Home State-of-the-art of polyoxometalate-based catalysts for catalytic removal of vesicants
Article Open Access

State-of-the-art of polyoxometalate-based catalysts for catalytic removal of vesicants

  • Aiping Gao ORCID logo EMAIL logo , Si-Han Zhao , Jiajia Wang and Hai-Lou Li EMAIL logo
Published/Copyright: July 30, 2025

Abstract

Vesicants are one frequently-used kind of chemical warfare agents due to their low-boiling point, high gaseous diffusion, and accessibility. Sulfur mustard is one of the main vesicant that has since been banned by international organization, but there are still large stocks in some countries and its production is relatively easy, requiring no tedious steps or expensive equipment. Thus far, catalytic oxidation, alongside physical adsorption and hydrolysis has been proven one of the most effective methods for decontaminating the HD simulant, 2-chloroethyl ethyl sulfide (CEES). POMs have exhibited outstanding catalytic oxidation performance toward CEES owing to their unique redox properties, robust structural stability, flexible metal valence and the presence of numerous active sites. This review summarizes the advances of POM-based materials for catalytic decontamination of vesicants simulant CEES, including POMs, POM-based MOFs and POMs composites. This review provides new sights for engineering robust POM-based catalysts for the catalytic decontamination of CWAs.

1 Introduction

Chemical warfare agents (CWAs) still can bring a potential threat to mankind owing to their abuse in wars and terrorist attacks. 1 , 2 , 3 Vesicants as a main category of CWAs, such as sulfur mustard (HD), are commonly used CWAs owing to their ready and cheap procurement, these can cause serious skin blisters, irritation to the respiration system and eyes, DNA damage through alkylation, and even death. 4 , 5 , 6 Therefore, it is essential to develop novel and efficient materials for the decontamination of vesicants. According to molecular structure characteristics of vesicants, the primary decomposition strategy is oxidation. 7 A series of different materials including metal-organic frameworks (MOFs), 8 metal-coordinated porous organic polymer, 9 metal NHC complexes, 10 covalent organic frameworks, 11 covalent-triazine frameworks, 12 TiO2-based materials, 13 zeolites, 14 and metal oxides, 15 , 16 and polyoxometalates (POMs) 17 have been designed for the removal of vesicants based on adsorption or catalytic decontamination. Comparing with adsorption removal mode, catalytic decontamination has been prevailed due to its outstanding feature of high efficiency, celerity, and repeatability. For vesicants such as HD, selective oxidization into non-toxic sulfoxide products is the primary means. Furthermore, exposure incidents to CWAs can occur via the gas, aerosol, liquid, or gel-like “thickened” composites of the agent.

POM chemistry has developed for almost 200 years after Berzelius reported the first POM [PMo12O40]3− in 1826 . 18 The desirable physical and chemical properties of POMs such as size, structure, photochemistry, redox ability, and charge distribution can be regulated through the introduction of different heteroatoms, organic components and varying structural dimensions from one-to three-dimensional forms. 19 , 20 , 21 , 22 POMs based on Mo, W, Nb, V, Ta, Al, and other elements have been widely prepared. 23 , 24 , 25 , 26 POMs are broadly used in many fields such as proton conductive materials, 27 removal of environmental pollutants, 28 biology as anticancer 29 and catalysis, 30 , 31 supramolecular materials, 32 colloid science, 33 energy, 34 electro/photochromic systems, 35 , 36 sensors, 37 and magnetism 38 through interdisciplinary. Due to their tunable acid-base and reversible multi-electron redox properties of POMs, POMs and the post-modified POMs-based composite exhibit notable catalytic performance in a wide range of reactions including esterification, hydrolysis, and oxidation of organic compounds. 39 , 40 Different kinds of POM-based materials have been designed and prepared for the catalytic decontamination of vesicants, but that still have some challenges even though great progress has been made (Figure 1). It is of great necessity to review the advances in the catalytic decontamination of vesicants by POM-based catalysts for aiming to propel further progress.

Figure 1: 
POM-based materials for catalytic decontamination of vesicants.
Figure 1:

POM-based materials for catalytic decontamination of vesicants.

2 POM-based compounds for catalytic decontamination of vesicants

Vesicants were frequently-used CWAs due to their low-boiling point, high gaseous diffusion, and accessibility. 41 Although not as lethal as other CWAs, their use causes prolonged pain that would often be followed by death. The heavy use of vesicants led to extreme suffering in the Second World War. 42 Sulfur mustard is one of the main vesicant that has since been banned by international organization, but there were still large stocks in some countries and its production is relatively easy, requiring no tedious steps or expensive equipment. Thus far, catalytic oxidation, alongside physical adsorption and hydrolysis has been proven on of the most effective methods for decontaminating the HD simulant, 2-chloroethyl ethyl sulfide (CEES). POMs have exhibited outstanding catalytic oxidation performance toward CEES owing to their uniqueredox properties, robust structural stability, flexible metal valence and the presence of numerous active sites. This section summarizes the advances of POM-based materials for catalytic decontamination of vesicants simulant CEES, including POMs, POM-based MOFs and POMs composites.

2.1 POMs in firsthand catalytic decontamination of vesicants

An’s group designed four carboxylic acid modified POMs in the form of Cs4 [M(H2O)4][PMo6O21(PABA)3]2·nH2O (M = Co, Mn, Ni, Zn, PABA = p-aminobenzoic acid) (CoPMo 6 O 21 , MnPMo 6 O 21 , NiPMo 6 O 21 , ZnPMo 6 O 21 ) for selective catalytic oxidation of thioethers including CEES (Figure 2). 43 These four hybrid POMs displayed similar dimeric isostructural architectures. The synthesis of corresponding POMs using K+, Cs+, and Rb+ were attempted, but only Cs worked. Further, different metal cations and organic ligands were used to synthesize different POMs, and only Co2+, Mn2+, Ni2+, and Zn2+ and PABA were suitable. Four POMs were crystallized in space group P1 constructed from polyoxoanion units {PMo6O21}. Using methyl phenyl sulfide model, CoPMo 6 O 21 could efficiently catalyse its oxidation into the corresponding sulfoxide with the selectivity of 98 % in 20 min, which displayed the highest activity of the catalysts. Minimal methyl phenyl sulfide was converted by both PABA and CoCl2, and Cs3 [PMo6O21(PABA)3]·nH2O could catalyse 72 % of methyl phenyl sulfide with a selectivity of 89 %, far inferior to CoPMo 6 O 21 . This result showed the synergistic effect of Co site and PABA-modified POMs together was responsible for the rapid reaction rate and excellent selectivity. Similar results were found using {CoAsMo6(PABA)3} and {CoTeMo6(PABA)3}, indicating that the central heteroatoms affected the selectivity (in the order of PMo > AsMo > TeMo). Red shifts and a new band at 760 nm in the UV–vis absorption spectrum of the interaction between CoPMo 6 O 21 and H2O2 indicated the generation of active peroxomolybdenum and peroxocobalt species, which attacked the S atom of the sulfides to yield the corresponding product. The conversion of various organic sulfides did not obviously decrease with the electronic and steric-hindrance changes. The oxidation reaction by CoPMo 6 O 21 , MnPMo 6 O 21 , NiPMo 6 O 21 , and ZnPMo 6 O 21 were first-order reactions with corresponding kinetics constants of 0.23842, 0.1624, 0.20015, and 0.17483 min−1, respectively. When methyl phenyl sulfide was replaced with the mustard gas simulant, CEES, 98.8 % was oxidized by CoPMo 6 O 21 into the nontoxic CEESO with 99.0 % selectivity in 12 min. The recyclability and stability of this system was also well demonstrated during the selective oxidation of CEES.

Figure 2: 
The schematic diagram of catalytic oxidation of thioethers including CEES by MPMo6O21 (M = co, Mn, Ni, Zn). Reproduced with permission from ref. 43]. Copyright 2020 The royal society of Chemistry.
Figure 2:

The schematic diagram of catalytic oxidation of thioethers including CEES by MPMo6O21 (M = co, Mn, Ni, Zn). Reproduced with permission from ref. 43]. Copyright 2020 The royal society of Chemistry.

Zhou’s group employed four nanopolyoxometalate clusters Na8K14(VO)2 [K10⊂{Mo(Mo)5O21(H2O)3(SO4)}12 {(VO)30(H2O)20] (denoted {Mo 72 V 30 }), 44 ([{Na10(H2O)12⊂{Mo72Cr30O252(CH3COO)19(H2O)94}]·ca.120·H2O (denoted {Mo 72 Cr 30 }), 45 [Mo72Fe30O252(CH3COO)10{Mo2O7(H2O)}{H2Mo2O8(H2O)}3(H2O)91]·ca.140·H2O (denoted {Mo 72 Fe 30 }), 46 [{Mo(Mo)5O21(H2O)4CH3COO}12{MoO(H2O)}30]·ca. 150·H2O (denoted {Mo 102 }) 47 to catalyse the decontamination of HD and CEES. 48 The conversion rate of CEES (0.042 mmol, 5 μL) reached 99.45 % in 10 min using 10.0 mg (0.53 μmol) of {Mo 72 Cr 30 } at approximately 20 °C. The optimal molar ratio of H2O2: CEES was determined to be 1:1. Within the reaction time of 1.0 min, the conversion of CEES was 51.42 %, 45.65 %, 38.38 %, and 34.95 % for {Mo 72 V 30 }, {Mo 72 Cr 30 }, {Mo 72 Fe 30 }, and {Mo 102 }, respectively. A good linear relationship between ln (Ct/C 0 ) (C0 and Ct was the concentrations of CEES at the start and time t) and the reaction time indicated first-order reaction kinetics with the kinetic constants of {Mo 72 V 30 }, {Mo 72 Cr 30 }, {Mo 72 Fe 30 }, and {Mo 102 } of 0.56, 0.53, 0.47, and 0.44 min−1, respectively and the following catalytic activity order: V > Cr > Fe > Mo. The reversible oxidation peak potentials with the addition of H2O2 and the further addition of the substrate CEES demonstrated that the intermediates generated from nodal metal interaction with H2O2 implemented the selective oxidation of CEES to CEESO. The unchanged MoVI3d X-ray photoelectron spectroscopy (XPS) peaks of {Mo 72 V 30 } indicated that MoVI did not participate in the catalytic reaction. Conversely, the binding energies of V 2p at 516.48 and 523.96 eV assigned to VIV 2p3/2 and VIV 2p1/2 were shifted to 516.40 eV (VIV 2p3/2), 517.49 eV (VV 2p3/2), 523.37 eV (VIV 2p1/2), and 524.71 eV (VV 2p1/2) with H2O2 addition, which showed that V was oxidized to a higher valence state by H2O2. 49 With the simultaneous addition of H2O2 and CEES to the solution of {Mo 72 V 30 }, XPS peaks of V 2p were recovered, demonstrating that the catalyst recovered its original state. A similar situation was observed with the three other catalysts. The desired fast degradation of CEES by four catalysts benefitted from the large oil−water interface area due to agitation and the full exposure of active sites in {Mo 72 M 30 }. All four catalysts exhibited outstanding reusability, and the conversion rates of CEES still reached 93.41 %, 90.50 %, 90.13 %, and 89.70 % for {Mo 72 V 30 }, {Mo 72 Cr 30 }, {Mo 72 Fe 30 }, and {Mo 102 }, respectively after five cycles.

Zhou’s group prepared an organic functional POM [C44H28N4Cl4]1·5 [H2PMo10V2O40]·2C2H6O (H 2 TClPP-H 2 PVMo) by ion-exchange between 5,10,15,20-tetra(4-chlorine)phenylporphyrin (TClPP) and H5PV2Mo10O40 (H 5 PVMo) for photocatalytic degradation of CEES . 50 A higher singlet oxygen generation capacity was observed for H 2 TClPP-H 2 PVMo than for TClPP, which laid solid foundation for the selective oxidization of CEES to nontoxic CEESO. The red-shift of characteristic bands of P-Oa and the blue-shift of characteristic bands of Mo-Ob-Mo and Mo-Oc-Mo in the Fourier-transform infrared (FTIR) spectrum of H 2 TClPP-H 2 PVMo indicated the formation of hybrid H 2 TClPP-H 2 PVMo, which was further verified by the 31P solid-state nuclear magnetic resonance (NMR) spectrum. Spectrometric titration results of H5PVMo solution by TClPP showed a nonequality to the simple sum of the UV spectrum of TClPP and H 5 PVMo, which indicated that both had strong interactions and TClPP was converted to a deprotonated salt. An obvious turning point indicated that a stable hybrid was formed between TClPP and H 5 PVMo at the stoichiometric ratio of 1:1.5; elemental analysis and thermogravimetric analysis (TGA) results confirmed this result. Within 180 min, the degradation rate of CEES by H 2 TClPP-H 2 PVMo in methanol could reach 99.52 % with irradiation and 28.24 % without irradiation, showing that light significantly accelerated the decontamination of CEES by H 2 TClPP-H 2 PVMo. At the same catalytic conditions, the degradation rates of CEES were approximately 20 % and 55.31 % for H 5 PVMo and TClPP, respectively. In view of the hydrolysis effect on HD and CEES, the degradation of CEES by H 2 TClPP-H 2 PVMo in methanol/water (1:1, v/v) was also explored. The degradation rate of CEES by H 2 TClPP-H 2 PVMo reached 99.14 % with irradiation and 91.05 % within 90 min (under irradiation). The degradation rate of 98.1 % in methanol/water (1:1, v/v) was still achieved after the fifth in situ circulation, indicating the good repeatability of H 2 TClPP-H 2 PVMo. Under irradiation but no catalyst in air, CEES could be alcoholised to 2-methoxyethyl ethyl sulfide in methanol, and hydrolysed to ethyl 2-hydroxyethyl sulfide (HEES) because of a small amount of water in methanol. Active oxygen generated from H 2 TClPP-H 2 PVMo selectively oxidized CEES and HEES to CEESO and ethyl hydroxyethyl sulfoxide (HEESO) under irradiation. In addition, 2-methoxyethane and ethyl sulfide were possibly produced in methanol/water (1:1, v/v) via the alcoholysis and hydrolysis of CEES. 51 , 52 Active oxygen trapping experiments using benzoquinone (BQ) or NaN3, revealed that CEES was degraded by O2·−and 1O2 producing from H 2 TClPP-H 2 PVMo in methanol with visible light irradiation. The 1O2 quantum yield of TClPP was significantly increased from 0.35 to 0.73 of H 2 TClPP-H 2 PVMo in methanol. The fluorescence intensity of H 2 TClPP-H 2 PVMo was greatly reduced because it dissipated more energy to complete energy transfer from the singlet state to the excited triplet state under irradiation. 53 This material could degrade CEES under visible light irradiation.

Yang’s group successfully prepared a novel 12-Ti-substituted POM K10H15 [{K2Na(H2O)3}@{(Ti2–O)2(Ti4O4)2 (A-α-1,3,5-GeW9O36)2 (A-α-2,3,4-GeW9O36)2}]·45H2O (TiGeWO) with the key characteristic of chiral trivacant Keggin-type GeW9O36 (A-α-1,3,5-GeW 9 O 36 and A-α-2,3,4-GeW 9 O 36 ) units. 54 Two A-α-1,3,5-GeW 9 O 36 and two A-α-2,3,4-GeW 9 O 36 were connected via two Ti2O and two Ti4O4 alternate bridging cluster cores, which was the first time for the coexistence of Ti2O and Ti4O4 cores as the bridges in POM. Each Ti atom in the Ti4O4 tetrad was coordinated with three μ2-O, one μ4-O in the GeW9O36 unit, and two μ2-O in the Ti4O4 core. The two Ti atoms of the Ti2O core were coordinated by four μ2-O, one μ4-O in the GeW9O36 unit, and one μ2-O in the Ti2O core. Although the trivacant A-α-1,3,5-GeW 9 O 36 and A-α-2,3,4-GeW 9 O 36 segments were chiral enantiomers, they coexisted in TiGeWO, resulting in a non-chiral overall structure. TiGeWO has been applied in the heterogeneous catalytic oxidation of various aromatic thioethers with H2O2 as an oxidant in acetonitrile at 60 °C. TiGeWO could efficiently catalyze the oxidation of various substrates into sulfoxides or sulfones. Substrates with electron-donating groups on the aromatic ring exhibited a higher conversion rates than those with electron-withdrawing groups. Additionally, larger steric hindrance could reduce the conversion and selectivity. The catalytic decontamination of CEES by TiGeWO was investigated under H2O2 as oxidizer at 30 °C. Nearly 100 % of CEES could be oxidized to CEESO in 90 min. The linear relation between ln (Ct/C0) suggested that this was a first-order reaction with a kinetic constant k1 of 0.05197 min−1 and a half-life of approximately 13 min. After three cycles, the high catalytic conversion of aromatic thioethers and CEES was retained, and the unanimous IR spectra of TiGeWO after three cycles demonstrated its good reusability.

Overcoming the inactivity of oxoboron clusters and the repulsion between negative polyoxotungstates and oxoboron clusters has been a challenge in the creation of B-containing polyoxotungstates. recently, Zheng and Yang’s group developed B30-incorporated polyoxotungstates [(B18Si3Ln6O36(OH)14){B4Ni4O10(OH) (A-α-SiWO34)}3]44− (1Ln) [(B19Si2Ln7O35(O–H)15(H2O)){B4Ni4O10(OH)4 (A-α-SiW9O34)}2{B3Ni4O9 (OH)3 (A-α-SiW9O34)}]41− (2Ln, Ln = Gd, Tb and Dy), and B22-incorporated POTs [(B22O42){LnNi3(OH)3(B-α-SiW9O34)}4]34− (3Ln; Ln = Sm, Gd and Tb) through multicomponent charge and symmetry matching among Ln3+, TM2+, borates, and polyoxotungstates. 55 To date, 1Ln and 2Ln were the largest B-containing polyoxotungstates. B-rich polyoxotungstates 1Ln, 2Ln and 3Ln were prepared using NaBO3·4H2O, Li2B4O7 and K2B10O16·8H2O as the sources of B, respectively. 1Gd, 2Gd, and 3Gd were discussed in detail as representatives (Figure 3). In 1Gd, 30 B atoms were dispersed over two B9O15(OH)6 ({B9}) nonamers, three B2O4(OH)2 ({B2}) dimers, and six BO3 monomers with a new B–O nonamer appearing in {B9}. A unique trigonal-prism cage [B18Si3O36(OH)12]18− ({B18Si3}) was formed by two {B9} assemblies in a foot-to-foot orientation and three SiO4 tetrahedra connected by six B–O–Si bridges. The anionic, hollow {B18Si3} cage was encapsulated by two cationic planar triangular [Gd3 (μ3-OH)]8+ units in its inner cavities through coordination with O from the {B18Si3} cage, and forming the [B18Si3Gd6O36(OH)14]2− ({B18Si3Gd6}) cluster. In this cluster, the innermost Gd-shell Gd6O12(OH)2 ({Gd6}) with trigonal prismatic shape was formed. Six BO3 units in 1Gd served as bridges, bonding to both Ni2+ and Gd3+ ions. For the 12 Ni2+ ions in the Ni-shell, every four Ni2+ ions were arranged in a pyramidal array, and two μ3-OH groups, two BO3 units, and one {B2} dimer are bridged by adjacent Ni2+ ions to form B4Ni4O10(OH)4 ({B4Ni4}). Three Ni-substituted polyoxotungstate motifs [B4Ni4O10(OH)4](A-α-SiW9O34) ({B4Ni4(A-SiW9)}) were formed after the introduction of {B4Ni4} clusters in the vacant sites of [A-SiW9]. The B30-incorporated 1Gd was obtained with the {B18Si3Gd6} cluster surrounded by three {B4Ni4(A-SiW9)} motifs. Owing to the multicomponent cooperative charge-matching interactions and matched geometries between components, a multi-shell configuration of Ln-shell@B-shell@Ni-shell@W-shell ({Gd6}@{B24Si3}@{Ni4}3@{SiW9}3) was existed in 1Ln polyoxotungstates series. The robust coordination ability and flexible coordination modes of Gd3+ ions used as templates possessed fewer structural restrictions on the formation of oxoboron motifs and their architectures. The long Gd–O bond, high coordination number and positive charge of Gd3+ ion were crucial for constructing larger B–O clusters and reducing the outer steric hindrance of the polyoxotungstate. Similar structures were observed in 2Ln. The tetrameric 3Gd exhibited a Gd-shell@B-shell@Ni-shell@W-shell ({Gd4}@{B22}@{Ni3}4 @{B–SiW9}4) configuration featuring a cage-like 22-nuclearity Td-symmetry B–O cluster [B22O42]18− ({B22}). This was the first observation of the {B22} cluster topology in oxoboron cluster chemistry. Within the {B22} cluster, six B3O7 three-rings are located in the tetrahedral arrays defined by four Gd3+ ions, forming a hollow tetrahedral cage [Gd4B22O42]6− ({Gd4}@{B22}≡{Gd4B22}) with a diameter of approximately 1.2 nm through the connections of the Gd- and B-shells. A slightly larger tetrahedral cluster [Ni12Gd4B22O46(OH)12]6− {Ni12Gd4B22} was formed by capping each corner of the tetrahedral {Gd4B22} cage with three edge-shared NiO6 octahedra (Ni3O(OH)3 core). Under the precondition of the isomerization from {A-SiW9} to [B-α-SiW9O34]10− ({B–SiW9}), four {B–SiW9} attached to four corners of {Ni12Gd4B22} cluster, resulting in the Td-symmetrical 3Gd. 1Ln, 2Ln, and 3Ln possessed 39, 21, and 24 protons, respectively, which were delocalized on the overall structures. All these POMs were soluble in water, but not in many solvents. 1Gd was tested for the catalytic decontamination CEES, where it could selectively convert about 96 % of CEES into CEESO with 97 % selectivity in 2 h and a half-life of approximately 10 min. In a control experiment without 1Gd, only around 66 % CEES was converted. Consistent IR and powder X-ray diffraction (PXRD) experimental results along with the maintained high activity of 1Gd in six catalytic reactions confirmed its structural stability and recyclability.

Figure 3: 
Crystal structures of POMs 1Gd, 2Gd and 3Gd. Reproduced with permission from ref. 55]. Copyright 2021 The Chinese chemical Society.
Figure 3:

Crystal structures of POMs 1Gd, 2Gd and 3Gd. Reproduced with permission from ref. 55]. Copyright 2021 The Chinese chemical Society.

Our group designed and synthesized a series of Ln-containing polytungstate ([H2N(CH3)2]8Na8Cs4H9 [Ln2 -SeW4O11(OH) (H2O)4(SbW9O33) (SeW9O33) (Se1/2Sb1/2W9O33)]2·32H2O [Ln = Tb (1), Dy (2), Ho(3), Er (4)] (Figure 4) 56 and Cs18Na8H20 [Ce3(H2O)10W8Bi4O28(B-α-BiW9 O33)4]2·64H2O{SeO2 (OH)} (5) 57 for catalytic oxidation CEES. POMs 14 were in the monoclinic space group P21/c and the hexameric architectures. In POMs 14, two equivalent trimeric subunits Ln2W4O9(H2O)4(SbW9O33) (SeW9O33) (Se1/2Sb1/2W9O33) were linked through two μ2-{SeO2(OH)} bridges. Three different types of trivacant Keggin fragments {SbW9}, {SeW9}, and {Se1/2Sb1/2W9} were existed in compounds 14, and {SeW9} was formed in situ, not from the raw materials. It was a rare occurrence that POM was constructed from different fragments with the central heteroatoms of trigonal pyramidal SbIII and SeIV. From the bond lengths of Se1/2Sb1/2−O located in the scope of 1.892–1.924, which were between that of Se–O and Sb–O bonds, indicating that 1/2 S b and 1/2 Se atoms co-occupied the central position of Se1/2Sb1/2W9O33 fragment. POM three was selected as a representative to investigate their catalytic oxidation ability of CEES. POM three exhibited that the remarkable oxidation activity of CEES with 99 % percent conversion of CEES into CEESO and a high selectivity of 95 % within 30 min. Without 3, only 30 % conversion was observed. In addition, POM three exhibited a desirable heterogeneous catalytic behavior, and remained stable conversion and selectivity of CEES after five cycles. POM five was in monoclinic C2/c space group and contained, and it was the first case that tungstobismuthate building blocks self-assembled into unprecedented octameric Ln-containing POM([Ce3(H2O)10W8Bi4O28(B-α-BiW9 O33)4]246−) with an extended structure. The octameric [Ce3(H2O)10W8Bi4O28(B-α -BiW9O33)4]246− was constructed by two tetrameric [Ce3(H2O)10W8Bi4O28(B-α-BiW9O33)4]23− via two Ce–O–W bonds as linker, which was formed by a heterometallic cluster core Ce3(H2O)10W8Bi4O28 surrounded with four B-α-BiW9O33 fragments. POM five also demonstrated its.

Figure 4: 
The crystal structures of ([H2N(CH3)2]8Na8Cs4H9 [Ln2SeW4O11(OH) (H2O)4(SbW9O33) (SeW9O33) (Se1/2Sb1/2W9O33)]2·32H2O [ln = Tb, dy, Ho, Er]. Reproduced with permission from ref. 56]. Copyright 2021 American chemical Society.
Figure 4:

The crystal structures of ([H2N(CH3)2]8Na8Cs4H9 [Ln2SeW4O11(OH) (H2O)4(SbW9O33) (SeW9O33) (Se1/2Sb1/2W9O33)]2·32H2O [ln = Tb, dy, Ho, Er]. Reproduced with permission from ref. 56]. Copyright 2021 American chemical Society.

brilliant catalytic oxidation performance of CEES. The conversion of CEES and selectivity to CEESO could reach 98 % and 96 % within 35 min, and the pseudo first-order dynamic reaction rate constant k 1 was 0.11632 min−1 with a half-life of about 6 min. After five cycles, POM five still rrretained high catalytic activity and selectivity to CEESO, and meaning that POM five could be as a robust heterogeneous catalyst.

2.2 POMs-based MOFs for catalytic decontamination of vesicants

Liu’s group designed two polyoxovanadate-based MOFs ([Co(bib)]{V2O6} (V–Co-MOF) and [Ni(en) (bib)]{V2O6}·2H2O (V–Ni-MOF) (bib = 1,4-bis(1H-imidazoly-1-yl)benzene, en = ethanediamine) for two-site synergetic catalytic decontamination of CEES. 58 One-dimensional (1D) {V2O6}2- polyanion chains (abbreviated as {V2O6}) were formed into two 2D bimetallic oxide layers, {CoV2O6} and {NiV2O6}, using {CoN2O2} tetrahedra and {NiN4O2} octahedra as linkers, and further assembled into 3D frameworks. The four-coordinated Co (II) ions in the V–Co-MOF were unsaturated metal sites, in contrast to the saturated six-coordinated Ni (II) ions of V–Co-MOF. The catalytic activities of both MOFs for CEES oxidation were studied. CEES was selectively oxidized to CEESO without 2-chloroethyl ethyl sulfone (CEESO2) byproduct of. V–Co-MOF exhibited a higher catalytic activity than V–Ni-MOF. A pseudo-first order reaction of V–Co-MOF and V–Ni-MOF was deduced with the corresponding rate constants k1 and k2 of 0.27044 and 0.06273 min−1, and the half-lives of 2.6 and 11.0 min, respectively. Through evaluating the catalytic activities of en, bib, NiCl2, CoCl2, NaK{V2O6}, NaK{V2O6} + NiCl2, NaK{V2O6} + CoCl2, V–Ni-MOF and V–Co-MOF, V–Ni-MOF possessed the highest catalytic activity, possibly owing to the generation of active peroxovanadium from the reaction of Vv with H2O2 and the CEES + via the coordination of Co(II) and S. V–Co-MOF could rapidly capture CEES with a maximum absorption of 14 mg g−1, then activating CEES. 59 The binding energy of Co 2p in the XPS of CEES-absorbing V–Co-MOF showed a notable downward shift, indicating electron transfer from CEES to Co ions and a significant increase in electron density of Co ions in V–Co-MOF. Similar experimental results were observed with V–Ni-MOF. The formation of O–O in the peroxovanadium was verified by new peaks at 860 cm−1 and 870 cm−1 in the FTIR and Raman spectra. Throughout the reaction, the weight and structure of V–Co-MOF remained largely unchanged, demonstrating its high stability and recyclability. The possible mechanism was speculated based on the Experimental results and shown in Figure 5. This work was the first example of a dual-active-site POM-based MOF catalyst for mustard gas detoxification.

Figure 5: 
The proposed catalytic mechanism of CEES by V–Co-MOF. Reproduced with permission from ref. 58]. Copyright 2020 The royal society of Chemistry.
Figure 5:

The proposed catalytic mechanism of CEES by V–Co-MOF. Reproduced with permission from ref. 58]. Copyright 2020 The royal society of Chemistry.

Another POV-based MOF [Ni(bib)2]{V2O6} ({V 6 }-MOF) was successfully synthesized by Liu’s group for the catalytic decontamination of CEES. 60 A hexanuclear vanadium anion cluster, {V6O18}6− ({V6}) in {V 6 }-MOF presented an unprecedented hollow Lindqvist-like structure compared with reported {V6} clusters. 61 Bimetallic decanuclear cluster {Ni4–V6} was generated through the coordination between the {V6} cluster and four Ni atoms. Through alternating {V6} clusters and octahedral Ni atoms, a 2D bimetallic oxide layer of {Ni3V6O18} emerged, which was further assembled into a 3D framework using bib ligand as the linker (Figure 6). {V 6 }-MOF exhibited excellent thermal stability up to 450 °C and solvent stability, which was crucial to its sustainability. The catalytic activity of {V 6 }-MOF on the oxidative detoxification of CEES was evaluated, whereby CEES was completely converted into nontoxic CEESO in 40 min with a half-life of 10.5 min. This was significantly faster than the 240 min reaction time of Lindqvist-type POV-based MOF catalyst also containing hexanuclear vanadium clusters, which was attributed to the more accessible V sites in {V 6 }-MOF {V 6 }-MOF compared to the Lindqvist-type {V6} cluster. Feasible ratios of {V 6 }-MOF: CEES and CEES: H2O2 was determined to be 0.05:1 and 1.0:1.2, respectively. Various raw materials and intermediates were used as catalysts in a series of control tests to verify the active components of {V 6 }-MOF. {V6}-mIM could completely catalyze the conversion of CEES to CEESO with 100 % selectivity, but bib and [NiII(H2O)4(L)2] exhibited no activity, verifying that the main active site of {V 6 }-MOF was the {V6} cluster. A new Raman spectrum peak at 870 cm−1 assigned to the O–O stretch of {V 6 }-MOF treated with H2O2 indicated the production of active peroxo species. These species attacked S atom of CEES, generating nontoxic CEESO. The conversion and selectivity were retained after four successive runs. The consistent PXRD patterns, FTIR and XPS spectra of the cycled catalyst indicated that {V 6 }-MOF could be applied as a highly effective catalyst for the detoxification of mustard gas due to its excellent stability and sustainability.

Figure 6: 
The crystal structure of {V6}-MOF. Reproduced with permission from ref. 60]. Copyright 2021 American chemical Society.
Figure 6:

The crystal structure of {V6}-MOF. Reproduced with permission from ref. 60]. Copyright 2021 American chemical Society.

Liu’s group developed a hollow POM@MOF with a pseudo-homoepitaxial structure [Cu2(BTC)4/3(H2O)2]6 [H3PW12O40]·(C4H12N)2 (NENU-3a) for promoting the adsorption and catalytic performance toward CEES. 62 H3PW12O40 (PW 12 ) as guest with ideal catalytic activity was perfectly encapsulated in the [Cu3BTC2] (BTC = 1,3,5-benzenetricarboxylate) host through supramolecular assembly and formed, which exhibited a perfectly matched lattice/structure for [Cu3BTC2] and PW 12 , yielding desirable catalytic performance. Scanning electron microscopy (SEM) images of NENU-3a showed the clear transits from cuboctahedral to octahedral morphology, indicative of epitaxial growth. XRD patterns and FTIR peaks of NENU-3a confirmed the presence of [Cu3BTC2] and PW 12 in NENU-3a. Formation of an internal mesoporous structure was evident due to the noticeable interior collapse of the crystalline particle, leading to a clear hollow structure with a diameter of 700 nm after 1.5 min. Energy dispersive X-ray (EDX) elemental mapping revealed the distribution of Cu and W, confirming the hollow structure and the uniform epitaxial growth of NENU-3a. The stable existence of [Cu3BTC2] in the PW 12 /Cu(NO3)2 solution was attributed to the inhibition of PW 12 dissociation in weak polar chloroform/methanol and lowering H+ concentration. HNO3, released from reactions between H3BTC and Cu(NO3)2 in the NENU-3a assembly contributed to the formation of the hollow structure via in-diffusion etching. The concentration difference drove HNO3 enriched in the confined space to diffuse inwardly through hydrogen bonding, etching the [Cu3(BTC)2] cores. Without this, HNO3 directly diffused to the exterior owing to the lack of hydrogen bonds, and relatively dilute HNO3 could not etch the inside [Cu3(BTC)2]. Transmission electron microscopy (TEM) images and the concentration-dependent etching experiment of HNO3 on [Cu3(BTC)2] both confirmed the proposed internal diffusion mechanism in Figure 7. Etching occurred on the [Cu3(BTC)2] surface only when the HNO3 concentration increased to about 70 mM, which indicated that the interior was more readily etched than the exterior. Compared to [Cu3(BTC)2] and solid NENU-3a, the intermediate Brunauer–Emmett–Teller (BET) surface area of hollow NENU-3a indicated that micropores were present only in the hollow layer of the surface, with some [Cu3(BTC)2] remaining inside. Similar thermal stability was observed on hollow NENU-3a, mainly due to the stabilization and protective effect of its outer shell layers for the interior [Cu3(BTC)2], in addition to its robust single-crystalline nature. Another POM@MOF ([Cu12(BTC)8][PMo 12 ]) (NENU-5a) was constructed using the pseudo-homoepitaxial growth strategy, demonstrating the versatility of the strategy. The catalytic oxidation performance of CEES was tested on hollow NENU-3a with H2O2 as an oxidant. Hollow NENU-3a exhibited a faster catalytic rate than solid NENU-3a. The CEES conversions were almost 100 % and 86 % for hollow and solid NENU-3a, respectively, in 30 min. Moreover, hollow NENU-3a showed the higher selectivity toward CEESO than the solid form due to the rapid departure of CEESO from the shell structure, preventing over-oxidation. Hollow NENU-3a retained its catalytic activity, selectivity, morphology and cavity structure upon reuse.

Figure 7: 
The proposed pseudo-homoepitaxial growth strategy of hollow POM@MOF. Reproduced with permission from ref. 62]. Copyright 2022 Chinese chemical Society.
Figure 7:

The proposed pseudo-homoepitaxial growth strategy of hollow POM@MOF. Reproduced with permission from ref. 62]. Copyright 2022 Chinese chemical Society.

Wang’s group synthesized a novel and multifunctional 2D POMOF [CoII(bcbpy)2 (θ-Mo8O26)0.5]·5H2O (BHU-1) through hydrothermal reactions of HbcbpyCl (1-(4-carboxybenzyl)-4,4′-bipyridinium chloride [Mo8O26]4−, and Co2+. 63 BHU-1 was the first 2D [θ-Mo8O26]4−-based cobalt–viologen framework with a triclinic system and the space group P-1. The consistency of experimental and simulated PXRD confirmed the purity of BHU-1. 64 BHU-1 exhibited two broad absorption bands, from 200 to 497 nm, and from 497 to 1,621 nm and one absorption peak at 1,663 nm, which were assigned to Mo–O ligand-to-metal charge transfer (LMCT), Co-bcbpy LMCT and the phenyl C–H stretching overtone band, respectively. Mott–Schottky (MS) electrochemical tests revealed that BHU-1 was a typical n-type semiconductor with the Fermi level (Ef) of −1.15 eV versus Ag/AgCl. According to the valence band (VB)-XPS method, the gap between Ef and the VB potential of BHU-1 (0.73 eV, vs. NHS) was found to be 1.88 eV. Under visible light irradiation, Nyquist experiments indicated that BHU-1 exhibited the fast separation and high transfer efficiency of photogenerated charge. BHU-1 catalyzed the oxidation of 98 % of CEES with 97 % selectivity within 5 min in ethanol under visible light. Under full-spectrum and near-infrared (NIR) light irradiation, the conversions of CEES were 99 % and 93 %, respectively, and the selectivities were both 96 % within 5 min, which meant that the catalytic activity was retained under different light regimes. Using the raw materials as homogeneous catalysts, CEES conversions of 17 %, 21 %, and 18 % were achieved with CoCl2·6H2O, HbcbpyCl and Na2MoO4·2H2O, respectively, indicating that the photocatalytic oxidation of CEES was enhanced by the synergy of the three components in BHU-1. The excellent cycling stability and durability was confirmed by recycling experiments and consistent PXRD and IR spectra. The catalytic reaction was simplified to a pseudo-first-order reaction using the Langmuir–Hinshelwood model. 65 The photogenerated electrons were separated from the hole with the help of C2H5OH in catalytic process of CEES, and the OH radical was easily produced through the reaction of H2O2 with photo-induced holes. The conversion rate of CEES was decreased by the introduction of BQ or tert-butyl alcohol (TBA), which indicated that CEES was oxidized to CEESO by O2− and OH generated from the reaction between photogenerated electrons with oxygen. H2O2 could also facilitate O2 circulation: H2O2 + ·O2 → O2 + e → O2 + CEES → peroxide intermediate + CEES → CEESO.

To leverage the complementary advantages of MOFs and POMs and achieve excellent catalytic performance, Zhou’s group designed a composite material by encapsulating H5PV2Mo10O40 in MIL-101(Cr) for efficient CEES decontamination. 66 The α-Keggin H5PV2Mo10O40 showed promise for the amient decontamination of HD. 67 Using the impregnation method, the maximum loading amount of H5PV2Mo10O40 by MIL-101(Cr) was approximately 33.28 wt% even with additional H5PV2Mo10O40. Given the large 34 Å cages of MIL-101(Cr) and the diameter of 13 Å and volume of 2,250 Å3 of PV2Mo10O405− anion, four H5PV2Mo10O4 per cage was speculated for the maximum loading sample. Analysis via IR spectra, XRD, TG, and N2 adsorption/desorption experiments confirmed the successful preparation and intact crystalline structure of the MOF. Furthermore, the absence of characteristic peaks of H5PV2Mo10O40 in the XRD pattern of the composite and the elemental mapping analysis (Cr, O, P, Mo, and V) in the EDX spectra indicated a uniform distribution of H5PV2Mo10O40 in MIL-101(Cr). The maximum decontamination efficiency of CEES reached 90 % using 20 mg MIL-101(Cr). For the composite materials, the decontamination efficiency gradually increased with the loading amount of H5PV2Mo10O40 increasing. The maximum decontamination efficiency of HD was 97.39 % by the composite with the loading amount of 120 mg H5PV2Mo10O40. The decontamination reaction of HD exhibited pseudo-first order kinetics with a half-life of 26.38 min by theoretical fitting. Compared to H5PV2Mo10O40, MIL-101(Cr), or their physical mixtures, the composite showed superior performance, highlighting a synergetic effect. No products from the suspension in the decontamination process of MIL-101(Cr) were detected by gas chromatography/mass spectrometry (GC/MS) except for CEES, implying the decontamination was adsorption. However, in the composite, CEES and CEESO were detected with the relative amounts of 64.93 % and 32.16 %, indicating a combined adsorption and oxidation degradation. Additionally, the composite outperformed activated clay, nano-Al2O3 and nano-MgO under the same conditions. Its decontamination efficiency remained at 93.83 % and its structure was intact even after four consecutive recycles using the composite. Notably, the negligible leakage of H5PV2Mo10O40 from MIL-101(Cr) in the recycle process was attributed to the insolubility of highly polar H5PV2Mo10O40 in nonpolar petroleum ether.

2.3 POMs composite for catalytic decontamination of vesicants

Three composites Zn2Cr-LDH-PW11M(H2O)O39 were constructed through exfoliation-reassembly of mono-transition metal (Ni, Co, Cu)–substituted POMs [PW11M(H2O)O39]5- (PW 11 M) into layered double hydroxides (LDHs) Zn 2 Cr-LDH-NO 3 by Chi’s group. 68 Increased in spacing layers (1.06 nm matching the diameter of cluster) observed in PXRD patterns of the composites indicated the replacement of small nitrate with large PW 11 M and the intercalation of PW 11 M into the LDH. Broad hump diffraction peaks (2θ = 32-42°) and FTIR characteristic peaks of composites confirmed the preservation of the LDH lamellar structure and the integrity of POMs. SEM and TEM images showed that the intrinsic LDH nanosheet morphology retained in the composite. The loading amount of PW 11 Ni was 52 wt% by inductively coupled plasma atomic emission spectroscopy (ICP-AES) analyses. Solid state 31P NMR of composite Zn 2 Cr-LDH-PW 11 Ni(H 2 O)O 39 still exhibited the characteristic peak of PW11Ni (−11.3 ppm). The BET surface area and average pore size of composite Zn 2 Cr-LDH-PW 11 Ni(H 2 O)O 39 was enlarged to 24.92 m2/g and 42.48 nm, respectively, from 13.56 m2/g and 14.14 nm of the LDH host. CEES was used to evaluate the catalytic abilities of the composites, and Zn 2 Cr-LDH-PW 11 Ni exhibited the best catalytic performance with 98 % conversion. Similar selectivity (94 %) for nontoxic CEESO was observed for all three composites. At the same conditions, the catalytic conversions of CEES were 81 %, 78 %, 73 % and 30 % for PW 11 Ni, PW 11 Co, PW 11 Cu, and Zn 2 Cr-LDH, respectively, highlighting POM as the primary catalytic active species and the cooperative effect between POM and LDHs. The low conversions of 21 % and 58 % for plenary PW 12 and lacunary K7PW11O39, respectively, revealed the important roles of tungstophosphate and substituted transition metals in the catalytic reaction. The catalytic activity order of the composites was consistent with that of PW 11 M. CEES conversion was increased with the amount of the composite Zn 2 Cr-LDH-PW 11 Ni and oxidant increasing. Initially, toxic CEESO2 was formed within the first 30 min, and remained stable until the reaction was complete. Zn 2 Cr-LDH-PW 11 Ni maintained the good stability over nine cycles. Introducing several radical scavengers had no effect on the CEES conversion, indicating a non-radical reaction process for catalytic decontamination.

Due to the large accessible areas and workability of polymers with intrinsic microporosity (PIMs), Farha’s group constructed PIM-based hybrid materials for CEES degradation using PIMs carriers for supporting POM (H3PW12O40, PW 12 ) (Figure 8). 69 Three kinds of functionalized PIMs with cyano (PIM-1), carboxy (PIM-1-COOH) and amidoxime (PIM-1-AO) groups were prepared and used to evaluate their POM loading performances. The characteristic stretching vibration peaks of POM appearing in the FTIR spectrum of PW 12 @PIM-1-AO were not observed in that of PW 12 @PIM-1 and PW 12 @PIM-1-COOH, indicating that PIM-1-AO was a promising POM support. ICP-OES analysis indicated that about 80 % of POM was supported on PIM-1-AO, but only ∼1 wt % on the others, revealing a strong interaction between POMs and PIM-1-AO. 31P solid state NMR showed a chemical shift at −15.7 ppm, and 1H NMR displayed –OH and –NH2 shifts at 9.4 and 5.8 ppm, indicating the presence of free PW 12 and interactions between POMs and the amidoxime of PIM-1-AO. The BET surface area of PW 12 @PIM-1-AO decreased with the POM loading increasing. SEM images with EDX mapping confirmed uniform POM distribution in PIM-1-AO. Two XPS peaks of 4f7/2 and 4f5/2 at 36.3 and 38.5 eV in the W 4f region of PW 12 were slightly shifted to lower binding energies for PW 12 @PIM-1-AO. Combined with a broader peak at 399.8 eV in the N 1 s region and two O 1 s peaks at 532.8 and 530.1 eV, it was concluded that POMs were attached to PIM-1-AO via electrostatic interactions between the protonated amine groups and anionic POMs. 70 Uniform fibers with an average diameter of 1.2 μm were prepared from a mixed solution of POM and PIM-1-AO in DMF (40 % w/v) via the electro-spinning. Analyses of the 31P cross-polarization magic-angle-spinning (CP/MAS) NMR spectrum, FTIR spectrum and EDX mappings of the composite fiber suggested that the POM was intact and uniform distributed within the fiber. PW 12 @PIM-1-AO composite fibers exhibited complete catalytic conversion of CEES within 60 min with a selectivity of 86 % toward CEESO. However, full conversion of CEES by PW 12 @PIM-1-AO composite powders required 90 min with 87 % selectivity toward CEESO. Compared with the full conversion and 95 % selectivity by homogeneous PW 12 in 60 min, the trapping of CEESO in the pores of the composite possibly slightly reduced its selectivity. At the same conditions, negligible conversion of CEES occurred without an active catalyst or with PIM-1-AO fiber. The composite fibers with variable POM loading amounts could be prepared by immersing PIM-1-AO fibers into POM solution with different concentrations. Marginal conversion in the leaching tests further demonstrated the heterogeneous nature of the catalytic process. After three cycles, the conversion of CEES reached 72 ± 3.3 % after 60 min.

Figure 8: 
The synthesis routine of PIM-based composite. Reproduced with permission from ref. 69]. Copyright 2022 American chemical Society.
Figure 8:

The synthesis routine of PIM-based composite. Reproduced with permission from ref. 69]. Copyright 2022 American chemical Society.

Wang’s group designed a graphene supported polyoxotungstate K1·5Cs5.5 [γ-SiW10O39Cu2(N3)2] (SiW 10 Cu 2 (N 3 ) 2 ) composite through a “click” reaction between the azido group of POM and the alkynyl groups on functionalized graphene, which boasted several advantages redox properties, higher thermal stability, electronic band structure, special hydrophobicity, and monodisperse sites (Figure 9). 71 Alkyne group-functionalized graphene (GO−C ≡ CH) was obtained through the reaction between graphene functionalized with a –COOH group and 4-(2-propynyloxy)aniline, which then reacted with SiW 10 Cu 2 (N 3 ) 2 through a “click” reaction. The Raman spectrum of SiW 10 Cu 2 /GO with the peaks at 935 and 717 cm−1 confirmed the retention of the SiW 10 Cu 2 structure. XPS analysis of SiW 10 Cu 2 /GO showed binding energy peaks at ∼935 and 932 eV corresponding to Cu2+ and Cu+, respectively. This facilitated the electron transfer between graphene and SiW 10 Cu 2 and promoted the dismutation of H2O2 to O2 and the oxidation reaction. TEM and EDX images of SiW 10 Cu 2 /GO revealed a single molecular layer of particles, approximately 1.2 nm in size (close to 10 Å of POMs) on the surface of SiW 10 Cu 2 /GO samples. SiW 10 Cu 2 /GO was used to catalyze the oxidation of CEES with H2O2. The conversion of CEES without SiW 10 Cu 2 /GO was 34.0 % with the selectivities of 50.9 % and 46.2 % toward CEESO and CEESO2, respectively, after 60 min. A similar result was observed with the coexistence of graphene and H2O2. The addition of POM increased the conversion to 64.5 % with 100 % selectivity toward CEESO. The turnover frequency (TOF) increased from 0.748 min−1 for SiW 10 Cu 2 (N 3 ) 2 to 51.556 for SiW 10 Cu 2 /GO. Under SiW 10 Cu 2 /GO catalysis, CEES was almost completely and selectively oxidized to CEESO. Radical scavenger experiments indicated no OH or superoxide free radicals in this catalytic system. The time-dependent absorbance decrease at 410 nm of 1,3-diphenylisobenzofuran (DPBF) (50 μM) with SiW 10 Cu 2 /GO catalyst (1 mg) and H2O2 (1.5 μM) indicated the emergence of singlet oxygen in this catalytic reaction. The CEES conversion of 72.0 % from the mixture of SiW 10 Cu 2 (N 3 ) 2 and graphene after 60 min highlighted synergistic effects. Silence in the UV−vis band at 300−400 nm for SiW 10 Cu 2 (N 3 ) 2 with H2O2 suggested that no peroxidation occurred, except for H2O2 dismutation. A decomposition efficiency of 35.9 % at 5 min and 61.5 % at 30 min for H2O2 on SiW 10 Cu 2 /GO was observed, with a decomposition rate of 1.35 × 10−1 μmol s−1 for H2O2, exceeding the 1.89 × 10−2 μmol s−1 rate of SiW 10 Cu 2 (N 3 ) 2  at 30 min. This efficiency benefited from SiW10Cu2 being bonded as single molecules to the graphene surface, unlike the assembly of SiW 10 Cu 2 (N 3 ) 2 . XPS result revealed that the molar ratio of Cu2+ and Cu+ in SiW 10 Cu 2 /GO was 1.3:1, which possibly enhanced the dismutation of H2O2 to O2. A slightly higher reaction rate for heterogeneous CuCl compared to CuCl2 indicated a higher activity for Cu+. SiW 10 Cu 2 (N 3 ) 2 achieved a 48.1 % conversion of CEES at 6 h through aerobic oxidation. For SiW 10 Cu 2 /GO, the conversion of CEES was significantly increased with increasing concentrations of O2 and CEES around the graphene interface. The adsorption amount of SiW 10 Cu 2 /GO for CEES and O2 was much larger than that of SiW 10 Cu 2 (N 3 ) 2 . The hydrophobic environment of SiW 10 Cu 2 /GO readily adsorbed more oleophilic molecules, but not polar molecules. When oleophilic CEES was.

Figure 9: 
The schematic diagram of simultaneous dismutation of H2O2 to O2 and oxidation of CEES by SiW10Cu2/GO catalyst. Reproduced with permission from ref. 71]. Copyright 2020 American chemical Society.
Figure 9:

The schematic diagram of simultaneous dismutation of H2O2 to O2 and oxidation of CEES by SiW10Cu2/GO catalyst. Reproduced with permission from ref. 71]. Copyright 2020 American chemical Society.

oxidized to polar CEESO, CEESO was repelled and escaped from SiW 10 Cu 2 /GO, preventing peroxidation to CEESO2, as confirmed by XPS and IR experiments. Electron transfer and synergistic effects between SiW 10 Cu 2 and graphene promoted the aerobic oxidation of CEES. A selectivity of 100 % toward CEESO was still achieved with physically mixed hybrid SiW 10 Cu 2 (N 3 ) 2 /GO, although the conversion was only 49.0 % at 6 h. 72 Near-perfect CEES conversion and minimal leaching were observed after 10 cycles of SiW 10 Cu 2 (N 3 ) 2 /GO. The maintained structure and morphology confirmed by IR, 29Si MAS NMR, XPS and TEM images of SiW 10 Cu 2 /GO after the cycle experiment demonstrated its potential as a catalyst for mustard gas analogue removal via aerobic oxidation at room temperature without additives.

Farha’s group prepared a composite material PW 12 @NU-1000 via the direct impregnation method for catalytic oxidation of CEES leveraging the advantages of MOF NU-1000 and POM H3PW12O40 (PW 12 ), such as small triangular channels (12 Å) and larger hexagonal channels (31 Å) for fast substrate diffusion, the suitable size and high catalytic activity. 73 About 0.8 [PW12O40]3- was loaded per Zr6 node in NU–1000 as confirmed by ICP-OES. SEM images with EDS mapping showed the intact size and morphology of NU-1000 and the uniform dispersion of PW 12 within the MOF, not just on the surface. Compared to bare NU-1000, the BET surface area and the length of the large hexagonal pores of PW 12 @NU-1000 were greatly decreased from 2,100 m2/g and 31 Å to 850 m2/g and 25 Å, respectively. The absence of non-H bonded–OH and H2O characteristic peaks on the NU-1000 node and the presence of H-bonded H2O and–OH peaks in diffuse reflectance infrared fourier-transform spectroscopy (DRIFTS) of PW 12 @NU-1000 indicated that the non-H-bonded H2O and hydroxyl ligands were replaced by POM. PW 12 @NU-1000 exhibited good stability in HCl solution (pH 2–7) and only a 13 % loss of POMs after one week at pH 6, with less lost at lower pH values. PW 12 @NU-1000 effectively catalyzed the oxidation of CEES with a TOF of 10.4 min−1 in acetonitrile (1 mL) at 45 °C. In contrast, PW 12 and NU-1000 alone had lower TOFs of 9.3 and 3.1 min−1, respectively. For instance, full conversion of CEES was obtained by PW 12 within 90 min, whereas PW 12 @NU-1000 needed only 20 min under the same conditions. The cooperative effect between the POM and MOF endowed PW 12 @NU-1000 with a faster initial rate and higher conversion efficiency than either component alone. However, the selectivity of PW 12 to CEESO was superior to that of PW 12 @NU-1000. The 10 % conversion and absence of a phosphorous NMR signal after the filtration of PW 12 @NU-1000 suggested negligible leaching of POMs from the MOF during catalysis. The selectivity and reactivity were maintained upon recycling under these catalytic conditions, and the integrity of the composite material was unchanged in the cyclic catalytic process.

Farha’s group synthesized PW 12 @NU-1000-scCO 2 and PW 12 @NU-1000-120 o C by the incorporation of H3PW12O40 (PW 12 ) into the MOF, NU-1000, and the influence of the POM on catalytic performance was studied. 74 PW 12 @NU-1000-scCO 2 was obtained via activation by supercritical CO2 drying, and PW 12 @NU-1000-120 o C was synthesized in 120 °C oven for over 1 h. ICP-OES analysis indicated a maximum loading of 0.8 POMs per Zr6 node in both composites, with near uniform PW 12 distribution (Figure 10). In situ variable temperature PXRD patterns showed that PW 12 @NU-1000-scCO 2 was changed into PW 12 @NU-1000-120 o C via heating to and maintaining at 120 °C, and the process was not reversible. Through differential envelope density (DED) analyses, POMs of PW 12 @NU-1000-scCO 2 were located in the hexagonal mesoporous channels, as opposed to the triangular microporous channels of PW 12 @NU-1000-120 o C. N2 adsorption isotherms and DFT results indicated a reduction in mesopore volume for PW12@NU-1000-scCO2 and micropore volume for PW 12 @NU-1000-120 o C, with the BET surface area decreasing from 1,020 m2 cm−3 to 700, and 850 m2 cm−3, respectively. The difference in BET was due to 5 wt% water present during the activated process of PW 12 @NU-1000-scCO 2 . At 170 °C, and a slight exothermic change in the differential scanning calorimetry (DSC) curve of PW 12 @NU-1000-scCO 2 indicated the loss of water H-bonded to the acidic protons of PW 12 @NU-1000-scCO 2 , correlating to POM movement from mesopore to micropore under the increasing van der Waals forces. In the micropore, POM was encircled by three pyrene linkers, as opposed to one pyrene linker in the mesopore. The broader and shifted of 31P signal indicated the loss of phosphorous symmetry and that POM strongly interacted with the MOF. The shift of reduction events to more positive potentials in the cyclic voltammograms (CVs) of PW 12 @NU-1000-120 o C from H3PW12O40 also indicated a stronger interaction between PW 12 and pyrene, making it more easily reducible than POM alone. Fluorescence quenching of bare NU-1000 by POM about 95–99 % indicated the electron transfer from pyrene linkers to POM. Catalytic oxidation of CEES to CEESO was used to evaluate the catalytic performance of both composites. The half-lifes of catalytic oxidation of CEES were 13, 5, 3, and 1 min for NU-1000, PW 12 , PW 12 @NU-1000-120 o C, and PW 12 @NU-1000-scCO 2 , respectively. The selectivity of PW 12 @NU-1000-scCO 2 toward CEESO was 90.5 %, which was higher than that of PW 12 @NU-1000-120 o C (59 ± 7 %). Similar to its selectivity toward CEESO, the TOF of PW 12 @NU-1000-scCO 2 was about three times that of PW 12 @NU-1000-120 o C. POMs situated in the mesopores facilitated sulfide diffusion to POM and the near-exclusive generation of singly-oxidized product. Leach tests confirmed the heterogeneity of the reaction and the stability of POMs in the composite during catalysis.

Figure 10: 
Visual structures of PW12@NU-1000-scCO2 and PW12@NU-1000-120 °C. Reproduced with permission from ref. 74]. Copyright 2018 The royal society of Chemistry.
Figure 10:

Visual structures of PW12@NU-1000-scCO2 and PW12@NU-1000-120 °C. Reproduced with permission from ref. 74]. Copyright 2018 The royal society of Chemistry.

Compared with PW 12 @NU-1000, PV 2 Mo 10 @NU-1000 was designed for catalytic degradation of CEES through immobilising decamolybdodivanado phosphoric acid (H5PV2Mo10O40, PV 2 Mo 10 ) in NU-1000. 75 ICP-OES revealed a loading of 1.3 PV 2 Mo 10 per Zr6 node. SEM images and EDS scans showed that the NU-1000 crystallite was retained, with POM homogeneously distributed throughout the crystallite. The BET surface areas of PV 2 Mo 10 @ NU-1000-scCO 2 and PV 2 Mo 10 @NU-1000-80°C slightly increased to 1,160 and 1,190 m2 cm−3 from the original 1,020 m2 cm−3 of NU-1000. The mesoporous peak was reduced and the microporous peak was retained for PV 2 Mo 10 @NU-1000-scCO 2 . The opposite phenomenon was observed for PV 2 Mo 10 @NU -1,000-80°C. These results suggested that POMs were located in the mesopores of PV 2 Mo 10 @NU -1000-scCO 2 and in the micropores of PV 2 Mo 10 @NU-1000-80°C, which was verified by comparison of PXRD pattern of NU-1000 and PV 2 Mo 10 @NU-1000-scCO 2 . 76 In situ temperature-dependent PXRD experiments revealed that POM migration from mesopores to micropores occurred earlier in PV 2 Mo 10 @NU-1000-scCO 2 than in PW 12 @NU-1000. The 31P NMR spectra of both composites broadened with slight shifts, indicating slight distortion of POM due to its interaction with the MOF framework. Compared to PMo 12 @NU-1000, vanadium atoms were found to play a crucial role in aerobic activity, consistent with previous reports. 73 CEESO2 was almost the only product even after CEES consumption. Despite optimising the reaction conditions, over-oxidized CEESO2 was still formed and the reaction rate was severely decreased. Although approximately 20 % POM leached from both composites, 100 % CEES conversion by PV 2 Mo 10 @NU-1000 was achievable in a second catalytic trial. In contrast, 5 % conversion was obtained using recycled H5PV2Mo10O40 due to POM decomposition. The in situ PXRD patterns of the composite, either as a dry powder or as a suspension in cyclohexane, indicated that solvent molecules could hinder migration of POM from mesopore to micropore. The MOF support strengthened the stability of POM in the composite.

An’s group designed two novel POMOFs {[ε-PMo8VMo4VIO37(OH)3][Zn2(C10N2H8) (H2–O)2]2}2·8H2O (PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β) for the selective catalytic oxidation of multitudinous sulfides in which POMs was well dispersed at the molecular level and more active sites were exposed. 77 PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β were prepared at a pH = 5.3 ± 0.1 and 4.7 ± 0.1, respectively, containing four MoV and two MoVI in the asymmetric unit. One-dimensional inorganic chains were initially formed by linking one hexacoordinated Zn4-ε-Keggin POM unit with two adjacent Zn4-ε-Keggin POM units perpendicular to the ab plane, creating a 3D framework through coordination with bpy ligands along the c axis, parallel to the ab plane. The difference between PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β was the rotation of approximately 39° clockwise acround the b axis. The XPS spectra differences for P 2p and Zn 2p in PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β indicated that P5+ and Zn+2 were in two distinct POMs. According to the deconvolution of the spectra, the ratio of MoV/MoVI was approximately 2:1 for POMOFs PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β. The broad absorption band at 400–800 nm indicated the electronic transition of reduced molybdenum ions. PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β were used as heterogeneous catalysts in the catalytic oxidation of CEES. Both exhibited outstanding catalytic oxidation performance of CEES under conditions of nCEES/npre-catalyst/noxidant = 1:0.005:1.08, at room temperature, in 0.5 mL of methanol. Over 99 % of CEES was oxidized into CEESO with more than 99 % selectivity by PMo 12 O 37 Zn 4 -α in 15 min and PMo 12 O 37 Zn 4 -β in 18 min (Figure 11). The catalytic efficiency decreased slifhtly after four consecutive trials. The structure remained unchanged before and after the catalytic reaction, demonstrating remarkable stability and recyclability. The catalytic oxidation reactions showed quasi-first order kinetics, with kinetic constants of 0.1561 and 0.2484 min−1 for PMo 12 O 37 Zn 4 -α and PMo 12 O 37 Zn 4 -β, respectively. The synergistic effect of the raw materials along with their unique structure significantly contributed to the excellent catalytic activity and stability. The appearance of a new absorption peak at 237 nm and Raman spectrum peak at 760 cm−1 of a PMo 12 O 37 Zn 4 -α solution treated with H2O2 assigned to O–O stretching indicated the formation of peroxo species in the catalytic process, which attacked the electron-rich S atom of the sulfide to produce the target product. The addition of multiple scavengers did not effectively prevent the oxidation reaction, further confirming the generation of peroxo species in the catalytic process.

Figure 11: 
The catalytic performance of CEES by PMo
12
O
37
Zn
4
-α (a) and PMo
12
O
37
Zn
4
-β (b). Reproduced with permission from ref. 77]. Copyright 2021 American chemical Society.
Figure 11:

The catalytic performance of CEES by PMo 12 O 37 Zn 4 -α (a) and PMo 12 O 37 Zn 4 -β (b). Reproduced with permission from ref. 77]. Copyright 2021 American chemical Society.

Yang’s group designed four polyoxomolybdate-based inorganic−organic hybrid materials for the catalytic oxidation of HD stimulant [Mn (TMR4A) (H2O)4][Mo6O19]·0.5CH3CH2OHH2O [Ni(TMR4A) (H2O)4][Mo6–O19]·0.5CH3CH2OHH2O [Zn (TMR4A) (H2O)4][Mo6O19]·0.5CH3CH2OHH2O, and Co2(TMR4A)2(H2O)4 (β–Mo8O26)]·CH3CN·12H2O using a self-assembly strategy (Figure 10). 78 Synthesis in different solvents, these hybrids exhibited distinct crystal structures. The first three samples contained [Mo6O19]2− anions and the last one contained [β-Mo8O26]4− anions. All four materials remained stable in various organic solvents and H2O (pH = 0–12) for 12 h. Without catalyst, only 12 % of CEES was converted to CEESO and CEESO2 at 40 °C for 25 min, using 0.25 mmol CEES, 4 ml CH3OH and 1 mmol H2O2. With the addition of [Mn (TMR4A) (H2O)4][Mo6–O19]·0.5CH3CH2OHH2O (1.25 μM), the conversions of CEES could reach 93 %and 99 % with 97 % selectivity at 30 and 40 °C, respectively, after 25 min, demonstrating its superiority as a catalyst for CEES. No catalytic activity was observed after filtering the catalyst, and the absence of MnII cations in the filtrate indicated a heterogeneous reaction. FTIR spectra and PXRD patterns of the catalyst after four cycles confirmed its structural integrity. Catalytic experiments using multiple components indicated that polyoxomolybdate was the primary catalytic active site.

Jiao’s group fabricated four nylon-6/POMs composite nanofibrous membranes (H3PW12O40·xH2O, H4PMo11VO40·xH2O, H5PMo10V2O40·xH2O and H6PMo9V3O40·xH2O, abbreviated as nylon-6/PW, PMoV1, PMoV2 and PMoV3) by electrospinning. 79 FTIR absorption peaks at 1,080–1,060 cm−1, ca. 960 cm−1, and 860–900 cm−1, 780 cm−1 assigned to the corresponding stretching vibrations of P–O, Mo=O and Mo–O–Mo bonds, confirmed the successful preparation of PMoV1, PMoV2 and PMoV3. SEM images showed that the diameter of nylon-6/PW nanofibers was 500–700 nm, whereas that of nylon-6/PMoV1, nylon-6/PMoV2 and nylon-6/PMoV3 nanofibers ranged from 50 to 100 nm. The higher viscosity of nylon-6/PW precursor solution led to the larger fibers. The nylon-6/POMs membranes exhibited decent thermal stability below 200 °C. Compared with the other composite membranes, nylon-6/PMoV3 exhibited the highest HD degradation rate, which was 41.55 % in 6 h. Increasing the V content in POMs was beneficial for redox catalysis and resulted in the appearance of more reactive lattice oxygen associated with the Mo–O–V species. For nylon-6/PMoV1, the relative lower degradation efficiency was attributed to reduced mobility due to water loss. Considering the aerosol form of HD at ambient conditions, the filtration behaviors of these composite membranes were evaluated. Using nylon-6/PMoV3 as a typical example and PP nonwoven fabric as a support the aerosol filtration efficiency increased significantly to 98.35 % in 3 h, much greater than that of PP nonwoven fabric alone (11.45 %). The mass per square meter of the nylon-6/PMoV3 fibrous membrane was about 2.6 times that used in typical HD degradation experiments. Most HD was blocked by the membrane, and penetrating HD was rapidly degraded, indicating the potential of these composite nanofibrous membranes in soft protecting materials for in situ degradation of HD.

Babri’s group prepared H3PW12O40/TiO2 and H6P2W16O62/TiO2 via a sol-gel technique for photocatalytic degradation of HD. 80 To understand solvent effect, photocatalytic oxidation of HD by illuminated TiO2 (150 mg) was investigated in relatively low viscosity solvents: hexane, acetonitrile, and acetone. After 90 min of irradiation, degradation rates of HD in the above solvents were 95 %, 90 %, and 88 %, respectively. The lower degradation rates in acetonitrile and acetone were due to reduced adsorption of the agent on TiO2 surfaces, caused by the good solubility of HD in these solvents. Photocatalytic degradation of HD by POMs/TiO2 was carried out in hexane because the physically adsorbed HD on the nTiO2 surface could not be completely washed by hexane and hexane was inert and transparent. However, POMs/TiO2 did not exhibit the ideal photocatalytic decontamination efficiency toward HD, indicating that POMs embedded in TiO2 did not significantly enhance the catalytic effect. Under no radiation filtering, HD was possibly photolysed by the radiation. By only UV irradiation, the decontamination efficiency of HD could reach to 90 % via direct photolysis in 360 min. However, nearly 95 % HD was eliminated by nanoTiO2 in 90 min. GC-MS analysis revealed that sulfoxide was initially formed, followed by sulfone, and elimination reactions occurred on both oxidation products, producing monovinyl and divinyl sulfoxides and sulfones. During photocatalytic degradation, no products or intermediates were detected, suggesting that the degradation products were adsorbed on the TiO2 surface. These products were detected when acetone was used as the extracting solvent for the TiO2 surface. Due to the presence of hydroxyl groups on the surface of nanoTiO2, hydrolytic reactions still occurred, but the incomplete hydrolysis products were less toxic than HD itself.

3 Conclusions

We summarized the recent advances in POM-based catalysts for removal of vesicants in the last decade. These POM-based catalysts exhibited the outstanding catalytic performance to vesicants benefited from their tunable solubility, high thermal and redox stability, variable valences of metals, and Lewis base and acidity properties. The synergistic effect of valence states inter-conversion of metals and the doping of other metal ions enhance intermolecular electron transfer, that is salutary to oxidate removal of vesicants. Additionally, the dual-anion POMs can possess multifunctional catalytic capacities for different kinds of CWAs. Up to now, it is the greatest challenge to POM-based catalysts is catalytic oxidation decontamination of vesicants without H2O2. Hence, it may be the future development direction to design high active POM-based catalysts, that can selectively oxidate vesicants only depending on oxygen in the air. On the other hand, the combination of the wearable fabric and POM-based catalysts is necessary to subsequent practical application.


Corresponding author: Aiping Gao and Hai-Lou Li, College of Chemistry and Chemical Engineering & Henan Province Key Laboratory of Utilization of Non-Metallic Mineral in the South of Henan, Green Catalysis and Synthesis Key Laboratory of Xinyang, Xinyang Normal University, Xinyang, Henan, 464000, P.R. China, E-mail: (A. Gao), (H.-L. Li)

Award Identifier / Grant number: U1704164

  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: Aiping Gao: Writing–original draft. Si-Han Zhao and Jiajia Wang: Investigation. Hai-Lou Li: Supervision, Writing–review & editing. All authors approved the final manuscript.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: All other authors state no conflict of interest.

  6. Research funding: The authors declare no conflict of interest, financial or otherwise, that could influence the research findings or interpretation presented in this review.

  7. Data availability: Not applicable.

  8. Copyright and Permissions: Figures/tables reproduced from prior publications are duly credited with permission from the original copyright holders (if applicable).

References

1. DeCoste, J. B.; Peterson, G. W. Chem. Rev. 2014, 111, 5695–5727.Search in Google Scholar

2. Kumar, V.; Rana, H.; Raviraju, G.; Gupta, A. K. Chemo Dosimeter for Selective and Sensitive Chromogenic and Fluorogenic Detection of Mustard Gas for Real Time Analysis. Anal. Chem. 2018, 90, 1417–1422; https://doi.org/10.1021/acs.analchem.7b04882.Search in Google Scholar PubMed

3. Saladi, R. N.; Smith, E.; Persaud, A. N. Mustard: a Potential Agent of Chemical Warfare and Terrorism. Clin. Exp. Dermatol. 2006, 31, 1–5; https://doi.org/10.1111/j.1365-2230.2005.01945.x.Search in Google Scholar PubMed

4. Baršienė, M. H.; Karjalainen, M.; Niemikoski, H.; Vanninen, P.; Broeg, K.; Lehtonen, K. K.; Berglind, R.; Baršienė, J. Toxic Effects of Chemical Warfare Agent Mixtures on the Mussel Mytilus Trossulus in the Baltic Sea: a Laboratory Exposure Study. Mar. Environ. Res. 2019, 145, 112–122; https://doi.org/10.1016/j.marenvres.2019.02.001.Search in Google Scholar PubMed

5. Smith, B. M. Catalytic Methods for the Destruction of Chemical Warfare Agents Under Ambient Conditions. Chem. Soc. Rev. 2008, 37, 470–478; https://doi.org/10.1039/b705025a.Search in Google Scholar PubMed

6. Szinicz, L. History of Chemical and Biological Warfare Agents. Toxicology 2005, 214, 167–181; https://doi.org/10.1016/j.tox.2005.06.011.Search in Google Scholar PubMed

7. DeCoste, J. B.; Peterson, G. W. Metal–Organic Frameworks for Air Purification of Toxic Chemicals. Chem. Rev. 2014, 114, 5695–5727; https://doi.org/10.1021/cr4006473.Search in Google Scholar PubMed

8. Ko, Y.; Bae, E. J.; Chitale, S. K.; Soares, C. V.; Leitão, A. A.; Kim, M.-K.; Chang, J.-S.; Maurin, G.; Ryu, S. G.; Lee, U.-H. ACS Appl. Nano Mater. 2022, 5, 9657–9665.10.1021/acsanm.2c01895Search in Google Scholar

9. Das, N.; Paul, R.; Tomar, S.; Biswas, C.; Chakraborty, S.; Mondal, J. Catching an Oxo Vanadate Porous Acetylacetonate Covalent Adaptive Catalytic Network that Renders Mustard-Gas Simulant Harmless. Inorg. Chem. 2024, 63, 6092–6102; https://doi.org/10.1021/acs.inorgchem.4c00519.Search in Google Scholar PubMed

10. Weetman, C.; Notman, S.; Arnold, P. L. Destruction of Chemical Warfare Agent Simulants by Air and Moisture Stable Metal NHC Complexes. Dalton Trans. 2018, 47, 2568–2574; https://doi.org/10.1039/c7dt04805j.Search in Google Scholar PubMed

11. Meng, Y.; Luo, Y.; Shi, J.-L.; Ding, H.; Lang, X.; Chen, W.; Zheng, A.; Sun, J.; Wang, C. 2D and 3D Porphyrinic Covalent Organic Frameworks: the Influence of Dimensionality on Functionality. Angew. Chem., Int. Ed. 2020, 59, 3624–3629; https://doi.org/10.1002/anie.201913091.Search in Google Scholar PubMed

12. Wu, B.; Liu, Y.; Zhang, Y.; Fan, L.; Li, Q.-Y.; Yu, Z.; Zhao, X.; Zheng, Y.-C.; Wang, X.-J. Molecular Engineering of Covalent Triazine Frameworks for Highly Enhanced Photocatalytic Aerobic Oxidation of Sulfides. J. Mater. Chem. A 2022, 10, 12489–12496; https://doi.org/10.1039/d2ta01441f.Search in Google Scholar

13. Sheng, W.; Shi, J.-L.; Hao, H.; Li, X.; Lang, X. Selective Aerobic Oxidation of Sulfides by Cooperative polyimide-Titanium Dioxide Photocatalysis and Triethylamine Catalysis. J. Colloid Interface Sci. 2020, 565, 614–622; https://doi.org/10.1016/j.jcis.2020.01.046.Search in Google Scholar PubMed

14. Meng, Q.; Doetschman, D. C.; Rizos, A. K.; Lee, M.-H.; Schulte, J. T.; Spyros, A.; Kanyi, C. W. Adsorption of Organophosphates Into Microporous and Mesoporous Nax Zeolites and Subsequent Chemistry. Environ. Sci. Technol. 2011, 45, 3000–3005; https://doi.org/10.1021/es1030678.Search in Google Scholar PubMed

15. Aubert, S. D.; Li, Y.; Raushel, F. M. Mechanism for the Hydrolysis of Organophosphates by the Bacterial Phosphotriesterase. Biochemistry 2004, 43, 5707–5715; https://doi.org/10.1021/bi0497805.Search in Google Scholar PubMed

16. Wagner, G. W. Nanoscale Materials in Chemistry. Environ. Appl. 2010, 125–136, Chapter 7.10.1021/bk-2010-1045.ch007Search in Google Scholar

17. Oheix, E.; Gravel, E.; Doris, E. Catalytic Processes for the Neutralization of Sulfur Mustard. Chem. Eur J. 2021, 27, 54–68; https://doi.org/10.1002/chem.202003665.Search in Google Scholar PubMed

18. Pope, M. T. Heteropoly and Isopoly Oxometalates; Springer-Verlag: Berlin, 1983; pp 26.10.1007/978-3-662-12004-0Search in Google Scholar

19. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962; https://doi.org/10.1021/cr500390v.Search in Google Scholar PubMed

20. Anyushin, A. V.; Kondinski, A.; Parac-Vogt, T. N. Hybrid Polyoxometalates as Post-Functionalization Platforms: from Fundamentals to Emerging Applications. Chem. Soc. Rev. 2020, 49, 382–432; https://doi.org/10.1039/c8cs00854j.Search in Google Scholar PubMed

21. Ma, T.; Yan, R.; Wu, X.; Wang, M.; Yin, B.; Li, S.; Cheng, C.; Thomas, A. Polyoxometalate-Structured Materials: Molecular Fundamentals and Electrocatalytic Roles in Energy Conversion. Adv. Mater. 2024, 36, 2310283; https://doi.org/10.1002/adma.202310283.Search in Google Scholar PubMed

22. Fabre, B.; Falaise, C.; Cadot, E. Polyoxometalates-Functionalized Electrodes for (Photo)Electrocatalytic Applications: Recent Advances and Prospects. ACS Catal. 2022, 12, 12055–12091; https://doi.org/10.1021/acscatal.2c01847.Search in Google Scholar

23. Cui, L.-M.; Fang, W.-H.; Zhang, J. Polyoxometalates Containing Aluminum Atoms. Chin. Chem. Lett. 2024, 110386. https://doi.org/10.1016/j.cclet.2024.110386.Search in Google Scholar

24. Li, K.; Liu, T.; Ying, J.; Tian, A.; Wang, X. Recent Research Progress on Polyoxometalate-Based Electrocatalysts in Energy Generation. J. Mater. Chem. A 2024, 12, 13576–13604; https://doi.org/10.1039/d4ta01636j.Search in Google Scholar

25. Wang, Y.; Yu, Z.; Xiong, J.; Yan, K.; Lu, X. Temperature-Responsive Polyoxometalates-Based Materials: From Underlying Mechanism to Promising Applications. Adv. Func. Mater. 2024, 34, 2405880; https://doi.org/10.1002/adfm.202405880.Search in Google Scholar

26. Ma, X.-Q.; Xiao, H.-P.; Chen, Y.; Lai, Q.-S.; Li, X.-X.; Zheng, S.-T. Polyoxometalate-Based Macrocycles and Their Assembly. Coord. Chem. Rev. 2024, 510, 215818; https://doi.org/10.1016/j.ccr.2024.215818.Search in Google Scholar

27. Liu, R.-L.; Chen, Y.; Su, X.; Zhu, W.; Liu, Z.; Chen, Y.; Wang, D.-Y.; Li, G. Proton Conductive Polyoxometalates. Coord. Chem. Rev. 2025, 522, 216224; https://doi.org/10.1016/j.ccr.2024.216224.Search in Google Scholar

28. Aramesh, N.; Bagheri, A. R.; Zhang, Z.; Yadollahi, B.; Lee, H. K. Coord. Chem. Rev. 2024, 507, 215767.10.1016/j.ccr.2024.215767Search in Google Scholar

29. Xu, X.; Guo, Y.; Li, B.; Lv, Y.; Wu, Z.; Liang, S.; He, L.; Song, Y.-F. Polyoxometalates Emerging as Multifunctional Powerhouses in the Battle Against Cancer. Coord. Chem. Rev. 2025, 522, 216210; https://doi.org/10.1016/j.ccr.2024.216210.Search in Google Scholar

30. Yang, G.; Liu, Y.; Wei, Y. Application of Polyoxometalates in Biomass Conversion. Coord. Chem. Rev. 2024, 521, 216172; https://doi.org/10.1016/j.ccr.2024.216172.Search in Google Scholar

31. Zhong, B.; Liu, J.; Liu, G.; Zhang, Z.; Chen, Y.; Wang, X. J. Mol. Struct. 2025, 1332, 141679.10.1016/j.molstruc.2025.141679Search in Google Scholar

32. Li, Z.; Li, H.; Lin, Y.; Shi, W.; Wang, X. Cluster-Cation Pairs Mediated Assembly of Subnanometer Polyoxometalates Superstructures. J. Am. Chem. Soc. 2024, 146, 28874–28884; https://doi.org/10.1021/jacs.4c09034.Search in Google Scholar PubMed

33. Zhang, S.; Shi, W.; Wang, X. Science 2022, 7574.Search in Google Scholar

34. Tang, W.; Zhang, L.; Qiu, T.; Tan, H.; Wang, Y.; Liu, W.; Li, Y. Efficient Conversion of Biomass to Formic Acid Coupled with Low Energy Consumption Hydrogen Production from Water Electrolysis. Angew. Chem. Int. Ed. 2023, 62, e202305843; https://doi.org/10.1002/anie.202305843.Search in Google Scholar PubMed

35. Zheng, K.; Ye, Y.; Shi, Y.; Xu, Y.; Yang, Z.; Ma, P.; Wang, J.; Niu, J. dl-Serine Covalently Ornamented and Ln3+-Incorporated Arsenotungstates with Fast-Responsive Photochromic and Photoinduced Luminescent Switchable Behaviors. Inorg. Chem. 2022, 61, 15871–15879; https://doi.org/10.1021/acs.inorgchem.2c01806.Search in Google Scholar PubMed

36. Li, S.; Ma, Y.; Zhao, Y.; Liu, R.; Zhao, Y.; Dai, X.; Ma, N.; Streb, C.; Chen, X. Hydrogenation Catalysis by Hydrogen Spillover on Platinum-Functionalized Heterogeneous Boronic Acid-Polyoxometalates. Angew. Chem. Int. Ed. 2023, 62, e202314999; https://doi.org/10.1002/anie.202314999.Search in Google Scholar PubMed

37. Song, P.; Wang, T. Application of Polyoxometalates in Chemiresistive Gas Sensors: a Review. ACS Sens. 2022, 7, 3634–3643; https://doi.org/10.1021/acssensors.2c02341.Search in Google Scholar PubMed

38. Clemente-Juan, J. M.; Coronado, E.; Gaita-Ariño, A. Chem. Soc. Rev. 2012, 41, 7464–7478.10.1039/c2cs35205bSearch in Google Scholar PubMed

39. Zhang, Z.-R.; Sui, H.-Y.; Shi, W.-X.; Ren, J.; Yao, S.; Lu, T.-B.; Zhang, Z.-M. Polyoxometalate-Based Single-Atom Catalyst with Precise Structure and Extremely Exposed Active Site for Efficient H2 Evolution. Angew. Chem. Int. Ed. 2025, 64, e202416711; https://doi.org/10.1002/anie.202416711.Search in Google Scholar PubMed

40. J. Liu, N. Jiang, J.-M. Lin, Z.-B. Mei, L.-Z. Dong, Y. Kuang, J.-J. Liu, S.-J. Yao, S.-L. Li, Y.-Q. Lan, Structural Evolution of Giant Polyoxometalate: from “Keplerate” to “Lantern” Type Mo132 for Improved Oxidation Catalysis. 2023, 62, e202304728, https://doi.org/10.1002/anie.202304728.Search in Google Scholar PubMed

41. Meng, W.-Q.; Sedgwick, A. C.; Kwon, N.; Sun, M.; Xiao, K.; He, X.-P.; Anslyn, E. V.; James, T. D.; Yoon, J. Fluorescent Probes for the Detection of Chemical Warfare Agents. Chem. Soc. Rev. 2023, 52, 601–662; https://doi.org/10.1039/d2cs00650b.Search in Google Scholar PubMed

42. Bobbitt, N. S.; Mendonca, M. L.; Howarth, A. J.; Islamoglu, T.; Hupp, J. T.; Farha, O. K.; Snurr, R. Q. Metal-Organic Frameworks for the Removal of Toxic Industrial Chemicals and Chemical Warfare Agents. Chem. Soc. Rev. 2017, 46, 3357–3385; https://doi.org/10.1039/c7cs00108h.Search in Google Scholar PubMed

43. An, H.; Hou, Y.; Chang, S.; Zhang, J.; Zhu, Q. Highly Efficient Oxidation of Various Thioethers Catalyzed by Organic ligand-modified Polyoxomolybdates. Inorg. Chem. Front. 2020, 7, 169–176; https://doi.org/10.1039/c9qi01098j.Search in Google Scholar

44. Müller, A.; Todea, A. M.; van Slageren, J.; Dressel, M.; Bogge, H.; Schmidtmann, M.; Luban, M.; Engelhardt, L.; Rusu, M. Triangular Geometrical and Magnetic Motifs Uniquely Linked on a Spherical Capsule Surface. Angew. Chem., Int. Ed. 2005, 44, 3857–3861; https://doi.org/10.1002/anie.200500697.Search in Google Scholar

45. Todea, A. M.; Merca, A.; Bögge, H.; van Slageren, J.; Dressel, M.; Engelhardt, L.; Luban, M.; Glaser, T.; Henry, M.; Müller, A. Extending the {(Mo)MO5}12M30 capsule keplerate sequence: a {Cr30} cluster of S=3/2 metal centers with a {Na(H2O)12} encapsulate. Angew. Chem., Int. Ed. 2007, 46, 6106–6110; https://doi.org/10.1002/anie.200700795.Search in Google Scholar

46. Müller, A.; Das, S. K.; Krickemeyer, E.; Kögerler, P.; Bögge, H.; Schmidtmann, M. Solid State Sci. 2000, 2, 847–854.10.1016/S1293-2558(00)01082-7Search in Google Scholar

47. Müller, A.; Shah, S. Q. N.; Bögge, H.; Schmidtmann, M.; Paul, K.; Hauptfleisch, B.; Leiding, S.; Wittler, K. Thirty Electrons “Trapped” in a Spherical Matrix: A Molybdenum Oxide-based Nanostructured Keplerate Reduced by 36 Electrons. Angew. Chem., Int. Ed. 2000, 39, 1614–1616; https://doi.org/10.1002/(sici)1521-3773(20000502)39:9<1614::aid-anie1614>3.0.co;2-9.10.1002/(SICI)1521-3773(20000502)39:9<1614::AID-ANIE1614>3.0.CO;2-9Search in Google Scholar

48. Yin, J.; Huang, C.; Zhou, Y.; Zhang, L.; Li, N.; Sun, R. Selective Oxidation of 2-Chloroethyl Ethyl Sulfide in Aqueous Media Catalyzed by {Mo72M30} Nano-polyoxometalate Clusters Differentiating the Catalytic Activity of Nodal Metals. Ind. Eng. Chem. Res. 2022, 61, 7699–7708; https://doi.org/10.1021/acs.iecr.2c00479.Search in Google Scholar

49. Silversmit, G.; Depla, D.; Poelman, H.; Marin, G. B.; De Gryse, R. Determination of the V2p XPS Binding Energies for Different Vanadium Oxidation States (V5+ to V0+). J. Electron. Spectrosc. 2004, 135, 167–175; https://doi.org/10.1016/j.elspec.2004.03.004.Search in Google Scholar

50. Yang, Y.; Tao, F.; Zhang, L.; Zhou, Y.; Zhong, Y.; Tian, S.; Wang, Y. Preparation of a Porphyrin-Polyoxometalate Hybrid and Its Photocatalytic Degradation Performance for Mustard Gas Simulant 2-chloroethyl Ethyl Sulfide. Chin. Chem. Lett. 2022, 33, 2625–2629; https://doi.org/10.1016/j.cclet.2021.09.093.Search in Google Scholar

51. Bartlett, P. D.; Swain, C. G. Kinetics of Hydrolysis and Displacement Reactions of ‘β,β’-Dichlorodiethyl Sulfide (Mustard Gas) and of “β-Chloro-β”-Hydroxydiethyl Sulfide (Mustard Chlorohydrin). J. Am. Chem. Soc. 1949, 71, 1406–1415; https://doi.org/10.1021/ja01172a076.Search in Google Scholar

52. Narske, R. M.; Klabunde, K. J.; Fultz, S. Solvent Effects on the Heterogeneous Adsorption and Reactions of (2-Chloroethyl)ethyl Sulfide on Nanocrystalline Magnesium Oxide. Langmuir 2002, 18, 4819–4825; https://doi.org/10.1021/la020195j.Search in Google Scholar

53. Mojarrad, A. G.; Zakavi, S. Simple Low Cost Porphyrinic Photosensitizers for Large Scale Chemoselective Oxidation of Sulfides to Sulfoxides Under Green Conditions: Targeted Protonation of Porphyrins. Catal. Sci. Technol. 2018, 8, 768–781; https://doi.org/10.1039/c7cy02308a.Search in Google Scholar

54. 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; https://doi.org/10.1007/s11426-022-1215-1.Search in Google Scholar

55. Chen, Y.; Guo, Z.-W.; Li, X.-X.; Zheng, S.-T.; Yang, G.-Y. CCS Chem. 2021, 3, 1232–1241.Search in Google Scholar

56. Li, H.-L.; Zhao, S.-H.; Gao, A.; Lian, C.; Cao, X. {SeO2(OH)} Bridging Lanthanide-Containing Antimono-Seleno-Tungstates. Inorg. Chem. 2024, 63, 9899–9906; https://doi.org/10.1021/acs.inorgchem.4c00767.Search in Google Scholar PubMed

57. Lian, C.; Zhao, S.-H.; Li, H.-L.; Cao, X. A Giant Ce-containing Poly(Tungstobismuthate): Synthesis, Structure and Catalytic Performance for the Decontamination of a Sulfur Mustard Simulant. Chin. Chem. Lett. 2024, 35, 109343; https://doi.org/10.1016/j.cclet.2023.109343.Search in Google Scholar

58. Tian, H.-R.; Zhang, Z.; Liu, S.-M.; Dang, T.-Y.; Li, X.-H.; Lu, Y.; Liu, S.-X. A Novel Polyoxovanadate-Based Co-Mof: Highly Efficient and Selective Oxidation of a Mustard Gas Simulant by Two-Site Synergetic Catalysis. J. Mater. Chem. A 2020, 8, 12398–12405; https://doi.org/10.1039/d0ta00537a.Search in Google Scholar

59. Vorontsov, A. V.; Lion, C.; Savinov, E. N.; Smirniotis, P. G. Pathways of Photocatalytic Gas Phase Destruction of HD Simulant 2-Chloroethyl Ethyl Sulfide. J. Catal. 2003, 220, 414–423; https://doi.org/10.1016/s0021-9517-03-00293-8.Search in Google Scholar

60. Tian, H.-R.; Zhang, Z.; Dang, T.-Y.; Liu, S.-M.; Lu, Y.; Liu, S.-X. Hollow Lindqvist-like-Shaped {V6} Cluster-Based Metal–Organic Framework for the Highly Efficient Detoxification of Mustard Gas Simulant. Inorg. Chem. 2021, 60, 840–845; https://doi.org/10.1021/acs.inorgchem.0c02890.Search in Google Scholar PubMed

61. Li, J.-K.; Huang, X.-Q.; Yang, S.; Ma, H.-W.; Chi, Y.-N.; Hu, C.-W. Four Alkoxohexavanadate-based Pd-Polyoxovanadates as Robust Heterogeneous Catalysts for Oxidation of Benzyl-Alkanes. Inorg. Chem. 2015, 54, 1454–1461; https://doi.org/10.1021/ic502447b.Search in Google Scholar PubMed

62. Zhang, Z.; Ma, X.; Han, X.; Cui, H.; Lu, Y.; Liu, S.; Liu, Y. Straightforward Construction of Hollow Polyoxometalate-Based Metal-Organic Framework Via Pseudo-Homoepitaxial Growth. Sci. China: Chem. 2022, 65, 1921–1928; https://doi.org/10.1007/s11426-022-1295-3.Search in Google Scholar

63. Yang, L.; Zhang, Z.; Zhang, C.; Li, S.; Liu, G.; Wang, X. An Excellent Multifunctional Photocatalyst with a Polyoxometalate–Viologen Framework for CEES Oxidation, Cr(Vi) Reduction and Dye Decolorization Under Different Light Regimes. Inorg. Chem. Front. 2022, 9, 4824–4833; https://doi.org/10.1039/d2qi00838f.Search in Google Scholar

64. Gong, Y.; Wu, T.; Jiang, P. G.; Lin, J. H.; Yang, Y. X. Inorg. Chem. 2013, 52, 777–7849.10.1021/ic3018858Search in Google Scholar PubMed

65. Li, X.; Pi, Y.; Xia, Q.; Li, Z.; Xiao, J. TiO2 Encapsulated in Salicylaldehyde-NH2-MIL-101(Cr) for Enhanced Visible light-driven Photodegradation of MB. Appl. Catal., B 2016, 191, 192–201; https://doi.org/10.1016/j.apcatb.2016.03.034.Search in Google Scholar

66. Li, Y.; Gao, Q.; Zhang, L.; Zhou, Y.; Zhong, Y.; Ying, Y.; Zhang, M.; Huang, C.; Wang, Y. H5PV2Mo10O40 Encapsulated in MIL-101(Cr): Facile Synthesis and Characterization of Rationally Designed Composite Materials for Efficient Decontamination of Sulfur Mustard. Dalton Trans. 2018, 47, 6394–6403; https://doi.org/10.1039/c8dt00572a.Search in Google Scholar PubMed

67. Sharma, A.; Singh, B.; Saxena, A. Polyoxometalate Impregnated Carbon Systems for the insitu Degradation of Sulphur Mustard. Carbon 2009, 47, 1911–1915; https://doi.org/10.1016/j.carbon.2009.03.034.Search in Google Scholar

68. Sun, X.; Dong, J.; Li, Z.; Liu, H.; Jing, X.; Chi, Y.; Hu, C. Mono-Transition-Metal-Substituted Polyoxometalate Intercalated Layered Double Hydroxides for the Catalytic Decontamination of Sulfur Mustard Simulant. Dalton Trans. 2019, 48, 5285–5291; https://doi.org/10.1039/c9dt00395a.Search in Google Scholar PubMed

69. Jung, D.; Su, S.; Syed, Z. H.; Atilgan, A.; Wang, X.; Sha, F.; Lei, Y.; Gianneschi, N. C.; Islamoglu, T.; Farha, O. K. A Catalytically Accessible Polyoxometalate in a Porous Fiber for Degradation of a Mustard Gas Simulant. ACS Appl. Mater. Interfaces 2022, 14, 16687–16693; https://doi.org/10.1021/acsami.2c01584.Search in Google Scholar PubMed

70. Wang, X.-H.; Huang, Y.-B.; Lin, Z.-J.; Cao, R. Phosphotungstic Acid Encapsulated in the Mesocages of Amine-Functionalized Metal–Organic Frameworks for Catalytic Oxidative Desulfurization. Dalton Trans. 2014, 43, 11950–11958; https://doi.org/10.1039/c4dt01043d.Search in Google Scholar PubMed

71. 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; https://doi.org/10.1021/acsanm.9b01438.Search in Google Scholar

72. Klein, M.; Varvak, A.; Segal, E.; Markovsky, B.; Pulidindi, I. N.; Perkas, N.; Gedanken, A. Sonochemical Synthesis of HSiW/Graphene Catalysts for Enhanced Biomass Hydrolysis. Green Chem. 2015, 17, 2418–2425; https://doi.org/10.1039/c4gc02519a.Search in Google Scholar

73. Buru, C. T.; Li, P.; Mehdi, B. L.; Dohnalkova, A.; Platero-Prats, A. E.; Browning, N. D.; Chapman, K. W.; Hupp, J. T.; Farha, O. K. Adsorption of a Catalytically Accessible Polyoxometalate in a Mesoporous Channel-Type Metal–Organic Framework. Chem. Mater. 2017, 29, 5174–5181; https://doi.org/10.1021/acs.chemmater.7b00750.Search in Google Scholar

74. Buru, C. T.; Platero-Prats, A. E.; Chica, D. G.; Kanatzidis, M. G.; Chapman, K. W.; Farha, O. K. Thermally Induced Migration of a Polyoxometalate Within a metal–organic Framework and Its Catalytic Effects. J. Mater. Chem. A 2018, 6, 7389–7394; https://doi.org/10.1039/c8ta02562b.Search in Google Scholar

75. Buru, C. T.; Wasson, M. C.; Farha, O. K. H5PV2Mo10O40 Polyoxometalate Encapsulated in NU-1000 Metal–Organic Framework for Aerobic Oxidation of a Mustard Gas Simulant. ACS Appl. Nano Mater. 2020, 3, 658–664; https://doi.org/10.1021/acsanm.9b02176.Search in Google Scholar

76. Livingston, S. R.; Landry, C. C. Oxidation of a Mustard Gas Analogue Using an Aldehyde/O2 System Catalyzed by V-Doped Mesoporous Silica. J. Am. Chem. Soc. 2008, 130, 13214–13215; https://doi.org/10.1021/ja8056166.Search in Google Scholar PubMed

77. Chen, Y.; Chang, S.; An, H.; Li, Y.; Zhu, Q.; Luo, H.; Huang, Y. Two Polymorphic Polyoxometalate-Based Metal–Organic Frameworks for the Efficient Synthesis of Functionalized Sulfoxides and Detoxification of Mustard Gas Simulants. ACS Sustain. Chem. Eng. 20219, 15683–15693; https://doi.org/10.1021/acssuschemeng.1c06433.Search in Google Scholar

78. Yu, M.-Y.; Yang, J.; Guo, T.-T.; Ma, J.-F. Efficient Catalytic Oxidative Desulfurization Toward Thioether and Sulfur Mustard Stimulant by Polyoxomolybdate–Resorcin[4]Arene-Based Metal–Organic Materials. Inorg. Chem. 2020, 59, 4985–4994; https://doi.org/10.1021/acs.inorgchem.0c00225.Search in Google Scholar PubMed

79. Liu, F.; Lu, Q.; Jiao, X.; Chen, D. Fabrication of Nylon-6/POMs Nanofibrous Membranes and the Degradation of Mustard Stimulant Research. RSC Adv. 2014, 4, 41271–41276; https://doi.org/10.1039/c4ra06800a.Search in Google Scholar

80. M. T. Naseri, M. Sarabadani, D. Ashrafi, H. Saeidian and M. Babri, Environ. Sci. Pollut. Res. 2013, 20, 907–916, https://doi.org/10.1007/s11356-012-0997-7.Search in Google Scholar PubMed


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/revic-2025-0047).


Received: 2025-05-19
Accepted: 2025-07-21
Published Online: 2025-07-30

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 22.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/revic-2025-0047/html
Scroll to top button