polym. Int. 2017, 66, 93. [Google Scholar] [CrossRef]
  4. Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nat. Cell Biol. 1999, 402, 276–279. [Google Scholar] [CrossRef] [Green Version]
  5. Schaus, S.E.; Brandes, B.D.; Larrow, J.F.; Tokunaga, M.; Hansen, K.B.; Gould, A.E.; Furrow, M.E.; Jacobsen, E.N. Highly Selective Hydrolytic Kinetic Resolution of Terminal Epoxides Catalyzed by Chiral (salen)CoIII Complexes. Practical Synthesis of Enantioenriched Terminal Epoxides and 1,2-Diols. J. Am. Chem. Soc. 2002, 124, 1307. [Google Scholar] [CrossRef]
  6. Stock, N.; Biswas, S. Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2011, 112, 933. [Google Scholar] [CrossRef]
  7. Campagnol, N.; Souza, E.R.; 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, 12545–12547. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, Y.; Bo, X.; Nsabimana, A.; Han, C.; Li, M.; Guo, L. Electrocatalytically active cobalt-based metal–organic framework with incorporated macroporous carbon composite for electrochemical applications. J. Mater. Chem. A 2014, 3, 732. [Google Scholar] [CrossRef]
  9. Masoomi, M.Y.; Morsali, A.; Junk, P.C. Rapid mechanochemical synthesis of two new Cd(ii)-based metal–organic frameworks with high removal efficiency of Congo red. CrystEngComm 2014, 17, 686. [Google Scholar] [CrossRef]
  10. Wu, J.-Y.; Chao, T.-C.; Zhong, M.-S. Influence of Counteranions on the Structural Modulation of Silver–Di(3-pyridylmethyl)amine Coordination Polymers. Cryst. Growth Des. 2013, 13, 2953. [Google Scholar] [CrossRef]
  11. Khan, N.A.; Jhung, S.H. Synthesis of metal-organic frameworks (MOFs) with microwave or ultrasound: Rapid reaction, phase-selectivity, and size reduction. Coord. Chem. Rev. 2015, 285, 11. [Google Scholar] [CrossRef]
  12. Maurin, G.; Serre, C.; Cooper, A.; Férey, G. The new age of MOFs and of their porous-related solids. Chem. Soc. Rev. 2017, 46, 3104–3107. [Google Scholar] [CrossRef]
  13. Murray, J.L.; Mircea, D.; Long, R.J. Hydrogen storage in metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1294–1314. [Google Scholar] [CrossRef]
  14. Sun, Y.; Zheng, L.; Yang, Y.; Qian, X.; Fu, T.; Li, X.; Yang, Z.; Yan, H.; Cui, C.; Tan, W. Metal–Organic Framework Nanocarriers for Drug Delivery in Biomedical Applications. Nano-Micro Lett. 2020, 12, 103. [Google Scholar] [CrossRef]
  15. Wang, H.-S. Metal–organic frameworks for biosensing and bioimaging applications. Coord. Chem. Rev. 2017, 349, 139–155. [Google Scholar] [CrossRef]
  16. Gao, X.; Dong, Y.; Li, S.; Zhou, J.; Wang, L.; Wang, B. MOFs and COFs for Batteries and Supercapacitors. Electrochem. Energy Rev. 2019, 3, 81–126. [Google Scholar] [CrossRef]
  17. Stavila, V.; Talin, A.A.; Allendorf, M.D. MOF-based electronic and opto-electronic devices. Chem. Soc. Rev. 2014, 43, 5994. [Google Scholar] [CrossRef] [Green Version]
  18. Xu, W.; Yaghi, O.M. Metal–Organic Frameworks for Water Harvesting from Air, Anywhere, Anytime. ACS Central Sci. 2020, 6, 1348–1354. [Google Scholar] [CrossRef]
  19. Zhang, L.; Li, H.; He, H.; Yang, Y.; Cui, Y.; Qian, G. Structural Variation and Switchable Nonlinear Optical Behavior of Metal–Organic Frameworks. Small 2021, 17, 2006649. [Google Scholar] [CrossRef]
  20. Dhakshinamoorthy, A.; Opanasenko, M.; Čejka, J.; Garcia, H. Metal organic frameworks as heterogeneous catalysts for the production of fine chemicals. Catal. Sci. Technol. 2013, 3, 2509–2540. [Google Scholar] [CrossRef]
  21. Vanesa, C.-C.; Rosa, M.M.-A. Advances in Metal-Organic Frameworks for Heterogeneous Catalysis. Recent Pat. Chem. Eng. 2011, 4, 1–16. [Google Scholar]
  22. Farrusseng, D.; Aguado, S.; Pinel, C. Metal-Organic Frameworks: Opportunities for Catalysis. Angew. Chem. Int. Ed. 2009, 48, 7502–7513. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, Y.; Ma, S. Biomimetic catalysis of metal–organic frameworks. Dalt. Trans. 2016, 45, 9744–9753. [Google Scholar] [CrossRef]
  24. Corma, A.; García, H. Lewis Acids as Catalysts in Oxidation Reactions: From Homogeneous to Heterogeneous Systems. Chem. Rev. 2002, 102, 3837–3892. [Google Scholar] [CrossRef]
  25. Kravchenko, D.E.; Tyablikov, I.A.; Kots, P.A.; Kolozhvari, B.A.; Fedosov, D.A.; Ivanova, I.I. Olefin Epoxidation over Metal-Organic Frameworks Modified with Transition Metals. Pet. Chem. 2019, 58, 1255–1262. [Google Scholar] [CrossRef]
  26. Arai, H.; Uehara, K.; Kinoshita, S.I.; Kunugi, T. Olefin Oxidation-Mercuric Salt-Active Charcoal Catalysis. Ind. Eng. Chem. Prod. Res. Dev. 1972, 11, 308–312. [Google Scholar] [CrossRef]
  27. Liniger, M.; Liu, Y.; Stoltz, B.M. Sequential Ruthenium Catalysis for Olefin Isomerization and Oxidation: Application to the Synthesis of Unusual Amino Acids. J. Am. Chem. Soc. 2017, 139, 13944–13949. [Google Scholar] [CrossRef]
  28. Gao, J.; Bai, L.; Zhang, Q.; Li, Y.; Rakesh, G.; Lee, J.-M.; Yang, Y.; Zhang, Q. Co63-OH)6 cluster based coordination polymer as an effective heterogeneous catalyst for aerobic epoxidation of alkenes. Dalton Trans. 2014, 43, 2559–2565. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, X.; Zhang, Y.-Z.; Jin, Y.-Q.; Geng, L.; Zhang, D.-S.; Hu, H.; Li, T.; Wang, B.; Li, J.-R. Pillar-Layered Metal–Organic Frameworks Based on a Hexaprismane [Co63-OH)6] Cluster: Structural Modulation and Catalytic Performance in Aerobic Oxidation Reaction. Inorg. Chem. 2020, 59, 11728–11735. [Google Scholar] [CrossRef]
  30. Yu, F.; Xiong, X.; Huang, K.; Zhou, Y.; Li, B. 2D Co-based coordination polymer with a histidine derivative as an efficient heterogeneous catalyst for the oxidation of cyclohexene. CrystEngComm 2017, 19, 2126–2132. [Google Scholar] [CrossRef]
  31. Zhang, H.; He, J.; Lu, X.; Yang, L.; Wang, C.; Yue, F.; Zhou, D.; Xia, Q. Fast-synthesis and catalytic property of heterogeneous Co-MOF catalysts for the epoxidation of α-pinene with air. New J. Chem. 2020, 44, 17413–17421. [Google Scholar] [CrossRef]
  32. Lu, H.-S.; Bai, L.; Xiong, W.-W.; Li, P.; Ding, J.; Zhang, G.; Wu, T.; Zhao, Y.; Lee, J.-M.; Yang, Y.; et al. Surfactant Media To Grow New Crystalline Cobalt 1,3,5-Benzenetricarboxylate Metal–Organic Frameworks. Inorg. Chem. 2014, 53, 8529–8537. [Google Scholar] [CrossRef]
  33. Sha, S.; Yang, H.; Li, J.; Zhuang, C.; Gao, S.; Liu, S. Co(II) coordinated metal-organic framework: An efficient catalyst for heterogeneous aerobic olefins epoxidation. Catal. Commun. 2014, 43, 146–150. [Google Scholar] [CrossRef]
  34. Fu, Y.; Sun, D.; Qin, M.; Huang, R.; Li, Z. Cu(ii)-and Co(ii)-containing metal–organic frameworks (MOFs) as catalysts for cyclohexene oxidation with oxygen under solvent-free conditions. RSC Adv. 2012, 2, 3309–3314. [Google Scholar] [CrossRef]
  35. Lashanizadegan, M.; Ashari, H.A.; Sarkheil, M.; Anafcheh, M.; Jahangiry, S. New Cu(II), Co(II) and Ni(II) azo-Schiff base complexes: Synthesis, characterization, catalytic oxidation of alkenes and DFT study. Polyhedron 2021, 200, 115148. [Google Scholar] [CrossRef]
  36. Junghans, U.; Suttkus, C.; Lincke, J.; Lässig, D.; Krautscheid, H.; Gläser, R. Selective oxidation of cyclooctene over copper-containing metal-organic frameworks. Microporous Mesoporous Mater. 2015, 216, 151–160. [Google Scholar] [CrossRef]
  37. Wang, B.; Jin, J.; Ding, B.; Han, X.; Han, A.; Liu, J. General Approach to Metal-Organic Framework Nanosheets With Controllable Thickness by Using Metal Hydroxides as Precursors. Front. Mater. 2020, 7, 37. [Google Scholar] [CrossRef] [Green Version]
  38. Shi, D.; Ren, Y.; Jiang, H.; Cai, B.; Lu, J. Synthesis, Structures, and Properties of Two Three-Dimensional Metal–Organic Frameworks, Based on Concurrent Ligand Extension. Inorg. Chem. 2012, 51, 6498–6506. [Google Scholar] [CrossRef] [PubMed]
  39. Li, J.; Ren, Y.; Qi, C.; Jiang, H. Fully meta-Substituted 4,4′-Biphenyldicarboxylate-Based Metal–Organic Frameworks: Synthesis, Structures, and Catalytic Activities. Eur. J. Inorg. Chem. 2017, 11, 1478–1487. [Google Scholar] [CrossRef]
  40. Saha, D.; Gayen, S.; Koner, S. Cu(II)/Cu(II)-Mg(II) containing pyridine-2,5-dicarboxylate frameworks: Synthesis, structural diversity, inter-conversion and heterogeneous catalytic epoxidation. Polyhedron 2018, 146, 93–98. [Google Scholar] [CrossRef]
  41. Dhakshinamoorthy, A.; Portillo, A.S.; Asiri, A.M.; Garcia, H. Engineering UiO-66 Metal Organic Framework for Heterogeneous Catalysis. ChemCatChem 2019, 11, 899–923. [Google Scholar] [CrossRef]
  42. Vandichel, M.; Hajek, J.; Vermoortele, F.; Waroquier, M.; De Vos, D.E.; Van Speybroeck, V. Active site engineering in UiO-66 type metal–organic frameworks by intentional creation of defects: A theoretical rationalization. CrystEngComm 2014, 17, 395–406. [Google Scholar] [CrossRef] [Green Version]
  43. Arrozi, U.S.; Wijaya, H.W.; Patah, A.; Permana, Y. Efficient acetalization of benzaldehydes using UiO-66 and UiO-67: Substrates accessibility or Lewis acidity of zirconium. Appl. Catal. A Gen. 2015, 506, 77–84. [Google Scholar] [CrossRef]
  44. Zheng, H.-Q.; Zeng, Y.-N.; Chen, J.; Lin, R.-G.; Zhuang, W.-E.; Cao, R.; Lin, Z.-J. Zr-Based Metal–Organic Frameworks with Intrinsic Peroxidase-Like Activity for Ultradeep Oxidative Desulfurization: Mechanism of H2O2 Decomposition. Inorg. Chem. 2019, 58, 6983–6992. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, Y.; Klet, R.C.; Hupp, J.T.; Farha, O. Probing the correlations between the defects in metal–organic frameworks and their catalytic activity by an epoxide ring-opening reaction. Chem. Commun. 2016, 52, 7806–7809. [Google Scholar] [CrossRef] [PubMed]
  46. Zalomaeva, O.V.; Evtushok, V.Y.; Ivanchikova, I.D.; Glazneva, T.S.; Chesalov, Y.A.; Larionov, K.P.; Skobelev, I.Y.; Kholdeeva, O.A. Nucleophilic versus Electrophilic Activation of Hydrogen Peroxide over Zr-Based Metal–Organic Frameworks. Inorg. Chem. 2020, 59, 10634–10649. [Google Scholar] [CrossRef] [PubMed]
  47. Payne, G.B.; Deming, P.H.; Williams, P.H. Reactions of Hydrogen Peroxide. VII. Alkali-Catalyzed Epoxidation and Oxidation Using a Nitrile as Co-reactant. J. Org. Chem. 1961, 26, 659–663. [Google Scholar] [CrossRef]
  48. Huang, K.; Yu, S.; Li, X.; Cai, Z. One-pot synthesis of bimetal MOFs as highly efficient catalysts for selective oxidation of styrene. J. Chem. Sci. 2020, 132, 139. [Google Scholar] [CrossRef]
  49. Wang, F.; Meng, X.-G.; Wu, Y.-Y.; Huang, H.; Lv, J.; Yu, W.-W. A Highly Efficient Heterogeneous Catalyst of Bimetal-Organic Frameworks for the Epoxidation of Olefin with H2O2. Molecules 2020, 25, 2389. [Google Scholar] [CrossRef]
  50. Santibáñez, L.; Escalona, N.; Torres, J.; Kremer, C.; Cancino, P.; Spodine, E. CuII- and CoII-based MOFs: {[La2Cu3(µ-H2O)(ODA)6(H2O)3]·3H2O}n and {[La2Co3(ODA)6(H2O)6]·12H2O}n. The Relevance of Physicochemical Properties on the Catalytic Aerobic Oxidation of Cyclohexene. Catalysts 2020, 10, 589. [Google Scholar] [CrossRef]
  51. Tuci, G.; Giambastiani, G.; Kwon, S.; Stair, P.C.; Snurr, R.Q.; Rossin, A. Chiral Co(II) Metal–Organic Framework in the Heterogeneous Catalytic Oxidation of Alkenes under Aerobic and Anaerobic Conditions. ACS Catal. 2014, 4, 1032–1039. [Google Scholar] [CrossRef]
  52. Otake, K.-I.; Ahn, S.; Knapp, J.; Hupp, J.T.; Notestein, J.M.; Farha, O.K. Vapor-Phase Cyclohexene Epoxidation by Single-Ion Fe(III) Sites in Metal–Organic Frameworks. Inorg. Chem. 2021, 60, 2457–2463. [Google Scholar] [CrossRef] [PubMed]
  53. Ahn, S.; Thornburg, N.; Li, Z.; Wang, T.C.; Gallington, L.; Chapman, K.W.; Notestein, J.M.; Hupp, J.T.; Farha, O.K. Stable Metal–Organic Framework-Supported Niobium Catalysts. Inorg. Chem. 2016, 55, 11954–11961. [Google Scholar] [CrossRef]
  54. Thornburg, N.E.; Nauert, S.L.; Thompson, A.B.; Notestein, J.M. Synthesis−Structure–Function Relationships of Silica-Supported Niobium(V) Catalysts for Alkene Epoxidation with H2O2. ACS Catal. 2016, 6, 6124–6134. [Google Scholar] [CrossRef]
  55. Wang, S.; Liu, Y.; Zhang, Z.; Li, X.; Tian, H.; Yan, T.; Zhang, X.; Liu, S.; Sun, X.; Xu, L.; et al. One-Step Template-Free Fabrication of Ultrathin Mixed-Valence Polyoxovanadate-Incorporated Metal–Organic Framework Nanosheets for Highly Efficient Selective Oxidation Catalysis in Air. ACS Appl. Mater. Interfaces 2019, 11, 12786–12796. [Google Scholar] [CrossRef] [PubMed]
  56. Hui, J.; Chu, H.; Zhang, W.; Shen, Y.; Chen, W.; Hu, Y.; Liu, W.; Gao, C.; Guo, S.; Xiao, G.; et al. Multicomponent metal–organic framework derivatives for optimizing the selective catalytic performance of styrene epoxidation reaction. Nanoscale 2018, 10, 8772–8778. [Google Scholar] [CrossRef] [PubMed]
  57. Saedi, Z.; Safarifard, V.; Morsali, A. Dative and covalent-dative postsynthetic modification of a two-fold interpenetration pillared-layer MOF for heterogeneous catalysis: A comparison of catalytic activities and reusability. Microporous Mesoporous Mater. 2016, 229, 51–58. [Google Scholar] [CrossRef]
  58. Berijani, K.; Morsali, A.; Hupp, J.T. An effective strategy for creating asymmetric MOFs for chirality induction: A chiral Zr-based MOF for enantioselective epoxidation. Catal. Sci. Technol. 2019, 9, 3388–3397. [Google Scholar] [CrossRef]
  59. Noh, H.; Cui, Y.; Peters, A.W.; Pahls, D.; Ortuño, M.A.; Vermeulen, N.A.; Cramer, C.J.; Gagliardi, L.; Hupp, J.T.; Farha, O.K. An Exceptionally Stable Metal–Organic Framework Supported Molybdenum(VI) Oxide Catalyst for Cyclohexene Epoxidation. J. Am. Chem. Soc. 2016, 138, 14720–14726. [Google Scholar] [CrossRef]
  60. Kardanpour, R.; Tangestaninejad, S.; Mirkhani, V.; Moghadam, M.; Mohammadpoor-Baltork, I.; Zadehahmadi, F. Efficient alkene epoxidation catalyzed by molybdenyl acetylacetonate supported on aminated UiO-66 metal−organic framework. J. Solid State Chem. 2015, 226, 262–272. [Google Scholar] [CrossRef]
  61. Tang, J.; Dong, W.; Wang, G.; Yao, Y.; Cai, L.; Liu, Y.; Zhao, X.; Xu, J.; Tan, L. Efficient molybdenum(vi) modified Zr-MOF catalysts for epoxidation of olefins. RSC Adv. 2014, 4, 42977–42982. [Google Scholar] [CrossRef]
  62. Liu, T.-F.; Vermeulen, N.A.; Howarth, A.J.; Li, P.; Sarjeant, A.A.; Hupp, J.T.; Farha, O.K. Adding to the Arsenal of Zirconium-Based Metal–Organic Frameworks: The Topology as a Platform for Solvent-Assisted Metal Incorporation. Eur. J. Inorg. Chem. 2016, 27, 4349. [Google Scholar] [CrossRef]
  63. Pereira, C.; Biernacki, K.; Rebelo, S.L.; Magalhães, A.L.; Carvalho, A.P.; Pires, J.; Freire, C. Designing heterogeneous oxovanadium and copper acetylacetonate catalysts: Effect of covalent immobilisation in epoxidation and aziridination reactions. J. Mol. Catal. A Chem. 2009, 312, 53–64. [Google Scholar] [CrossRef]
  64. Li, Z.; Wu, S.; Ding, H.; Lu, H.; Liu, J.; Huo, Q.; Guan, J.; Kan, Q. Oxovanadium(iv) and iron(iii) salen complexes immobilized on amino-functionalized graphene oxide for the aerobic epoxidation of styrene. New J. Chem. 2013, 37, 4220–4229. [Google Scholar] [CrossRef]
  65. Ben Zid, T.; Khedher, I.; Ghorbel, A. Chiral vanadyl salen catalyst immobilized on mesoporous silica as support for asymmetric oxidation of sulfides to sulfoxides. React. Kinet. Mech. Catal. 2010, 100, 131–143. [Google Scholar] [CrossRef] [Green Version]
  66. Zamanifar, E.; Farzaneh, F. Immobilized vanadium amino acid Schiff base complex on Al-MCM-41 as catalyst for the epoxidation of allyl alcohols. React. Kinet. Mech. Catal. 2011, 104, 197–209. [Google Scholar] [CrossRef]
  67. Pourkhosravani, M.; Dehghanpour, S.; Farzaneh, F.; Sohrabi, S. Designing new catalytic nanoreactors for the regioselective epoxidation of geraniol by the post-synthetic immobilization of oxovanadium(IV) complexes on a ZrIV-based metal–organic framework. React. Kinet. Mech. Catal. 2017, 122, 961–981. [Google Scholar] [CrossRef]
  68. Tan, C.; Han, X.; Li, Z.; Liu, Y.; Cui, Y. Controlled Exchange of Achiral Linkers with Chiral Linkers in Zr-Based UiO-68 Metal–Organic Framework. J. Am. Chem. Soc. 2018, 140, 16229–16236. [Google Scholar] [CrossRef]
  69. Huang, K.; Guo, L.L.; Wu, D.F. Synthesis of Metal Salen@MOFs and Their Catalytic Performance for Styrene Oxidation. Ind. Eng. Chem. Res. 2019, 58, 4744–4754. [Google Scholar] [CrossRef]
  70. Xia, Q.; Li, Z.; Tan, C.; Liu, Y.; Gong, W.; Cui, Y. Multivariate Metal–Organic Frameworks as Multifunctional Heterogeneous Asymmetric Catalysts for Sequential Reactions. J. Am. Chem. Soc. 2017, 139, 8259–8266. [Google Scholar] [CrossRef]
  71. Berijani, K.; Morsali, A. Construction of an Asymmetric Porphyrinic Zirconium Metal–Organic Framework through Ionic Postchiral Modification. Inorg. Chem. 2021, 60, 206–218. [Google Scholar] [CrossRef] [PubMed]
  72. Martínez, H.; Cáceres, M.F.; Martínez, F.; Páez-Mozo, E.A.; Valange, S.; Castellanos, N.J.; Molina, D.; Barrault, J.; Arzoumanian, H. Photo-epoxidation of cyclohexene, cyclooctene and 1-octene with molecular oxygen catalyzed by dichloro dioxo-(4,4′-dicarboxylato-2,2′-bipyridine) molybdenum(VI) grafted on mesoporous TiO2. J. Mol. Catal. A Chem. 2016, 423, 248–255. [Google Scholar] [CrossRef]
  73. Portillo, A.S.; Navalón, S.; Cirujano, F.G.; I Xamena, F.X.L.; Alvaro, M.; Garcia, H. MIL-101 as Reusable Solid Catalyst for Autoxidation of Benzylic Hydrocarbons in the Absence of Additional Oxidizing Reagents. ACS Catal. 2015, 5, 3216–3224. [Google Scholar] [CrossRef]
  74. Kou, J.; Sun, L.-B. Fabrication of Metal–Organic Frameworks inside Silica Nanopores with Significantly Enhanced Hydrostability and Catalytic Activity. ACS Appl. Mater. Interfaces 2018, 10, 12051–12059. [Google Scholar] [CrossRef] [PubMed]
  75. Zhou, Z.; Li, X.; Wang, Y.; Luan, Y.; Li, X.; Du, X. Growth of Cu-BTC MOFs on dendrimer-like porous silica nanospheres for the catalytic aerobic epoxidation of olefins. New J. Chem. 2020, 44, 14350–14357. [Google Scholar] [CrossRef]
  76. Zhao, J.; Wang, W.; Tang, H.; Ramella, D.; Luan, Y. Modification of Cu2+ into Zr-based metal–organic framework (MOF) with carboxylic units as an efficient heterogeneous catalyst for aerobic epoxidation of olefins. Mol. Catal. 2018, 456, 57–64. [Google Scholar] [CrossRef]
  77. Wu, Y.; Wang, H.; Guo, S.; Zeng, Y.; Ding, M. MOFs-induced high-amphiphilicity in hierarchical 3D reduced graphene oxide-based hydrogel. Appl. Surf. Sci. 2021, 540, 148303. [Google Scholar] [CrossRef]
  78. Canioni, R.; Roch-Marchal, C.; Sécheresse, F.; Horcajada, P.; Serre, C.; Hardi-Dan, M.; Férey, G.; Grenèche, J.-M.; Lefebvre, F.; Chang, J.-S.; et al. Stable polyoxometalate insertion within the mesoporous metal organic framework MIL-100(Fe). J. Mater. Chem. 2011, 21, 1226–1233. [Google Scholar] [CrossRef]
  79. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef]
  80. Song, X.; Hu, D.; Yang, X.; Zhang, H.; Zhang, W.; Li, J.; Jia, M.; Yu, J. Polyoxomolybdic Cobalt Encapsulated within Zr-Based Metal–Organic Frameworks as Efficient Heterogeneous Catalysts for Olefins Epoxidation. ACS Sustain. Chem. Eng. 2019, 7, 3624–3631. [Google Scholar] [CrossRef]
  81. Villabrille, P.; Romanelli, G.; Gassa, L.; Vazquez, P.; Caceres, C. Synthesis and characterization of Fe- and Cu-doped molybdovanadophosphoric acids and their application in catalytic oxidation. Appl. Catal. A Gen. 2007, 324, 69–76. [Google Scholar] [CrossRef]
  82. Jin, M.; Niu, Q.; Liu, G.; Lv, Z.; Si, C.; Guo, H. Encapsulation of ionic liquids into POMs-based metal–organic frameworks: Screening of POMs-ILs@MOF catalysts for efficient cycloolefins epoxidation. J. Mater. Sci. 2020, 55, 8199–8210. [Google Scholar] [CrossRef]
  83. Ke, F.; Guo, F.; Yu, J.; Yang, Y.; He, Y.; Chang, L.; Wan, X. Highly Site-Selective Epoxidation of Polyene Catalyzed by Metal–Organic Frameworks Assisted by Polyoxometalate. J. Inorg. Organomet. Polym. Mater. 2017, 27, 843–849. [Google Scholar] [CrossRef]
  84. Li, N.; Mu, B.; Lv, L.; Huang, R. Assembly of new polyoxometalate–templated metal–organic frameworks based on flexible ligands. J. Solid State Chem. 2015, 226, 88–93. [Google Scholar] [CrossRef]
  85. Zhang, Y.; Li, Y.-X.; Liu, L.; Han, Z.-B. Palladium nanoparticles supported on UiO-66-NH2 as heterogeneous catalyst for epoxidation of styrene. Inorg. Chem. Commun. 2019, 100, 51–55. [Google Scholar] [CrossRef]
  86. Limvorapitux, R.; Chou, L.-Y.; Young, A.P.; Tsung, C.-K.; Nguyen, S.T. Coupling Molecular and Nanoparticle Catalysts on Single Metal–Organic Framework Microcrystals for the Tandem Reaction of H2O2 Generation and Selective Alkene Oxidation. ACS Catal. 2017, 7, 6691–6698. [Google Scholar] [CrossRef]
  87. Jin, T.; Yang, Q.; Meng, C.; Xu, J.; Liu, H.; Hu, J.; Ling, H. Promoting desulfurization capacity and separation efficiency simultaneously by the novel magnetic Fe3O4@PAA@MOF-199. RSC Adv. 2014, 4, 41902–41909. [Google Scholar] [CrossRef]
  88. Li, J.; Gao, H.; Tan, L.; Luan, Y.; Yang, M. Superparamagnetic Core-Shell Metal-Organic Framework Fe3O4/Cu3(btc)2Microspheres and Their Catalytic Activity in the Aerobic Oxidation of Alcohols and Olefins. Eur. J. Inorg. Chem. 2016, 2016, 4906–4912. [Google Scholar] [CrossRef]
  89. Hou, J.; Luan, Y.; Yu, J.; Qi, Y.; Wang, G.; Lu, Y. Fabrication of hierarchical composite microspheres of copper-doped Fe3O4@P4VP@ZIF-8 and their application in aerobic oxidation. New J. Chem. 2016, 40, 10127–10135. [Google Scholar] [CrossRef]
  90. Qi, Y.; Luan, Y.; Yu, J.; Peng, X.; Wang, G. Nanoscaled Copper Metal-Organic Framework (MOF) Based on Carboxylate Ligands as an Efficient Heterogeneous Catalyst for Aerobic Epoxidation of Olefins and Oxidation of Benzylic and Allylic Alcohols. Chem. A Eur. J. 2015, 21, 1589–1597. [Google Scholar] [CrossRef]
  91. Yang, X.; Yuan, S.; Zou, L.; Drake, H.; Zhang, Y.; Qin, J.; Alsalme, A.; Zhou, H.C. One-Step Synthesis of Hybrid Core–Shell Metal–Organic Frameworks. Angew. Chem. 2018, 130, 3991–3996. [Google Scholar] [CrossRef]
  92. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the Valorization of Exhaust Carbon: From CO2 to Chemicals, Materials, and Fuels. Technological Use of CO2. Chem. Rev. 2013, 114, 1709–1742. [Google Scholar] [CrossRef]
  93. North, M.; Pasquale, R. Mechanism of Cyclic Carbonate Synthesis from Epoxides and CO2. Angew. Chem. Int. Ed. 2009, 48, 2946–2948. [Google Scholar] [CrossRef] [Green Version]
  94. Shi, L.; Xu, S.; Zhang, Q.; Liu, T.; Wei, B.; Zhao, Y.; Meng, L.; Li, J. Ionic Liquid/Quaternary Ammonium Salt Integrated Heterogeneous Catalytic System for the Efficient Coupling of Carbon Dioxide with Epoxides. Ind. Eng. Chem. Res. 2018, 57, 15319–15328. [Google Scholar] [CrossRef]
  95. Shi, X.-L.; Chen, Y.; Duan, P.; Zhang, W.; Hu, Q. Conversion of CO2 into Organic Carbonates over a Fiber-Supported Ionic Liquid Catalyst in Impellers of the Agitation System. ACS Sustain. Chem. Eng. 2018, 6, 7119–7127. [Google Scholar] [CrossRef]
  96. Ziaee, M.A.; Tang, Y.; Zhong, H.; Tian, D.; Wang, R. Urea-Functionalized Imidazolium-Based Ionic Polymer for Chemical Conversion of CO2 into Organic Carbonates. ACS Sustain. Chem. Eng. 2018, 7, 2380–2387. [Google Scholar] [CrossRef]
  97. Yang, Z.; Yu, B.; Zhang, H.; Zhao, Y.; Chen, Y.; Ma, Z.; Ji, G.; Gao, X.; Han, B.; Liu, Z. Metalated Mesoporous Poly(triphenylphosphine) with Azo Functionality: Efficient Catalysts for CO2 Conversion. ACS Catal. 2016, 6, 1268–1273. [Google Scholar] [CrossRef]
  98. Sun, Q.; Aguila, B.; Perman, J.A.; Nguyen, N.T.-K.; Ma, S. Flexibility Matters: Cooperative Active Sites in Covalent Organic Framework and Threaded Ionic Polymer. J. Am. Chem. Soc. 2016, 138, 15790–15796. [Google Scholar] [CrossRef]
  99. Beyzavi, M.H.; Stephenson, C.J.; Liu, Y.; Karagiaridi, O.; Hupp, J.T.; Farha, O.K. Metal-organic framework-based catalysts: Chemical fixation of CO2 with epoxides leading to cyclic organic carbonates. Front. Energy Res. 2015, 3, 63. [Google Scholar] [CrossRef] [Green Version]
  100. Bon, V.; Senkovskyy, V.; Senkovska, I.; Kaskel, S. Zr(IV) and Hf(IV) based metal–organic frameworks with reo-topology. Chem. Commun. 2012, 48, 8407–8409. [Google Scholar] [CrossRef] [Green Version]
  101. Beyzavi, M.H.; Klet, R.C.; Tussupbayev, S.; Borycz, J.; Vermeulen, N.A.; Cramer, C.J.; Stoddart, J.F.; Hupp, J.T.; Farha, O.K. A Hafnium-Based Metal–Organic Framework as an Efficient and Multifunctional Catalyst for Facile CO2 Fixation and Regioselective and Enantioretentive Epoxide Activation. J. Am. Chem. Soc. 2014, 136, 15861–15864. [Google Scholar] [CrossRef]
  102. Tran, Y.B.N.; Nguyen, P.T.K.; Luong, Q.T.; Nguyen, K.D. Series of M-MOF-184 (M = Mg, Co, Ni, Zn, Cu, Fe) Metal-Organic Frameworks for Catalysis Cycloaddition of CO2. Inorg. Chem. 2020, 59, 16747–16759. [Google Scholar] [CrossRef]
  103. Agarwal, R.A.; Gupta, A.K.; De, D. Flexible Zn-MOF Exhibiting Selective CO2 Adsorption and Efficient Lewis Acidic Catalytic Activity. Cryst. Growth Des. 2019, 19, 2010–2018. [Google Scholar] [CrossRef]
  104. Hayashi, H.; Côté, A.P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O.M. Zeolite A imidazolate frameworks. Nat. Mater. 2007, 6, 501–507. [Google Scholar] [CrossRef] [PubMed]
  105. Huang, A.; Chen, Y.; Wang, N.; Hu, Z.; Jiang, J.; Caro, J. A highly permeable and selective zeolitic imidazolate framework ZIF-95 membrane for H2/CO2 separation. Chem. Commun. 2012, 48, 10981–10983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Bhin, K.M.; Tharun, J.; Roshan, K.R.; Kim, D.-W.; Chung, Y.; Park, D.-W. Catalytic performance of zeolitic imidazolate framework ZIF-95 for the solventless synthesis of cyclic carbonates from CO2 and epoxides. J. CO2 Util. 2017, 17, 112–118. [Google Scholar] [CrossRef]
  107. Liang, J.; Chen, R.-P.; Wang, X.-Y.; Liu, T.-T.; Wang, X.-S.; Huang, Y.-B.; Cao, R. Postsynthetic ionization of an imidazole-containing metal–organic framework for the cycloaddition of carbon dioxide and epoxides. Chem. Sci. 2017, 8, 1570–1575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Ding, L.-G.; Yao, B.-J.; Jiang, W.-L.; Li, J.-T.; Fu, Q.-J.; Li, Y.-A.; Liu, Z.-H.; Ma, J.-P.; Dong, Y.-B. Bifunctional Imidazolium-Based Ionic Liquid Decorated UiO-67 Type MOF for Selective CO2 Adsorption and Catalytic Property for CO2 Cycloaddition with Epoxides. Inorg. Chem. 2017, 56, 2337–2344. [Google Scholar] [CrossRef]
  109. Kurisingal, J.F.; Rachuri, Y.; Pillai, R.S.; Gu, Y.; Choe, Y.; Park, D. Ionic-Liquid-Functionalized UiO-66 Framework: An Experimental and Theoretical Study on the Cycloaddition of CO2 and Epoxides. ChemSusChem 2019, 12, 1033–1042. [Google Scholar] [CrossRef]
  110. Ding, M.; Jiang, H.-L. Incorporation of Imidazolium-Based Poly(ionic liquid)s into a Metal–Organic Framework for CO2 Capture and Conversion. ACS Catal. 2018, 8, 3194–3201. [Google Scholar] [CrossRef]
  111. Pander, M.; Janeta, M.; Bury, W. Quest for an Efficient 2-in-1 MOF-Based Catalytic System for Cycloaddition of CO2 to Epoxides under Mild Conditions. ACS Appl. Mater. Interfaces 2021, 13, 8344–8352. [Google Scholar] [CrossRef] [PubMed]
  112. Zhu, J.; Usov, P.M.; Xu, W.; Celis-Salazar, P.J.; Lin, S.; Kessinger, M.C.; Landaverde-Alvarado, C.; Cai, M.; May, A.M.; Slebodnick, C.; et al. A New Class of Metal-Cyclam-Based Zirconium Metal–Organic Frameworks for CO2 Adsorption and Chemical Fixation. J. Am. Chem. Soc. 2018, 140, 993–1003. [Google Scholar] [CrossRef] [PubMed]
  113. Vitillo, J.G.; Savonnet, M.; Ricchiardi, G.; Bordiga, S. Tailoring Metal-Organic Frameworks for CO2 Capture: The Amino Effect. ChemSusChem 2011, 4, 1281–1290. [Google Scholar] [CrossRef] [PubMed]
  114. Torrisi, A.; Bell, R.G.; Mellot-Draznieks, C. Functionalized MOFs for Enhanced CO2 Capture. Cryst. Growth Des. 2010, 10, 2839–2841. [Google Scholar] [CrossRef]
  115. Senthilkumar, S.; Maru, M.S.; Somani, R.S.; Bajaj, H.C.; Neogi, S. Unprecedented NH2-MIL-101(Al)/n-Bu4NBr system as solvent-free heterogeneous catalyst for efficient synthesis of cyclic carbonates via CO2 cycloaddition. Dalton Trans. 2018, 47, 418–428. [Google Scholar] [CrossRef]
  116. Zhu, M.; Carreon, M.A. Porous crystals as active catalysts for the synthesis of cyclic carbonates. J. Appl. Polym. Sci. 2014, 131. [Google Scholar] [CrossRef]
  117. Cai, T.; Liu, J.; Cao, H.; Cui, C. Synthesis of bio-based cyclic carbonate from vegetable oil methyl ester by CO2 fixation with acid-base pair MOFs. Ind. Crop. Prod. 2020, 145, 112155. [Google Scholar] [CrossRef]
  118. Mercuri, G.; Moroni, M.; Domasevitch, K.V.; Di Nicola, C.; Campitelli, P.; Pettinari, C.; Giambastiani, G.; Galli, S.; Rossin, A. Carbon Dioxide Capture and Utilization with Isomeric Forms of Bis(amino)-Tagged Zinc Bipyrazolate Metal–Organic Frameworks. Chem. A Eur. J. 2021, 27, 4746–4754. [Google Scholar] [CrossRef]
  119. Norouzi, F.; Khavasi, H.R. Diversity-Oriented Metal Decoration on UiO-Type Metal-Organic Frameworks: An Efficient Approach to Increase CO2 Uptake and Catalytic Conversion to Cyclic Carbonates. ACS Omega 2019, 4, 19037–19045. [Google Scholar] [CrossRef]
  120. Li, J.; Ren, Y.; Yue, C.; Fan, Y.; Qi, C.; Jiang, H. Highly Stable Chiral Zirconium–Metallosalen Frameworks for CO2 Conversion and Asymmetric C–H Azidation. ACS Appl. Mater. Interfaces 2018, 10, 36047–36057. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of MOFs frameworks with different dimensionalities (3D, 2D, 1D).
Figure 1. Schematic representation of MOFs frameworks with different dimensionalities (3D, 2D, 1D).
Inorganics 09 00081 g001
Figure 2. General mechanism of epoxidation of alkenes with peroxycarboxylic acid as co-catalyst.
Figure 2. General mechanism of epoxidation of alkenes with peroxycarboxylic acid as co-catalyst.
Inorganics 09 00081 g002
Figure 3. Representation of cycloaddition reaction of CO2, captured by MOFs, to epoxides.
Figure 3. Representation of cycloaddition reaction of CO2, captured by MOFs, to epoxides.
Inorganics 09 00081 g003
Figure 4. Schematic representation of the formation of tert-butylperoxyl (tBuOO·) and tert-butoxyl radicals (tBuO) catalyzed by the Cu(II) sites of MOF.
Figure 4. Schematic representation of the formation of tert-butylperoxyl (tBuOO·) and tert-butoxyl radicals (tBuO) catalyzed by the Cu(II) sites of MOF.
Inorganics 09 00081 g004
Figure 5. (a) Crystal structures of NU-1000-Fe-NO3 (b) Structures of the inorganic Zr6-nodes.
Figure 5. (a) Crystal structures of NU-1000-Fe-NO3 (b) Structures of the inorganic Zr6-nodes.
Inorganics 09 00081 g005
Figure 6. (a) The preparation of [C-NU-1000-Mo] catalyst with solvent-assisted ligand incorporation (SALI); (b) simplification of the chirality induction mechanism of [C-NU-1000-Mo] when olefins approach by pro-S- or R-face to the catalytic active center.
Figure 6. (a) The preparation of [C-NU-1000-Mo] catalyst with solvent-assisted ligand incorporation (SALI); (b) simplification of the chirality induction mechanism of [C-NU-1000-Mo] when olefins approach by pro-S- or R-face to the catalytic active center.
Inorganics 09 00081 g006
Figure 7. (a) The 2D structure of [Cu6(bip)12(PMoVI12O40)2(PMoVMoVI11O40O2)]·8H2O; (b) the coordination environment of the Cu(II) cations. Hydrogens and hydroxyls are omitted for clarity. Light-blue polyhedral correspond to the (PMo12) polyanion.
Figure 7. (a) The 2D structure of [Cu6(bip)12(PMoVI12O40)2(PMoVMoVI11O40O2)]·8H2O; (b) the coordination environment of the Cu(II) cations. Hydrogens and hydroxyls are omitted for clarity. Light-blue polyhedral correspond to the (PMo12) polyanion.
Inorganics 09 00081 g007
Figure 8. Proposed mechanism for the cycloaddition of CO2 to epoxide with Lewis acid and base catalysts.
Figure 8. Proposed mechanism for the cycloaddition of CO2 to epoxide with Lewis acid and base catalysts.
Inorganics 09 00081 g008
Figure 9. Proposed mechanism for the cycloaddition of styrene oxide and CO2 using tetrabutylammonium bromide.
Figure 9. Proposed mechanism for the cycloaddition of styrene oxide and CO2 using tetrabutylammonium bromide.
Inorganics 09 00081 g009
Table 3. MOFs with functionalized organic linker in olefin epoxidation.
Table 3. MOFs with functionalized organic linker in olefin epoxidation.
MOFSubstrateReaction Data
T (°C) P (atm) Time (h)
Oxidant/Cocatalyst/
Solvent a
Conversion
%
Epoxide
Selectivity%
Ref.
UiO-66-SI/VO(acac)Geraniol4011TBHP/-/CH2Cl2100100[67]
UiO-66-N/VO(acac)2Geraniol4012TBHP/-/CH2Cl2100100[67]
UiO-66-sal-MoDcis-Cyclooctene80241TBHP/-/CH3CN9999[61]
PCN-224-Mn(tart)Styrene6014O2/IBA/CH3CN10089[71]
trans-Stilbene6014O2/IBA/CH3CN80100[71]
1-Phenyl-1-cyclohexene6014O2/IBA/CH3CN75100[71]
1-Octene6014O2/IBA/CH3CN70100[71]
UiO-67-Mo(CO)3Cyclooctene5531TBHP/-/toluene10099[33]
UiO-66-Mo(CO)3Cyclooctene5531TBHP/-/toluene9299[33]
[C-NU-1000-Mo]Styrene1200.035H2O2/-/CH2CH2Cl210086[58]
1-Octene 1200.038H2O2/-/CH2CH2Cl272100[58]
UiO-66-NH2-SA-Mo Cyclooctene830.751TBHP/-/CH2CH2Cl297100[60]
Cyclohexene831.51TBHP/-/CH2CH2Cl293100[60]
Styrene8351TBHP/-/CH2CH2Cl28792[60]
1-Octene8381TBHP/-/CH2CH2Cl278100[60]
1-Decene 83101TBHP/-/CH2CH2Cl279100[60]
UiO-66-NH2-TC-MoCyclooctene8311TBHP/-/CH2CH2Cl294100[60]
Cyclohexene8321TBHP/-/CH2CH2Cl290100[60]
Styrene8351TBHP/-/CH2CH2Cl28690[60]
1-Octene8381TBHP/-/CH2CH2Cl275100[60]
1-Decene 8310.51TBHP/-/CH2CH2Cl275100[60]
[Zn4O(L5Cu,Fe)3]2,2-Dimethyl-2H-chromene−20136MesPhIO/-/CHCl39487 ee[70]
[Zn4O(L5Cu,Mn,Co)3]3-Chloropropene0110sPhIO/-/CHCl392-[70]
Styrene0124sPhIO/-/CH2Cl263-[70]
UiO-66-PC-MoDcis-Cyclooctene80241TBHP/-/CH3CN90.799[61]
Mo-SIMCyclohexene6017TBHP/-/toluene9399[59]
(R)-UiO-68-Mn 2,2-Dimethyl-2H-chromene0110sPhIO/-/CH2Cl29188 ee[68]
UiO-66-PI-MoDcis-Cyclooctene80241TBHP/-/CH3CN77.599[61]
(R)-UiO-68-Fe2,2-Dimethyl-2H-chromene−20136MesPhIO/-/CHCl38486 ee[68]
Cusalen@NH2-MIL-101(Cr)Styrene 8016TBHP/-/CH3CN 98.7889.58[69]
[Zn4O(L5Cu,Mn)3]2,2-Dimethyl-2H-chromene−20136MesPhIO/-/CH2Cl28686 ee[70]
PCN-224-Mn(tart)Styrene60 °C14O2/IBA/CH3CN10089[71]
trans-Stilbene60 °C14O2/IBA/CH3CN80100[71]
1-Phenyl-1-cyclohexene60 °C14O2/IBA/CH3CN75100[71]
1-Octene60 °C14O2/IBA/CH3CN70100[71]
TMU-16-NH2Cyclohexene60140TBHP/-/CHCl36674[57]
Styrene60151TBHP/-/CHCl38898[57]
Cyclooctene60124TBHP/-/CHCl38383[57]
a sPhIO = 2-(tertbutylsulfonyl)iodosylbenzene; IBA = isobutyraldehyde; CHP = cumene hydroperoxide; TBHP = tert-butylhydroperoxide.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tombesi, A.; Pettinari, C. Metal Organic Frameworks as Heterogeneous Catalysts in Olefin Epoxidation and Carbon Dioxide Cycloaddition. Inorganics 2021, 9, 81. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9110081

AMA Style

Tombesi A, Pettinari C. Metal Organic Frameworks as Heterogeneous Catalysts in Olefin Epoxidation and Carbon Dioxide Cycloaddition. Inorganics. 2021; 9(11):81. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9110081

Chicago/Turabian Style

Tombesi, Alessia, and Claudio Pettinari. 2021. "Metal Organic Frameworks as Heterogeneous Catalysts in Olefin Epoxidation and Carbon Dioxide Cycloaddition" Inorganics 9, no. 11: 81. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9110081

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