Next Article in Journal
Process Optimization of Phytoantioxidant and Photoprotective Compounds from Carob Pods (Ceratonia siliqua L.) Using Ultrasonic Assisted Extraction Method
Next Article in Special Issue
Special Issue: Research on Polyoxometalate Materials
Previous Article in Journal
Anti-Inflammatory Activity of Velvet Bean (Mucuna pruriens) Substances in LPS−Stimulated RAW 264.7 Macrophages
Previous Article in Special Issue
Proton Affinity in the Chemistry of Beta-Octamolybdate: HPLC-ICP-AES, NMR and Structural Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances of Ti/Zr-Substituted Polyoxometalates: From Structural Diversity to Functional Applications

1
Center for Advanced Materials Research, Zhongyuan University of Technology, Zhengzhou 450007, China
2
MOE Key Laboratory of Cluster Science, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 102488, China
*
Authors to whom correspondence should be addressed.
Submission received: 20 November 2022 / Revised: 7 December 2022 / Accepted: 9 December 2022 / Published: 12 December 2022
(This article belongs to the Special Issue Research on Polyoxometalate Materials)

Abstract

:
Polyoxometalates (POMs), a large family of anionic polynuclear metal–oxo clusters, have received considerable research attention due to their structural versatility and diverse physicochemical properties. Lacunary POMs are key building blocks for the syntheses of functional POMs due to their highly active multidentate O-donor sites. In this review, we have addressed the structural diversities of Ti/Zr-substituted POMs based on the polymerization number of POM building blocks and the number of Ti and Zr centers. The synthetic strategies and relevant catalytic applications of some representative Ti/Zr-substituted POMs have been discussed in detail. Finally, the outlook on the future development of this area is also prospected.

1. Introduction

Polyoxometalates (POMs), as anionic metal-oxide clusters with diverse nuclearities, elemental compositions and physicochemical properties, have usually been constructed through the self-assembly of reactive oxometallate precursors in aqueous or organic reaction systems. [1,2,3,4] POMs can serve as crucial intermediates in the reaction pathway from water-soluble metal ions to insoluble metal oxides, and isolation of these molecular intermediates enable insightful elucidation on the formation mechanism and control over reaction pathways. POMs exhibit special characteristics of high negative charges, rich redox properties, good thermal stability, and readily available organic grafting [5,6], leading to wide applications in catalysis [7], magnetism [8], material science [9], electrochemistry [10], luminescence [11], etc.
As an important derivative of plenary POMs, lacunary POMs can be easily formed by removing one to several [MO6] (M = Mo, W) building blocks from prototypal architectures such as the Keggin or Wells–Dawson type POMs [12]. These lacunary POMs usually show high coordination reactivity and oxidative and thermal stability. Their high negative charge and nucleophilic oxygen-enriched surfaces render them suitable inorganic, diamagnetic, multidentate nucleophilic ligands toward the electrophilic center. Transition metals (TM) or lanthanide (Ln) cations can be easily incorporated into the defect sites of lacunary POM ligands to construct metal-substituted POMs, which can exhibit unique physicochemical properties depending on the types of incorporated metal ions [13,14,15,16,17,18,19,20]. Metal-substituted POMs (MSPs) typically possess a higher negative charge density than that of the plenary POMs due to the substitution of a high oxidation state M6+ ion (e.g., W6+, Mo6+) with a low oxidation state Mn+ ion (usually n = 1–3) [21]. To date, a wide variety of MSPs have been prepared, especially by the transition metals like manganese, iron, cobalt, nickel, copper and zinc in the fourth period and the lanthanides in the sixth period of the periodic table [22,23]. In contrast, the research on the syntheses of titanium- and zirconium-substituted POMs is still in a very early stage, which could be mainly attributed to the following two reasons: (a) the easy hydrolysis of Ti4+/Zr4+ salts in aqueous synthesis, and (b) the high tendency to formation of isolate oligomeric structures through intermolecular dehydration of terminal hydroxyls.
In this review, we have mainly focused on the structural diversities of Ti/Zr-substituted POMs according to the polymerization number of POM building blocks and the number of titanium or zirconium atoms. The representative catalytic application of Ti/Zr-substituted POMs has also been discussed. Finally, a perspective of this research area is also proposed. It is expected that this review could provide research insights into the controllable design and syntheses of Ti/Zr-substituted POMs derivatives with interesting catalytic properties.

2. The Syntheses and Structures of Ti/Zr-substituted POMs

2.1. Ti/Zr-Substituted Monomeric POMs

It is well known that titanium/zirconium-based compounds (e.g., TiO2, ZrO2) have been widely used in the fields of energy conversion, catalysis, and environmental treatment. As interesting molecular models of TiO2 and ZrO2 structures, the syntheses of titanium-/zirconium-substituted POMs could be dated back to the 1980s. In 1983, Knoth et al. reported the first case of Ti-substituted Keggin-type monomeric [TiW11PO40]5− polyoxoanion cluster (Figure 1a), which was prepared by the reaction of monovacant (Bu4N)4H3W11PO39 with titanium tetrachloride in dichloroethane solution [24]. In 2000, Kholdeeva and co-workers also reported two similar cases of [PTiW11O40]5− and [PTiW11O41]7− polyoxoanions [25]. Qu et al. reported Ti-substituted Dawson-type monomeric α2-[P2W17(TiO2)O61]8− polyoxoanion (Figure 1b), which was synthesized from vacant heteropolytungstate precursors α2-[P2W17O61]10− and Ti(SO4)2 using an aqueous solution-based synthetic approach [26]. Successively, Ti3-substituted monomeric POM has also been reported with multi-lacunary POMs α-1,2,3-[P2W15O56]12− as precursors. For instance, Nomiya et. al. successfully reported a tris-[peroxotitanium(IV)]-substituted α-Dawson monomeric [α-1,2,3-P2W15(TiO2)3O56(OH)3]9− polyoxoanion (1, Figure 1c), which are derived from {[α-1,2,3-P2W15Ti3O59(OH)3]4[µ3-Ti(OH)3]4Cl}33− (2a) in 30% aqueous hydrogen peroxide solution. Thereinto, the four bridging Ti octahedral groups of 2a were considered as the crucial roles for the synthesis of polyoxoanion 1 [27]. Subsequently, they prepared [[{Ti(OH)(ox)}2(μ-O)](α-PW11O39)]7− (Figure 1d) by using the tri-lacunary species of [A-PW9O34]9− and the anionic titanium(IV) complex as precursors with the molar ratio of 1:2 under acidic conditions. The molecular structure can be recognized as a hybrid containing one mono-lacunary POM ligand and two octahedral Ti-oxo moieties [28]. Then, a tetra-Ti-substituted di-lacunary α-Keggin monomeric [[{Ti(ox)(H2O)}4(µ-O)3](α-PW10O37)]7− polyoxoanion (Figure 1e) was also prepared, which was constructed by using the dimeric dititanium(IV)-substituted POM [(α-1,2-PW10Ti2O39)2]10− as precursor under acidic conditions [29]. Additionally, [[{Ti(H2O)3}2{Ti(H2O)2}2(μ-O)3(SO4)](PW10O37)] polyoxoanion (Figure 1f) was synthesized through the reaction of Ti(SO4)2 with [(α-1,2-PW10Ti2O38)2O2]10− and [(α-1,2,3-PW9Ti3O37)2O3]12− under strongly acidic conditions, where the tetra-titanium(IV) oxide cluster was anchored onto the binding sites of lacunary Keggin POM [30]. In 2018, An et al. reported two organic-inorganic hybrid POMs, [(H2O)4(3-Hpic)2Ln][(H2O)5(3-Hpic)2Ln][PW10Ti2O40] (Ln= Ce, Nd, Sm), where binuclear Ti-substituted Keggin-type [PW10Ti2O40]7− polyoxoanion was further modified by four Ln-3-Hpic coordinating groups [31]. Recently, Poblet et al. successfully prepared the other organic-inorganic hybrid POM, [B-α-SbW9O33(tBuSiO)3Ti(OiPr)]3−, by anchoring Ti(OiPr) moiety on the silanol-functionalized an antimony-containing trilacunary POM ligand. The resulting complex has been utilized as a catalyst for the catalytic epoxidation of alkenes [32].
Compared to Ti-substituted POMs, the exploration of Zr-containing POMs has seldom been studied. In 1985, Chauveau et al. reported a compound of Zr-containing POM, [ZrW5O19H2]2−, which was considered as the Lindqvist-type structure. However, it is still doubted about the exact structure given the presented low signal-to-noise ratio and incorrect intensity ratio of 183W NMR data [33]. Subsequently, Villanneau et al. has been unambiguously determined the structure as [W5O18Zr(H2O)3]2− by using EXFAS data [34,35]. Meanwhile, the same group also reported a similar compound with the structural formula of [{W5O18Zr(µ-OH)}2]6− (Figure 2a). In 2009, Sokolov et al. reported two mono-Zr-substituted Dawson-type monomeric polyoxoanion clusters [{(H2O)2ZrP2W17O61}]6− and [Zr(L-OOCCH(OH)CH2COO)P2W17O61]8− (Figure 2b) through the reaction of mono-vacant α2-[P2W17O61]10− ligand and Zr salt under aqueous solution conditions [36]. The latter complex exhibited chirality due to the presence of chiral L-malic acid ligand. Later on, other organic ligand-modified Zr-POM, (tpp)-Zr-(PW11O39)[TBA]5 (tpp referring to ternary porphyrin) (Figure 2c) [37] and (Pc)-Zr-(PW11O39)[TBA]5 (Pc referring to Phthalocyanine) [38], has been reported by Drain et al. in 2009 and 2013, respectively.

2.2. Ti/Zr-Substituted Dimeric POMs

In addition to the monomeric POMs, some presentative Ti/Zr-substituted dimeric POMs have also been reviewed herein in detail. In 1993, Finke et al. reported the first hexa-Ti-substituted sandwich-type dimeric silicotungstate, [A-β-Si2W18Ti6O77]14− (Figure 3a), which was prepared by the reaction of [A-β-HSiW9O34]9− with Ti(O)(C2O4)22− or Ti(O)SO4 in a regulated pH environment. The structural formulation has been lately corrected as [A-β-(SiW9O37)2(Ti-O-Ti)3]14−, implying the dimerization of two hypothetical “[A-β-SiW9(TiOH)3O37]7−” Keggin units through the linkage of Ti-O-Ti bridges [39]. In 2000, Kholdeeva and co-workers reported a Ti2-substituted dimeric POMs [(PTiW11O39)2OH]7− which was prepared using [PTiW11O40]5− subunit as reaction materials [25]. Then, Nomiya and co-workers prepared a similar hexa-Ti-substituted sandwich-type dimeric POMs except for the replacement of {A-SiW9O34} with {A-PW9O34} [40]. Subsequently, Cronin and co-workers reported a hexa-Ti-substituted tungstoarsenate, K6[Ti4(H2O)10(AsTiW8O33)2]·30H2O, where two {AsTiW8O33}fragments were used to encapsulate a {Ti4(H2O)10}16+ moiety [41]. Additionally, they also reported the first mono-Ti-substituted tungstoantimonate [TiO(SbW9O33)2]16−, which was two B-α-{SbIIIW9O33} fragments linked by five sodium cations and an unprecedented square pyramidal Ti(O)O4 group with a terminal Ti = O bond. In 2013, Kortz et al. reported two Ti-substituted phosphotungstates, [Ti8(C2O4)8P2W18O76(H2O)4]18− (Figure 3b) and [Ti6(C2O4)4P4W32O124]20− (Figure 3c). The former is the Ti8-substituted Keggin-type phosphotungstates, consisting of two {PW9} units encapsulating eight titanium centers bridged with two Ti–O–Ti bonds. The latter represents the first Ti6-substituted Dawson-type phosphotungstates, which are constructed by two di-Ti-substituted {P2W16} units connected via two Ti(C2O4) moietes [42]. In 2015, Nomiya reported the first hexa-Ti-substituted Well–Dawson phosphotungstates [{α-P2W15Ti3O60(OH)2}2(Cp*Rh)2]16− (Figure 3d). The polyoxoanion was constructed with two tri-Ti-substituted protonated Wells−Dawson subunits “[P2W15Ti3O60(OH)2]10−” bridged by the two organometallic Cp*Rh2+ groups [43]. Additionally, they also reported the first tetra-TiIV-1,2-substituted α-Keggin polyoxotungstate in aqueous solution, [α,α-P2W20Ti4O78]10− (Figure 3e). The polyoxoanion consisted of a dimeric anhydride form of two [α-1,2-PW10Ti2O40]7− Keggin units linked with two Ti–O–Ti bonds [44]. Similar structures have also been reported by Mizuno and Wang’s groups, respectively [45,46]. With continuous research, more di-Ti-substituted POMs have been further developed. In 2007, Kortz et al. reported a special di-Ti-substituted tungstodiarsenate (III) [Ti2(OH)2As2W19O67(H2O)]8− (Figure 3f), prepared by the reaction of TiOSO4 and K14[As2W19O67(H2O)] in a 2:1 molar ratio in acidic media (pH 2). The polyoxoanion was a sandwich-type structure with nominal C2v symmetry, which was constructed with two (B-α-AsIIIW9O33) Keggin moieties linked by an octahedral {WO5(H2O)} fragment and two unprecedented square-pyramidal {TiO4(OH)} groups [47]. Nomiya et al. further reported the synthesis of a novel molecular solid Brønsted acid based on the Dawson-type sandwich POM [Ti2{P2W15O54(OH2)2}2]8− (Figure 3g) [48]. Subsequently, they synthesized a similar structure using mono-lacunary Dawson precursor K10[α2-P2W17O61]·23H2O [49]. In 2015, Li’s group synthesized two isomorphic di-Ti-substituted Keggin-type phosphotungstate ([(Ti2O)(PW11O39)2]8−, Figure 3h) containing dissimilar copper under hydrothermal condition. The resulting organic–inorganic hybrid assemblies contained a rare corner-sharing double-Keggin type POM architecture in the Ti-POM species, which was further connected with the butterfly-type [CuIILo] units to form a 1-D chain and a square plane, respectively [50].
In contrast to the diverse structures of Ti-substituted POMs, Zr-substituted dimeric POMs have been far less reported. Some presentative examples are summarized below. In 2003, May et al. reported an example of a dimeric structure of mono-Zr-substituted Kegging-type POM, [Zr(PMo12O40)(PMo11O39)]6− (Figure 4a), which also represented the first crystallographic determination of the [PMo11O39]7− anion [51]. The similar Keggin-type chiral phosphotungstate and borotungstate were also reported by Liu and Xue’s groups in 2009, respectively [52,53]. In 2006, Nomiya et al. also reported the first Zr-substituted Well-Dawson phosphotungstate [Zr(α2-P2W17O61)2]16− (Figure 4b), deriving from mono-lacunary precursor [α2-P2W17O61]10− [54]. Similar mono-Zr-substituted Well-Dawson phosphotungstate POMs ([{P2W15O54(H2O)2}2Zr]12−, and [{P2W15O54(H2O)2}Zr{P2W17O61}]14−) have been further reported by Hill et al. in 2007 [55]. To increase the nuclearity of Zr centers, Kholdeeva et al. prepared three Zr2-substituted Keggin-type phosphotungstate ([{PW11O39Zr(µ-OH)}2]8−, [{PW11O39Zr(µ-OH)}2]8−, and [{PW11O39Zr}2(µ-OH)(µ-O)]9−) [56]. Then, Mizuno et al. synthesized a di-Zr-substituted Keggin-type silicotungstate [(γ-SiW10O36)2Zr2(μ-OH)2]10− (Figure 4c) [57], and Sokolov et al. reported di-Zr-substituted Dawson-type phosphotungstate [{(H2O)Zr(μ2-OH)(P2W17O61)}2]14− [58]. In 2011, Villanneau et al. reported two Zr2-containing POMs derivatives [{PW9O34{PO(R)}2}2{Zr(H2O)(μ-OH)}2]4− and [{PW9O34{PO(R)}2}2{Zr(DMF)(μ-OH)}2]4− (R = Ph, tBu), which were obtained by using [(nBu4N)3Na2[PW9O34{PO(R)}2] and ZrOCl2⋅8H2O [59]. Subsequently, the same group also reported a similar structure in 2013 [60]. In 2005, Hill et al. reported the first chiral tri-Zr-substituted Dawson-type phosphotungstate {[α-P2W15O55(H2O)]Zr3(μ3-O)(H2O)(L-tartH)[α-P2W16O59]}15− (Figure 4d) and {[α-P2W15O55(H2O)]Zr3(μ3-O)(H2O)(D-tartH)[α-P2W16O59]}15− [61]. After that, Cadot et. al. reported a tri-Zr(IV)-substituted sandwich-type Keggin POM [Zr3O(OH)2(SiW9O34)2]12−, which consists of a [Zr3O(OH)2] triangular central cluster closely embedded between two A-α-[SiW9O34]10− subunits [62]. Subsequently, three cases of isomorphic compounds were also reported by Xue, Nomiya and Yang’s groups, respectively [63,64,65]. Among these tri-Zr-substituted POMs, it is worth mentioning that Yang’s group reported the first tri-ZrIV-substituted POM [Zr3(µ2-OH)2(µ2-O)(A-α-GeW9O34)(1,4,9-α-P2W15O56)]14− (Figure 4e), where the tri-Zr centers were stabilized by mixed types of tri-lacunary POM ligands including Keggin-type [A-α-GeW9O34]10− and Dawson-type [1,4,9-α-P2W15O56]12− units [66]. In addition, a number of tetra-Zr substituted POMs have also been prepared. For instance, Pope et al. reported a Zr4-substituted phosphotungstate, [Zr4(µ3-O)2(µ2-OH)2(H2O)4(P2W16O59)2]14− (Figure 4f). Therein, the divacant lacunary {P2W16O59} ligands were derived from the plenary Wells–Dawson (α-P2W18O62) polyoxoanion [67]. Then, similar structures were reported by Hill and Li’s groups in 2005 and 2013, respectively [68,69]. Subsequently, tetra-Zr substituted Keggin-type silicotungstates [(γ-SiW10O36)2Zr4(µ4-O)(µ-OH)6]8− (Figure 4g) and five other similar structures were successively reported [70,71,72,73,74,75,76,77]. The nuclearity of Zr substitution has been further improved to six by Kortz and co-workers. they reported the first hexa-Zr-substituted dimeric tungstoarsenates [Zr6O4(OH)4(H2O)2(CH3COO)5(AsW9O33)2]11− (Figure 4h). In the polyoxoanion, the unprecedented hexa-Zr unit is perfectly located at the cavity formed by two (B-α-AsW9O33) fragments lying at an angle of about 74° with respect to each other [78].

2.3. Ti/Zr-Substituted Trimeric POMs

In contrast to the dimeric POM structures, the syntheses of Ti/Zr-substituted trimeric POMs have rarely been reported. The early reported Zr-substituted trimeric POM is Zr6O2(OH)4(H2O)3(β-SiW10O37)3]14− (Figure 5a), which was synthesized using an aqueous solution-based method by Kortz et al. in 2006. This polyoxoanion consists of three β23-SiW10O37 units linked by an unprecedented Zr6O2(OH)4(H2O)3 cluster with C1 point group symmetry [79]. The similar polyoxoanion [Zr6(μ3-O)3(OH)3(OAc)(H2O)(β-GeW10O37)3]16− was also synthesized via hydrothermal method by Yang’s group in 2019, except that {β-SiW10O37} was replaced by {β-GeW10O37} [80]. Late on, Kortz et al. reported another Zr6-substituted silicotungstate [Zr6(O2)6(OH)6(γ-SiW10O36)3]18− (Figure 5b) in 2008, where 6-Peroxo-6-Zr Crown embedded in a triangular polyoxoanion. This polyoxoanion is composed of three [γ-SiW10O36]8− units encapsulating the unprecedented [Zr6(O2)6(OH)6]6+ wheel, while it can also be considered as a cyclic assembly of three fused {Zr2(O2)2(OH)2(γ-SiW10O36)} monomers. This work belongs to the first structurally characterized Zr-peroxo POM with side-on, bridging peroxo units [81]. Additionally, Kortz et al. reported the first examples of Ti-containing trimeric polytungstates, two cyclic Ti9-containing trimeric POMs, [(α-Ti3PW9O38)3(PO4)]18− (Figure 5c) and [(α-Ti3SiW9O37OH)3(TiO3(OH2)3)]17− (Figure 5d) using the solution-based synthetic method. Both compounds were composed of three (Ti3XW9O37) units (X = P or Si) bridged with three Ti-O-Ti bonds and a capping group (tetrahedral PO4 or octahedral TiO6) [82]. Liu’s group reported two similar trimeric nine-TiIV contained tungstogermanates {K⸦[(Ge(OH)O3)(GeW9Ti3O38H2)3]}14− (Figure 5e) and {K⸦[(SO4)(GeW9Ti3O38H3)3]10−. The two compounds were obtained from the reactions between K8[γ-GeW10O36] [83] and TiO(SO4) under different pH conditions. The former polyoxoanion consisted of three tri-TiIV-substituted Keggin fragments [GeW9Ti3O38] and a GeO4 tetrahedral linker bridged with both Ti–O–Ti and Ti–O–Ge bonds, and the structure of the latter one is similar except for the replacement of SO4 with GeO4 [84]. Recently, a novel trimer compound [{Ca6(CO3)(μ3-OH)(OH2)18}(P2W15Ti3O61)3Ca(OH2)3]19− (Figure 5f) was reported that contains a hexacalcium cluster cation, one carbonate anion, and one calcium cation assembled on a trimeric tri-Ti-substituted Wells–Dawson polyoxometalates. This complex was obtained through the reaction of calcium chloride with the monomeric trititanium(IV)-substituted Wells−Dawson POM species “[P2W15Ti3O59(OH)3]9−”. During the synthesis, the [Ca6(CO3)(μ3-OH)(OH2)18]9+ cluster cation, composed of six calcium cations linked by one μ6-carbonato anion and one μ3-OH- anion, assembled with one calcium ion, a trimeric “[P2W15Ti3O59(OH)3]9−” species to form the target product. The compound is an unprecedented POM species containing an alkaline-earth-metal cluster cation, and it is also the first example of alkaline-earth-metal ions clustered around a Ti-substituted POM [85].

2.4. Ti/Zr-Substituted Tetrameric POMs

In this section, a number of Ti/Zr-substituted tetrameric POMs will be briefly introduced. The early example is a dodeca-Ti-substituted Dawson-type tetrameric, [{Ti3P2W15O57.5(OH)3}4]24− (Figure 6a), representing a supramolecular phosphotungstate reported by Kortz’s group in 2003 [86]. The polyoxoanion was composed of four lacunary [P2W15O56]12− Well–Dawson building blocks linked with terminal Ti-O bonds, resulting in a structure with Td symmetry. The {Ti12O46} core of the polyoxoanion is composed of four groups of three edge-shared, corner-linked TiO6 octahedra. Such a rare arrangement resembles one set of the four corner-shared faces of an octahedron, described as a “reversed Keggin structure”, which is very similar to the [As4Mo12O50]8− geometry reported by Sasaki and Nishikawa [87]. Apart from this Ti12 cluster, they also discovered another deca-Ti-substituted tetrameric species [{Ti3P2W15O57.5(OH)3}2{Ti2P2W16O60(OH)}2]26− containing two {Ti3P2W15} and two {Ti2P2W16} fragments, therefore resulting in a structure with C2v symmetry. Meanwhile, Nomiya et al. also reported two multi-Ti-substituted tetrameric POMs. The first one is a giant “tetrapod”-shaped dodeca-Ti-substituted Dawson-type tetrameric phosphotungstate, [(α-1,2,3-P2W15Ti3O60.5)4Cl]37− (Figure 6b), which contains the four Wells-Dawson units fused together through Ti–O–Ti bonds. The structure exhibits an approximately Td symmetry, where the four Ti3O6 facets of “P2W15Ti3” occupied four alternate facets of an octahedron, and the one Cl- ion was encapsulated in the central octahedral cavity [88]. It is noted that a similar structure [(P2W15Ti3O60.5)4(NH4)]35− was also reported in 2011, except that Cl- was replaced by NH4+ [89]. The other example belongs to a “tetrapod”-shaped Ti16-substituted Dawson-type tetrameric phosphotungstate [(α-1,2,3-P2W15Ti3O62)4{μ3-Ti(OH)3}4Cl]45− (Figure 6c). The polyoxoanion, prepared by the reaction of [P2W15O56]12− with an excess amount of TiCl4 in aqueous solution, was composed of four tri-TiIV-1,2,3-substituted α-Dawson substructures, four Ti(OH)3 bridging groups, and one encapsulated chloride ion [90]. Subsequently, Nomiya et al. also reported three similar Ti16-substituted POMs, [(α-P2W15Ti3O59(OH)3)4{μ3-Ti(H2O)3}4X]21− (X = Br, I, and NO3), except that halogen atoms and some oxygen atoms are protonated [91]. In 2004, Kortz et al. synthesized a unique cyclic octa-Ti- substituted tetrameric tungstosilicate [{β-Ti2SiW10O39}4]24− (Figure 6d) assembly under mild, one-pot reaction conditions. The polyoxoanion is composed of four {β-Ti2SiW10O39} Keggin fragments bridged with Ti–O–Ti bonds, leading to a cyclic assembly. The successful preparation of this compound provides future possibilities for preparing even larger wheel-shaped polyoxotungstates and other discrete nanomolecular objects of similar size, structure, and function as those made with polyoxometalates [92]. Then, Kortz’s group also successfully prepared a hepta-Ti substituted arsenotungstate [Ti6(TiO6)(AsW9O33)4]20− (Figure 6e) in 2014 using a simple one-pot procedure. The polyoxoanion contains a novel Ti7-core consisting of a central TiO6 octahedron surrounded by six TiO5 square pyramids, which was further capped by four trilacunary {AsIIIW9} fragments, leading to an assembly with Td point-group symmetry [93]. In 2019, Yang’s group also reported a similar structure [Ti7O6(SbW9O33)4]20− by replacing {AsW9O33} with {SbW9O33} [94]. In 2022, Yang’s group further synthesized a ring-shaped 12-Ti-substituted poly(polyoxometalate) [{K2Na(H2O)3}@{(Ti2O)2(Ti4O4)2(A-α-1,3,5-GeW9O36)2(A-α-2,3,4-GeW9O36)2}]25− (Figure 6f) under hydrothermal conditions, which represents the highest number of Ti centers in Keggin-type poly(POM) family to date. In this structure, two types of novel chiral trivacant [GeW9O36]14− (A-α-1,3,5-GeW9O36 and A-α-2,3,4-GeW9O36) fragments have been first discovered in POM chemistry, and four [GeW9O36]14− fragments are alternately connected by two Ti2O and two Ti4O4 cores to form a ring-shaped poly(POM) [95].
In addition to these Ti-substituted POMs, some Zr-substituted tetrameric POMs have also been actively investigated. For instance, a Zr4-substituted tungstoselenites, [(α-SeW9O34){Zr(H2O)}{WO(H2O)}(WO2)(SeO3){α-SeW8O31Zr(H2O)}]212− (Figure 7a), has been firstly constructed using {α-SeW9} building blocks. Such a tetrameric structure can be divided into two same subunits, each containing a dimer sandwich-type structure that consists of a well-known trivacant Keggin-type {α-SeW9O34} building blocks [96]. Then, Yang’s group continuously reported eight tetrameric Zr-substituted POMs. The first example represents a new tetra-Zr-substituted tungstophosphate, {Zr2[SbP2W4(OH)2O21][α2-PW10O38]}220− (Figure 7b) through the hydrothermal reaction of the [B-α-SbW9O33]9− building block with Zr4+ cations and PO43− anions in the presence of dimethylamine hydrochloride in NaOAc-HOAc buffer solution. The compound exhibits a toroidal structure formed by two divacant [α2-PW10O38]11− units and two [SbP2W4(OH)2O21]7− fragments linked by four Zr4+ cations. It is noted that the triangular pyramidal SbO3 in the [B-α-SbW9O33]9− precursor was replaced by tetrahedral PO4 unit in the final compound, and the pendant SbO3 derives from the dissociation of the [B-α-SbW9O33]9− precursor [97]. Then, the same group also reported a Zr(IV)-substituted tetramer polyoxotungstate, [Zr4(β-GeW10O38)2(A-α-PW9O34)2]26− (Figure 7c) [98], which was obtained in a one-pot reaction of the hexalacunary polyanion [P2W12O48]12− and trilacunary [GeW9O34]10− polyanion precursors with Zr4+ in a slightly alkali aqueous solution in the presence of borates. Subsequently, a Zr9-substituted tetrameric germanotungstate, [{Zr5(μ3-OH)4(OH)2}@{Zr2(OAc)2(α-GeW10O38)2}2]22− (Figure 7d) was constructed by two novel sandwich-type dimers [Zr2(OAc)2(α2-GeW10O38)2]18− and one unique [Zr5(μ3-OH)4(OH)2]14+ core in an approximately orthogonal fashion, showing a staggering tetrahedral polyoxoanion [80]. In addition, Yang et al. also reported a series of ring-shaped Zr8-substituted silicotungstates, [{Zr2(OH)2(α-SiW10O38)}2{Zr2(OH)2(β-SiW10O38)}2]24− (Figure 7e) (dap = 1,3-diamino-propane), [(Zr2(OH)2)2(Zr2BO(OH)4)2(β-SiW10O38)4]26− (Figure 7f) and [(Zr2BO(OH)4)2(Zr2B2O2(OH)5)2(β-SiW10O38)4]28− [99]. The latter two compounds first provided the possibility of introducing Zr–B–O linkage to the lacunary sites. Very recently, they also reported a di-Zr-substituted polyoxotungstate [ZrSb4(OH)O2(A-α-PW8O32)(A-α-PW9O34)]218− (Figure 7g) [100] and hepta-Zr-incorporated polyoxometalate [SbZr7O6(OH)4(B-α-GeW9O34)2(B-α-GeW11O39)2]21− (Figure 7h) [101], these works greatly enriched the family of Zr-substituted POMs.

2.5. Ti/Zr-Substituted Multimeric POMs

Compared to those mono-, di-, tri- and tetrameric POM structures, there are very few reports on the preparation of multimeric Ti/Zr-substituted POMs. To date, only two related compounds have been reported. The first example belongs to an octa-Ti-substituted Dawson-type supramolecular polyoxoanion reported by Kortz’s group in 2003 [86]. However, the polyoxoanion was described by the preliminary and incomplete formula “Ti8P12W84” or “(Ti2P2W15)2(Ti2P2W16)2(P2W11)2” (Figure 8a) due to the poor quality of the crystallographic data. The second representative example is the gigantic Zr24-cluster-substituted Keggin-type germanotungstates [Zr24O22(OH)10(H2O)2(W2O10H)2(GeW9O34)4(GeW8O31)2]32− (Figure 8b) reported by Yang’s group in 2014 [102]. The polyoxoanion was successfully synthesized under hydrothermal conditions, which contains the largest [Zr24O22(OH)10(H2O)2] cluster among all reported Zr-based poly(polyoxometalate)s to date. Detailed structural analyses of this complex showed that the centrosymmetric Zr24-cluster-based hexamer contained two symmetry-related [Zr12O11(OH)5(H2O)(W2O10H)(GeW9O34)2(GeW8O31)]16− trimers linked via six μ3-oxo bridges, which were further encapsulated by different POM fragments including B-α-GeW9O34, B-α-GeW8O31, and W2O10. Catalytic experiments also showed that this compound worked as a good catalyst for the oxygenation of thioethers to sulfoxides/sulfones in the presence of H2O2, which could be attributed to the unique redox property of oxygen-enriched polyoxotungstate fragments as well as the Lewis acidity of the Zr24 cluster. Although the synthesis of multimeric POMs is rather difficult, these pioneering works provide some insights and future direction for the exploration of this specific research area.

3. The Applications of Representative Ti/Zr-Substituted POMs

It is well known that transition-metal clusters are a unique area in inorganic chemistry, considering their vital contribution to the blossom of modern chemistry as well as their potential application as structural models for various industrial and biological catalytic processes [103,104,105,106]. To date, the Ti/Zr-substituted POMs have been widely investigated as catalysts for the oxidation of organic substrates. For example, Poblet et al. investigated the oxidation of alkenes by H2O2 catalyzed using Ti(IV)-containing POMs, which were models of Ti single-site catalysts at the DFT computational level. The catalytic mechanism of the C2H4 epoxidation with H2O2 mediated by [PTi(OH)W11O39]4− and [Ti2(OH)2As2W19O67(H2O)]8− can be processed by following two main steps (Scheme 1): (i) H2O2 was activated to form the titanium-peroxo or -hydroperxo intermediate, and (ii) the reactive intermediate further attack alkene to form the epoxide and water [107]. Kholdeeva and co-workers also reported several cases of various organic catalytic studies with Ti-containing POMs. For example, they investigated the mechanism of thioether oxidation of (Bu4N)7{[PW11O39Ti]2OH} dimeric heteropolytungstate in 2000 [108]. In 2004, they also reported a protonated titanium peroxo complex [Bu4N]4[HPTi(O2)W11O39] and found that protonated titanium peroxo complex has a higher redox potential, which can improve the catalytic performance [109]. In the same year, Ti(IV)-monosubstituted Keggin-type POMs were reported to exhibit excellent catalytic oxidation properties with H2O2 [110,111,112]. Subsequently, Kholdeeva et al. also reported the epoxidation of a range of alkenes easily proceeds with aqueous H2O2 as oxidant and the dititanium-containing 19-tungstodiarsenate (III) as catalyst [113]. In 2012, the same group also published two works on alkene oxidation by Ti-containing POMs. They claimed that the energy barrier for the heterolytic oxygen transfer from the reactive Ti hydroperoxo intermediate was significantly reduced by the protonated Ti-containing POM, as revealed by the kinetic and DFT studies, thereby greatly enhancing the activity and selectivity of alkene oxidation [114,115]. Subsequently, Poblet and Guillemot reported the alkene epoxidation catalyzed by the hybrid [B-α-SbW9O33(tBuSiO)3Ti(OiPr)]3−, [PW9O34(tBuSiO)3Ti(OiPr)]3−, and Ti-complexe of silanol functionalized POMs [SbW9O33(RSiO)3Ti(OiPr)]3−, respectively [32,116,117]. Based on the research of alkene epoxidation catalysis, Kholdeeva et al. also revealed the mechanism of thioether oxidation of Ti-substituted POMs by kinetic modeling and DFT calculations. Two possible models regarding the active group were proposed: (1) the active group is the terminal Ti−OH group for the mononuclear, and (2) the active group is the bridging Ti2(μ-OH) moiety for the multinuclear [118]. Subsequently, Yang’s group also reported two cases of the catalytic oxidation of thioethers using Ti7- and Ti12-substituted POMs, respectively, which both exhibited good catalytic properties [94,95]. In addition, Li’s group investigated the photocatalytic degradation of MB with Ti2-substituted POM units under UV irradiation, which proved that Ti-substituted Keggin-type POMs showed better photocatalytic activities than that of typical Keggin-type POMs [50]. Some Ti-substituted POMs have also been reported with good electrocatalytic properties [119].
Compared with the applications of Ti-substituted POMs, a part of Zr-substituted POMs also exhibited good catalytic oxidation of thioether, electrocatalytic and nonlinear optical properties, etc. For example, Yang’s groups reported a few works on the oxidation of sulfide using Zr2-, Zr4-, Zr7- and Zr8-substituted POMs, respectively [76,99,100,101]. These compounds showed remarkable heterogeneous catalysts for the catalytic oxidation of sulfides into the corresponding sulfones with H2O2. The same groups also reported several Chiral Zr-substituted POMs which have excellent nonlinear optical properties [75,77]. In addition, transition-metal-substituted POMs are air- and water-stable Lewis acids that were often used in organic reactions [120], and in the last two decades, most Zr-substituted POMs also were reported to be used as optimal Lewis acid catalysts to hydrolyze the O = C-NH- bonds in proteins or peptides (Scheme 2), leading to the formation of amino acids [121,122,123,124,125,126]. These works implied the potential applications of Zr-substituted POMs in biological systems.

4. Conclusions and Perspectives

In summary, this review has mainly addressed the development of Ti/Zr-substituted POMs with an emphasis on structural diversity, synthetic approaches, and potential catalytic applications. According to the overview of these reported Ti/Zr-substituted POMs, we can conclude that (1) the solution-based synthetic approach is an effective method for the syntheses of Ti/Zr-substituted POMs; (2) interesting and attracting POM structures could be often obtained via the hydrothermal synthetic strategy. These successful synthetic strategies and the persistent and dedicated efforts of chemists over the past decades have greatly contributed to the vast and beautiful array of Ti/Zr-substituted POMs. However, there are still bottlenecks in Ti/Zr-substituted POM chemistry with respect to the controllable assembly of target POM structures, the exploration of novel synthetic approaches, as well as the insightful understanding of catalytic mechanisms using Ti/Zr-substituted POM catalysts. Therefore, we believe that the development of other synthetic methods, for instance, the mixed solvent diffusion method, ionothermal approach, templated modular assembly method or the combination with existing solution/hydrothermal approaches, would provide new blood to the synthetic chemistry of Ti/Zr-substituted POMs. Moreover, the ground-breaking exploration of new catalytic functionalities of these Ti/Zr-substituted POMs should also be strengthened in the future. Finally, we hope this critical review could provide research insights into the controllable design and syntheses of Ti/Zr-substituted POMs derivatives and, in the meantime, attract more researchers to join the research community of POM Chemistry or the related interdisciplinary research areas.

Author Contributions

Literature research, Z.N.; writing—original draft preparation, Z.N.; writing—review and editing, H.L. and G.Y.; funding acquisition, Z.N., H.L., and G.Y. All authors have read and agreed to the published version of the manuscript.

Funding

Project supported by Key Projects of Science and Technology of Henan Province (No. 212102210442), the National Natural Science Foundation of China (Nos. 21871025 and 21831001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors would like to express their gratitude to Cambridge Crystallographic Data Centre (CCDC) for providing access to the CSD Enterprise.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cronin, L.; Müller, A. Special POM themed issue. Chem. Soc. Rev. 2012, 41, 7325–7648. [Google Scholar]
  2. Misra, A.; Kozma, K.; Streb, C.; Nyman, M. Beyond charge balance: Counter-cations in polyoxometalate chemistry. Angew. Chem. Int. Ed. 2020, 59, 596–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Liu, J.-X.; Zhang, X.-B.; Li, Y.-L.; Huang, S.-L.; Yang, G.-Y. Polyoxometalate functionalized architectures. Coord. Chem. Rev. 2020, 414, 213260–213275. [Google Scholar] [CrossRef]
  4. Pope, M.T.; Müller, A. Polyoxometalates: From Platonic Solids to Anti-Retroviral Activity; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1994. [Google Scholar]
  5. Coronado, E.; Giménez-Saiz, C.; Gómez-García, C.J. Recent advances in polyoxometalate-containing molecular conductors. Coord. Chem. Rev. 2005, 249, 1776–1796. [Google Scholar] [CrossRef]
  6. Zhang, J.; Xiao, F.P.; Hao, J.; Wei, Y.G. The chemistry of organoimido derivatives of polyoxometalates. Dalton Trans. 2012, 41, 3599–3615. [Google Scholar] [CrossRef]
  7. Wang, S.S.; Yang, G.Y. Recent advances in polyoxometalate-catalyzed reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef]
  8. Clemente-Juan, J.M.; Coronado, E.; Gaita-Ariño, A. Magnetic polyoxometalates: From molecular magnetism to molecular spintronics and quantum computing. Chem. Soc. Rev. 2012, 41, 7464–7478. [Google Scholar] [CrossRef]
  9. Long, D.L.; Burkholder, E.; Cronin, L. Polyoxometalate clusters, nanostructures and materials: From self-assembly to designer materials and devices. Chem. Soc. Rev. 2007, 36, 105–121. [Google Scholar] [CrossRef]
  10. Gupta, R.; Khan, I.; Hussain, F.; Bossoch, A.M.; Mbomekalle, I.M.; Oliveira, P.D.; Sadakane, M.; Kato, C.; Ichihashi, K.; Inoue, K.; et al. Two new sandwich-type manganese {Mn5}-substituted polyoxotungstates: Syntheses, crystal structures, electrochemistry, and magnetic properties. Inorg. Chem. 2017, 56, 8759–8767. [Google Scholar] [CrossRef]
  11. Xu, B.B.; Lu, M.; Kang, J.; Wang, D.G.; Brown, J.; Peng, Z.H. Synthesis and optical properties of conjugated polymers containing polyoxometalate clusters as side-chain pendants. Chem. Mater. 2005, 17, 2841–2851. [Google Scholar] [CrossRef]
  12. Rong, C.; Pope, M.T. Lacunary polyoxometalate anions are. Pi.-acceptor ligands. Characteriza-tion of some tungstoruthenate (II, III, IV, V) heteropolyanions and their atom-transfer reactivity. J. Am. Chem. Soc. 1992, 114, 2932–2938. [Google Scholar] [CrossRef]
  13. Li, C.; Jimbo, A.; Yamaguchi, K.; Suzuki, K. A protecting group strategy to access stable lacunary polyoxomolybdates for introducing multinuclear metal clusters. Chem. Sci. 2021, 12, 1240–1244. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, Z.-W.; Zhao, Q.; Chen, C.-A.; Sun, J.-J.; Lv, H.J.; Yang, G.-Y. Chiral {Ni6PW9} cluste-organic framework: Synthesis, structure, and properties. Inorg. Chem. 2022, 61, 7477–7483. [Google Scholar] [CrossRef] [PubMed]
  15. Qin, L.; Wang, R.J.; Xin, X.; Zhang, M.; Liu, T.F.; Lv, H.J.; Yang, G.-Y. Synergy of nitrogen vacancies and intercalation of carbon species for enhancing sunlight photocatalytic hydrogen production of carbon nitride. Appl. Catal. B Environ. 2022, 314, 121497–121509. [Google Scholar]
  16. Yonesato, K.; Yamazoe, S.; Yokogawa, D.; Yamaguchi, K.; Suzuki, K. A molecular hybrid of an atomically precise silver nanocluster and polyoxometalates for H2 cleavage into protons and electrons. Angew. Chem. Int. Ed. 2021, 60, 16994–16998. [Google Scholar] [CrossRef]
  17. Chen, Y.; Guo, Z.-W.; Li, X.-X.; Zheng, S.-T.; Yang, G.-Y. Multicomponent cooperative assembly of nanoscale boron-rich polyoxotungstates {B30Si6Ni12Ln6W27(OH)26O168}, {B30Si5Ni12Ln7W27(OH)26O166(H2O)}, and {B22Si4Ni12Ln4W36(OH)12O178}. CCS Chem. 2021, 3, 1232–1241. [Google Scholar]
  18. Xie, S.S.; Jiang, J.J.; Wang, D.; Tang, Z.G.; Mi, R.F.; Chen, L.J.; Zhao, J.W. Tricarboxylic-ligand-decorated lanthanoid-inserted heteropolyoxometalates built by mixed-heteroatom-directing polyoxotungstate units: Syntheses, structures, and electrochemical. Inorg. Chem. 2021, 60, 7536–7544. [Google Scholar] [CrossRef]
  19. Cai, J.; Ye, R.; Jia, K.; Qiao, X.; Zhao, L.; Liu, J.; Sun, W. pH-controlled construction of lanthanide clusters from lacunary polyoxometalate with single-molecule magnet behavior. Inorg. Chem. Commun. 2020, 112, 107694. [Google Scholar] [CrossRef]
  20. Li, C.; Mizuno, N.; Yamaguchi, K.; Suzuki, K. Self-Assembly of Anionic Polyoxometalate–Organic Architectures Based on Lacunary Phosphomolybdates and Pyridyl Ligands. J. Am. Chem. Soc. 2019, 141, 7687–7692. [Google Scholar] [CrossRef]
  21. Khajavian, R.; Jodaian, V.; Taghipour, F.; Mague, J.T.; Mirzaei, M. Roles of organic fragments in redirecting crystal/molecular structures of inorganic-organic hybrids based on lacunary Keggin-type polyoxometalates. Molecules 2021, 26, 5994. [Google Scholar] [CrossRef]
  22. Zheng, S.-T.; Yang, G.-Y. Recent advances in paramagnetic-TM-substituted polyoxometalates (TM = Mn, Fe, Co, Ni, Cu). Chem. Soc. Rev. 2012, 41, 7623–7646. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, J.W.; Li, Y.Z.; Chen, L.J.; Yang, G.Y. Research progress on polyoxometalate-based transition-metal–rare-earth heterometallic derived materials: Synthetic strategies, structural overview and functional applications. Chem. Commun. 2016, 52, 4418–4445. [Google Scholar] [CrossRef]
  24. Knoth, W.H.; Domaille, P.J.; Roe, D.C. Halometal derivatives of W12PO403− and related 183W NMR studies. Inorg. Chem. 1983, 22, 198–201. [Google Scholar] [CrossRef]
  25. Kholdeeva, O.A.; Maksimov, G.M.; Maksimovskaya, R.I.; Kovaleva, L.A.; Fedotov, M.A.; Grigoriev, V.A.; Hill, C.L. A dimeric titanium-containing polyoxometalate. Synthesis, characterization, and catalysis of H2O2-based thioether oxidation. Inorg. Chem. 2000, 39, 3828–3837.26. [Google Scholar] [CrossRef] [PubMed]
  26. Qu, L.-Y.; Shan, Q.-J.; Gong, J.; Lu, R.-Q.; Wang, D.-R. Synthesis, properties and characterization of Dawson-type tungstophosphate heteropoly complexes substituted by titanium and peroxotitanium. J. Chem. Soc. Dalton Trans. 1997, 23, 4525–4528. [Google Scholar] [CrossRef]
  27. Sakai, Y.; Kitakoga, Y.; Hayashi, K.; Yoza, K.; Nomiya, K. Isolation and molecular structure of a monomeric, tris [peroxotitanium (IV)]-substituted α-Dawson polyoxometalate derived from the tetrameric anhydride form composed of four tris [titanium (IV)]-substituted α-Dawson substructures and four bridging titanium (IV) octahedral groups. Eur. J. Inorg. Chem. 2004, 23, 4646–4652. [Google Scholar]
  28. Hayashi, K.; Takahashi, M.; Nomiya, K. Novel Ti–O–Ti bonding species constructed in a metal-oxide cluster. Dalton Trans. 2005, 23, 3751–3756. [Google Scholar] [CrossRef]
  29. Hayashi, K.; Murakami, H.; Nomiya, K. Novel Ti−O−Ti bonding species constructed in a metal-oxide cluster: Reaction products of bis (oxalato) oxotitanate (IV) with the dimeric, 1,2-dititanium (IV)-substituted Keggin polyoxotungstate. Inorg. Chem. 2006, 45, 8078–8085. [Google Scholar] [CrossRef]
  30. Nomiya, K.; Mouri, Y.; Sakai, Y.; Matsunaga, S. Reaction products of titanium (IV) sulfate with the two, dimeric precursors, 1,2,3-tri-titanium (IV)- and 1,2-di-titanium (IV)-substituted α-Keggin polyoxometalates (POMs), under acidic conditions. A tetra-titanium (IV) oxide cluster and one coordinated sulfate ion grafted on a di-lacunary Keggin POM. Inorg. Chem. Commun. 2012, 19, 10–14. [Google Scholar]
  31. An, H.Y.; Zhang, Y.M.; Hou, Y.J.; Hu, T.; Yang, W.; Chang, S.Z.; Zhang, J.J. Hybrid dimers based on metal-substituted Keggin polyoxometalates (metal = Ti, Ln) for cyanosilylation catalysis. Dalton Trans. 2018, 47, 9079–9089. [Google Scholar] [CrossRef]
  32. 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]
  33. Chauveau, F.; Eberle, J.; Lefebvre, J. Synthèse de molécules élaborées à partir de métaux de transition: Un polyoxométallate mixte de tungstène VI et de zirconium IV. Nouv. J. Chim. 1985, 9, 315. [Google Scholar]
  34. Villanneau, R.; Carabineiro, H.; Carrier, X.; Thouvenot, R.; Herson, P.; Lemos, F.; Ribeiro, F.R.; Che, M. Synthesis and characterization of Zr (IV) polyoxotungstates as molecular analogues of zirconia-supported tungsten catalysts. J. Phys. Chem. B 2004, 108, 12465–12471. [Google Scholar] [CrossRef]
  35. 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]
  36. Sokolov, M.N.; Izarova, N.V.; Peresypkina, E.V.; Mainichev, D.A.; Fedin, V.P. Zirconium and hafnium aqua complexes [(H2O)3M(P2W17O61)]6−: Synthesis, characterization and substitution of water by chiral ligand. Inorg. Chim. Acta 2009, 362, 3756–3762. [Google Scholar] [CrossRef]
  37. Falber, A.; Burton-Pye, B.P.; Radivojevic, I.; Todaro, L.; Saleh, R.; Francesconi, L.C.; Drain, C.M. Ternary porphyrinato HfIV and ZrIV polyoxometalate complexes. Eur. J. Inorg. Chem. 2009, 2009, 2459–2466. [Google Scholar] [CrossRef]
  38. Radivojevic, I.; Ithisuphalap, K.; Burton-Pye, B.P.; Saleh, R.; Francesconi, L.C.; Drain, C.M. Ternary phthalocyanato Hf (IV) and Zr (IV) polyoxometalate complexes. RSC Adv. 2013, 3, 2174–2177. [Google Scholar] [CrossRef]
  39. Lin, Y.; Weakley, T.J.R.; Rapko, B.; Finke, R.G. Polyoxoanions derived from A-β-SiW9O3410−: Synthesis, single-crystal structural determination, and solution structural characterization by 183W NMR and IR of A-β-Si2W18Ti6O7714-. Inorg. Chem. 1993, 32, 5095–5101. [Google Scholar] [CrossRef]
  40. Nomiya, K.; Takahashi, M.; Ohsawa, K.; Widegren, J.A. Synthesis and characterization of tri-titanium (IV)-1,2,3-substituted α-Keggin polyoxotungstates with heteroatoms P and Si. Crystal structure of the dimeric, Ti–O–Ti bridged anhydride form K10H2[α,α-P2W18Ti6O77]⋅17H2O and confirmation of dimeric forms in aqueous solution by ultracentrifugation molecular weight measurements. J. Chem. Soc. Dalton Trans. 2001, 19, 2872–2878. [Google Scholar] [CrossRef]
  41. McGlone, T.; Vila-Nadal, L.; Miras, H.N.; Long, D.L.; Poblet, J.M.; Cronin, L. Assembly of titanium embedded polyoxometalates with unprecedented structural features. Dalton Trans. 2010, 39, 11599–11604. [Google Scholar] [CrossRef]
  42. Al-Kadamany, G.; Bassil, B.S.; Raad, F.; Kortz, U. The oxalato-titanium-containing tungstophosphate (V) dimers, [Ti8(C2O4)8P2W18O76(H2O)4]18− and [Ti6(C2O4)4P4W32O124]20−. J Clust. Sci. 2014, 25, 867–878. [Google Scholar] [CrossRef]
  43. Matsuki, Y.; Hoshino, T.; Takaku, S.; Matsunaga, S.; Nomiya, K. Synthesis and molecular structure of a water-soluble, dimeric tri-titanium (IV)-substituted Wells−Dawson polyoxometalate containing two bridging (C5Me5)Rh2+ groups. Inorg. Chem. 2015, 54, 11105–11113. [Google Scholar] [CrossRef] [PubMed]
  44. Nomiya, K.; Takahashi, M.; Widegren, J.A.; Aizawa, T.; Sakai, Y.; Kasuga, N.C. Synthesis and pH-variable ultracentrifugation molecular weight measurements of the dimeric, Ti–O–Ti bridged anhydride form of a novel di-TiIV-1,2-substituted α-Keggin polyoxotungstate. Molecular structure of the [(α-1,2-PW10Ti2O39)2]10− polyoxoanion. J. Chem. Soc. Dalton Trans. 2002, 19, 3679–3685. [Google Scholar] [CrossRef]
  45. Goto, Y.; Kamata, K.; Yamaguchi, K.; Uehara, K.; Hikichi, S.; Mizuno, N. Synthesis, structural characterization, and catalytic performance of dititanium-substituted γ-Keggin silicotungstate. Inorg. Chem. 2006, 45, 2347–2356. [Google Scholar] [CrossRef] [PubMed]
  46. Tan, R.X.; Li, D.L.; Wu, H.B.; Zhang, C.L.; Wang, X.H. Synthesis and structure of dititanium-containing 10-tungstogermanate [{γ-GeTi2W10O36(OH)2}2(μ-O)2]8−. Inorg. Chem. Commun. 2008, 11, 835–836. [Google Scholar] [CrossRef]
  47. Hussain, F.; Bassil, B.S.; Kortz, U.; Kholdeeva, O.A.; Timofeeva, M.N.; de Oliveira, P.; Keita, B.; Nadjo, L. Dititanium-containing 19-tungstodiarsenate (III) [Ti2(OH)2As2W19O67(H2O)]8−: Synthesis, structure, electrochemistry, and oxidation catalysis. Chem. Eur. J. 2007, 13, 4733–4742. [Google Scholar] [CrossRef]
  48. Murakami, H.; Hayashi, K.; Tsukada, I.; Hasegawa, T.; Yoshida, S.; Miyano, R.; Kato, C.N.; Nomiya, K. Novel solid-state 8H+-heteropolyacid. Synthesis and molecular structure of a free-acid form of a Dawson-Type sandwich complex, [Ti2{P2W15O54(OH2)2}2]8−. Bull. Chem. Soc. Jpn. 2007, 80, 2161–2169. [Google Scholar] [CrossRef]
  49. Yoshida, S.; Murakami, H.; Sakai, Y.; Nomiya, K. Syntheses, molecular structures and pH-dependent monomer-dimer equilibria of Dawson α2-monotitanium (IV)-substituted polyoxometalates. Dalton Trans. 2008, 34, 4630–4638. [Google Scholar] [CrossRef]
  50. Xu, L.J.; Zhou, W.Z.; Zhang, L.Y.; Li, B.; Zang, H.Y.; Wang, Y.H.; Li, Y.G. Organic-inorganic hybrid assemblies based on Ti-substituted polyoxometalates for photocatalytic dye degradation. CrystEngComm 2015, 17, 3708–3714. [Google Scholar] [CrossRef]
  51. Gaunt, A.J.; May, L.; Collison, D.; Fox, O.D. A novel zirconium polyoxometalate complex that contains both a coordinated saturated anion, [PMo12O40]3−, and a coordinated unsaturated anion, [PMo11O39]7−. Inorg. Chem. 2003, 42, 5049–5051. [Google Scholar] [CrossRef]
  52. Cai, L.L.; Li, Y.X.; Yu, C.J.; Ji, H.M.; Liu, Y.; Liu, S.X. Spontaneous resolution of a chiral polyoxometalate: Synthesis, crystal structures and properties. Inorg. Chim. Acta 2009, 362, 2895–2899. [Google Scholar] [CrossRef]
  53. Niu, Y.J.; Liu, B.; Xue, G.L.; Hu, H.M.; Fu, F.; Wang, J.W. A new sandwich polyoxometalate based on Keggin-type monolacunary polyoxotungstoborate anion, [Zr(α-BW11O39)2]14−. Inorg. Chem. Commun. 2009, 12, 853–855. [Google Scholar] [CrossRef]
  54. Kato, C.N.; Shinohara, A.; Hayashi, K.; Nomiya, K. Syntheses and X-ray crystal structures of zirconium (IV) and hafnium (IV) complexes containing monovacant Wells-Dawson and Keggin polyoxotungstates. Inorg. Chem. 2006, 45, 8108–8119. [Google Scholar] [CrossRef] [PubMed]
  55. Fang, X.; Hill, C.L. Multiple reversible protonation of polyoxoanion surfaces: Direct observation of dynamic structural effects from proton transfer. Angew. Chem. Int. Ed. 2007, 46, 3877–3880. [Google Scholar] [CrossRef] [Green Version]
  56. Kholdeeva, O.A.; Maksimov, G.M.; Maksimovskaya, R.I.; Vanina, M.P.; Trubitsina, T.A.; Naumov, D.Y.; Kolesov, B.A.; Antonova, N.S.; Carbó, J.J.; Poblet, J.M. ZrIV-monosubstituted Keggin-type dimeric polyoxometalates: Synthesis, characterization, catalysis of H2O2-based oxidations, and theoretical study. Inorg. Chem. 2006, 45, 7224–7234. [Google Scholar] [CrossRef]
  57. Yamaguchi, S.; Kikukawa, Y.; Tsuchida, K.; Nakagawa, Y.; Uehara, K.; Yamaguchi, K.; Mizuno, N. Synthesis and structural characterization of a γ-Keggin-type dimeric silicotungstate with a Bis (µ-hydroxo) dizirconium core [(γ-SiW10O36)2Zr2(µ-OH)2]10−. Inorg. Chem. 2007, 46, 8502–8504. [Google Scholar] [CrossRef]
  58. Sokolov, M.N.; Izarova, N.V.; Peresypkina, E.V.; Virovets, A.V.; Fedin, V.P. Synthesis and structures of dinuclear ZrIV and HfIV hydroxo complexes with the monolacunar Keggin and Dawson anions. Russ. Chem. Bull. 2009, 58, 507–512. [Google Scholar] [CrossRef]
  59. Villanneau, R.; Racimor, D.; Messner-Henning, E.; Rousselière, H.; Picart, S.; Thouvenot, R.; Proust, A. Insights into the coordination chemistry of phosphonate derivatives of heteropolyoxotungstates. Inorg. Chem. 2011, 50, 1164–1166. [Google Scholar] [CrossRef]
  60. Villanneau, R.; Djamâa, A.B.; Chamoreau, L.-M.; Gontard, G.; Proust, A. Bisorganophosphonyl and -organoarsenyl derivatives of heteropolytungstates as hard ligands for early-transition-metal and lanthanide cations. Eur. J. Inorg. Chem. 2013, 2013, 1815–1820. [Google Scholar] [CrossRef]
  61. Fang, X.; Anderson, T.M.; Hill, C.L. Enantiomerically pure polytungstates: Chirality transfer through zirconium coordination centers to nanosized inorganic clusters. Angew. Chem. Int. Ed. 2005, 44, 3540–3544. [Google Scholar] [CrossRef]
  62. Leclerc-Laronze, N.; Marrot, J.; Haouas, M.; Taulelle, F.; Cadot, E. Slow-proton dynamics within a zirconium-containing sandwich-like complex based on the trivacant anion α-[SiW9O34]10− synthesis, structure and NMR spectroscopy. Eur. J. Inorg. Chem. 2008, 31, 4920–4926. [Google Scholar] [CrossRef]
  63. Chen, L.L.; Li, L.L.; Liu, B.; Xue, G.L.; Hu, H.M.; Fu, F.; Wang, J.W. A zirconium-containing sandwich-type dimer based on trivacant α- and β-[GeW9O34]10− units, [Zr3O(OH)2(α-GeW9O34)(β-GeW9O34)]12−. Inorg. Chem. Commun. 2009, 12, 1035–1037. [Google Scholar] [CrossRef]
  64. Saku, Y.; Sakai, Y.; Shinohara, A.; Hayashi, K.; Yoshida, S.; Kato, C.N.; Yoza, K.; Nomiya, K. Sandwich-type HfIV and ZrIV complexes composed of tri-lacunary Keggin polyoxometalates: Structure of [M3(μ-OH)3(A-α-PW9O34)2]9− (M = Hf and Zr). Dalton Trans. 2009, 5, 805–813. [Google Scholar] [CrossRef]
  65. Wei, K.Y.; Yang, T.; Qin, S.J.; Ma, X.; Li, X.X.; Yang, G.Y. Hydrothermal synthesis, structural characterization and proton-conducting property of a 3-D framework based on Zr3Na3-substituted polyoxometalate building blocks. Chin. J. Struct. Chem. 2016, 35, 1461–1468. [Google Scholar]
  66. Zhang, Z.; Zhao, J.-W.; Yang, G.-Y. Tri-ZrIV substituted sandwiched polyoxometalate with mixed trilacunary Keggin-/Dawson-type polyoxotungstate units. Eur. J. Inorg. Chem. 2017, 2017, 3244–3247. [Google Scholar] [CrossRef]
  67. Gaunt, A.J.; May, I.; Collison, D.; Travis Holman, K.; Pope, M.T. Polyoxometal cations within polyoxometalate anions. Seven-coordinate uranium and zirconium heteroatom groups in [(UO2)12(μ3-O)4(μ2-H2O)12(P2W15O56)4]32− and [Zr4(μ3-O)2(μ2-OH)2(H2O)4(P2W16O59)2]14−. J. Mol. Struct. 2003, 656, 101–106. [Google Scholar] [CrossRef]
  68. Fang, X.; Anderson, T.M.; Hou, Y.; Hill, C.L. Stereoisomerism in polyoxometalates: Structural and spectroscopic studies of bis (malate)-functionalized cluster systems. Chem. Commun. 2005, 40, 5044–5046. [Google Scholar] [CrossRef] [PubMed]
  69. Li, D.; Han, H.Y.; Wang, Y.G.; Wang, X.; Li, Y.G.; Wang, E.B. Modification of tetranuclear zirconium-substituted polyoxometalates-syntheses, structures, and peroxidase-like catalytic activities. Eur. J. Inorg. Chem. 2013, 2013, 1926–1934. [Google Scholar] [CrossRef]
  70. Kikukawa, Y.; Yamaguchi, S.; Tsuchida, K.; Nakagawa, Y.; Uehara, K.; Yamaguchi, K.; Mizuno, N. Synthesis and catalysis of di- and tetranuclear metal sandwich-type silicotungstates [(γ-SiW10O36)2M2(µ-OH)2]10− and [(γ-SiW10O36)2M4(µ4-O)(µ-OH)6]8− (M = Zr or Hf). J. Am. Chem. Soc. 2008, 130, 5472–5478. [Google Scholar] [CrossRef]
  71. Chen, L.L.; Liu, Y.; Chen, S.H.; Hu, H.M.; Fu, F.; Wang, J.W.; Xue, G.L. Acetate-functionalized zirconium-substituted tungstogermanate, [Zr4O2(OH)2(CH3COO)2(α-GeW10O37)2]12−. J. Clust. Sci. 2009, 20, 331–340. [Google Scholar] [CrossRef]
  72. Zhang, W.; Liu, S.-X.; Zhang, C.-D.; Tan, R.-K.; Ma, F.-J.; Li, S.-J.; Zhang, Y.-Y. An acetate-functionalized tetranuclear zirconium sandwiching polyoxometalate complex. Eur. J. Inorg. Chem. 2010, 2010, 3473–3477. [Google Scholar] [CrossRef]
  73. Zhang, Z.; Wang, Y.-L.; Yang, G.-Y. Two inorganic-organic hybrid polyoxotungstates constructed from tetra-ZrIV-substituted sandwich-type germanotungstates functionalized by tris ligand. Inorg. Chem. Commun. 2017, 85, 32–36. [Google Scholar] [CrossRef]
  74. Ni, Z.-H.; Zhang, Z.; Yang, G.-Y. Two new tetra-Zr (IV)-substituted sandwich-type polyoxometalates functionalized by different organic amine ligands. J. Clust. Sci. 2018, 29, 1185–1191. [Google Scholar] [CrossRef]
  75. Ni, Z.-H.; Li, H.-L.; Li, X.-Y.; Yang, G.-Y. Zr4-substituted polyoxometalate dimers decorated by D-tartaric acid/glycolic acid: Syntheses, structures and optical/electrochemical properties. CrystEngComm 2019, 21, 876–883. [Google Scholar] [CrossRef]
  76. Wang, Y.-L.; Zhang, Z.; Li, H.-L.; Li, X.-Y.; Yang, G.-Y. A new oxalate-functionalized tetra-Zr-substituted sandwich-type silicotungstate: Hydrothermal synthesis, structural characterization and catalytic oxidation of thioethers. Eur. J. Inorg. Chem. 2019, 2019, 417–422. [Google Scholar] [CrossRef]
  77. Wang, Y.L.; Zhao, J.W.; Zhang, Z.; Sun, J.J.; Li, X.Y.; Yang, B.F.; Yang, G.Y. Enantiomeric polyoxometalates based on malate chirality-inducing tetra-ZrIV-substituted Keggin dimeric clusters. Inorg. Chem. 2019, 58, 4657–4664. [Google Scholar] [CrossRef]
  78. Al-Kadamany, G.; Mal, S.S.; Milev, B.; Donoeva, B.G.; Maksimovskaya, R.I.; Kholdeeva, O.A.; Kortz, U. Hexazirconium- and hexahafnium-containing tungstoarsenates (III) and their oxidation catalysis properties. Chem. Eur. J. 2010, 16, 11797–11800. [Google Scholar] [CrossRef]
  79. Bassil, S.B.; Dickman, M.H.; Kortz, U. Synthesis and structure of asymmetric zirconium-substituted silicotungstates, [Zr6O2(OH)4(H2O)3(β-SiW10O37)3]14− and [Zr4O2(OH)2(H2O)4(β-SiW10O37)2]10−. Inorg. Chem. 2006, 45, 2394–2396. [Google Scholar] [CrossRef]
  80. Zhang, Z.; Li, H.L.; Wang, Y.L.; Yang, G.Y. Syntheses, structures, and electrochemical properties of three new acetate-functionalized zirconium-substituted germanotungstates: From dimer to tetramer. Inorg. Chem. 2019, 58, 2372–2378. [Google Scholar] [CrossRef]
  81. Bassil, B.S.; Mal, S.S.; Dickman, M.H.; Kortz, U.; Oelrich, H.; Walder, L. 6-peroxo-6-zirconium crown and its hafnium analogue embedded in a triangular polyanion: [M6(O2)6(OH)6(γ-SiW10O36)3]18− (M = Zr, Hf). J. Am. Chem. Soc. 2008, 130, 6696–6697. [Google Scholar] [CrossRef]
  82. Al-Kadamany, G.A.; Hussain, F.; Mal, S.S.; Dickman, M.H.; Leclerc-Laronze, N.; Marrot, J.; Cadot, E.; Kortz, U. Cyclic Ti9 Keggin trimers with tetrahedral (PO4) or octahedral (TiO6) capping groups. Inorg. Chem. 2008, 47, 8574–8576. [Google Scholar] [CrossRef] [PubMed]
  83. Nsouli, N.H.; Bassil, B.S.; Dickman, M.H.; Kortz, U.; Keita, B.; Nadjo, L. Synthesis and structure of dilacunary decatungstogermanate, [γ-GeW10O36]8−. Inorg. Chem. 2006, 45, 3858–3860. [Google Scholar] [CrossRef]
  84. Ren, Y.-H.; Liu, S.-X.; Cao, R.-G.; Zhao, X.-Y. Cao, J.-F.; Gao, C.-Y. Two trimeric tri-TiIV-substituted Keggin tungstogermanates based on tetrahedral linkers. Inorg. Chem. Commun. 2008, 11, 1320–1322. [Google Scholar] [CrossRef]
  85. Hoshino, T.; Isobe, R.; Kaneko, T.; Matsuki, Y.; Nomiya, K. Synthesis and molecular structure of a novel compound containing a carbonate-bridged hexacalcium cluster cation assembled on a trimeric trititanium (IV)-substituted Wells-Dawson polyoxometalate. Inorg. Chem. 2017, 56, 9585–9593. [Google Scholar] [CrossRef]
  86. Kortz, U.; Hamzeh, S.S.; Nasser, N.A. Supramolecular structures of titanium (IV)-substituted Wells-Dawson polyoxotungstates. Chem. Eur. J. 2003, 9, 2945–2952. [Google Scholar] [CrossRef]
  87. Nishikawa, T.; Sasaki, Y. The crystal structure of monium dodecamolybdotetraarsenate (V) tetrahydrate, (NH4)4H4As4Mo12O50·4H2O. Chem. Lett. 1975, 4, 1185–1186. [Google Scholar] [CrossRef] [Green Version]
  88. Sakai, Y.; Yoza, K.; Katoa, C.N.; Nomiya, K. A first example of polyoxotungstate-based giant molecule. Synthesis and molecular structure of a tetrapod-shaped Ti–O–Ti bridged anhydride form of Dawson tri-titanium (IV)-substituted polyoxotungstate. Dalton Trans. 2003, 18, 3581–3586. [Google Scholar] [CrossRef]
  89. Sakai, Y.; Ohta, S.; Shintoyo, Y.; Yoshida, S.; Taguchi, Y.; Matsuki, Y.; Matsunaga, S.; Nomiya, K. Encapsulation of anion/cation in the central cavity of tetrameric polyoxometalate, composed of four trititanium (IV)-substituted α-Dawson subunits, initiated by protonation/deprotonation of the bridging oxygen atoms on the intramolecular surface. Inorg. Chem. 2011, 50, 6575–6583. [Google Scholar] [CrossRef]
  90. Sakai, Y.; Yoza, K.; Kato, C.N.; Nomiya, K. Tetrameric, trititanium (IV)-substituted polyoxo-tungstates with an α-Dawson substructure as soluble metal-oxide analogues: Molecular structure of the Giant “Tetrapod” [(α-1,2,3-P2W15Ti3O62)4{μ3-Ti(OH)3}4Cl]45−. Chem. Eur. J. 2003, 9, 4077–4083. [Google Scholar] [CrossRef]
  91. Sakai, Y.; Yoshida, S.; Hasegawa, T.; Murakami, H.; Nomiya, K. Tetrameric, tri-titanium (IV)-substituted polyoxometalates with an α-Dawson substructure as soluble metal oxide analogues. synthesis and molecular structure of three giant ‘‘tetrapods’’ encapsulating different anions (Br, I, and NO3). Bull. Chem. Soc. 2007, 80, 1965–1974. [Google Scholar] [CrossRef]
  92. Hussain, F.; Bassil, B.S.; Bi, L.H.; Reicke, M.; Kortz, U. Structural control on the nanomolecular scale: Self-assembly of the polyoxotungstate wheel [{β-Ti2SiW10O39}4]24−. Angew. Chem. Int. Ed. 2004, 43, 3485–3488. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, K.-Y.; Bassil, B.S.; Lin, Z.-G.; Haider, A.; Cao, J.; Stephan, H.; Viehweger, K.; Kortz, U. Ti7-containing, tetrahedral 36-tungsto-4-arsenate (III) [Ti6(TiO6)(AsW9O33)4]20−. Dalton Trans. 2014, 43, 16143–16146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Li, H.-L.; Lian, C.; Yin, D.-P.; Jia, Z.-Y.; Yang, G.-Y. A new hepta-nuclear Ti-oxo-cluster-substituted tungstoantimonate and its catalytic oxidation of thioethers. Cryst. Growth Des. 2019, 19, 376–380. [Google Scholar] [CrossRef]
  95. Li, H.-L.; Lian, C.; Yang, G.-Y. A ring-shaped 12-Ti-substituted poly (polyoxometalate): Synthesis, structure, and catalytic properties. Sci. China Chem. 2022, 65, 892–897. [Google Scholar] [CrossRef]
  96. Chen, W.-C.; Yan, L.-K.; Wu, C.-X.; Wang, X.-L.; Shao, K.-Z.; Su, Z.-M.; Wang, E.-B. Assembly of Keggin-/Dawson-type polyoxotungstate clusters with different metal units and SeO32− heteroanion templates. Cryst. Growth Des. 2014, 14, 5099–5110. [Google Scholar] [CrossRef]
  97. Li, H.L.; Wang, Y.L.; Zhang, Z.; Yang, B.F.; Yang, G.Y. A new tetra-Zr (IV)-substituted polyoxotungstate aggregate. Dalton Trans. 2018, 47, 14017–14024. [Google Scholar] [CrossRef]
  98. Zhang, Z.; Wang, Y.L.; Yang, G.Y. An unprecedented Zr containing polyoxometalate tetramer with mixed trilacunary/dilacunary Keggin-type polyoxotungstate units. Acta Crystallogr. Sect. C Struct. Chem. 2018, 74, 1284–1288. [Google Scholar] [CrossRef]
  99. Zhang, Z.; Wang, Y.L.; Liu, Y.; Hang, S.-L.; Yang, G.Y. Three ring-shaped Zr (IV)-substituted silicotungstates: Syntheses, structures and their properties. Nanoscale 2020, 12, 18333–18341. [Google Scholar] [CrossRef]
  100. Li, H.L.; Lian, C.; Yin, D.P.; Yang, G.Y. Three Zr (IV)-substituted polyoxotungstate aggregates: Structural transformation from tungstoantimonate to tungstophosphate induced by pH. Inorg. Chem. 2020, 59, 12842–12849. [Google Scholar] [CrossRef]
  101. Zhang, P.Y.; Wang, Y.; Yao, L.Y.; Yang, G.Y. Hepta-Zr-incorporated polyoxometalate assembly. Inorg. Chem. 2022, 61, 10410–10416. [Google Scholar] [CrossRef]
  102. Huang, L.; Wang, S.S.; Zhao, J.W.; Cheng, L.; Yang, G.Y. Synergistic combination of multi-ZrIV cations and lacunary Keggin germanotungstates leading to a gigantic Zr24-cluster-substituted polyoxometalate. J. Am. Chem. Soc. 2014, 136, 7637–7642. [Google Scholar] [CrossRef] [PubMed]
  103. Chou, P.-T.; Chi, Y.; Chung, M.-W.; Lin, C.-C. Harvesting luminescence via harnessing the photophysical properties of transition metal complexes. Coord. Chem. Rev. 2011, 255, 2653–2665. [Google Scholar] [CrossRef]
  104. Ma, D.-L.; Ma, V.P.-Y.; Chan, D.S.-H.; Leung, K.-H.; He, H.-Z.; Leung, C.-H. Recent advances in luminescent heavy metal complexes for sensing. Coord. Chem. Rev. 2012, 256, 3087–3113. [Google Scholar] [CrossRef]
  105. Daniel, C. Photochemistry and photophysics of transition metal complexes: Quantum chemistry. Coord. Chem. Rev. 2015, 282–283, 19–32. [Google Scholar] [CrossRef]
  106. Gómez-Coca, S.; Aravena, D.; Morales, R.; Ruiz, E. Large magnetic anisotropy in mononuclear metal complexes. Coord. Chem. Rev. 2015, 289–290, 379–392. [Google Scholar] [CrossRef] [Green Version]
  107. Antonova, N.S.; Carbo, J.J.; Kortz, U.; Kholdeeva, O.A.; Poblet, J.M. Mechanistic insights into alkene epoxidation with H2O2 by Ti-and other TM-containing polyoxometalates: Role of the metal nature and coordination environment. J. Am. Chem. Soc. 2010, 132, 7488–7497. [Google Scholar] [CrossRef]
  108. Kholdeeva, O.A.; Kovaleva, L.A.; Maksimovskaya, R.I.; Maksimov, G.M. Kinetics and mechanism of thioether oxidation with H2O2 in the presence of Ti (IV)-substituted heteropolytungstates. J. Mol. Catal. A Chem. 2000, 158, 223–229. [Google Scholar] [CrossRef]
  109. Kholdeeva, O.A.; Trubitsina, T.A.; Maksimovskaya, R.I.; Golovin, A.V.; Neiwert, W.A.; Kolesov, B.A.; López, X.; Poblet, J.M. First isolated active titanium peroxo complex:  Characterization and theoretical study. Inorg. Chem. 2004, 43, 2284–2292. [Google Scholar] [CrossRef]
  110. Kholdeeva, O.A.; Trubitsina, T.A.; Maksimov, G.M.; Golovin, A.V.; Maksimovskaya, R.I. Synthesis, characterization, and reactivity of Ti (IV)-monosubstituted Keggin polyoxometalates. Inorg. Chem. 2005, 44, 1635–1642. [Google Scholar] [CrossRef]
  111. Kholdeeva, O.A.; Trubitsina, T.A.; Timofeeva, M.N.; Maksimov, G.M.; Maksimovskaya, R.I.; Rogov, V.A. The role of protons in cyclohexene oxidation with H2O2 catalysed by Ti (IV)-monosubstituted Keggin polyoxometalate. J. Mol. Catal. A Chem. 2005, 232, 173–178. [Google Scholar] [CrossRef]
  112. 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]
  113. Donoeva, B.G.; Trubitsina, T.A.; Antonova, N.S.; Carbó, J.J.; Poblet, J.M.; Al-Kadamany, G.; Kortz, U.; Kholdeeva, O.A. Epoxidation of alkenes with H2O2 catalyzed by dititanium-containing 19-tungstodiarsenate (III): Experimental and theoretical studies. Eur. J. Inorg. Chem. 2010, 2010, 5312–5317. [Google Scholar] [CrossRef]
  114. Jimenez-Lozano, P.; Ivanchikova, I.D.; Kholdeeva, O.A.; Poblet, J.M.; Carbo, J.J. Alkene oxidation by Ti-containing polyoxometalates. unambiguous characterization of the role of the protonation state. Chem. Commun. 2012, 48, 9266–9268. [Google Scholar] [CrossRef]
  115. Jimenez-Lozano, P.; Skobelev, I.Y.; Kholdeeva, O.A.; Poblet, J.M.; Carbo, J.J. Alkene epoxidation catalyzed by Ti-containing polyoxometalates: Unprecedented beta-oxygen transfer mechanism. Inorg. Chem. 2016, 55, 6080–6084. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, T.; Mazaud, L.; Chamoreau, L.-M.; Paris, C.; Proust, A.; Guillemot, G. 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] [Green Version]
  117. Zhang, T.; Solé-Daura, A.; Fouilloux, H.; Poblet, J.M.; Proust, A.; Carbó, J.J.; Guillemot, G. Reaction pathway discrimination in alkene oxidation reactions by designed Ti-siloxy-polyoxometalates. ChemCatChem 2021, 13, 1220–1229. [Google Scholar] [CrossRef]
  118. Skobelev, I.Y.; Zalomaeva, O.V.; Kholdeeva, O.A.; Poblet, J.M.; Carbo, J.J. Mechanism of thioether oxidation over di- and tetrameric Ti centres: Kinetic and DFT studies based on model Ti-containing polyoxometalates. Chem. Eur. J. 2015, 21, 14496–14506. [Google Scholar] [CrossRef]
  119. Wang, K.-Y.; Lin, Z.G.; Bassil, B.S.; Xing, X.L.; Haider, A.; Keita, B.; Zhang, G.J.; Silvestru, C.; Kortz, U. Ti2--containing 18-tungsto-2-arsenate (III) monolacunary host and the incorporation of a phenylantimony (III) guest. Inorg. Chem. 2015, 54, 10530–10532. [Google Scholar] [CrossRef]
  120. Dupré, N.; Rémy, P.; Micoine, K.; Boglio, C.; Thorimbert, S.; Lacôte, E.; Hasenknopf, B.; Malacria, M. Chemoselective catalysis with organosoluble Lewis acidic polyoxotungstates. Chem. Eur. J. 2010, 16, 7256–7264. [Google Scholar] [CrossRef]
  121. Absillis, G.; Parac-Vogt, T.N. Peptide bond hydrolysis catalyzed by the Wells−Dawson Zr(α2--P2W17O61)2 polyoxometalate. Inorg. Chem. 2012, 51, 9902–9910. [Google Scholar] [CrossRef]
  122. Stroobants, K.; Absillis, G.; Moelants, E.; Proost, P.; Parac-Vogt, T.N. Regioselective hydrolysis of human serum albumin by Zr (IV)-substituted polyoxotungstates at the interface of positively charged protein surface patches and negatively charged amino acid residues. Chem. Eur. J. 2014, 20, 3894–3897. [Google Scholar] [CrossRef] [PubMed]
  123. Stroobants, K.; Goovaerts, V.; Absillis, G.; Bruylants, G.; Moelants, E.; Proost, P.; Parac-Vogt, T.N. Molecular origin of the hydrolytic activity and fixed regioselectivity of a Zr (IV)-substituted polyoxotungstate as artificial protease. Chem. Eur. J. 2014, 20, 9567–9577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ly, H.G.; Absillis, G.; Janssens, R.; Proost, P.; Parac-Vogt, T.N. Highly amino acid selective hydrolysis of myoglobin at aspartate residues as promoted by zirconium(IV)-substituted polyoxometalates. Angew. Chem. Int. Ed. 2015, 54, 7391–7394. [Google Scholar] [CrossRef] [PubMed]
  125. Ly, H.G.; Mihaylov, T.; Absillis, G.; Pierloot, K.; Parac-Vogt, T.N. Reactivity of dimeric tetrazirconium (IV) Wells-Dawson polyoxometalate toward dipeptide hydrolysis studied by a combined experimental and density functional theory approach. Inorg. Chem. 2015, 54, 11477–11492. [Google Scholar] [CrossRef] [PubMed]
  126. Sap, A.; De Zitter, E.; Van Meervelt, L.; Parac-Vogt, T.N. Structural characterization of the complex between hen egg-white lysozyme and Zr(IV) -substituted Keggin polyoxometalate as artificial protease. Chem. Eur. J. 2015, 21, 11692–11695. [Google Scholar] [CrossRef]
Figure 1. The ball-and-stick and polyhedral illustration of (a) [TiW11PO40]5− polyoxoanion, (b) α2-[P2W17(TiO2)O61]8− polyoxoanion, (c) [α-1,2,3-P2W15(TiO2)3O56(OH)3]9− polyoxoanion, (d) [[{Ti(OH)(ox)}2(μ-O)](α-PW11O39)]7− polyoxoanion; (e) [[{Ti(ox)(H2O)}4(µ-O)3](α-PW10O37)]7− polyoxoanion, and (f) [[{Ti(H2O)3}2{Ti(H2O)2}2(μ-O)3(SO4)](PW10O37)] polyoxoanion. Color codes: WO6, red; PO4, yellow; Ti, green; S, dark blue; C, gray; O, red.
Figure 1. The ball-and-stick and polyhedral illustration of (a) [TiW11PO40]5− polyoxoanion, (b) α2-[P2W17(TiO2)O61]8− polyoxoanion, (c) [α-1,2,3-P2W15(TiO2)3O56(OH)3]9− polyoxoanion, (d) [[{Ti(OH)(ox)}2(μ-O)](α-PW11O39)]7− polyoxoanion; (e) [[{Ti(ox)(H2O)}4(µ-O)3](α-PW10O37)]7− polyoxoanion, and (f) [[{Ti(H2O)3}2{Ti(H2O)2}2(μ-O)3(SO4)](PW10O37)] polyoxoanion. Color codes: WO6, red; PO4, yellow; Ti, green; S, dark blue; C, gray; O, red.
Molecules 27 08799 g001
Figure 2. The ball-and-stick and polyhedral illustration of (a) [W5O18Zr(μ-OH)2]6− polyanion, (b) [Zr(L-OOCCH(OH)CH2COO)P2W17O61]8− polyanion and (c) (tpp)Zr(PW11O39)[TBA]5. Color codes: WO6, red; PO4, yellow; Zr, bright green; N, blue; C, gray; O, red.
Figure 2. The ball-and-stick and polyhedral illustration of (a) [W5O18Zr(μ-OH)2]6− polyanion, (b) [Zr(L-OOCCH(OH)CH2COO)P2W17O61]8− polyanion and (c) (tpp)Zr(PW11O39)[TBA]5. Color codes: WO6, red; PO4, yellow; Zr, bright green; N, blue; C, gray; O, red.
Molecules 27 08799 g002
Figure 3. The ball-and-stick and polyhedral illustration of (a) [A-β-Si2W18Ti6O77]14− polyoxoanion, (b) [Ti8(C2O4)8P2W18O76(H2O)4]18− polyoxoanion, (c) [Ti6(C2O4)4P4W32O124]20− polyoxoanion, (d) P2W15Ti3O60(OH)2}2(Cp*Rh)2]16− polyoxoanion; (e) [α,α-P2W20Ti4O78]10− polyoxoanion; (f) [Ti2(OH)2As2W19O67(H2O)]8− polyoxoanion; (g) [Ti2{P2W15O54(OH2)2}2]8− polyoxoanion and (h) [(Ti2O)(PW11O39)2]8− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; Ti, green; As, lavender; C, gray; O, red; H, black.
Figure 3. The ball-and-stick and polyhedral illustration of (a) [A-β-Si2W18Ti6O77]14− polyoxoanion, (b) [Ti8(C2O4)8P2W18O76(H2O)4]18− polyoxoanion, (c) [Ti6(C2O4)4P4W32O124]20− polyoxoanion, (d) P2W15Ti3O60(OH)2}2(Cp*Rh)2]16− polyoxoanion; (e) [α,α-P2W20Ti4O78]10− polyoxoanion; (f) [Ti2(OH)2As2W19O67(H2O)]8− polyoxoanion; (g) [Ti2{P2W15O54(OH2)2}2]8− polyoxoanion and (h) [(Ti2O)(PW11O39)2]8− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; Ti, green; As, lavender; C, gray; O, red; H, black.
Molecules 27 08799 g003
Figure 4. The ball-and-stick and polyhedral illustration of (a) [Zr(PMo12O40)(PMo11O39)]6− polyoxoanion, (b) [Zr(α2-P2W17O61)2]16− polyoxoanion, (c) [(γ-SiW10O36)2Zr2(μ-OH)2]10− polyoxoanion, (d) {[α-P2W15O55(H2O)]Zr3(μ3-O)(H2O)(L-tartH)[α-P2W16O59]}15− polyoxoanion; (e) [Zr3(µ2-OH)2(µ2-O)(A-α-GeW9O34)(1,4,9-α-P2W15O56)]14− polyoxoanion; (f) [Zr4(µ3-O)2(µ2-OH)2(H2O)4(P2W16O59)2]14− polyoxoanion; (g) [(γ-SiW10O36)2Zr4(µ4-O)(µ-OH)6]8− polyoxoanion and (h) [Zr6O4(OH)4(H2O)2(CH3COO)5(AsW9O33)2]11− polyoxoanion. Color codes: W/MoO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Zr, bright green; As, lavender; C, gray; O, red; H, black.
Figure 4. The ball-and-stick and polyhedral illustration of (a) [Zr(PMo12O40)(PMo11O39)]6− polyoxoanion, (b) [Zr(α2-P2W17O61)2]16− polyoxoanion, (c) [(γ-SiW10O36)2Zr2(μ-OH)2]10− polyoxoanion, (d) {[α-P2W15O55(H2O)]Zr3(μ3-O)(H2O)(L-tartH)[α-P2W16O59]}15− polyoxoanion; (e) [Zr3(µ2-OH)2(µ2-O)(A-α-GeW9O34)(1,4,9-α-P2W15O56)]14− polyoxoanion; (f) [Zr4(µ3-O)2(µ2-OH)2(H2O)4(P2W16O59)2]14− polyoxoanion; (g) [(γ-SiW10O36)2Zr4(µ4-O)(µ-OH)6]8− polyoxoanion and (h) [Zr6O4(OH)4(H2O)2(CH3COO)5(AsW9O33)2]11− polyoxoanion. Color codes: W/MoO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Zr, bright green; As, lavender; C, gray; O, red; H, black.
Molecules 27 08799 g004
Figure 5. The ball-and-stick and polyhedral illustration of (a) Zr6O2(OH)4(H2O)3(β-SiW10O37)3]14− polyoxoanion, (b) [Zr6(O2)6(OH)6(γ-SiW10O36)3]18− polyoxoanion, (c) [(α-Ti3PW9O38)3(PO4)]18− polyoxoanion, (d) [(α-Ti3SiW9O37OH)3(TiO3(OH2)3)]17− polyoxoanion, (e) {K⸦[(Ge(OH)O3)(GeW9Ti3O38H2)3]}14− polyoxoanion and (f) [{Ca6(CO3)(μ3-OH)(OH2)18}(P2W15Ti3O61)3Ca(OH2)3]19− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Ti, green; Zr, bright green; Ca, orange; C, gray; O, red.
Figure 5. The ball-and-stick and polyhedral illustration of (a) Zr6O2(OH)4(H2O)3(β-SiW10O37)3]14− polyoxoanion, (b) [Zr6(O2)6(OH)6(γ-SiW10O36)3]18− polyoxoanion, (c) [(α-Ti3PW9O38)3(PO4)]18− polyoxoanion, (d) [(α-Ti3SiW9O37OH)3(TiO3(OH2)3)]17− polyoxoanion, (e) {K⸦[(Ge(OH)O3)(GeW9Ti3O38H2)3]}14− polyoxoanion and (f) [{Ca6(CO3)(μ3-OH)(OH2)18}(P2W15Ti3O61)3Ca(OH2)3]19− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Ti, green; Zr, bright green; Ca, orange; C, gray; O, red.
Molecules 27 08799 g005
Figure 6. The ball-and-stick and polyhedral illustration of (a) [{Ti3P2W15O57.5(OH)3}4]24− polyoxoanion, (b) [(α-1,2,3-P2W15Ti3O60.5)4Cl]37− polyoxoanion, (c) [(α-1,2,3-P2W15Ti3O62)4{μ3-Ti(OH)3}4Cl]45− polyoxoanion, (d) [{β-Ti2SiW10O39}4]24− polyoxoanion, (e) [Ti6(TiO6)(AsW9O33)4]20− polyoxoanion and (f) [{K2Na(H2O)3}@{(Ti2O)2-(Ti4O4)2(A-α-1,3,5-GeW9O36)2(A-α-2,3,4-GeW9O36)2}]25− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Ti, green; As, lavender; O, red.
Figure 6. The ball-and-stick and polyhedral illustration of (a) [{Ti3P2W15O57.5(OH)3}4]24− polyoxoanion, (b) [(α-1,2,3-P2W15Ti3O60.5)4Cl]37− polyoxoanion, (c) [(α-1,2,3-P2W15Ti3O62)4{μ3-Ti(OH)3}4Cl]45− polyoxoanion, (d) [{β-Ti2SiW10O39}4]24− polyoxoanion, (e) [Ti6(TiO6)(AsW9O33)4]20− polyoxoanion and (f) [{K2Na(H2O)3}@{(Ti2O)2-(Ti4O4)2(A-α-1,3,5-GeW9O36)2(A-α-2,3,4-GeW9O36)2}]25− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Ti, green; As, lavender; O, red.
Molecules 27 08799 g006
Figure 7. The ball-and-stick and polyhedral illustration of (a) [(α-SeW9O34){Zr(H2O)}{WO(H2O)}(WO2)(SeO3){α-SeW8O31Zr(H2O)}]212− polyoxoanion, (b) {Zr2[SbP2W4(OH)2O21][α2-PW10O38]}220− polyoxoanion, (c) [Zr4(β-GeW10O38)2(A-α-PW9O34)2]26− polyoxoanion; (d) [{Zr5(μ3-OH)4(OH)2}@{Zr2(OAc)2(α-GeW10O38)2}2]22− polyoxoanion, (e) [{Zr2(OH)2(α-SiW10O38)}2{Zr2(OH)2(β-SiW10O38)}2]24− polyoxoanion, (f) [(Zr2(OH)2)2(Zr2BO(OH)4)2(β-SiW10O38)4]26− polyoxoanion, (g) [ZrSb4(OH)O2(A-α-PW8O32)(A-α-PW9O34)]218− polyanion and (h) [SbZr7O6(OH)4(B-α-GeW9O34)2(B-α-GeW11O39)2]21− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Se, pale blue; Sb, violet; Zr, bright green; B, rose; O, red.
Figure 7. The ball-and-stick and polyhedral illustration of (a) [(α-SeW9O34){Zr(H2O)}{WO(H2O)}(WO2)(SeO3){α-SeW8O31Zr(H2O)}]212− polyoxoanion, (b) {Zr2[SbP2W4(OH)2O21][α2-PW10O38]}220− polyoxoanion, (c) [Zr4(β-GeW10O38)2(A-α-PW9O34)2]26− polyoxoanion; (d) [{Zr5(μ3-OH)4(OH)2}@{Zr2(OAc)2(α-GeW10O38)2}2]22− polyoxoanion, (e) [{Zr2(OH)2(α-SiW10O38)}2{Zr2(OH)2(β-SiW10O38)}2]24− polyoxoanion, (f) [(Zr2(OH)2)2(Zr2BO(OH)4)2(β-SiW10O38)4]26− polyoxoanion, (g) [ZrSb4(OH)O2(A-α-PW8O32)(A-α-PW9O34)]218− polyanion and (h) [SbZr7O6(OH)4(B-α-GeW9O34)2(B-α-GeW11O39)2]21− polyoxoanion. Color codes: WO6, red; PO4, yellow; SiO4, light blue; GeO4, turquoise; Se, pale blue; Sb, violet; Zr, bright green; B, rose; O, red.
Molecules 27 08799 g007
Figure 8. The ball-and-stick and polyhedral illustration of (a) (Ti2P2W15)2(Ti2P2W16)2(P2W11)2 polyoxoanion, (b) [Zr24O22(OH)10(H2O)2(W2O10H)2(GeW9O34)4(GeW8O31)2]32− polyoxoanion. Color codes: WO6, red; GeO4, turquoise; Zr, bright green; O, red. Adapted with permission from ref. [86], copyright 2003 John Wiley and Sons.
Figure 8. The ball-and-stick and polyhedral illustration of (a) (Ti2P2W15)2(Ti2P2W16)2(P2W11)2 polyoxoanion, (b) [Zr24O22(OH)10(H2O)2(W2O10H)2(GeW9O34)4(GeW8O31)2]32− polyoxoanion. Color codes: WO6, red; GeO4, turquoise; Zr, bright green; O, red. Adapted with permission from ref. [86], copyright 2003 John Wiley and Sons.
Molecules 27 08799 g008
Scheme 1. Schematic diagram of the proposed mechanistic processes.
Scheme 1. Schematic diagram of the proposed mechanistic processes.
Molecules 27 08799 sch001
Scheme 2. Schematic diagram of hydrolysis of GG in the Presence of POMs.
Scheme 2. Schematic diagram of hydrolysis of GG in the Presence of POMs.
Molecules 27 08799 sch002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ni, Z.; Lv, H.; Yang, G. Recent Advances of Ti/Zr-Substituted Polyoxometalates: From Structural Diversity to Functional Applications. Molecules 2022, 27, 8799. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27248799

AMA Style

Ni Z, Lv H, Yang G. Recent Advances of Ti/Zr-Substituted Polyoxometalates: From Structural Diversity to Functional Applications. Molecules. 2022; 27(24):8799. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27248799

Chicago/Turabian Style

Ni, Zhihui, Hongjin Lv, and Guoyu Yang. 2022. "Recent Advances of Ti/Zr-Substituted Polyoxometalates: From Structural Diversity to Functional Applications" Molecules 27, no. 24: 8799. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27248799

Article Metrics

Back to TopTop