Next Article in Journal
Similarities and Differences between Site-Selective Acylation and Phosphorylation of Amphiphilic Diols, Promoted by Nucleophilic Organocatalysts Decorated with Outer-Sphere Appendages
Next Article in Special Issue
Depolymerization of Polyesters by Transesterification with Ethanol Using (Cyclopentadienyl)titanium Trichlorides
Previous Article in Journal
Innovative Electrocatalysts for Fuel Cell and Battery Applications
Previous Article in Special Issue
Phenyl Formate as a CO Surrogate for the Reductive Cyclization of Organic Nitro Compounds to Yield Different N-Heterocycles: No Need for Autoclaves and Pressurized Carbon Monoxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

H2O2-Based Selective Oxidations Catalyzed by Supported Polyoxometalates: Recent Advances

by
Nataliya V. Maksimchuk
and
Oxana A. Kholdeeva
*
Boreskov Institute of Catalysis, Lavrentieva Ave. 5, Novosibirsk 630090, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(2), 360; https://doi.org/10.3390/catal13020360
Submission received: 16 December 2022 / Revised: 27 January 2023 / Accepted: 2 February 2023 / Published: 6 February 2023
(This article belongs to the Special Issue Exclusive Papers of the Editorial Board Members (EBMs) of Catalysts)

Abstract

:
Polyoxometalates (POMs) are transition metal oxygen anionic clusters that are oxidatively and thermally robust due to their inorganic, metal oxide-like nature. The versatility of their structures and compositions ensures tunable acid and redox properties, solubility, and functionality. The potential of POMs as homogeneous catalysts and building blocks for the construction of heterogeneous selective oxidations catalysts is being intensively investigated. POM catalysts immobilized on solid supports have the clear advantages of easy separation and reuse and, thus, better meet the requests of sustainable chemistry, provided that they are leaching-resistant under the reaction conditions. Here, we give a brief overview of recent advances in the field of liquid-phase selective oxidation of organic compounds using supported POMs and the green oxidant–hydrogen peroxide, with a focus on the critical issues of the catalyst stability and reusability. The scope and limitations of various approaches to POM immobilization are discussed.

1. Introduction

In the modern chemical industry, the role of liquid-phase selective oxidation in the production of bulk and fine chemicals is permanently increasing [1,2,3]. The development of cost-effective and environmentally friendly oxidation processes employing safe and readily available oxidants is a challenging goal, as requested by the pursuit of sustainable and green manufacturing of valuable chemicals. In this context, hydrogen peroxide, the molecule that can provide 47% of potentially active oxygen, leaving water as the only byproduct, remains one of the most ecologically and economically attractive oxidants [4,5,6]. Significant progress achieved on the way to the direct synthesis of H2O2 from molecular hydrogen and oxygen contributes to the expansion of interest in this oxidant [7,8,9,10,11,12].
The use of hydrogen peroxide in oxidation reactions requires catalysts because it is inert toward the majority of organic substrates. Heterogeneous catalysts have clear advantages of facile recovery, recycling, and compatibility with flow reactors and, thus, better meet the needs of sustainable chemistry, provided that the solid catalyst is stable to the leaching of active metal in the reaction medium [2,3].
A serious problem of the currently existing selective oxidation catalysts is their relatively low productivity caused by the damage of active sites under conditions of a large excess of H2O2. While catalysts bearing organic and organometallic ligands (both homogeneous and supported) suffer from oxidative degradation, most of metal-containing zeolites and other zeotype solids reveal hydro(solvo)lytic instability in the presence of aqueous H2O2, which results in the aggregation of active sites and may be accompanied by metal leaching into the solution. A rare exception among them is the well-known catalyst titanium–silicalite-1 (TS-1), developed by the ENI group in the 1980s [13,14]. The invention of TS-1 appeared to be a real breakthrough in the field of liquid-phase oxidation and greatly expanded interest in the use of aqueous hydrogen peroxide as oxidant. Nowadays, TS-1 is employed in large-scale industrial processes [15,16,17,18], but its scope is limited to small organic substrates capable of penetrating the micropores (0.53 × 0.56 nm) of the catalyst. Significant scientific efforts have been directed to the development of larger pore analogues of TS-1, in particular, mesoporous Ti-silicates, and several catalytic materials with a fairly good hydro(solvo)lytic stability have been elaborated [19,20,21,22,23]. However, unlike TS-1, these materials are endowed with a hydrophilic nature, and the preferential adsorption of H2O2/H2O, rather than the organic substrate, favors the unproductive H2O2 decomposition on the catalyst surface, which results in a large contribution of homolytic oxidation pathways and negative effects on the oxidation selectivity [22,23]. Some limited achievements in this area are related to the partial hydrophobization of the surface of titanium silicate catalysts [24,25,26] and the introduction of other metals, first of all, niobium, into silicate matrices [27,28,29,30]. Meanwhile, the search for alternative types of catalysts capable of heterolytic activation of aqueous hydrogen peroxide remains a challenging goal of oxidation catalysis.
Early transition metal oxygen anionic clusters, also called polyoxometalates (POMs; general formula [XxYyOz]q− (x ≤ y), where X is a main group element or transition metal and Y are addenda atoms, most often MoVI or WVI) have long been attracting the research attention as oxidation catalysts thanks to a unique combination of properties, which includes, first of all, the inorganic nature and metal oxide-like structure, ensuring thermodynamic stability to oxidation along with good thermal and hydro(solvo)lytic stability [31,32,33,34,35,36,37,38,39,40,41]. The obvious structural analogy between POMs and the surface of metal oxides makes it possible to consider POMs as discrete soluble molecular models of heterogeneous catalysts, which can be systematically investigated at the atomic level by experimental and computational techniques [34,42,43,44,45].
The solubility, redox, and acid–base properties of POMs and, as a result, their catalytic performance can be finely tuned by choosing the POM structure and chemical composition of both the polyanion and its counter cations. In particular, interesting objects for oxidation catalysis are M-substituted polyoxotungstates that comprise various heterometals (M) strongly bound to the molecular tungsten oxide through multiple M–O–W bonds, which potentially prevents the active site from hydrolysis and di(oligo)merization. In addition, a covalent functionalization of POMs can be used to control the surrounding of the active site, specifically its steric hindrance [46,47]. Some structural types of POMs that will be mentioned in this review, in relation to their use in oxidation catalysis with H2O2, are shown in Figure 1.
A comprehensive review literature, including several thematic issues and book chapters [32,33,34,35,38,48,49,50,51,52,53,54,55], covers different aspects of POM chemistry, in particular, their application in various fields of catalysis [31,32,33,35,36,37,38,39,40,41,49,53,54,55]. A book chapter of Kholdeeva and Hill covered the literature (until 2012) related to POM immobilization and use in selective oxidations [38]. Wu discussed the progress of using supramolecular interactions for the construction of encapsulated POM catalysts [54]. Enferadi-Kerenkan et al. reviewed catalysis by solidified polyoxotungstates [56], while Evtushok et al. surveyed the POM immobilization on carbon nanotubes and their use in selective oxidations [57]. Some modern trends in the environmentally benign oxidation catalysis by POMs and H2O2, specifically new approaches to biphasic catalysis with POMs and POM immobilization, were touched on in a book chapter of Kholdeeva [55]. However, the topic of selective oxidation using supported POMs has greatly expanded in recent years, and our aim here is to give a brief overview of the new advances in this field, with a focus on the use of hydrogen peroxide as oxidant. Given that the resistance to the leaching of active species under turnover conditions of liquid-phase oxidation is a critical issue for solid catalysts [58], we paid special attention to this problem and singled out those works for which the heterogeneous nature of catalysis was thoroughly validated.

2. Supported POMs as Heterogeneous Catalysts for Selective Oxidations with H2O2

To date, a range of different approaches to the immobilization of catalytically active POMs on solid supports have been proposed [38,59,60,61,62,63,64,65,66,67,68,69,70]. Conventional methodologies include the synthesis of insoluble POM salts (Cs+, Ag+, K+, NH4+, or some organic polycations) [61,68,69,71,72,73,74,75], embedding into silica by means of sol–gel synthesis [59,65,69,76,77,78,79,80], electrostatic attachment via anion exchange with layered double hydroxides [81,82,83,84], or modified surface of silica materials [65,68,69,76,80,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99], anchoring through the formation of dative bonds between POM and surface ligands [100,101,102,103], and finally, the covalent binding of organo-functionalized POMs [61,63,66,69,104,105]. In the last decade, some novel approaches that involve combinations of various types of interactions, including supramolecular ones, between the POM and surface, have been developed and successfully employed for the preparation of stable POM-based oxidation catalysts.

2.1. Embedding via Template Synthesis

One of the efficient and convenient approaches to POM heterogenization is encapsulation in an inert matrix during the matrix synthesis. Izumi was the first who suggested inserting a catalytically active POM into silica by means of the sol–gel method [59]. The discovery of ordered mesoporous materials stimulated the further development and modifications of this technique, using various templates as structure-directing agents [38] and references cited therein. However, while microporous POM/SiO2 composites could be relatively stable [65,76,77], mesoporous composite materials often showed insufficient stability to POM leaching in polar media.
The evaporation-induced self-assembly (EISA) procedure was employed for a one-pot preparation of a series of ordered meso-macroporous titania–silica–polyoxometalate materials using phosphotungstic acid as a POM and non-ionic surfactant (P123) and monodisperse polystyrene microspheres as co-structure-directing agents [106,107]. Catalysts PW12/SiO2–TiO2 exhibited a good performance in oxidative desulfurization (ODS) reactions: dibenzothiophene (DBT, 500 ppm S as model fuel), which is considered to be the main sulfide pollutant in fuel oil, was completely removed within 1 h under optimized conditions. The catalytic performance of PW12/SiO2–TiO2 was superior to that of PW12/SiO2, which was explained by the presence of surface Lewis acidic sites in the former [107]. No obvious degradation of the catalytic performance was observed upon catalyst recycling during six runs. The retention of the Keggin structure of PW12 was confirmed by FT-IR spectra of the spent catalyst, while XRD measurements and TEM images indicated maintenance of the orderly morphological and microscopic structure. However, the PW12 loading in the reused catalyst, as determined by ICP, dropped from 19.7 to 17.8%, indicating that the leaching of active tungsten species could be possible during the reaction course.
Nogueira and co-workers have suggested the synthesis of core/shell nanoparticles comprising POM as a core encapsulated within silica shell through the hydrolysis of tetraethoxysilane in the presence of POM [108]. Following a similar approach, a PW12@TiO2 core–shell microsphere material was prepared via one-step template synthesis using K3PW12O40 as a core and TiO2 as a shell [109]. Multiple characterization methods (FT-IR, XRD, XPS, Raman, SEM, and TEM) implicated that PW12 was encapsulated into the mesoporous TiO2 phase as the core. The resultant composite was found to be efficient catalyst for ODS of a model oil using H2O2 as an oxidant with 99.9% S removal of DBT after 60 min under the optimum conditions. Additionally, leaching and recycling experiments coupled with catalyst filtration tests confirmed that the PW12@TiO2 is a truly heterogeneous catalyst with excellent recyclability during, at least, seven reuses.

2.2. Electrostatic Binding via Anion-Exchange

Electrostatic binding with NH3+-group of amine-functionalized silica supports continues to be one of the most common methods for POM immobilization. A La-containing POM, [LaW10O36]9−·(LaW10) attached to a commercial mesoporous silica functionalized with (3-aminopropyl)triethoxysilane (aptes) [110] and a sandwich-type [Eu(PW11O39)2]11− (Eu(PW11)2) supported on an aptes-modified SBA-15 [111] behaved as active and recyclable catalysts in ODS of a multicomponent model diesel containing the most refractory sulfur compounds (1-benzothiophene, dibenzothiophene, 4,6-dimethyldibenzothiophene, and 4-methyldibenzothiophene) using H2O2 as the oxidant. Eu(PW11)2@aptesSBA-15 revealed not only a high recycling ability without loss of activity during ten consecutive ODS runs, but also the high robustness and stability of the composite material as confirmed by FT-IR, Raman, 31P MAS NMR, XRD, and SEM/EDS techniques [111].
A PW12-based cross-linked copolymer catalyst was synthesized via the anion exchange of phosphotungstic acid with ionic copolymer obtained through the radical copolymerization of dicationic ([1,1′-(butane-1,4-diyl)-bis(3-vinylimidazolium)]Br2) and amine-functionalized ([3-aminoethyl-1-vinylimidazolium]Br·HBr) ionic liquids [112]. The resulting hybrid material catalyzed the efficient epoxidation of a range of alkenes (cyclooctene, cyclohexene, 1-hexene, and 1-octene) with up to 100% selectivity to the corresponding epoxides under the conditions of oxidant deficiency. The ICP-AES analysis of the filtrates indicated that ca. 3 wt.% of W leached into the reaction medium during the first operation cycle, while leaching became negligible upon further catalyst reuse. The hot catalyst filtration test confirmed the truly heterogeneous nature of the observed catalysis.
Doherty et al. have used cation-decorated polymeric supports based on ionic liquids (PIILPs) prepared by ring-opening metathesis polymerization of a pyrrolidinium-functionalised norbornene-based monomer with cyclooctene to immobilize the Venturello complex [PO4{WO(O2)2}4]3− (PW4) [113]. The resulting solid was active in the epoxidation of a range of alkenes and allylic alcohols using H2O2 with the corresponding epoxides yields up to 94%. The catalyst recovered by centrifugation could be reused at least three times, with a minor reduction in the initial reaction rate. ICP-OES analysis revealed no tungsten in the reaction mixture after the catalyst recovery, while the solid state 31P NMR confirmed stability of the PW4 structure in the recycled catalyst. However, the TEM images of the reused PW4-PIILP have shown some aggregation of the peroxophosphotungstate on the surface. Such aggregation can lead to a reduced accessibility of the active species, which, in turn, may cause decrease in catalytic activity after several reuses.
The Keggin polyanion [PMo10V2O40]5− (PMo10V2) could be successfully immobilized by electrostatic interactions with the positively charged surface of the hydrophilic porous block copolymer membranes on the basis of polystyrene-block-poly(N,N-dimethylaminoethyl methacrylate [114]. The POMbranes thus obtained were active in oxidation of anthracene and 9,10-dihydroanthracene with H2O2, producing anthraquinone with 71–100% yields, as well as oxidized tetrahydrothiophene (THT), to a mixture of sulfoxide and sulfone. Despite the fact that no POM leaching was observed under the turnover conditions, the recycling experiments have shown a steady loss of catalytic activity. The catalyst deactivation could be explained by a strong adsorption of the reaction products on the catalyst surface. Indeed, the reactivation by drying the membrane in a vacuum improved the catalytic performance upon reuse in THT oxidation.
Lindqvist-type Na9LaW10O36·32H2O intercalated into layered double hydroxides modified with ionic liquids (ILs) demonstrated high activity and selectivity in the epoxidation of a range of allylic alcohols with H2O2 [115]. The highest activity was exhibited by a Mg3Al-ILs-C8-LaW10 material based on [Mg0.75Al0.25(OH)2](NO3)0.25·2H2O (Mg3Al-NO3) modified with 1-octyl-3-(3-triethoxy-silylpropyl)-4,5-dihydroimidazolium hexafluorophosphate ((EtO)3Si-ILs-C8). The oxidation of trans-2-hexen-1-ol in the presence of Mg3Al-ILs-C8-LaW10 showed 99% epoxide selectivity at 96% substrate conversion after 2.5 h at 25 °C. The catalyst could be readily isolated from the reaction mixture by centrifugation and reused for, at least, five times without a loss of the initial catalytic activity. The stability of the Mg3Al-ILs-C8-LaW10 structure under turnover conditions was confirmed by FTIR, 29Si NMR CP/MAS, and XPS measurements, while the heterogeneous nature of the catalysis was verified by the catalyst filtration test performed during the reaction course.

2.3. Inclusion in Metal–Organic Frameworks

In recent years, the immobilization of POMs into the cages of metal–organic frameworks (MOFs) has attracted much research attention [70,116,117,118,119,120,121,122,123,124,125,126]. Férey and co-workers were the first who demonstrated that POM molecules can penetrate MOF, and a large cage (3.4 nm diameter with windows of 1.2 and 1.6 nm) of the mesoporous chromium terephthalate MIL-101 can accommodate up to five Keggin-type POMs [127]. The subsequent studies by Maksimchuk et al. revealed that only one polyanion per cage is strongly attached to the MIL-101(Cr) surface (Figure 2), and the POM attachment occurs through anion exchange with nitrate anions present in the MIL-101 framework, while all extra POM molecules easily leach into solution [118,128,129]. The binding between polyanion and MIL-101 is, therefore, electrostatic, and POM can be re-extracted by an anion exchange procedure, for example, using a solution of a perchlorate or nitrate salt [118,128,129].
The combination of MIL-101 and a POM capable of activating H2O2 has made it possible to use hydrogen peroxide as oxidant in a range of selective oxidations (Table 1). Specifically, Figure 3 demonstrates the application of PW12@MIL-101(Cr) in the epoxidation of various alkenes [118,129]. Fairly good yields of epoxides could be obtained for monoterpenes, the substrates prone to various rearrangements, using equimolar amount of H2O2. Although the activity of the PW12/H2O2 catalytic system is commonly attributed to the lower nuclearity peroxotungstates (first of all, PW4 and PW2) formed in situ under the conditions of H2O2 excess [130,131,132,133], the 31P NMR MAS spectrum of the PW12/MIL-101(Cr) catalyst remained practically unchanged after five runs of cyclohexene oxidation. This implies that immobilization on MIL-101(Cr) could increase the stability of the polyanion, even if the transformation of the MIL-included PW12 to low-nuclearity species occurs in the presence of H2O2, such transformation is apparently reversible.
Interestingly, the hybrid materials POM@MIL-101 (POM = [PW11TiO40]5− (PW11Ti) or PW12) demonstrated a pronounced improvement of both epoxidation selectivity and alkene conversion upon enlarging concentration of the oxidant (Figure 4), despite the increasing water concentration in the system (H2O2 was taken as a 30% aqueous solution) [118,128,129]. The reverse trend was observed for the corresponding homogeneous POMs, which is what one would expect, given that the increased concentration of H2O2 (H2O) is favorable for epoxide ring opening and overoxidation processes. The unusual behavior of the POM@MIL-101 catalysts in alkene epoxidation with aqueous H2O2 was explained by the specific sorption properties of the MOF support [118,120]. It was suggested that the hydrophobic part of the terephthalate linker favors adsorption of nonpolar hydrocarbons and, oppositely, disfavors the adsorption of water. This suggestion is supported by the character of the water adsorption isotherm reported for MIL-101, which shows that H2O uptake by MIL-101 starts at high relative pressures (p/p0 = 0.4) [134], indicating that the MOF surface behaves as a hydrophobic one when the concentration of water in the organic solvent is low [120].
Following these first works, a large number of papers concerning POM immobilization within MOFs have been published [55,70,120,121,124,125,126]. Hatton and co-workers reported a one-pot synthesis of MIL-101(Cr)-immobilized phosphotungstic acid, which was active in the epoxidation of caryophyllene with H2O2 (Table 1), behaved as truly heterogeneous catalyst, and could be recycled without loss of activity [135]. Nonetheless, the PW12@MIL-101 material showed a reduction of the POM content upon the recycling. Balula and co-workers inserted a range of POMs (Tb-substituted Keggin-type dimer [Tb(PW11O39)2]11− (Tb(PW11)2) [136], sandwich-type Co4(PW9)2 (see Figure 1 for the structure) [137], trivacant ([A-PW9O34]9− (PW9), [138], Venturello peroxotungstate PW4 [139], and europium-substituted Lindqvist dimer [Eu(W5O18)2]9− (EuW10) [140]) by post-synthetic impregnation of MOF (MIL-101(Cr) or ZIF-8) with a POM solution. PW9@MIL-101 revealed catalytic activity in oxidation of monoterpenes and S-compounds with H2O2. Most of the hybrid materials exhibited high activity in the oxidative desulfurization [124,137,138,139,140] (Table 2), whereas Co4(PW9)2@MIL-101 also catalyzed oxidation of alkenes (geraniol, limonene, styrene) and cyclooctane with H2O2 (Table 1) [137]. While the oxidation of geraniol gave only 1,2-epoxygeraniol, limonene was transformed mainly to 1,2-epoxide and the corresponding 1,2-diol, whereas benzaldehyde was the main product obtained from styrene [137]. Although the structure of both POM (confirmed by EDX, FT-IR, and Raman measurements) and MOF (confirmed by XRD and SEM) was retained, the catalyst kept its activity during at least three consecutive runs, and some POM leaching was documented [137].
A one-pot mechanochemical synthesis was suggested as an efficient strategy to insert PMoV heteropolyacids H3+nPMo12−nVnO40 (n = 2–4) into the large cavities of a metastable MOF, rho-ZIF [141]. The PMoV@rho-ZIF composites were effective catalysts for the selective oxidation of a series of organic sulfides to sulfoxides with aqueous H2O2 (Table 1). The heterogeneous nature of the catalysis was confirmed by the catalyst filtration experiment, and the PMoV@rho-ZIF materials could be recycled at least four times without a significant loss of activity, with retention of the POM and MOF structures. Wang et al. immobilized an ionic liquid–phosphotungstate inside the cages of pre-activated MIL-101(Cr) with unsaturated Cr3+ active sites via formation of Cr–N dative bonds [142]. The catalytic material was active in the H2O2-based oxidation of cyclohexene to produce adipic acid (AA) and maintained activity for at least six reuses. Yields of AA up to 78% under solvent-free conditions have been claimed (Table 1). Unfortunately, the absence of leaching and the nature of the catalysis were not addressed by the authors.
In order to suppress the leaching of immobilized POM, Qiu and co-workers proposed an encapsulation strategy based on an ionic liquid bridging [143,144]. Thus, phosphotungstic acid was successfully encapsulated into the Zr-based MOF UiO-66 by (i) in situ solovothermal procedure using carboxyl-functionalized methylimidazole as bridging molecule [143] or (ii) manual grinding technique using 4-aminopyridine as bridging compound [144]. A range of physicochemical methods (XRD, EDS, and XPS) confirmed the successful introduction and high dispersion of POM. The PW12@UiO-66 composites exhibited high activity in oxidative desulfurization of DBT (as 1000 ppm S model oil) with H2O2 (Table 2). The authors suggested that such activity is attributed to the high dispersion of the active component on the surface of UiO-66 and synergy between the Zr(IV) open sites in the MOF and active W(VI) sites in the POM. Indeed, IL-functionalized PW12, POM-free UiO-66, and even a mixture of POM and MOF showed lower activity in DBT desulfurization, as compared to the composite catalyst [143].
Phosphotungstic and phosphomolybdic (PMo12) heteropolyacids were encapsulated within the MOF-808 by in situ synthesis based on a ‘ship-in-a-bottle’ approach [145,146]. Lin et al. used MOF-808X with tunable window diameters [145]. Both PMo12@MOF-808 and PW12@MOF-808X revealed high activity and stability in ODS of model fuels and could be used repeatedly without a loss of catalytic properties (Table 2). A few UiO-66(67)-incorporated hybrid materials with a range of POMs (PW12 or PMo12 heteropolyacids and salts), and they were synthesized via a direct solvothermal reaction [147,148,149]. A composite PW12@UiO-66 with 35 wt% of POM was highly active in the selective oxidation of cyclopentene to glutaraldehyde with H2O2 (78% yield at 95% substrate conversion, see Table 1), did not suffer active metal leaching, and could be used repeatedly [147].
Table 1. H2O2-based oxidations over MOF-included POMs.
Table 1. H2O2-based oxidations over MOF-included POMs.
POMMOFSynthesis MethodSubstrate/Conv. %Product/ Selectivity, %Nature of
Catalysis a
Ref.
PW12MIL-101(Cr)Ads. bCatalysts 13 00360 i00155Catalysts 13 00360 i00289Heterog.[129]
Catalysts 13 00360 i00350Catalysts 13 00360 i00461
Catalysts 13 00360 i00572Catalysts 13 00360 i00676
PW4MIL-101(Cr)Ads.Catalysts 13 00360 i00775Catalysts 13 00360 i00877Heterog.[129]
PW12MIL-101(Cr)Solv. cCatalysts 13 00360 i009n.d. dCatalysts 13 00360 i01092 eHeterog.[135]
IL-PW12MIL-101(Cr)Post-synth. fCatalysts 13 00360 i011n.d.Catalysts 13 00360 i01278 en.d.[142]
PW11TiMIL-101(Cr)Ads.Catalysts 13 00360 i01388Catalysts 13 00360 i014100Heterog. g[128]
Co4(PW9)2MIL-101(Cr)Ads.Catalysts 13 00360 i015˃99Catalysts 13 00360 i016˃99n.d.[137]
Catalysts 13 00360 i01757Catalysts 13 00360 i018˃99
PW9MIL-101(Cr)Imp. iCatalysts 13 00360 i01993Catalysts 13 00360 i02050 hHeterog.[138]
Catalysts 13 00360 i021˃99Catalysts 13 00360 i022˃99
PMo11Vrho-ZIFMechanochem.Catalysts 13 00360 i02383Catalysts 13 00360 i02489Heterog.[141]
Catalysts 13 00360 i02597Catalysts 13 00360 i02696
PW12UiO-66Solv.Catalysts 13 00360 i02795Catalysts 13 00360 i02883n.d.[147]
Co-PMo12 jUiO-67Solv.Catalysts 13 00360 i02982Catalysts 13 00360 i030˃99Heterog.[149]
PMo12UiO-67Solv.75˃99n.d.
a Nature of catalysis (truly heterogeneous catalysis was confirmed by hot catalyst filtration tests). b Synthesis by adsorption of POM from acetonitrile. c Solvothermal one-pot synthesis. d Not determined. e Isolated yield. f Immobilization by post-synthetic strategy through a Cr–N dative bond. g Depended on the reaction conditions. h Diepoxide and 1,2-diol were also formed. i Impregnation of MOF with aqueous solution of PW9. j Presumably CoHPMo12O40.
Table 2. H2O2-based ODS over MOF-immobilized POMs.
Table 2. H2O2-based ODS over MOF-immobilized POMs.
POMMOFSynthesis
Method
O/S,
mol. Ratio
Sulfur
Content, ppm
Desulfurization %Nature of Catalysis aRef.
Tb(PW11)2MIL-101(Cr)Imp. b211500˃99Heterog.[136]
PW9MIL-101(Cr)Imp.211700˃99Heterog.[138]
PW4MIL-101(Cr)Imp.52000˃99n.d. c[139]
EuW10ZIF-8Imp. 150095n.d.[140]
PW12MOF-808Solv. d51000 e100Heterog.[145]
PMo12MOF-808Solv.21150092n.d.[146]
PMo12UiO-66Solv.3500100n.d.[150]
(mim(CH2)3COO)-PW12 fUiO-66Solv.51000 e100n.d.[143]
PW12UiO-66-D gSolv.7800100n.d.[151]
PW12UiO-66Mechanochem.51000 e100Heterog.[144]
PW12UiO-67Solv.131000100Heterog.[148]
BMIM-PMo12 hZIF-8Solv.16200092n.d.[152]
a Nature of catalysis (truly heterogeneous catalysis was confirmed by hot catalyst filtration tests). b Synthesis by impregnation of MOF with solution of POM. c Not determined. d Solvothermal one-pot synthesis. e DBT was used as model fuel. f 1-Carboxypropyl-3-methyl imidazolium phosphotungstate [mim(CH2)3COOH]3[PW12O40]. g UiO-66 with defects. h 1-Butyl-3-methylimidazolium phosphomolybdate [BMIM]3[PMo12O40].
Another method of immobilization is related to the use of POMs as building blocks for the construction of POM-based frameworks (so-called POMOFs) [119,153]. The family of diverse POMOF materials combines the advantages of POM chemistry with the specific MOF’s properties; however, just a few of these materials were assessed as catalysts for aerobic- or tBuOOH-based oxidations [154,155,156,157,158], and no H2O2-based oxidation was reported.
The main limitation for the use of MOF-supported POMs in oxidation catalysis, especially in H2O2-based oxidations, is related to the stability of the chosen framework. For example, in the case of MIL-101-supported POMs, the reaction conditions have a great impact on the catalyst stability. At temperatures below 50 °C and [H2O2] = 0.2 M or lower, POM/MIL-101 composites behave as true heterogeneous catalysts and can be recycled with a retention of the catalytic properties, while at higher H2O2 concentrations and/or temperatures, the MIL-101 matrix starts to destroy, leading to deactivation of the POM/MIL-101 catalysts [128].

2.4. Covalent Anchoring in Hybrid Materials

The covalent approach to the immobilization of active complexes offers the obvious advantage of improved stability of the hybrid material. To perform the covalent post-functionalization of POMs, the appropriate organic functions need to be grafted on the polyanion. Proust and co-workers elaborated a synthetic approach applicable to the functionalization of catalytically active vacant Keggin-type POMs with organo-silyl or organo-phosphonyl moieties [62,66,159]. Later on, Villanneau and co-workers have proposed an efficient strategy for the covalent linkage of POM via functionalization by complementary organic functions of both vacant polyoxometalate species and mesoporous silica support (Figure 5) [160]. A few combinations of functional groups have been tested, including carboxylic acid or alkylamine/amine functions for either vacant POM or silica support modification [160,161]. This strategy allowed for an excellent nanostructuration of the POMs shell at the surface of silica support to be created, as determined by HR-TEM (Figure 5). The resulting hybrid materials exhibited a high resistance to leaching in ionic and polar solvent mixtures [160] and, moreover, in oxidative reaction medium of H2O2-based liquid-phase oxidations [161]. The main limitation of this approach was the low control of the quantity of POM grafted, which hardly corresponded to the amount of POM engaged during the preparation.
The catalytic performance of POM-CO2H@SBA-NH2, POM-CO2H@MCF-NH2, and POM-NH2@SBA-CO2H materials (POM = (tBu4N)3NaH[AsIIIW9O33{P(O)(CH2CH2CO2H)}2] (AsW9-P(O)COOH) or (tBu4N)4H[PW9O34{As(O)(C6H4NH2)}2] (PW9-As(O)NH2)) was assessed in cyclooctene epoxidation with H2O2 at room temperature, and the resulting substrate conversions varied in the range of 19–76% (Table 3) [161]. The material with the most regular distribution of POM inside the support channels (POM-CO2H@SBA-NH2) showed the best activity among others, although the reaction rate was lower than in homogeneous conditions. This catalyst was also active in cyclohexene epoxidation, but similarly to cyclooctene oxidation, a significant decrease in the initial reaction rate was observed, as compared to the corresponding homogeneous catalyst. Meanwhile, the final conversion was at the same level (ca. 75%), and the selectivity to epoxy cyclohexane was even higher for the immobilized POM than with the homogeneous one (94 vs. 81%, Table 3).
The covalent immobilization methodology was further amended using cross-linking agents that are able to form specific covalent bonds with both a surface and the desired molecule. The hybrid derivative of a heteropolytungstate with two aniline groups, (Bu4N)3[NaHPW9O34{As(O)C6H4-p-NH2}2] was post-functionalized by 1,4-phenylene diisothiocyanate and then covalently grafted onto the surface of an amino-functionalized SBA-15 via formation of thiourea bonds [162]. The formation of thiourea link was confirmed by 13C CP-MAS NMR spectroscopy, whereas the dispersion of POM along the channels of the SBA-15 support, without the formation of large aggregates, was corroborated by HR-TEM. The catalytic behavior of the grafted catalyst was tested in a model reaction of cyclooctene epoxidation with H2O2. However, the immobilization of POM has led to a deterioration of the catalytic performance, most likely caused by the steric hindrance around the active site.
Kozhevnikov and co-workers elaborated an immobilization technique of POM binding onto the surface of a phosphazene (RPN, R = benzyl, iso-butyl, or iso-propyl) functionalized silica via alkylaminocyclophosphazene tethers, which could potentially prevent POM from leaching [163]. The POM/RPN-SiO2 (POM = [PW12O40]3− or [PMo12O40]3−) catalysts were active in the oxidative desulfurization of benzothiophenes with H2O2 in a heptane-H2O reaction medium, and PMo12/BzPN-SiO2 exhibited higher activity than homogeneous PMo-BzPN counterpart.
Liu et al. grafted poly(glycidyl methacrylate) (PGMA) onto the surface of SiO2 to prepare ‘hairy’ PGMA/SiO2 particles, which were further used for the covalent binding of the amino-modified vacant Wells-Dawson POM [α-P2W17O61]10− (P2W17) [164]. The resulting P2W17-PGMA/SiO2 material catalyzed oxidative desulfurization of tetrahydrothiophene with H2O2 (100% conversion after 60 min) behaved as a truly heterogeneous catalyst, according to the filtration experiment, and could be reused at least five times with the retention of its activity and structure.
While covalently bound POMs usually reveal superior stability to leaching, their activity is typically lower, relative to the homogeneous counterparts. Another shortcoming is associated with synthetic difficulties of covalent derivatization of many catalytically interesting POMs.

2.5. Encapsulation within Supramolecular Complexes

Surfactant-encapsulated polyoxometalates (SEPs) can be formed by exchanging counterions in POMs with cationic surfactants. Depending on the surfactant nature, the resulting SEPs can be soluble in traditional organic solvents acting as homogeneous catalysts or behave as heterogeneous catalysts. Amphiphilic POM-based, well-defined, and robust nanowires and nanotubes (the assembly shape depends on the nature of the encapsulating molecules) constructed by self-assembly of units consisting of POMs encapsulated by single-chain surfactant molecules (Figure 6) can provide an optimal catalytic environment, acting as a dual trapping catalyst [165]. The hydrophobic long alkyl chains on the surface of POM cluster adsorb the weakly polar sulfide molecules, where they are oxidized to sulfones by an active peroxo-POM-complex and then easily escape because of their polarity. The POM–amphiphile units constructed of phosphotungstic acid and a series of surfactants varying in alkyl chain length (dodecyltrimethylammonium (DDA), tetradecyltrimethylammonium, hexadecyltrimethylammonium, and octadecyltrimethylammonium) catalyzed oxidation of dibenzothiophene into corresponding sulfone with 100% selectivity at complete conversions (20–40 min, 50 °C) [165]. After the oxidation reaction, the catalyst could be recovered by filtration or centrifugation and reused with retention of the mesomorphology and catalytic properties. The 31P NMR investigations of the reaction solutions indicated no leaching of the active species.
Song and co-workers have prepared a series of amphiphilic lanthanide-containing POMs ([LnW10O36]9−, Ln = La, Ce, Y, Nd, Sm, Eu, Gd, Tb, Yb) by surfactant encapsulation using surfactants varying in alkyl chain lengths (DDA, trimethylstearylammonium, and dimethyldioctadecylammonium (DODA)) and tested the catalytic activity of the new SEPs in ODS reactions using H2O2 [166]. (DDA)9LaW10 in combination with [omim]PF6 (omim = 1-octyl-3-methyl-imidazolium) as a sulfone extractant exhibited 100% DBT conversion in 14 min under mild conditions (30 °C, O/S = 5).
An interesting POM immobilization strategy involving both covalent and non-covalent interactions has been developed by Wu and co-workers [54,167,168,169,170]. The concept involves the surfactant encapsulation of POM via replacing its counter ions by quaternary ammonium cations (di(11-hydroxyundecyl)dimethylammonium (DOHDA), DODA, or didodecyldimethylammonium) through electrostatic interactions, thus forming a structure similar to that of a reverse micelle system (Figure 7).
The resulting hybrid species can then be covalently fixed within a silica gel matrix through sol–gel condensation with tetraethyl orthosilicate. Such immobilized supramolecular hybrid catalysts possess a defined hydrophobic environment to capture the organic substrate, which could potentially maximize the catalytic efficiency. The supramolecular hybrid catalyst based on PW12 polyanion and DOHDA cation was active in a range of H2O2-based selective oxidations [169]. Thus, DBT gave the corresponding sulfone as the sole product. By analyzing 1H NMR spectra of DBT in the presence of pure silica, surfactant-encapsulated POM species, and the resulting hybrid catalyst, the authors have concluded that hydrophobic nano-environments around the POM in the catalyst tend to capture the substrates of low polarity (sulfides) and release the polar oxidized product–sulfone. The structural integrity of the developed hybrid material under turnover conditions was confirmed by 31P NMR, IR, and XPS measurements along with elemental analysis data. Moreover, the catalytic activity of the recovered catalyst was as high as that of the freshly prepared one for at least five consecutive runs. The elaborated hybrid catalyst was also active in epoxidation of cyclohexene with nearly 100% selectivity to the corresponding epoxide at 86% substrate conversion and revealed the complete oxidation of cyclohexanol to cyclohexanone [169].
Following the same methodology, Wu and co-workers synthesized chiral complexes via the electrostatic interaction of a catalytically efficient sandwich POM, [WZn3(H2O)2(ZnW9O34)2]12− and a new cationic surfactant bearing chiral head, which were further covalently immobilized into the silica matrix using a sol–gel procedure [171]. The supramolecular chirality around the POM made possible kinetic resolution of racemic alcohols via catalytic oxidation with H2O2.
Nogueira et al. prepared a composite material based on a cross-linked organic/inorganic hybrid core composed by the mono-substituted POM [PW11Zn(H2O)O39]5− (PW11Zn) and amine-organosilane (aptes) surrounded by a silica shell [172]. The resulting PW11Zn-aptes@SiO2 composite catalyzed the selective epoxidation of cyclooctene and geraniol, as well as oxidative desulfurization of a model oil with H2O2, and the catalytic performance of the immobilized POM was similar to that of homogeneous PW11Zn counterpart. Despite the fact that elemental analysis has shown the leaching of ca. 6% of the initial POM amount during an ODS run, the experiment with catalyst filtration confirmed the true heterogeneous nature of the observed catalysis.
If a charge transfer (CT) complex is employed for POM encapsulation, the resulting composite material can exhibit properties of a photothermal catalyst. Wu and co-workers synthesized a series of on-site heating catalytic composites through the electrostatic complexation of a cationic CT-complex with 3,3′,5,5′-tetramethylbenzidine (TMB) bearing strong NIR absorption and a range of Keggin polyanions (PMo10V2, PMo11V, PW11V, and PMo12) [173]. The resulting assembled materials displayed an integrated NIR photothermal effect and enhanced catalytic activity in the mild oxidation of toluene and its derivatives, as well as alkenes and organic sulfides with H2O2 [173]. Under NIR radiation, a substrate conversion increased ca. 25%, relative to the external heating. The CT complex plays the role of a local heater attaching to the POM catalyst, which enhances the catalytic reaction efficiency, while itself does not join the catalytic process. Filtration tests confirmed heterogeneous nature of the catalysis over PMo10V2@TMBCT, and besides, no increased conversion was detected in the filtrate, even under the external heating to 55 °C. The PMo10V2@TMBCT catalyst could be successfully recycled with retention of its structure, as confirmed by FT-IR and UV–Vis-NIR spectroscopy.

2.6. Immobilization on Carbon Materials

The entrapping of POMs inside the pores of activated carbons was first reported by Izumi and Urabe [174]. Later on, Van Bekkum et al. showed that the carbon nature and the activation method strongly affect the strength of adsorption and amount of adsorbed heteropolyacids [175]. A summary of the early literature on this topic can be found in the book chapter by Hill and Kholdeeva [38].
The mesoporous carbon Sibunit was found to be one of the best carbon supports that enabled strong irreversible adsorption of catalytically active POMs with retention of their structure and catalytic properties [65]. The use of Sibunit as a carrier made it possible to prepare catalysts that remained stable toward POM leaching, even if highly polar products, such as carboxylic acids, were formed in the reaction [65]. However, the main disadvantage of POM/Sibunit catalysts was their poor reusability caused by strong adsorption of oxidation products, which cannot be removed from the catalyst surface by washing, extraction, or evacuation [65,176].
In recent years, carbon nanomaterials (CNMs) have received great attention for the preparation of hybrid inorganic materials for energy storage and conversion, electronic and optic sensors, and catalysis [177,178,179,180]. The immobilization of POM on CNMs was accomplished using electrostatic [181,182,183,184] and π-π [185] interactions, as well as covalent binding [180,186].
Hybrid materials based on POMs and CNMs have been widely used in electrocatalysis and energy storage applications [181,182,184,185,186,187,188,189,190]. In recent years, such materials have also attracted attention as selective oxidation catalysts [176,183,191,192,193,194,195,196]. Salavati et al. reported POM immobilization by a one-step solid-state reaction of a sodium salt of PMo10V2 with multiwall carbon nanotubes (MWCNTs) [192]. The resulting materials catalyzed oxidation of a range of alkenes by H2O2 with selectivity to the corresponding epoxides of 58–100% at 47–80% conversions. The catalyst maintained its catalytic properties for at least four reuses, despite some leaching of vanadium during the reaction. A catalyst prepared by the deposition of Cs2.5H0.5PW12O40 on CNTs was used for the oxidative removal of DBT, with a desulfurization efficiency of up to 100%; however, no information on POM leaching was provided [193].
The functionalization of CNMs provides additional opportunities for the immobilization of active complexes, including POMs. Hajian and Alghour have developed a hybrid catalyst, following a sequence of MWCNT modifications [191]. Carboxylic acid-tailored MWCNTs (MWCNT-COOH) were first functionalized with thionyl chloride and then with 1-(3-aminopropyl) imidazole, followed by the reaction with 1-bromobutan (APIB) and the subsequent addition of tetrabutylammonium salt of [PZnMo2W9O39]5− (ZnMo2W9). Imidazolium (Im) groups present on the support bound the POM electrostatically. The resulting hybrid ZnMo2W9/MWCN-APIB exhibited high activity and selectivity in the oxidation of various alcohols, with H2O2 producing corresponding aldehydes. Although noticeable leaching was observed only during the first two reuses, catalytic activity staidly decreased during seven consecutive runs. Another group used Im-based ionic liquids to modify the surface of amphiphilic graphene oxide (GO) for the immobilization of PW12 via anion exchange [197]. The Im-PW12/GO material catalyzed oxidation of a range of alcohols with H2O2, while the hot catalyst filtration test confirmed the heterogeneous nature of catalysis.
Masteri-Farahani and Modarres immobilized PW4 through electrostatic interaction with the surface of graphene oxide modified with cysteamine hydrochloride as a linking agent [198]. The resulting PW4/GO catalyst was active in H2O2-based oxidation of alkenes and allylic alcohols with selectivity to the corresponding epoxides ˃99%. Moreover, the catalyst could be readily recovered by centrifugation and reused for at least five runs, without a significant decrease in activity. The catalyst filtration experiment exhibited almost no activity in the filtrate after the separation of the catalyst from the reaction mixture, approving the high stability of the PW4/GO catalyst to leaching and heterogeneous nature of catalysis. No PW4 species were detected in the solution by means of ICP-OES.
Gan et al. modified carbonized cellulose nanofiber (CCNF) with polyethyleneimine and γ-(2,3-epoxypropoxy)propyltrimethoxysilane, in order to load it with K4[β-SiMo3W9O40]·5H2O [199]. The resulting POM/CCNF material was active in ODS reactions: the removal efficiency of DBT, benzothiophene, and 4,6-dimethyldibenzothiophene was 99, 89, and 100%, respectively, while using model oils with the initial sulfur concentration of 2000 ppm. Recycling experiments have shown that the POM/CCNF catalyst kept its activity for at least three consecutive runs.
Wang and co-workers used a ‘click reaction’ between the azido group of K1.5Cs5.5[γ-SiW10O39Cu2(N3)2] (SiW10Cu2(N3)2) and the alkynyl groups on functionalized graphene for covalent immobilization of POM [200]. The SiW10Cu2/GO catalyst exhibited excellent activity in the oxidation of toxic 2-chloroethyl ethyl sulfide (mustard simulant) to much less toxic sulfoxide, with 100% selectivity at complete conversion, and it could be reused at least 10 times without losing catalytic properties. Experiment with catalyst filtration during the course of the reaction confirmed the heterogeneous nature of catalysis, while 29Si MAS NMR along with TEM measurements corroborated retention of the catalyst original structure and morphology after the oxidation reaction.
The doping of carbon nanomaterials with nitrogen in the process of their synthesis leads to the formation of different types of surface N-species (pyridine-like, pyrrole-like, or quaternary ones), which afford supplementary opportunities for immobilization of catalytically active complexes and may also affect the hydrophilic properties of the supported catalysts [201,202,203]. Evtushok et al. first reported the use of N-CNMs for immobilization of POM [176]. The di-V-substituted γ-Keggin phosphotungstate [γ-PW10O38V2(μO)(μ-OH)]4− (γ-PW10V2) was immobilized on bamboo-like N-doped carbon nanotubes (N-CNTs) [176] and N-doped carbon nanofibers (N-CNFs) having herring-bone packing of graphite layers [195] by adsorption from acetonitrile. Undoped CNTs were also used for comparison. It was shown that the presence of nitrogen is necessary to ensure the strong attachment and quasi-molecular dispersion of the POM on the carbon surface, which, in turn, affects the catalytic behavior and catalyst stability under turnover conditions. Other key factors that favor the successful and maximum adsorption of POM are pre-drying of the supports and addition of mineral acid. The addition of acid makes possible irreversible binding of γ-PW10V2 to the surface with retention of the POM structure and catalytic properties. Using an optimal catalyst PW10V2/N-CNTs (15 wt.% POM, 1.8 at.% N), 2,3,6-trimethylphenol (TMP) was oxidized with H2O2 to afford trimethyl-p-benzoquinone (TMBQ, vitamin E precursor) in a nearly quantitative yield and with 80% oxidant utilization efficiency, retaining characteristics of homogeneous γ-PW10V2 (Figure 8). It is noteworthy that γ-PW10V2 requires retention of its structure and protonation state to perform heterolytic activation of H2O2 [204]. The first attempts of the immobilization of γ-PW10V2 onto commercially available Fe2O3 led to catalysts that were stable in a 1:1 mixture of EtOAc/tBuOH, but lost POMs in CH3CN [205]. Immobilization using other approaches, such as embedding into silica, electrostatic attachment to amine-modified SiO2 or MIL-101, and adsorption on Sibunit, resulted in a pronounced decrease in catalytic activity and selectivity of γ-PW10V2 (Figure 8), which demonstrated the significant leaching of POM, leading to high contribution of homogeneous catalysis (γ-PW10V2@SiO2), or revealed poor recyclability (γ-PW10V2/Sibunit). The confined space inside nanotube tangles endowed the catalyst with superior activity (TOF = 500 h−1) and space–time yield (450 g L−1 h−1) [176,206]. Moreover, the catalyst did not suffer POM leaching under the turnover conditions and showed excellent recycling performance, in contrast to γ-PW10V2 immobilized on N-free CNTs.
The scope of N-free and N-doped CNTs as supports was also explored in the immobilization of the Venturello complex PW4 [194] and Nb-substituted Lindqvist tungstates [Nb(L)W5O18]n− (Nb(L)W5, L = O, OCH3, (O)2 or OOH) and [(NbW5O18)2O]4− ((NbW5)2O) [196]. In contrast to the immobilization of γ-PW10V2, neither N-doping nor acid additives were imperative to providing strong binding and quasi-molecular dispersion of PW4 on the surface of CNTs. Moreover, N-free CNTs have proven to be even more preferable than N-CNTs as a support for PW4 and the Lindqvist Nb-POM because, in contrast to N-CNTs, they are inert for H2O2 unproductive decomposition and, thus, enable a higher oxidant utilization efficiency. The addition of acid during the immobilization process allows the POM content in the catalyst and catalytic activity to be increased, but acidity may be objectionable for product selectivity. In the oxidation of cyclohexene and other acid-sensitive alkenes, acidity deteriorates epoxide selectivity because of epoxide-ring opening and overoxidation processes, whereas selectivity to the sum of heterolytic oxidation products remains high (Table 4). On the contrary, a marked improvement in sulfoxide selectivity and yield is observed for the thioether oxidation (Figure 9), most likely due to an increased electrophilicity of the active peroxo species, which favors sulfoxidation and disfavors subsequent oxidation to sulfone [57,194,207,208]. The catalyst 15 wt.% PW4/CNTs prepared using 2 equiv. of HClO4 was effective and truly heterogeneous for the epoxidation of cyclooctene, which is not an acid-sensitive substrate, and sulfoxidation of various thioethers, but the catalytic activity gradually decreased during recycling. The deactivation was, most likely, caused by the loss of catalyst acidity and/or diffusion of PW4 inside the inner channels of CNT (the latter could be observed in HR-TEM images [194]). The scope of the acid-free catalyst 5 wt.% PW4/CNTs for alkene epoxidation was broader (Table 4), and it did not lose activity after reuse. Dimethyl carbonate was the solvent of choice in terms of catalyst activity and stability to POM leaching.
The mmobilization of the Lindqvist-type Nb-POM onto the surface of both CNTs and N-CNT enabled stabilization of the most reactive monomeric form Nb(OH)W5, thus preventing its dimerization and subsequent deactivation [196]. As a result, (N)-CNTs-supported Nb(L)W5 exhibited higher catalytic activity than their homogeneous counterparts. The addition of acid upon the immobilization increased the Nb-POM loading [196] and, moreover, facilitated the formation of active hydroperoxo species Nb(OOH)W5 responsible for the heterolytic oxygen transfer to C=C bond of alkene [207]. Similar to γ-PW10V2, N-doping facilitated a quasi-molecular dispersion of Nb(L)W5 on the surface, but it also favored unproductive decomposition of H2O2, thereby leading to the worsening of the catalytic performance in terms of alkene conversion and epoxide selectivity (see Table 4). Therefore, the use of undoped CNTs for the preparation of epoxidation catalysts seems to be advantageous.
Importantly, CNT-supported PW4 and Nb(L)W5 did not suffer from POM leaching if dimethylcarbonate was employed as solvent and demonstrated a truly heterogeneous nature of the catalysis and excellent recycling performance [57,194,196].

3. Conclusions

Polyoxometalates have great potential as catalysts for the liquid-phase selective oxidation of organic compounds using green oxidants, in particular, dilute hydrogen peroxide. During the last decade, some new highly active and selective POM catalysts have been discovered, and various approaches have been proposed for POM immobilization on solid supports to make easy catalyst separation and recycling possible. The most developed and prospective methodologies that enable the strong binding of POM to the surface and preventing leaching into the solution involve POM immobilization by means of different types of chemical bonding and supramolecular interactions. The proper choice of the POM, support, and immobilization method depends on the type of oxidative transformation and may also depend on the nature of the organic substrate and target oxidation product.
MOFs are ‘smart’ supports that can enhance catalytic properties of POMs by different synergetic effects; however, the reaction conditions, especially in the case of oxidation with aqueous H2O2, should be properly controlled, in order to keep the structure of MOF within its stability limit. The organic/inorganic hybrid derivatives of POMs are a new class of catalysts that offer an undisputable advantage of ability to covalent bonding using different linking strategies with an excellent dispersion along the surface of an inorganic support and with the retention of the catalytic properties. However, synthetic difficulties for derivatization of some POMs may be a shortcoming of this approach. Encapsulation within the supramolecular complexes may provide an optimized catalytic environment based on dual polar properties. Moreover, if a charge transfer complex is employed for POM encapsulation, the resulting composite material can exhibit properties of a photothermal catalyst. For those POMs, which require retention of the structure coupled with the specific protonation state to keep catalytic activity, the best choice might be use of carbon-based nanomaterials, in particular, carbon nanotubes. A blend of electrostatic forces and hydrogen bonding ensures the excellent stability and recyclability of the CNTs-supported POM catalysts, while a balance between the activity and selectivity can be tuned through the careful control of the amount of acid added during the POM immobilization.
So far, most of the elaborated supported POM catalysts were evaluated for the selective oxidation of S-compounds, olefins, phenols, and alcohols to produce sulfoxides, sulfones, epoxides, diols, quinones, aldehydes, ketones, and a few highly efficient catalyst systems have been suggested. However, the potential of immobilized POMs for the selective oxidation of alkanes and aromatics still remains very little explored. The development of supported POM catalysts for oxidative cleavage of C=C and C–C bonds and production of highly polar products–carboxylic and dicarboxylic acids–continues to be a challenge for oxidation catalysis researchers and POM chemists.

Author Contributions

Writing—review and editing, O.A.K. and N.V.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the Ministry of Science and Higher Education of the Russian Federation within the governmental order for the Boreskov Institute of Catalysis (project АААА-А21-121011390008-4).

Data Availability Statement

Not applicable. Refer to the original papers for data availability.

Acknowledgments

The authors thank all co-authors of the joint papers published on H2O2-based selective oxidation catalysis by supported polyoxometalates.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AA: adipic acid; APIB, 1-bromobutan; aptes, (3-aminopropyl)triethoxysilane; AsW9-P(O)COOH, (tBu4N)3NaH[AsIIIW9O33{P(O)(CH2CH2CO2H)}2]; BMIM, 1-butyl-3-methylimidazolium; CCNF, carbonized cellulose nanofiber; CNMs, carbon nanomaterials; Co-PMo12, CoHPMo12O40; Co4(PW9)2, [Co4(H2O)2(PW9O34)2]10−; CT, charge transfer; DBT, dibenzothiophene; DDA, dodecyltrimethylammonium; DODA, dimethyldioctadecylammonium; DOHDA, di(11-hydroxyundecyl)dimethylammonium; EISA, evaporation-induced self-assembly; (EtO)3Si-ILs-C8, 1-octyl-3-(3-triethoxy-silylpropyl)-4,5-dihydroimidazolium hexafluorophosphate; EuW10, [Eu(W5O18)2]9−; Eu(PW11)2, [Eu(PW11O39)2]11−; GO, graphene oxide; Im, imidazolium; LaW10, [LaW10O36]9−; Mg3Al-NO3, [Mg0.75Al0.25(OH)2](NO3)0.25·2H2O; mim, 3-methyl imidazolium; MOFs, metal–organic frameworks; MWCNTs, multiwall carbon nanotubes; Nb(L)W5, [Nb(L)W5O18]n−; (NbW5)2O, [(NbW5O18)2O]4−; N-CNFs, N-doped carbon nanofibers; N-CNTs, N-doped carbon nanotubes; ODS, oxidative desulfurization; omim, 1-octyl-3-methyl-imidazolium; PGMA, poly(glycidyl methacrylate; PIILPs, polymer immobilized ionic liquid phase; PMo12, [PMo12O40]3−; PMo10V2, [PMo10V2O40]5−; PMo11V, [PMo11VO40]4−; POMs, polyoxometalates; PW4, [PO4{WO(O2)2}4]3−; PW9, [A-PW9O34]9−; PW9-As(O)NH2, (tBu4N)4H[PW9O34{As(O)(C6H4NH2)}2]; PW11Ti, [PW11TiO40]5−; PW11Zn, [PW11Zn(H2O)O39]5−; PW12, [PW12O40]3−; P2W17, [α-P2W17O61]10−; PY11M, [PY11MO40]n−; RPN (R = benzyl, iso-butyl or iso-propyl), phosphazene; SEPs, surfactant-encapsulated polyoxometalates; SiW10Cu2(N3)2, K1.5Cs5.5[γ-SiW10O39Cu2(N3)2]; Tb(PW11)2, [Tb(PW11O39)2]11−; THT, tetrahydrothiophene; TMB, 3,3′,5,5′-tetramethylbenzidine; TMP, 2,3,6-trimethylphenol; TMBQ, trimethyl-p-benzoquinone; γ-XW10V2 (X = P, Si), [γ-XW10O38V2(μ-O)(μ-OH)]n−; ZnMo2W9, [PZnMo2W9O39]5−.

References

  1. Sheldon, R.; Arends, I.W.C.E.; Hanefeld, U. Green Chemistry and Catalysis; Wiley VCH: Weinheim, Germany, 2007. [Google Scholar]
  2. Clerici, M.G.; Kholdeeva, O.A. (Eds.) Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Wiley: Hoboken, NJ, USA, 2013. [Google Scholar]
  3. Duprez, D.; Cavani, F. (Eds.) Handbook of Advanced Methods and Processes in Oxidation Catalysis; Imperial College Press: London, UK, 2014. [Google Scholar]
  4. Jones, C.W. Application of Hydrogen Peroxide and Derivatives; Royal Society of Chemistry: Cambridge, UK, 1999. [Google Scholar]
  5. Strukul, G. (Ed.) Catalytic Oxidations with Hydrogen Peroxide as Oxidant; Kluwer: Dordrecht, The Netherlands, 1992. [Google Scholar]
  6. Strukul, G.; Scarso, A. Environmentally benign oxidants. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 1–20. [Google Scholar]
  7. Strukul, G.; Menegazzo, F. (Eds.) Catalysts 2019, Special Issue on Direct Synthesis of Hydrogen Peroxide. Available online: https://www.mdpi.com/si/catalysts/hydrogen_peroxide (accessed on 10 December 2022).
  8. Campos-Martin, J.M.; Blanco-Brieva, G.; Fierro, J.L.G. Hydrogen peroxide synthesis: An outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef] [PubMed]
  9. Edwards, J.K.; Freakley, S.J.; Lewis, R.J.; Pritchard, J.C.; Hutchings, G.J. Advances in the direct synthesis of hydrogen peroxide from hydrogen and oxygen. Catal. Today 2015, 248, 3–9. [Google Scholar] [CrossRef]
  10. Ciriminna, R.; Albanese, L.; Meneguzzo, F.; Pagliaro, M. Hydrogen peroxide: A key chemical for today’s sustainable development. ChemSusChem 2016, 9, 3374–3381. [Google Scholar] [CrossRef]
  11. Ranganathan, S.; Sieber, V. Recent advances in the direct synthesis of hydrogen peroxide using chemical catalysis—A review. Catalysts 2018, 8, 379. [Google Scholar] [CrossRef]
  12. Menegazzo, F.; Signoretto, M.; Ghedini, E.; Strukul, G. Looking for the “dream catalyst” for hydrogen peroxide production from hydrogen and oxygen. Catalysts 2019, 9, 251. [Google Scholar] [CrossRef]
  13. Taramasso, M.; Perego, G.; Notari, B. Preparation of Porous Crystalline Synthetic Material Comprised of Silicon and Titanium Oxides. US Patent Application 4 410 501, 18 December 1983. [Google Scholar]
  14. Clerici, M.G.; Domine, M.E. Oxidation reactions catalyzed by transition-metal-substituted zeolites. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 21–126. [Google Scholar]
  15. Romano, U.; Ricci, M. The hydroxylation of phenol to hydroquinone and catechol. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 451–462. [Google Scholar]
  16. Rivetti, F.; Buzzoni, R. The greening of nylon: The ammoximation process. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 462–474. [Google Scholar]
  17. Forlin, A.; Bergamo, M.; Lindner, J. Production of propylene oxide. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 474–496. [Google Scholar]
  18. Cavani, F.; Teles, J.H. Sustainability in catalytic oxidation: An alternative approach or a structural evolution? ChemSusChem 2009, 2, 508–534. [Google Scholar] [CrossRef] [PubMed]
  19. Xiao, F.-S. Ordered mesoporous materials with improved stability and catalytic activity. Top. Catal. 2005, 35, 9–24. [Google Scholar] [CrossRef]
  20. Kholdeeva, O.A.; Mel’gunov, M.S.; Shmakov, A.N.; Trukhan, N.N.; Kriventsov, V.V.; Zaikovskii, V.I.; Malyshev, M.E.; Romannikov, V.N. A new mesoporous titanium-silicate Ti-MMM-2: A highly active and hydrothermally stable catalyst for H2O2-based selective oxidations. Catal. Today 2004, 91–92, 205–209. [Google Scholar] [CrossRef]
  21. Ivanchikova, I.D.; Kovalev, M.K.; Mel’gunov, M.S.; Shmakov, A.N.; Kholdeeva, O.A. User-friendly synthesis of highly selective and recyclable mesoporous titanium-silicate catalysts for the clean production of substituted p-benzoquinones. Catal. Sci. Technol. 2014, 4, 200–207. [Google Scholar] [CrossRef]
  22. Kholdeeva, O.A. Selective oxidations catalyzed by mesoporous metal silicates. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 127–219. [Google Scholar]
  23. Kholdeeva, O.A. Recent developments in liquid-phase selective oxidation using environmentally benign oxidants and mesoporous metal silicates. Catal. Sci. Technol. 2014, 4, 1869–1889. [Google Scholar] [CrossRef]
  24. Bisio, C.; Carniato, F.; Guidotti, M. The control of the coordination chemistry for the genesis of heterogeneous catalytically active sites in oxidation reactions. Angew. Chem. Int. Ed. 2022, 61, e202209894. [Google Scholar] [CrossRef] [PubMed]
  25. Guidotti, M.; Batonneau-Gener, I.; Gianotti, E.; Marchese, L.; Mignard, S.; Psaro, R.; Sgobba, M.; Ravasio, N. The effect of silylation on titanium-containing silica catalysts for the epoxidation of functionalised molecules. Micropor. Mesopor. Mater. 2008, 111, 39–47. [Google Scholar] [CrossRef]
  26. Guidotti, M.; Psaro, R.; Batonneau-Gener, I.; Gavrilova, E. Heterogeneous catalytic epoxidation: High limonene oxide yields by surface silylation of Ti-MCM-41. Chem. Eng. Technol. 2011, 34, 1924–1927. [Google Scholar] [CrossRef]
  27. Gallo, A.; Tiozzo, C.; Psaro, R.; Carniato, F.; Guidotti, M. Niobium metallocenes deposited onto mesoporous silica via dry impregnation as catalysts for selective epoxidation of alkenes. J. Catal. 2013, 298, 77–83. [Google Scholar] [CrossRef]
  28. Ivanchikova, I.D.; Maksimchuk, N.V.; Skobelev, I.Y.; Kaichev, V.V.; Kholdeeva, O.A. Mesoporous niobium-silicates prepared by evaporation-induced self-assembly as catalysts for selective oxidations with aqueous H2O2. J. Catal. 2015, 332, 138–148. [Google Scholar] [CrossRef]
  29. Kholdeeva, O.A.; Ivanchikova, I.D.; Maksimchuk, N.V.; Skobelev, I.Y. H2O2-based selective oxidations: Nb(V) versus Ti(IV). Catal. Today 2019, 333, 63–70. [Google Scholar] [CrossRef]
  30. Thornburg, N.E.; Thompson, A.B.; Notestein, J.M. Periodic trends in highly dispersed groups IV and V supported metal oxide catalysts for alkene epoxidation with H2O2. ACS Catal. 2015, 5, 5077–5088. [Google Scholar] [CrossRef]
  31. Hill, C.L.; Prosser-McCartha, C.M. Homogeneous catalysis by transition metal oxygen anion clusters. Coord. Chem. Rev. 1995, 143, 407–455. [Google Scholar] [CrossRef]
  32. Neumann, R. Polyoxometalates as catalysts for oxidation with hydrogen peroxide and molecular oxygen. In Transition Metals for Organic Synthesis, 2nd ed.; Beller, M., Bolm, C., Eds.; Wiley-VCH: Weinheim, Germany, 2004; Volume 2, pp. 415–426. [Google Scholar]
  33. Hill, C.L. Polyoxometalates: Reactivity. In Comprehensive Coordination Chemistry II; Wedd, A.G., Ed.; Elsevier Science: New York, NY, USA, 2004; Volume 4, pp. 679–759. [Google Scholar]
  34. Pope, M.T. Polyoxo anions: Synthesis and structure. In Comprehensive Coordination Chemistry II; Wedd, A.G., Ed.; Elsevier Science: New York, NY, USA, 2004; Volume 4, pp. 635–678. [Google Scholar]
  35. Mizuno, N.; Kamata, K.; Uchida, S.; Yamaguchi, K. Liquid-phase oxidations with hydrogen peroxide and molecular oxygen catalyzed by polyoxometalate-based compounds. In Modern Heterogeneous Oxidation Catalysis: Design, Reactions and Characterization; Mizuno, N., Ed.; WILEY-VCH Verlag GmbH & Co., KGaA: Weinheim, Germany, 2009; pp. 185–216. [Google Scholar]
  36. Neumann, R. Activation of molecular oxygen, polyoxometalates, and liquid-phase catalytic oxidation. Inorg. Chem. 2010, 49, 3594–3601. [Google Scholar] [CrossRef]
  37. Mizuno, N.; Yamaguchi, K.; Kamata, K. Epoxidation of olefins with hydrogen peroxide catalyzed by polyoxometalates. Coord. Chem. Rev. 2005, 249, 1944–1956. [Google Scholar] [CrossRef]
  38. Hill, C.L.; Kholdeeva, O.A. Selective liquid phase oxidations in the presence of supported polyoxometalates. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 263–319. [Google Scholar]
  39. Sun, M.; Zhang, J.; Putaj, P.; Caps, V.; Lefebvre, F.; Pelletier, J.; Basset, J.-M. Catalytic oxidation of light alkanes (C1−C4) by heteropoly compounds. Chem. Rev. 2014, 114, 981–1019. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, S.-S.; Yang, G.-Y. Recent advances in polyoxometalate-catalyzed reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef] [PubMed]
  41. Weinstock, I.A.; Schreiber, R.E.; Neumann, R. Dioxygen in polyoxometalate mediated reactions. Chem. Rev. 2018, 118, 2680–2717. [Google Scholar] [CrossRef]
  42. Carabineiro, H.; Villanneau, R.; Carrier, X.; Herson, P.; Lemos, F.; Ribeiro, F.R.; Proust, A.; Che, M. Zirconium-substituted isopolytungstates: Structural models for zirconia-supported tungsten catalysts. Inorg. Chem. 2006, 45, 1915–1923. [Google Scholar] [CrossRef]
  43. Kholdeeva, O.A. Titanium-monosubstituted polyoxometalates: Relation between homogeneous and heterogeneous Ti-single-site-based catalysis. Top. Catal. 2006, 40, 229–243. [Google Scholar] [CrossRef]
  44. Kholdeeva, O.A.; Maksimovskaya, R.I. Titanium- and Zirconium-monosubstituted polyoxometalates as molecular models for studying mechanisms of oxidation catalysis. J. Mol. Catal. A Chem. 2007, 262, 7–24. [Google Scholar] [CrossRef]
  45. Kholdeeva, O.A. Hydrogen peroxide activation over Ti(IV): What have we learned from studies on Ti-containing polyoxometalates? Eur. J. Inorg. Chem. 2013, 2013, 1595–1605. [Google Scholar] [CrossRef]
  46. Zhang, T.; Mazaud, L.; Chamoreau, L.-M.; Paris, C.; Guillemot, G.; Proust, A. Unveiling the active surface sites in heterogeneous titanium-based silicalite epoxidation catalysts: Input of silanol-functionalized polyoxotungstates as soluble analogues. ACS Catal. 2018, 8, 2330–2342. [Google Scholar] [CrossRef]
  47. Solé-Daura, A.; Zhang, T.; Fouilloux, H.; Robert, C.; Thomas, C.M.; Chamoreau, L.-M.; Carbó, J.J.; Proust, A.; Guillemot, G.; Poblet, J.M. Catalyst design for alkene epoxidation by molecular analogues of heterogeneous titanium-silicalite catalysts. ACS Catal. 2020, 10, 4737–4750. [Google Scholar] [CrossRef]
  48. Pope, M.T.; Muüller, A. (Eds.) Polyoxometalate Chemistry: From Topology via Self-Assembly to Applications; Kluwer: Dordrecht, The Netherlands, 2001. [Google Scholar]
  49. Kozhevnikov, I.V. Catalysis by Polyoxometalates; Wiley: Chichester, UK, 2002. [Google Scholar]
  50. Hill, C.L. (Ed.) Special Issue on Polyoxometalates in Catalysis. J. Mol. Catal. A Chem. 2007, 262, 1–242. [Google Scholar]
  51. Weinstock, I.A. (Ed.) Thematic issue on Frontiers in Metal Oxide Cluster Science. Isr. J. Chem. 2011, 51, 169–302. [Google Scholar]
  52. Cronin, L.; Muüller, A. (Eds.) Themed collection on Polyoxometalate Cluster Science. Chem. Soc. Rev. 2012, 41, 7325–7648. [Google Scholar]
  53. Carraro, M.; Fiorani, G.; Sartorel, A.; Bonchio, M. Polyoxometalates catalysts for sustainable oxidations and energy applications. In Handbook of Advanced Methods and Processes in Oxidation Catalysis; Duprez, D., Cavani, F., Eds.; Imperial College Press: London, UK, 2014; pp. 586–630. [Google Scholar]
  54. Wu, L. Organically encapsulated polyoxometalate catalysts: Supramolecular composition and synergistic catalysis. In Encapsulated Catalysts; Sadjadi, S., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–33. [Google Scholar]
  55. Kholdeeva, O.A. Recent progress in selective oxidations with hydrogen peroxide catalyzed by polyoxometalates. In Frontiers of Green Catalytic Selective Oxidations. Green Chemistry and Sustainable Technology; Bryliakov, K.P., Ed.; Springer: Singapore, 2019; pp. 61–91. [Google Scholar]
  56. Enferadi-Kerenkan, A.; Do, T.-O.; Kaliaguine, S. Heterogeneous catalysis by tungsten-based heteropoly compounds. Catal. Sci. Technol. 2018, 8, 2257–2284. [Google Scholar] [CrossRef]
  57. Evtushok, V.Y.; Lopatkin, V.A.; Podyacheva, O.Y.; Kholdeeva, O.A. Immobilization of polyoxometalates on carbon nanotubes: Tuning catalyst activity, selectivity and stability in H2O2-based oxidations. Catalysts 2022, 12, 472. [Google Scholar] [CrossRef]
  58. Sheldon, R.A.; Wallau, M.; Arends, I.W.C.E.; Schuchardt, U. Heterogeneous catalysts for liquid-phase oxidations: Philosophers’ stones or Trojan horses? Acc. Chem. Res. 1998, 31, 485–493. [Google Scholar] [CrossRef]
  59. Izumi, Y. Recent advances in immobilization of heteropolyacids. Res. Chem. Intermediat. 1998, 24, 461–471. [Google Scholar] [CrossRef]
  60. Vazylyev, M.; Sloboda-Rozner, D.; Haimov, A.; Maayan, G.; Neumann, R. Strategies for oxidation catalyzed by polyoxometalates at the interface of homogeneous and heterogeneous catalysis. Top. Catal. 2005, 34, 93–99. [Google Scholar] [CrossRef]
  61. Mizuno, N.; Kamata, K.; Yamaguchi, K. Liquid-phase oxidations catalyzed by polyoxometalates. In Surface and Nanomolecular Catalysis; Richards, R., Ed.; CRC Press LLC: Boca Raton, FL, USA, 2006; pp. 463–492. [Google Scholar]
  62. Proust, A.; Thouvenot, R.; Gouzerh, P. Functionalization of polyoxometalates: Towards advanced applications in catalysis and materials science. Chem. Commun. 2008, 2008, 1837–1852. [Google Scholar] [CrossRef]
  63. Mizuno, N.; Kamata, K.; Yamaguchi, K. Green oxidation reactions by polyoxometalate-based catalysts: From molecular to solid catalysts. Top. Catal. 2010, 53, 876–893. [Google Scholar] [CrossRef]
  64. Dolbecq, A.; Dumas, E.; Mayer, C.R.; Mialane, P. Hybrid organic−inorganic polyoxometalate compounds: From structural diversity to applications. Chem. Rev. 2010, 110, 6009–6048. [Google Scholar] [CrossRef]
  65. Kholdeeva, O.A.; Maksimchuk, N.V.; Maksimov, G.M. Polyoxometalate-based heterogeneous catalysts for liquid phase selective oxidations: Comparison of different strategies. Catal. Today 2010, 157, 107–113. [Google Scholar] [CrossRef]
  66. Proust, A.; Matt, B.; Villanneau, R.; Guillemot, G.; Gouzerh, P.; Izzet, G. Functionalization and post-functionalization: A step towards polyoxometalate-based materials. Chem. Soc. Rev. 2012, 41, 7605–7622. [Google Scholar] [CrossRef]
  67. Song, Y.F.; Tsunashima, R. Recent advances on polyoxometalate-based molecular and composite materials. Chem. Soc. Rev. 2012, 41, 7384–7402. [Google Scholar] [CrossRef]
  68. Zhou, Y.; Chen, G.; Long, Z.; Wang, J. Recent advances in polyoxometalate-based heterogeneous catalytic materials for liquid-phase organic transformations. RSC Adv. 2014, 4, 42092–42113. [Google Scholar] [CrossRef]
  69. Rafiee, E.; Eavani, S. Heterogenization of heteropoly compounds: A review of their structure and synthesis. RSC Adv. 2016, 6, 46433–46466. [Google Scholar] [CrossRef]
  70. Cherevan, A.S.; Nandan, S.P.; Roger, I.; Liu, R.; Streb, C.; Eder, D. Polyoxometalates on functional substrates: Concepts, synergies, and future perspectives. Adv. Sci. 2020, 7, 1903511. [Google Scholar] [CrossRef] [PubMed]
  71. Rhule, J.T.; Neiwert, W.A.; Hardcastle, K.I.; Do, B.T.; Hill, C.L. Ag5PMo10V2O40, a heterogeneous catalyst for air-based selective oxidation at ambient temperature. J. Am. Chem. Soc. 2001, 123, 12101–12102. [Google Scholar] [CrossRef] [PubMed]
  72. Okuhara, T. Water-tolerant solid acid catalysts. Chem. Rev. 2002, 102, 3641–3666. [Google Scholar] [CrossRef] [PubMed]
  73. Plault, L.; Hauseler, A.; Nlate, S.; Astruc, D.; Ruiz, J.; Gatard, S.; Neumann, R. Synthesis of dendritic polyoxometalate complexes assembled by ionic bonding and their function as recoverable and reusable oxidation catalysts. Angew. Chem. Int. Ed. 2004, 43, 2924–2928. [Google Scholar] [CrossRef]
  74. Mizuno, N.; Uchida, S.; Kamata, K.; Ishimoto, R.; Nojima, S.; Yonehara, K.; Sumida, Y. A flexible nonporous heterogeneous catalyst for size-selective oxidation through a bottom-up approach. Angew. Chem. Int. Ed. 2010, 49, 9972–9976. [Google Scholar] [CrossRef]
  75. Mirante, F.; Dias, L.; Silva, M.; Ribeiro, S.O.; Corvo, M.C.; de Castro, B.; Granadeiro, C.M.; Balula, S.S. Efficient heterogeneous polyoxometalate-hybrid catalysts for the oxidative desulfurization of fuels. Catal. Commun. 2018, 104, 1–8. [Google Scholar] [CrossRef]
  76. Kholdeeva, O.A.; Vanina, M.P.; Timofeeva, M.N.; Maksimovskaya, R.I.; Trubitsina, T.A.; Melgunov, M.S.; Burgina, E.B.; Mrowiec-Bialon, J.; Jarzebski, A.B.; Hill, C.L. Co-containing polyoxometalate-based heterogeneous catalysts for the selective aerobic oxidation of aldehydes under ambient conditions. J. Catal. 2004, 226, 363–371. [Google Scholar] [CrossRef]
  77. Maksimchuk, N.V.; Melgunov, M.S.; Mrowiec-Białoń, J.; Jarzębski, A.B.; Kholdeeva, O.A. H2O2-based allylic oxidation of α-pinene over different single site catalysts. J Catal. 2005, 235, 175–183. [Google Scholar] [CrossRef]
  78. Dufaud, V.; Lefebvre, F.; Niccolai, G.P.; Aouine, M. New insights into the encapsulation and stabilization of heteropolyacids inside the pore walls of mesostructured silica materials. J. Mater. Chem. 2009, 19, 1142–1150. [Google Scholar] [CrossRef]
  79. Sousa, J.L.C.; Santos, I.C.M.S.; Simões, M.M.Q.; Cavaleiro, J.A.S.; Nogueira, H.I.S.; Cavaleiro, A.M.V. Iron(III)-substituted polyoxotungstates immobilized on silica nanoparticles: Novel oxidative heterogeneous catalysts. Catal. Commun. 2011, 12, 459–463. [Google Scholar] [CrossRef]
  80. Jalkh, C.; Ghazaly, C.; El-Rassy, H. Entrapment vs. immobilization of polyoxometalates into silica and titania aerogels: Application in heterogeneous oxidation catalysis. Mater. Chem. Phys. 2020, 252, 123296. [Google Scholar] [CrossRef]
  81. Yun, S.K.; Pinnavaia, T.J. Layered double hydroxides intercalated by polyoxometalate anions with Keggin (α-H2W12O406−), Dawson (α-P2W18O626−), and Finke (Co4(H2O)2(PW9O34)210−) structures. Inorg. Chem. 1996, 35, 6853–6860. [Google Scholar] [CrossRef]
  82. Jana, S.K.; Kubota, Y.; Tatsumi, T. Cobalt-substituted polyoxometalate pillared hydrotalcite: Synthesis and catalysis in liquid-phase oxidation of cyclohexanol with molecular oxygen. J. Catal. 2008, 255, 40–47. [Google Scholar] [CrossRef]
  83. Omwoma, S.; Chen, W.; Tsunashima, R.; Song, Y.F. Recent advances on polyoxometalates intercalated layered double hydroxides: From synthetic approaches to functional material applications. Coord. Chem. Rev. 2014, 258–259, 58–71. [Google Scholar] [CrossRef]
  84. Li, T.; Miras, H.N.; Song, Y.-F. Polyoxometalate (POM)-layered double hydroxides (LDH) composite materials: Design and catalytic applications. Catalysts 2017, 7, 260. [Google Scholar] [CrossRef]
  85. Neumann, R.; Miller, H.J. Alkene oxidation in water using hydrophobic silica particles derivatized with polyoxometalates as catalysts. J. Chem. Soc. Chem. Commun. 1995, 1995, 2277–2278. [Google Scholar] [CrossRef]
  86. Kamada, M.; Kominami, H.; Kera, Y. Deposition and interaction of phosphododecatungstate on a silica gel surface modified with a silane coupling agent having anilino groups. J. Colloid Interface Sci. 1996, 182, 297–300. [Google Scholar] [CrossRef]
  87. Okun, N.M.; Anderson, T.M.; Hill, C.L. [(FeIII(OH2)2)3(A-α-PW9O34)2]9− on cationic silica nanoparticles, a new type of material and efficient heterogeneous catalyst for aerobic oxidations. J. Am. Chem. Soc. 2003, 125, 3194–3195. [Google Scholar] [CrossRef] [PubMed]
  88. Kasai, J.; Nakagawa, Y.; Uchida, S.; Yamaguchi, K.; Mizuno, N. [γ-1,2-H2SiV2W10O40] immobilized on surface-modified SiO2 as a heterogeneous catalyst for liquid-phase oxidation with H2O2. Chem. Eur. J. 2006, 12, 4176–4184. [Google Scholar] [CrossRef] [PubMed]
  89. Bordoloi, A.; Lefebvre, F.; Halligudi, S.B. Selective oxidation of anthracene using inorganic–organic hybrid materials based on molybdovanadophosphoric acids. J. Catal. 2007, 247, 166–175. [Google Scholar] [CrossRef]
  90. Yamaguchi, K.; Yoshida, C.; Uchida, S.; Mizuno, N. Peroxotungstate immobilized on ionic liquid-modified silica as a heterogeneous epoxidation catalyst with hydrogen peroxide. J. Am. Chem. Soc. 2005, 127, 530–531. [Google Scholar] [CrossRef]
  91. Kato, C.N.; Tanabe, A.; Negishi, S.; Goto, K.; Nomiya, K. An efficient PMo11VVO404−/silica material having cationic ammonium moiety: Synthesis, characterization, and catalytic performance for oxidation of alcohols with dioxygen. Chem. Lett. 2005, 34, 238–239. [Google Scholar] [CrossRef]
  92. Maksimchuk, N.V.; Melgunov, M.S.; Chesalov, Y.A.; Mrowiec-Białoń, J.; Jarzebski, A.B.; Kholdeeva, O.A. Aerobic oxidations of α-pinene over cobalt-substituted polyoxometalate supported on amino-modified mesoporous silicates. J. Catal. 2007, 246, 241–248. [Google Scholar] [CrossRef]
  93. Inumaru, K.; Ishihara, T.; Kamiya, Y.; Okuhara, T.; Yamanaka, S. Water-tolerant, highly active solid acid catalysts composed of the Keggin-type polyoxometalate H3PW12O40 immobilized in hydrophobic nanospaces of organomodified mesoporous silica. Angew. Chem. Int. Ed. 2007, 46, 7625–7676. [Google Scholar] [CrossRef]
  94. Chen, L.; Zhu, K.; Bi, L.-H.; Suchopar, A.; Reicke, M.; Mathys, G.; Jaensch, H.; Kortz, U.; Richards, R.M. Solvent-free aerobic oxidation of n-alkane by iron(iii)-substituted polyoxotungstates immobilized on SBA-15. Inorg. Chem. 2007, 46, 8457–8459. [Google Scholar] [CrossRef]
  95. Panchenko, V.N.; Borbáth, I.; Timofeeva, M.N.; Góbölös, S. Amine-modified silica NH2–(CH2)x–SiO2 (x = 0, 2, 3) as support for cobalt-substituted polyoxometalate TBA4HPW11CoO39: Effect of the nature of the support on the oxidation activity. J. Mol. Catal. A Chem. 2010, 319, 119–125. [Google Scholar] [CrossRef]
  96. Kovalchuk, T.V.; Sfihi, H.; Zaitsev, V.N.; Fraissard, J. Preparation and characterization of catalysts based on oniumsilica-immobilized Keggin acids. Catal. Today 2011, 169, 138–149. [Google Scholar] [CrossRef]
  97. Balula, S.S.; Cunha-Silva, L.; Santos, I.C.M.S.; Estrada, A.C.; Fernandes, A.C.; Cavaleiro, J.A.S.; Pires, J.; Freirea, C.; Cavaleiro, A.M.V. Mono-substituted silicotungstates as active catalysts for sustainable oxidations: Homo- and heterogeneous performance. New J. Chem. 2013, 37, 2341–2350. [Google Scholar] [CrossRef]
  98. Mirante, F.; Ribeiro, S.O.; de Castro, B.; Granadeiro, C.M.; Balula, S.S. Sustainable desulfurization processes catalyzed by titanium-polyoxometalate@TM-SBA-15. Top. Catal. 2017, 60, 1140–1150. [Google Scholar] [CrossRef]
  99. Ribeiro, S.O.; Granadeiro, C.M.; Almeida, P.L.; Pires, J.; Valenca, R.; Campos-Martin, J.M.; Ribeiro, J.C.; de Castro, B.; Balula, S.S. Effective zinc-substituted Keggin composite to catalyze the removal of sulfur from real diesels under a solvent-free system. Ind. Eng. Chem. Res. 2019, 58, 18540–18549. [Google Scholar] [CrossRef]
  100. Johnson, B.J.S.; Stein, A. Surface modification of mesoporous, macroporous, and amorphous silica with catalytically active polyoxometalate clusters. Inorg. Chem. 2001, 40, 801–808. [Google Scholar] [CrossRef]
  101. Yang, Y.; Guo, Y.; Hu, C.; Wang, Y.; Wang, E. Preparation of surface modifications of mesoporous titania with monosubstituted Keggin units and their catalytic performance for organochlorine pesticide and dyes under UV irradiation. Appl. Catal. A General 2004, 273, 201–210. [Google Scholar] [CrossRef]
  102. Errington, R.J.; Petkar, S.S.; Horrocks, B.R.; Houlton, A.; Lie, L.H.; Patole, S.N. Covalent immobilization of a TiW5 polyoxometalate on derivatized silicon surfaces. Angew. Chem. Int. Ed. 2005, 44, 1254–1257. [Google Scholar] [CrossRef]
  103. Zhou, Y.; Bao, R.; Yue, B.; Gu, M.; Pei, S.; He, H. Synthesis, characterization and catalytic application of SBA-15 immobilized rare earth metal sandwiched polyoxometalates. J. Mol. Catal. A Chemical 2007, 270, 50–55. [Google Scholar] [CrossRef]
  104. Joo, N.; Renaudineau, S.; Delapierre, G.; Bidan, G.; Chamoreau, L.-M.; Thouvenot, R.; Gouzerh, P.; Proust, A. Organosilyl/-germyl polyoxotungstate hybrids for covalent grafting onto silicon surfaces: Towards molecular memories. Chem. Eur. J. 2010, 16, 5043–5051. [Google Scholar] [CrossRef]
  105. Luo, X.; Yang, C. Photochromic ordered mesoporous hybrid materials based on covalently grafted polyoxometalates. Phys. Chem. Chem. Phys. 2011, 13, 7892–7902. [Google Scholar] [CrossRef] [PubMed]
  106. Du, Y.; Zhou, L.; Guo, Z.; Du, X.; Lei, J. Preparation of ordered meso/macroporous HPW/titania–silica catalyst for efficient oxidative desulfurization of model fuel. J. Porous Mater. 2019, 26, 1069–1077. [Google Scholar] [CrossRef]
  107. Yan, X.-M.; Mei, P.; Xiong, L.; Gao, L.; Yang, Q.; Gong, L. Mesoporous titania–silica–polyoxometalate nanocomposite materials for catalytic oxidation desulfurization of fuel oil. Catal. Sci. Technol. 2013, 3, 1985–1992. [Google Scholar] [CrossRef]
  108. Granadeiro, C.M.; Ferreira, R.A.S.; Soares-Santos, P.C.R.; Carlos, L.D.; Trindadea, T.; Nogueira, H.I.S. Lanthanopolyoxotungstates in silica nanoparticles: Multi-wavelength photoluminescent core/shell materials. J. Mater. Chem. 2010, 20, 3313–3318. [Google Scholar] [CrossRef]
  109. Gao, H.; Wu, X.; Sun, D.; Niu, G.; Guan, J.; Meng, X.; Liu, C.; Xia, W.; Song, X. Preparation of core–shell PW12@TiO2 micro-spheres and oxidative desulfurization performance. Dalton Trans. 2019, 48, 5749–5755. [Google Scholar] [CrossRef]
  110. Chen, Y.; Zhao, S.; Song, Y.-F. An efficient heterogeneous catalyst based on highly dispersed Na7H2LaW10O36·32H2O nanoparticles on mesoporous silica for deep desulfurization. Appl. Catal. A Gen. 2013, 466, 307–314. [Google Scholar] [CrossRef]
  111. Ribeiro, S.O.; Nogueira, L.S.; Gago, S.; Almeida, P.L.; Corvd, M.C.; de Castro, B.; Granadeiro, C.M.; Balula, S.S. Desulfurization process conciliating heterogeneous oxidation and liquid extraction: Organic solvent or centrifugation/water? Appl. Catal. A Gen. 2017, 542, 359–367. [Google Scholar] [CrossRef]
  112. Leng, Y.; Zhang, W.; Wang, J.; Jiang, P. A novel heteropolyanion-based amino-containing cross-linked ionic copolymer catalyst for epoxidation of alkenes with H2O2. Appl. Catal. A Gen. 2012, 445–446, 306–311. [Google Scholar] [CrossRef]
  113. Doherty, S.; Knight, J.G.; Ellison, J.R.; Weekes, D.; Harrington, R.W.; Hardacre, C.; Manya, H. An efficient recyclable peroxometalate-based polymer-immobilised ionic liquid phase (PIILP) catalyst for hydrogen peroxide-mediated oxidation. Green Chem. 2012, 14, 925–929. [Google Scholar] [CrossRef]
  114. Romanenko, I.; Lechner, M.; Wendler, F.; Horenz, C.; Streb, C.; Schacher, F.H. POMbranes: Polyoxometalate-functionalized block copolymer membranes for oxidation catalysis. J. Mater. Chem. A 2017, 5, 15789–15796. [Google Scholar] [CrossRef]
  115. Li, T.; Wang, Z.; Chen, W.; Miras, H.N.; Song, Y.-F. Rational design of a polyoxometalate intercalated layered double hydroxide: Highly efficient catalytic epoxidation of allylic alcohols under mild and solvent-free conditions. Chem. Eur. J. 2017, 23, 1069–1077. [Google Scholar] [CrossRef] [PubMed]
  116. Hwang, Y.K.; Férey, G.; Lee, U.H.; Chang, J.S. Liquid phase oxidation of organic compounds by metal-organic frameworks. In Liquid Phase Oxidation via Heterogeneous Catalysis: Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; pp. 371–409. [Google Scholar]
  117. Yu, R.; Kuang, X.-F.; Wua, X.-Y.; Lua, C.-Z.; Donahue, J.P. Stabilization and immobilization of polyoxometalates in porous coordination polymers through host–guest interactions. Coord. Chem. Rev. 2009, 253, 2872–2890. [Google Scholar] [CrossRef]
  118. Maksimchuk, N.V.; Kholdeeva, O.A.; Kovalenko, K.A.; Fedin, V.P. MIL-101 Supported polyoxometalates: Synthesis, characterization, and catalytic applications in selective liquid-phase oxidation. Isr. J. Chem. 2011, 51, 281–289. [Google Scholar] [CrossRef]
  119. Du, D.Y.; Qin, J.S.; Li, S.L.; Su, Z.M.; Lan, Y.Q. Recent advances in porous polyoxometalate based metal–organic framework materials. Chem. Soc. Rev. 2014, 43, 4615–4632. [Google Scholar] [CrossRef]
  120. Kholdeeva, O.A. Liquid-phase selective oxidation catalysis with metal-organic frameworks. Catal. Today 2016, 278, 22–29. [Google Scholar] [CrossRef]
  121. Freire, C.; Fernandes, D.M.; Nunes, M.; Abdelkader, V.K. POM & MOF-based electrocatalysts for energy-related reactions. Chem. Cat. Chem. 2018, 10, 1703–1730. [Google Scholar]
  122. Samaniyan, M.; Mirzaei, M.; Khajavian, R.; Eshtiagh-Hosseini, H.; Streb, C. Heterogeneous catalysis by polyoxometalates in metal−organic frameworks. ACS Catal. 2019, 9, 10174–10191. [Google Scholar] [CrossRef]
  123. Buru, C.T.; Farha, O.K. Strategies for incorporating catalytically active polyoxometalates in metal−organic frameworks for organic transformations. ACS Appl. Mater. Interfaces 2020, 12, 5345–5360. [Google Scholar] [CrossRef]
  124. Piscopo, C.G.; Granadeiro, C.M.; Balula, S.S.; Bošković, D. Metal-organic framework-based catalysts for oxidative desulfurization. ChemCatChem 2020, 12, 4721–4731. [Google Scholar] [CrossRef]
  125. Sun, J.; Abednatanzi, S.; Van Der Voort, P.; Liu, Y.-Y.; Leus, K. POM@MOF hybrids: Synthesis and applications. Catalysts 2020, 10, 578. [Google Scholar] [CrossRef]
  126. Mialane, P.; Mellot-Draznieks, C.; Gairola, P.; Duguet, M.; Benseghir, Y.; Oms, O.; Dolbecq, A. Heterogenisation of polyoxometalates and other metal-based complexes in metal–organic frameworks: From synthesis to characterisation and applications in catalysis. Chem. Soc. Rev. 2021, 50, 6152–6220. [Google Scholar] [CrossRef] [PubMed]
  127. Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A chromium terephthalate-based solid with unusually large pore volumes and surface area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  128. Maksimchuk, N.V.; Timofeeva, M.N.; Melgunov, M.S.; Shmakov, A.N.; Chesalov, Y.A.; Dybtsev, D.N.; Fedin, V.P.; Kholdeeva, O.A. Heterogeneous selective oxidation catalysts based on coordination polymer MIL-101 and transition metal substituted polyoxometalates. J. Catal. 2008, 257, 315–323. [Google Scholar] [CrossRef]
  129. Maksimchuk, N.V.; Kovalenko, K.A.; Arzumanov, S.S.; Chesalov, Y.A.; Stepanov, A.G.; Fedin, V.P.; Kholdeeva, O.A. Hybrid polyoxotungstate/MIL-101 materials: Synthesis, characterization, and catalysis of H2O2-based alkene epoxidation. Inorg. Chem. 2010, 49, 2920–2930. [Google Scholar] [CrossRef] [PubMed]
  130. Aubry, C.; Chottard, G.; Platzer, N.; Bregeault, J.-M.; Thouvenot, R.; Chauveau, F.; Huet, C.; Ledon, H. Reinvestigation of epoxidation using tungsten-based precursors and hydrogen peroxide in a biphase medium. Inorg. Chem. 1991, 30, 4409–4415. [Google Scholar] [CrossRef]
  131. Duncan, D.C.; Chambers, R.C.; Hecht, E.; Hill, C.L. Mechanism and dynamics in the H3[PW12O40]-catalyzed selective epoxidation of terminal olefins by H2O2. Formation, reactivity, and stability of {PO4[WO(O2)2] 4}3−. J. Am. Chem. Soc. 1995, 117, 681–691. [Google Scholar] [CrossRef]
  132. Bailey, A.J.; Griffith, W.P.; Parkin, B.C.; Bailey, A.J.; Griffith, W.P.; Parkin, B.C. Heteropolyperoxo- and isopolyperoxo-tungstates and-molybdates as catalysts for the oxidation of tertiary amines, alkenes and alcohols. J. Chem. Soc. Dalton Trans. 1995, 1995, 1833–1837. [Google Scholar] [CrossRef]
  133. Bregeault, J.-M. Transition-metal complexes for liquid-phase catalytic oxidation: Some aspects of industrial reactions and of emerging technologies. Dalton Trans. 2003, 2003, 3289–3302. [Google Scholar] [CrossRef]
  134. Küsgens, P.; Rose, M.; Senkovska, I.; Fröde, H.; Henschel, A.; Siegle, S.; Kaskel, S. Characterization of metal-organic frameworks by water adsorption. Micropor. Mesopor. Mater. 2009, 120, 325–330. [Google Scholar] [CrossRef]
  135. Bromberg, L.; Diao, Y.; Wu, H.; Speakman, T.A.; Hatton, S.A. Chromium(III) terephthalate metal organic framework (MIL-101): HF-free synthesis, structure, polyoxometalate composites, and catalytic properties. Chem. Mater. 2012, 24, 1664–1675. [Google Scholar] [CrossRef]
  136. Ribeiro, S.; Granadeiro, C.M.; Silva, P.; Paz, F.A.A.; de Biani, F.F.; Cunha-Silva, L.; Balula, S.S. An efficient oxidative desulfurization process using terbium-polyoxometalate@MIL-101(Cr). Catal. Sci. Technol. 2013, 3, 2404–2414. [Google Scholar] [CrossRef]
  137. Balula, S.S.; Granadeiro, C.M.; Barbosa, A.D.S.; Santos, I.C.M.S.; Cunha-Silva, L. Multifunctional catalyst based on sandwich-type polyoxotungstate and MIL-101 for liquid phase oxidations. Catal. Today 2013, 210, 142–148. [Google Scholar] [CrossRef]
  138. Granadeiro, C.M.; Barbosa, A.D.S.; Ribeiro, S.; Santos, I.C.M.S.; de Castro, B.; Cunha-Silva, L.; Balula, S.S. Oxidative catalytic versatility of a trivacant polyoxotungstate incorporated into MIL-101(Cr). Catal. Sci. Technol. 2014, 4, 1416–1425. [Google Scholar] [CrossRef]
  139. Gao, Y.; Mirante, F.; de Castro, B.; Zhao, J.; Cunha-Silva, L.; Balula, S.S. An effective hybrid heterogeneous catalyst to desulfurize diesel: Peroxotungstate@metal–organic framework. Molecules 2020, 25, 5494. [Google Scholar] [CrossRef] [PubMed]
  140. Fernandes, S.; Flores, D.; Silva, D.; Santos-Vieira, I.; Mirante, F.; Granadeiro, C.M.; Balula, S.S. Lindqvist@nanoporous MOF-based catalyst for effective desulfurization of fuels. Nanomaterials 2022, 12, 2887. [Google Scholar] [CrossRef] [PubMed]
  141. Zhao, X.; Duan, Y.; Yang, F.; Wei, W.; Xu, Y.; Hu, C. Efficient mechanochemical synthesis of polyoxometalate⊂ZIF complexes as reusable catalysts for highly selective oxidation. Inorg. Chem. 2017, 56, 14506–14512. [Google Scholar] [CrossRef]
  142. Wang, L.; Lub, J.; Wang, Y.; Wang, H.; Wang, J.; Ren, T. Preparation and characterization of novel cyclohexene-to-adipic acid catalyst with ionic liquid phosphotungstate immobilized on MIL-101 nanocages based on Cr-N coordination. J. Mol. Struct. 2023, 1271, 133973. [Google Scholar] [CrossRef]
  143. Qi, Z.; Huang, Z.; Wang, H.; Li, L.; Ye, C.; Qiu, T. In situ bridging encapsulation of a carboxyl-functionalized phosphotungstic acid ionic liquid in UiO-66: A remarkable catalyst for oxidative desulfurization. Chem. Eng. Sci. 2020, 225, 115818. [Google Scholar] [CrossRef]
  144. Wu, C.; Sun, Z.; Ye, C.; Qi, Z.; Chen, J.; Huang, Z.; Qiu, T. Encapsulation of HPW and preparation of composites rich in Zr-defects by manual grinding: Synergistic catalysis for efficient oxidative desulfurization at room temperature. Chem. Eng. J. 2023, 451, 138906. [Google Scholar] [CrossRef]
  145. Lin, Z.-J.; Zheng, H.-Q.; Chen, J.; Zhuang, W.-E.; Lin, Y.-X.; Su, J.-W.; Huang, Y.-B.; Cao, R. Encapsulation of phosphotungstic acid into metal−organic frameworks with tunable window sizes: Screening of PTA@MOF catalysts for efficient oxidative desulfurization. Inorg. Chem. 2018, 57, 13009–13019. [Google Scholar] [CrossRef]
  146. Fernandes, S.C.; Viana, A.M.; de Castro, B.; Cunha-Silva, L.; Balula, S.S. Synergistic combination of the nanoporous system of MOF-808 with a polyoxomolybdate to design an effective catalyst: Simultaneous oxidative desulfurization and denitrogenation processes. Sustain. Energy Fuels 2021, 5, 4032–4040. [Google Scholar] [CrossRef]
  147. Yang, X.-L.; Qiao, L.-M.; Dai, W.-L. Phosphotungstic acid encapsulated in metal-organic framework UiO-66: An effective catalyst for the selective oxidation of cyclopentene to glutaraldehyde. Micropor. Mesopor. Mater. 2015, 211, 73–81. [Google Scholar] [CrossRef]
  148. Peng, Y.-L.; Liu, J.; Zhang, H.-F.; Luo, D.; Li, D. A size-matched POM@MOF composite catalyst for highly efficient and recyclable ultra-deep oxidative fuel desulfurization. Inorg. Chem. Front. 2018, 5, 1563–1569. [Google Scholar] [CrossRef]
  149. 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]
  150. Zhang, X.; Zhang, Z.; Zhang, B.; Yang, X.; Chang, X.; Zhou, Z.; Wang, D.-H.; Zhang, M.-H.; Bu, X.-H. Synergistic effect of Zr-MOF on phosphomolybdic acid promotes efficient oxidative desulfurization. Appl. Catal. B Environ. 2019, 256, 117804. [Google Scholar] [CrossRef]
  151. Ma, Y.; Li, A.; Wang, C.; Ge, X. Preparation of HPW@UiO-66 catalyst with defects and its application in oxidative desulfurization. Chem. Eng. J. 2021, 404, 127062. [Google Scholar] [CrossRef]
  152. Silva, D.F.; Viana, A.M.; Santos-Vieira, I.; Balula, S.S.; Cunha-Silva, L. Ionic liquid-based polyoxometalate incorporated at ZIF-8: A sustainable catalyst to combine desulfurization and denitrogenation processes. Molecules 2022, 27, 1711. [Google Scholar] [CrossRef]
  153. Miras, H.N.; Vilá-Nadal, L.; Cronin, L. Polyoxometalate based open-frameworks (POM-OFs). Chem. Soc. Rev. 2014, 43, 5679–5699. [Google Scholar] [CrossRef]
  154. Song, J.; Luo, Z.; Britt, D.K.; Furukawa, H.; Yaghi, O.M.; Hardcastle, K.I.; Hill, C.L. A multiunit catalyst with synergistic stability and reactivity: A polyoxometalatemetal organic framework for aerobic decontamination. J. Am. Chem. Soc. 2011, 133, 6839–16846. [Google Scholar] [CrossRef]
  155. Zou, C.; Zhang, Z.; Xu, X.; Gong, Q.; Li, J.; Wu, C.-D. A multifunctional organic–inorganic hybrid structure based on MnIII−porphyrin and polyoxometalate as a highly effective dye scavenger and heterogenous catalyst. J. Am. Chem. Soc. 2012, 134, 87–90. [Google Scholar] [CrossRef]
  156. Han, Q.; He, C.; Zhao, M.; Qi, B.; Niu, J.; Duan, C. Engineering chiral polyoxometalate hybrid metal−organic frameworks for asymmetric dihydroxylation of olefins. J. Am. Chem. Soc. 2013, 135, 10186–10189. [Google Scholar] [CrossRef] [PubMed]
  157. Li, D.; Xu, Q.; Li, Y.; Qiu, Y.; Ma, P.; Niu, J.; Wang, J. A stable polyoxometalate-based metal−organic framework as highly efficient heterogeneous catalyst for oxidation of alcohols. Inorg. Chem. 2019, 58, 4945–4953. [Google Scholar] [CrossRef] [PubMed]
  158. Wang, Q.; Xu, B.; Wang, Y.; Wang, H.; Hu, X.; Ma, P.; Niu, J.; Wang, J. Polyoxometalate-incorporated framework as a heterogeneous catalyst for selective oxidation of C−H bonds of alkylbenzenes. Inorg. Chem. 2021, 60, 7753–7761. [Google Scholar] [CrossRef] [PubMed]
  159. Gouzerh, P.; Proust, A. Main-group element, organic, and organometallic derivatives of polyoxometalates. Chem. Rev. 1998, 98, 77–112. [Google Scholar] [CrossRef] [PubMed]
  160. Villanneau, R.; Marzouk, A.; Wang, Y.; Djamaa, A.B.; Laugel, G.; Proust, A.; Launay, F. Covalent grafting of organic−inorganic polyoxometalates hybrids onto mesoporous SBA-15: A key step for new anchored homogeneous catalysts. Inorg. Chem. 2013, 52, 2958–2965. [Google Scholar] [CrossRef]
  161. Bentaleb, F.; Makrygenni, O.; Brouri, D.; Diogo, C.C.; Mehdi, A.; Proust, A.; Launay, F.; Villanneau, R. Efficiency of polyoxometalate-based mesoporous hybrids as covalently anchored catalysts. Inorg. Chem. 2015, 54, 7607–7616. [Google Scholar] [CrossRef]
  162. Makrygenni, O.; Brouri, D.; Proust, A.; Launay, F.; Villanneau, R. Immobilization of polyoxometalate hybrid catalysts onto mesoporous silica supports using phenylene diisothiocyanate as a cross-linking agent. Micropor. Mesopor. Mater. 2019, 278, 314–321. [Google Scholar] [CrossRef]
  163. Cravena, M.; Xiao, D.; Kunstmann-Olsena, C.; Kozhevnikova, E.F.; Blanca, F.; Steiner, A.; Kozhevnikov, I.V. Oxidative desulfurization of diesel fuel catalyzed by polyoxometalate immobilized on phosphazene-functionalized silica. Appl. Catal. B Environ. 2018, 231, 82–91. [Google Scholar] [CrossRef]
  164. Liu, Y.; Zuo, P.; Wang, R.; Liu, Y.; Jiao, W. Covalent immobilization of Dawson polyoxometalates on hairy particles and its catalytic properties for the oxidation desulfurization of tetrahydrothiophene. J. Cleaner Production 2020, 274, 122774. [Google Scholar] [CrossRef]
  165. Nisar, A.; Zhuang, J.; Wang, X. Construction of amphiphilic polyoxometalate mesostructures as a highly efficient desulfurization catalyst. Adv. Mater. 2011, 23, 1130–1135. [Google Scholar] [CrossRef]
  166. Xu, J.; Zhao, S.; Ji, Y.; Song, Y.-F. Deep desulfurization by amphiphilic lanthanide-containing polyoxometalates in ionic-liquid emulsion systems under mild conditions. Chem. Eur. J. 2013, 19, 709–715. [Google Scholar] [CrossRef] [PubMed]
  167. Qi, W.; Li, H.; Wu, L. A novel, luminescent, silica-sol–gel hybrid basedon surfactant-encapsulated polyoxometalates. Adv. Mater. 2007, 19, 1983–1987. [Google Scholar] [CrossRef]
  168. Qi, W.; Wu, L. Polyoxometalate/polymer hybrid materials: Fabrication and properties. Polym. Int. 2009, 58, 1217–1225. [Google Scholar] [CrossRef]
  169. Qi, W.; Wang, Y.; Li, W.; Wu, L. Surfactant-encapsulated polyoxometalates as immobilized supramolecular catalysts for highly efficient and selective oxidation reactions. Chem. Eur. J. 2010, 16, 1068–1078. [Google Scholar] [CrossRef]
  170. Yan, Y.; Wu, L. Polyoxometalate-incorporated supramolecular self assemblies: Structures and functional properties. Isr. J. Chem. 2011, 51, 181–190. [Google Scholar] [CrossRef]
  171. Shi, L.; Wang, Y.Z.; Li, B.; Wu, L.X. Polyoxometalate complexes for oxidative kinetic resolution of secondary alcohols: Unique effects of chiral environment, immobilization and aggregation. Dalton Trans. 2014, 43, 9177–9188. [Google Scholar] [CrossRef]
  172. Nogueira, L.S.; Ribeiro, S.; Granadeiro, C.M.; Pereira, E.; Feio, G.; Cunha-Silva, L.; Balula, S.S. Novel polyoxometalate silica nano-sized spheres: Efficient catalysts for olefin oxidation and the deep desulfurization process. Dalton Trans. 2014, 43, 9518–9528. [Google Scholar] [CrossRef]
  173. Chen, X.; Yang, A.; Wang, G.; Wei, M.; Liu, N.; Li, B.; Wu, L. Reinforced catalytic oxidation of polyoxometalate@charge transfer complex by on-site heating from photothermal conversion. Chem. Eng. J. 2022, 446, 137134. [Google Scholar] [CrossRef]
  174. Izumi, Y.; Urabe, K. Catalysis of heteropoly acids entrapped in activated carbon. Chem. Lett. 1981, 10, 663–666. [Google Scholar] [CrossRef]
  175. Schwegler, M.A.; Vinke, P.; van der Eijk, M.; van Bekkum, H. Activated carbon as a support for heteropolyanion catalysts. Appl. Catal. 1992, 80, 41–57. [Google Scholar] [CrossRef]
  176. Evtushok, V.Y.; Suboch, A.N.; Podyacheva, O.Y.; Stonkus, O.A.; Zaikovskii, V.I.; Chesalov, Y.A.; Kibis, L.S.; Kholdeeva, O.A. Highly efficient catalysts based on divanadium-substituted polyoxometalate and N-doped carbon nanotubes for selective oxidation of alkylphenols. ACS Catal. 2018, 8, 1297–1307. [Google Scholar] [CrossRef]
  177. Eder, D. Carbon nanotube−inorganic hybrids. Chem. Rev. 2010, 110, 1348–1385. [Google Scholar] [CrossRef]
  178. Shearer, C.J.; Cherevan, A.; Eder, D. Application of functional hybrids incorporating carbon nanotubes or graphene. In Carbon Nanotubes and Graphene; Tanaka, K., Iijima, S., Eds.; Elsevier Ltd.: Amsterdam, The Netherlands, 2014; Chapter 16; pp. 387–433. [Google Scholar]
  179. Shearer, C.J.; Cherevan, A.; Eder, D. Application and future challenges of functional nanocarbon hybrids. Adv. Mater. 2014, 26, 2295–2318. [Google Scholar] [CrossRef] [PubMed]
  180. Ji, Y.; Huang, L.; Hu, J.; Streb, C.; Song, Y.F. Polyoxometalate-functionalized nanocarbon materials for energy conversion, energy storage and sensor systems. Energy Environ. Sci. 2015, 8, 776–789. [Google Scholar] [CrossRef]
  181. Toma, F.M.; Sartorel, A.; Iurlo, M.; Carraro, M.; Parisse, P.; Maccato, C.; Rapino, S.; Gonzalez, B.R.; Amenitsch, H.; Da Ros, T.; et al. Efficient water oxidation at carbon nanotube–polyoxometalate electrocatalytic interfaces. Nat. Chem. 2010, 2, 826–831. [Google Scholar] [CrossRef]
  182. Akter, T.; Hu, K.; Lian, K. Investigations of multilayer polyoxometalates-modified carbon nanotubes for electrochemical capacitors. Electrochim. Acta 2011, 56, 4966–4971. [Google Scholar] [CrossRef]
  183. Gao, Y.; Gao, R.; Zhang, G.; Zheng, Y.; Zhao, J. Oxidative desulfurization of model fuel in the presence of molecular oxygen over polyoxometalate based catalysts supported on carbon nanotubes. Fuel 2018, 224, 261–270. [Google Scholar] [CrossRef]
  184. Jawale, D.V.; Fossard, F.; Miserque, F.; Geertsen, V.; Teillout, A.L.; de Oliveira, P.; Mbomekalle, I.M.; Gravel, E.; Doris, E. Carbon nanotube-polyoxometalate nanohybrids as efficient electro-catalysts for the hydrogen evolution reaction. Carbon 2022, 188, 523–532. [Google Scholar] [CrossRef]
  185. Ma, D.; Liang, L.; Chen, W.; Liu, H.; Song, Y.F. Covalently tethered polyoxometalate-pyrene hybrids for noncovalent side-wall functionalization of single-walled carbon nanotubes as high-performance anode material. Adv. Funct. Mater. 2013, 23, 6100–6105. [Google Scholar] [CrossRef]
  186. Chen, W.; Huang, L.; Hu, J.; Li, T.; Jia, F.; Song, Y.F. Connecting carbon nanotubes to polyoxometalate clusters for engineering high-performance anode materials. Phys. Chem. Chem. Phys. 2014, 16, 19668–19673. [Google Scholar] [CrossRef]
  187. Kawasaki, N.; Wang, H.; Nakanishi, R.; Hamanaka, S.; Kitaura, R.; Shinohara, H.; Yokoyama, T.; Yoshikawa, H.; Awaga, K. Nanohybridization of polyoxometalate clusters and single-wall carbon nanotubes: Applications in molecular cluster batteries. Angew. Chem. Int. Ed. 2011, 50, 3471–3474. [Google Scholar] [CrossRef]
  188. Ji, Y.; Hu, J.; Huang, L.; Chen, W.; Streb, C.; Song, Y.F. Covalent attachment of Anderson-type polyoxometalates to single-walled carbon nanotubes gives enhanced performance electrodes for lithium ion batteries. Chem. Eur. J. 2015, 21, 6469–6474. [Google Scholar] [CrossRef]
  189. Chen, H.-Y.; Al-Oweini, R.; Friedl, J.; Lee, C.Y.; Li, L.; Kortz, U.; Stimming, U.; Srinivasan, M. A novel SWCNT-polyoxometalate nanohybrid material as an electrode for electrochemical supercapacitors. Nanoscale 2015, 7, 7934–7941. [Google Scholar] [CrossRef]
  190. Hu, J.; Ji, Y.; Chen, W.; Streb, C.; Song, Y.F. “Wiring” redox-active polyoxometalates to carbon nanotubes using a sonication-driven periodic functionalization strategy. Energy Environ. Sci. 2016, 9, 1095–1101. [Google Scholar] [CrossRef]
  191. Hajian, R.; Alghour, Z. Selective oxidation of alcohols with H2O2 catalyzed by zinc polyoxometalate immobilized on multi-wall carbon nanotubes modified with ionic liquid. Chin. Chem. Lett. 2017, 28, 971–975. [Google Scholar] [CrossRef]
  192. Salavati, H.; Tangestaninejad, S.; Moghadam, M.; Mirkhani, V.; Mohammadpoor-Baltork, I. Sonocatalytic epoxidation of alkenes by vanadium-containing polyphosphomolybdate immobilized on multi-wall carbon nanotubes. Ultrason. Sonochemistry 2010, 17, 453–459. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, R.; Yu, F.; Zhang, G.; Zhao, H. Performance evaluation of the carbon nanotubes supported Cs2.5H0.5PW12O40 as efficient and recoverable catalyst for the oxidative removal of dibenzothiophene. Catal. Today 2010, 150, 37–41. [Google Scholar] [CrossRef]
  194. Evtushok, V.Y.; Ivanchikova, I.D.; Podyacheva, O.Y.; Stonkus, O.A.; Suboch, A.N.; Chesalov, Y.A.; Zalomaeva, O.V.; Kholdeeva, O.A. Carbon Nanotubes modified by Venturello complex as highly efficient catalysts for alkene and thioethers oxidation with hydrogen peroxide. Front. Chem. 2019, 7, 858. [Google Scholar] [CrossRef]
  195. Evtushok, V.Y.; Podyacheva, O.Y.; Suboch, A.N.; Maksimchuk, N.V.; Stonkus, O.A.; Kibis, L.S.; Kholdeeva, O.A. H2O2-based selective oxidations by divanadium-substituted polyoxotungstate supported on nitrogen-doped carbon nanomaterials. Catal. Today 2020, 354, 196–203. [Google Scholar] [CrossRef]
  196. Evtushok, V.Y.; Ivanchikova, I.D.; Lopatkin, V.A.; Maksimchuk, N.V.; Podyacheva, O.Y.; Suboch, A.N.; Stonkus, O.A.; Kholdeeva, O.A. Heterolytic alkene oxidation with H2O2 catalyzed by Nb-substituted Lindqvist tungstates immobilized on carbon nanotubes. Catal. Sci. Technol. 2021, 11, 3198–3207. [Google Scholar] [CrossRef]
  197. Zhang, W.-H.; Shen, J.-J.; Wu, J.; Liang, X.-Y.; Xu, J.; Liu, P.; Xue, B.; Li, Y.-X. An amphiphilic graphene oxide-immobilized polyoxometalate-based ionic liquid: A highly efficient triphase transfer catalyst for the selective oxidation of alcohols with aqueous H2O2. Mol. Catal. 2017, 443, 262–269. [Google Scholar] [CrossRef]
  198. Masteri-Farahani, M.; Modarres, M. Clicked graphene oxide supported Venturello catalyst: A new hybrid nanomaterial as catalyst for the selective epoxidation of olefins. Mater. Chem. Phys. 2017, 199, 522–527. [Google Scholar] [CrossRef]
  199. Gan, M.; Yang, G.; Wang, Z.; Sui, X.; Hou, Y. Highly efficient oxidative desulfurization catalyzed by a polyoxometalate/carbonized cellulose nanofiber composite. Energy Fuels 2020, 34, 778–786. [Google Scholar] [CrossRef]
  200. Zhang, X.; Li, Y.; Li, Y.; Wang, S.; Wang, X. Polyoxometalate immobilized on graphene via click reaction for simultaneous dismutation of H2O2 and oxidation of sulfur mustard simulant. ACS Appl. Nano Mater. 2019, 2, 6971–6981. [Google Scholar] [CrossRef]
  201. Podyacheva, O.Y.; Ismagilov, Z.R. Nitrogen-doped carbon nanomaterials: To the mechanism of growth, electrical conductivity and application in catalysis. Catal. Today 2015, 249, 12–22. [Google Scholar] [CrossRef]
  202. Arrigo, R.; Schuster, M.E.; Xie, Z.; Yi, Y.; Wowsnick, G.; Sun, L.L.; Hermann, K.E.; Friedrich, M.; Kast, P.; Hävecker, M.; et al. Nature of the N-Pd interaction in nitrogen-doped carbon nanotube catalysts. ACS Catal. 2015, 5, 2740–2753. [Google Scholar] [CrossRef]
  203. Cao, Y.; Mao, S.; Li, M.; Chen, Y.; Wang, Y. Metal/porous carbon composites for heterogeneous catalysis: Old catalysts with improved performance promoted by N-doping. ACS Catal. 2017, 7, 8090–8112. [Google Scholar] [CrossRef]
  204. Skobelev, I.Y.; Evtushok, V.Y.; Kholdeeva, O.A.; Maksimchuk, N.V.; Maksimovskaya, R.I.; Ricart, J.M.; Poblet, J.M.; Carbó, J.J. Understanding the regioselectivity of aromatic hydroxylation over divanadium-substituted γ-Keggin polyoxotungstate. ACS Catal. 2017, 7, 8514–8523. [Google Scholar] [CrossRef]
  205. Wang, Y.; Kamata, K.; Ishimoto, R.; Ogasawara, Y.; Suzuki, K.; Yamaguchi, K.; Mizuno, N. Composites of [γ-H2PV2W10O40]3− and [α-SiW12O40]4− supported on Fe2O3 as heterogeneous catalysts for selective oxidation with aqueous hydrogen peroxide. Catal. Sci. Technol. 2015, 5, 2602–2611. [Google Scholar] [CrossRef]
  206. Kholdeeva, O.A.; Zalomaeva, O.V. Recent advances in transition-metal-catalyzed selective oxidation of substituted phenols and methoxyarenes with environmentally benign oxidants. Coord. Chem. Rev. 2016, 306, 302–330. [Google Scholar] [CrossRef]
  207. Maksimchuk, N.V.; Maksimov, G.M.; Evtushok, V.Y.; Ivanchikova, I.D.; Chesalov, Y.A.; Maksimovskaya, R.I.; Kholdeeva, O.A.; Sole-Daura, A.; Poblet, J.M.; Carbo, J.J. Relevance of protons in heterolytic activation of H2O2 over Nb(V): Insights from model studies on Nb-substituted polyoxometalates. ACS Catal. 2018, 8, 9722–9737. [Google Scholar] [CrossRef]
  208. Zalomaeva, O.V.; Maksimchuk, N.V.; Maksimov, G.M.; Kholdeeva, O.A. Thioether oxidation with H2O2 catalyzed by Nb-substituted polyoxotungstates: Mechanistic insights. Eur. J. Inorg. Chem. 2019, 2019, 410–416. [Google Scholar] [CrossRef]
Figure 1. Structures of some POMs used for H2O2-based selective oxidations.
Figure 1. Structures of some POMs used for H2O2-based selective oxidations.
Catalysts 13 00360 g001
Figure 2. Incorporation of Keggin POM into the cages of MIL-101(Cr).
Figure 2. Incorporation of Keggin POM into the cages of MIL-101(Cr).
Catalysts 13 00360 g002
Figure 3. Alkene epoxidation with equimolar amount of H2O2 catalyzed by PW12@MIL-101(Cr).
Figure 3. Alkene epoxidation with equimolar amount of H2O2 catalyzed by PW12@MIL-101(Cr).
Catalysts 13 00360 g003
Figure 4. Effect of H2O2 concentration on alkene epoxidation catalyzed by POM@MIL-101(Cr).
Figure 4. Effect of H2O2 concentration on alkene epoxidation catalyzed by POM@MIL-101(Cr).
Catalysts 13 00360 g004
Figure 5. Covalent grafting of phosphonate derivatives of vacant POMs onto the walls of NH2-functionalized SBA-15 silica (left) and HR-TEM micrographs of POM-CO2H@SBA-NH2 at different (ac) magnifications (right). Reprinted with permission from ref. [161]. Copyright 2015 American Chemical Society.
Figure 5. Covalent grafting of phosphonate derivatives of vacant POMs onto the walls of NH2-functionalized SBA-15 silica (left) and HR-TEM micrographs of POM-CO2H@SBA-NH2 at different (ac) magnifications (right). Reprinted with permission from ref. [161]. Copyright 2015 American Chemical Society.
Catalysts 13 00360 g005
Figure 6. Schematic representation of preparation of (a) surfactant-encapsulated POM semitube and wire assemblies and (b) oxidation of sulfides to sulfones. Reproduced from ref. [165] with permission of WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim. Copyright 2011.
Figure 6. Schematic representation of preparation of (a) surfactant-encapsulated POM semitube and wire assemblies and (b) oxidation of sulfides to sulfones. Reproduced from ref. [165] with permission of WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim. Copyright 2011.
Catalysts 13 00360 g006
Figure 7. Schematic representation of preparation of supramolecular hybrid catalyst assembled on the basis of PW12 and DOHDA and its catalytic performance in H2O2-based oxidations. Adapted from ref. [54] with permission of Elsevier Inc. Copyright 2017.
Figure 7. Schematic representation of preparation of supramolecular hybrid catalyst assembled on the basis of PW12 and DOHDA and its catalytic performance in H2O2-based oxidations. Adapted from ref. [54] with permission of Elsevier Inc. Copyright 2017.
Catalysts 13 00360 g007
Figure 8. (a) Oxidation of TMP to TMBQ with H2O2 over γ-PW10V2-based catalysts. In case of POM encapsulation within SiO2 significant POM leaching was observed. (b) Recycling of 15 wt% γ-PW10V2/N-CNTs in TMP oxidation. Adapted with permission from ref. [176]. Copyright 2018, American Chemical Society.
Figure 8. (a) Oxidation of TMP to TMBQ with H2O2 over γ-PW10V2-based catalysts. In case of POM encapsulation within SiO2 significant POM leaching was observed. (b) Recycling of 15 wt% γ-PW10V2/N-CNTs in TMP oxidation. Adapted with permission from ref. [176]. Copyright 2018, American Chemical Society.
Catalysts 13 00360 g008
Figure 9. Oxidation of methyl phenyl sulfide with H2O2 in the presence of PW4/CNTs catalysts differed in the amount of HClO4 added during the catalyst preparation. Adapted from ref. [194]. Copyright 2019, Creative Commons Attribution License (CC BY).
Figure 9. Oxidation of methyl phenyl sulfide with H2O2 in the presence of PW4/CNTs catalysts differed in the amount of HClO4 added during the catalyst preparation. Adapted from ref. [194]. Copyright 2019, Creative Commons Attribution License (CC BY).
Catalysts 13 00360 g009
Table 3. Alkene epoxidation with H2O2 over AsW9-P(O)COOH- and PW9-As(O)NH2-based catalysts.
Table 3. Alkene epoxidation with H2O2 over AsW9-P(O)COOH- and PW9-As(O)NH2-based catalysts.
CatalystSubstrate/Conversion, %Epoxide
Selectivity, %
AsW9-P(O)COOHCatalysts 13 00360 i03196˃99
AsW9-P(O)COOH@SBA-NH276˃99
PW9-As(O)NH297˃99
PW9-As(O)NH2@SBA-COOH19˃99
AsW9-P(O)COOH@MCF-NH241˃99
AsW9-P(O)COOHCatalysts 13 00360 i0327381
AsW9-P(O)COOH@SBA-NH27594
Table 4. Alkene oxidation with H2O2 in the presence of CNT-supported POMs.
Table 4. Alkene oxidation with H2O2 in the presence of CNT-supported POMs.
POMSupportH+ Added, Equiv. aSubstrateSubstrate Conv. %Epoxide
Selectivity, b %
Ref.
PW4CNTs0.2Catalysts 13 00360 i0336679[194]
PW10V2 cCNTs08279 (100)[195]
HNb(O2)W5CNTs03239 (93)[196]
HNb(O2)W5CNTs25915 (90)[196]
HNb(O2)W5N-CNTs23023 (81)[196]
PW4CNTs2Catalysts 13 00360 i0349397[194]
HNb(O2)W5CNTs29498[196]
HNb(O2)W5N-CNTs27097[196]
PW4CNTs2Catalysts 13 00360 i03510065[194]
HNb(O2)W5CNTs210055 (100)[196]
HNb(O2)W5CNTs010060 (99)[196]
PW4CNTs0Catalysts 13 00360 i0365080[194]
HNb(O2)W5CNTs28025 (77)[196]
PW4CNTs0Catalysts 13 00360 i0378585[194]
HNb(O2)W5CNTs08067 (94)[196]
a The amount of HClO4 relative to POM added during the catalyst preparation. b In parentheses, total selectivity for heterolytic oxidation products (corresponding epoxide+diol and products of their further oxidation, diepoxide and ketol). c Hot catalyst filtration test along with ICP-OES analysis of the filtrate indicated some leaching of the active vanadium species.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Maksimchuk, N.V.; Kholdeeva, O.A. H2O2-Based Selective Oxidations Catalyzed by Supported Polyoxometalates: Recent Advances. Catalysts 2023, 13, 360. https://doi.org/10.3390/catal13020360

AMA Style

Maksimchuk NV, Kholdeeva OA. H2O2-Based Selective Oxidations Catalyzed by Supported Polyoxometalates: Recent Advances. Catalysts. 2023; 13(2):360. https://doi.org/10.3390/catal13020360

Chicago/Turabian Style

Maksimchuk, Nataliya V., and Oxana A. Kholdeeva. 2023. "H2O2-Based Selective Oxidations Catalyzed by Supported Polyoxometalates: Recent Advances" Catalysts 13, no. 2: 360. https://doi.org/10.3390/catal13020360

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