Next Article in Journal
Poly(dA:dT) Tracts Differentially Modulate Nucleosome Remodeling Activity of RSC and ISW1a Complexes, Exerting Tract Orientation-Dependent and -Independent Effects
Previous Article in Journal
Physical Pretreatments Applied in Three Commercial Kits for the Extraction of High-Quality DNA from Activated Sewage Sludge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Polyoxometalates as Electron Mediators for Photocatalytic Dye Degradation

Key Laboratory of Biomass Materials Science and Technology of Jilin Province, Beihua University, Binjiang East Road, Jilin 132013, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(20), 15244; https://doi.org/10.3390/ijms242015244
Submission received: 7 September 2023 / Revised: 6 October 2023 / Accepted: 11 October 2023 / Published: 17 October 2023
(This article belongs to the Section Physical Chemistry and Chemical Physics)

Abstract

:
The increasing concerns over the environment and the growing demand for sustainable water treatment technologies have sparked substantial interest in the field of photocatalytic dye removal. Polyoxometalates (POMs), known for their intricate metal–oxygen anion clusters, have received considerable attention due to their versatile structures, compositions, and efficient facilitation of photo-induced electron transfers. This paper provides an overview of the ongoing research progress in the realm of photocatalytic dye degradation utilizing POMs and their derivatives. The details encompass the compositions of catalysts, catalytic efficacy, and light absorption propensities, and the photocatalytic mechanisms inherent to POM-based materials for dye degradation are exhaustively expounded upon. This review not only contributes to a better understanding of the potential of POM-based materials in photocatalytic dye degradation, but also presents the advancements and future prospects in this domain of environmental remediation.

1. Introduction

As chemical entities, dyes are used in various industries such as textile manufacturing [1], leather production [2], papermaking [3], cosmetics [4,5], and food processing [6], playing an indispensable role in modern industrial processes and daily life. However, the rapid expansion of the textile printing and dyeing sector has led to a significant increase in the release of dye-laden wastewater, exacerbating water pollution [7]. The presence of aromatic rings, azo groups, and other chemical moieties within these dyes renders them particularly detrimental to the well-being of organisms [1,8]. Furthermore, dyes exhibit remarkable stability and solubility, thereby necessitating intricate, multi-stage treatment strategies for effective degradation [9].
Dyes are commonly classified into three distinct categories based on their structural configurations: azo, anthraquinone, and heterocyclic [10,11]. Azo dyes, exemplified by methyl orange (MO) [12], methyl blue [13], and Congo red (CR) [14], are extensively used in the synthesis of colorants within the textile industry. Anthraquinone dyes, characterized by vibrant and diverse color profiles, include red 3B [15], reduced yellow G (yellow anthraquinone), and reduced blue RS (blue anthraquinone). Heterocyclic dyes include both oxygen and nitrogen heterocyclic rings [16], encompassing examples like rhodamine B (RhB) [17] and methylene blue (MB). Moreover, charges can serve as an effective criterion for classifying dyes. This leads to the categorization of dyes into three primary classes: cationic dyes, anionic dyes, and nonionic dyes [18]. Cationic dyes (e.g., MB and RhB) carry positive charges and are conventionally applied to negatively charged fibers, such as cotton and linen [19]. The dyeing process with these dyes involves interactions with the fibers’ inherent negative charges. Conversely, anionic dyes (e.g., MO and CR) exhibit negative charges and find utility in positively charged fibers like wool and silk, interacting with the fibers’ positive charge during dyeing [20,21]. And nonionic dyes, exemplified by the Cibacron series, are devoid of charges, and their dyeing relies on nonionic interactions, such as hydrogen bonding [22]. Among the array of dyes, MB, RhB and MO are the frequently employed options. However, exposure to and the consumption or ingestion of these dyes could potentially result in harm to organisms [23]. Upon entering the body, MO undergoes metabolism by intestinal microorganisms into aromatic amines, involved in the initiation of intestinal cancer [24,25]. MB has the propensity to accumulate within the body through water or food chains, even at trace levels, potentially inducing carcinogenic and mutagenic effects [26,27,28]. Medically, RhB has been definitively shown to be neurotoxic and carcinogenic, capable of irritating the skin, respiratory tract, and eyes [29]. Consequently, the imperative removal of dyes from water holds substantial significance, as it is pivotal for the preservation of organismal health and the attainment of sustainable development.
Photocatalytic technology replicates natural photosynthesis, utilizing light as the energy source to activate the redox capabilities of photocatalysts across diverse light conditions. Under light irradiation, photocatalysts absorb energy, promoting the transition of electrons to higher energy levels and generating electron–hole pairs. These pairs then react with oxygen and water to produce active species like O2•− and OH. These active species work in tandem with electron–hole pairs to convert pollutants into water and carbon dioxide [30,31]. In comparison to conventional dye removal techniques, photocatalysis exhibits distinct advantages, characterized by its mild reaction conditions, environmental compatibility, non-toxic nature, and cost-effectiveness [30]. Despite the drawbacks of light dependence and the necessity for specific catalytic materials, photocatalysis retains its potential as a green and clean technology in water treatment and environmental management [32].
Polyoxometalates (POMs) are compounds consisting of metal–oxo clusters, wherein early transition metals (such as Mo, W, V, etc.) are connected via bridging oxygen linkages [33,34,35]. The amalgamation of various metals imparts POMs with a range of electronic structures, enabling them to absorb light across multiple wavelengths and display photocatalytic powers [36,37,38]. When incident light energy surpasses the bandgap energy, a process of electron excitation transpires within POMs, propelling electrons from the ground state (valence band) to the excited state (conduction band) [39]. Particularly, electrons of oxygen atom within metal–oxygen clusters become excited, subsequently transiting to the conduction band (CB) of metals and occupying the vacant d orbitals of metals in metal–oxygen clusters. This transition encompasses the shift from the highest occupied molecular orbital (HOMO) of O2− to the lowest unoccupied molecular orbital (LUMO) associated with W6+/Mo6+/V5+ [40,41,42]. This progression generates electron–hole pairs conducive to pollutant oxidation or radical generation [43]. Additionally, the interplay among distinct metals and the formation of oxygen cluster structures within POMs yield expanded reaction pathways [44,45,46]. Furthermore, the fine-tuning of POMs’composition and structure facilitates the creation of precise electronic configurations, aligning the material’s bandgap with the energy spectrum of incident light, thereby heightening photocatalytic efficiency [42].
Endowed with robust photocatalytic activity facilitated by efficient charge separation capabilities, POMs emerge as a promising avenue for sustainable water treatment [47,48,49,50]. However, a notable gap exists in the documentation of POMs’ effectiveness and mechanisms in dye removal from aqueous solutions. By scrutinizing the current research, this study has designated MO, RhB, and MB as representative dyes, and the composition of catalysts, removal efficiency, and photocatalytic mechanisms involving POM-based catalysts in connection with these dyes are critically investigated. The objective is to furnish an interdisciplinary comprehension of their potential in confronting contemporary challenges associated with water pollution. Through a systematic review of existing literature, this study aims to provide valuable references for the design of POM-based photocatalysts and offer insights into sustainable strategies for the remediation of dye-contaminated wastewater using POM-based photocatalytic materials.

2. Photocatalytic Degradation of Different Dyes with POM-Based Materials

2.1. Methylene Blue (MB)

Keggin-type polyanions [XM12O40]n− (X = B, P, Si, Ge, As; M = W, Mo, V), as one of the fundamental polyanions, possess oxygen-rich surfaces and robust coordination capabilities [51]. Vanadium atoms establish connections with the framework through metal–oxygen bonds, giving rise to vanadium-terminated or vanadium-substituted structures, thereby enhancing the compositional and property diversity of Keggin-type POMs [52]. Fengrui Li et al. [53] obtained a tri-vanadium-capped phosphomolybdate ([Na3PMo12V3O43]·4H2O, denoted as PMo12V3) through a hydrothermal reaction for the removal of MB. This compound exhibited distinct characteristic absorption peaks of POMs at 227 nm (O→Mo) and 280 nm (Mo→O→Mo). Under UV irradiation, it achieved a remarkable 99.3% MB removal within 65 min using 500 mg/L PMo12V3 and 10 mg/L MB, and the removal efficiency remained unchanged after five cycles (Table 1). The photocatalytic performance of PMo12V3O433− was amplified through vanadium substitution, leading to a reduced band gap of 2.8 eV. When subjected to UV radiation, electron transfer occurred from the HOMO of O to the LUMO of both molybdenum (Mo) and vanadium (V). This process propelled PMo12V3O433− into an excited state, generating pairs of electrons (e) and holes (h+) (Equation (1)) [53]. Concurrently, vanadium facilitated the dispersion of photo-generated electrons owing to its heightened redox characteristics and facilitated electron transfer [54]. In this sequence, photogenerated electrons, functioning as reducing agents, engaged with O2 to generate superoxide radicals (O2•−) (Equation (2)), while photo-generated holes served as oxidants or combined with hydroxides/water to yield hydroxyl radicals (OH) (Equation (3)) [55]. The collective action of O2•−, OH, and hole–electron pairs contributed to the removal of MB (Equation (4)).
[Mo6+ − O2− − V5+] → [Mo6+ − O2− − V4+]*/[Mo5+ − O2− − V5+]* (h+ + e)
e + O2 → O2•−
h+ + H2O → OH + H+
MB + h+, OH, O2•−→ degradation intermediates → CO2 + H2O
The enhancement of photocatalytic properties can be achieved through the immobilization of POMs onto specific supports. The Keggin-type POM (NH4)4[PMo11VO40] was immobilized onto the surface of g-C3N4 through the dipping technique, resulting in the fabrication of (NH4)4[PMo11VO40]/g-C3N4 (PMo11V/g-C3N4) [56]. In comparison to g-C3N4, the band intensity of PMo11V/g-C3N4 exhibited an extension from 200 to 500 nm to 200 to 580 nm. Notably, when exposed to visible light irradiation, the removal efficiency reached 94.7% within 120 min using 400 mg/L of PMo11V/g-C3N4 and 10 mg/L of MB (Table 1). In contrast, the corresponding removal efficiency of MB were 29.7% for single g-C3N4 and 50.8% for (NH4)4[PMo11VO40] under visible light irradiation. Benefitting from the favorable electronic properties of (NH4)4[PMo11VO40], the combination of (NH4)4[PMo11VO40] with g-C3N4 not only facilitated the generation of holes and electrons but also contributed to the attenuation of hole–electron recombination [56]. This synergy was pivotal in the formation and utilization of h+, e, OH, and O2•−, thus facilitating the efficient degradation of MB. Similarly, the incorporation of P2W18Sn3 into Nd-TiO2 through one-pot synthesis induced a blue shift in absorption, intensifying the ultraviolet absorption capacity [57]. Impressively, a mere 5 min exposure to UV light resulted in 91.0% MB removal (10 mg/L) when subjected to P2W18Sn3/Nd-TiO2 (Table 1) [57]. Although both PMo11V/g-C3N4 [56] and P2W18Sn3/Nd-TiO2 [57] have achieved high removal efficiencies, there was no indication of stability.
POMOFs, formed by integrating POMs into metal–organic frameworks (MOFs), not only embody the versatile catalytic traits intrinsic to POMs but also encompass the amplified porosity, expansive surface area, and modifiable pore dimensions characteristic of MOFs [58,59,60,61]. Mn-BTC@Ag5[BW12O40] (Mn-BTC@Ag5[BW12]; BTC: 1,3,5-benzenetricarboxylic acid) was synthesized using the grinding method, resulting in a core-shell structure [62]. It achieved 95.6% MB removal within 140 min, utilizing 500 mg/L of Mn-BTC@Ag5[BW12] and 15 mg/L of MB under UV irradiation. Notably, even after undergoing five cycles, the removal efficiency of MB remained unchanged (Table 1) [62]. Upon exposure to UV light, electrons transitioned from the valence band (VB) of both Mn-BTC and Ag5[BW12] to their conduction band (CB), resulting in the creation of electron–hole pairs (Figure 1) [63]. Additionally, photogenerated holes migrated from the VB of Mn-BTC to that of BW12 due to the latter′s more positive VB potential. Furthermore, electron migration occurred from the CB of BW12 to that of Mn-BTC, which was facilitated by Mn-BTC’s more negative potential [64]. This process induced the separation of energy levels and reduced the electron transfer time [65]. Among the generated active species, OH played a pivotal role in the degradation (Figure 1). Using the same synthesis method as Mn-BTC@Ag5[BW12], Ag5[BW12]@[Ag3(µ-HBTC)(µ-H2BTC)]n (termed Ag-BTC@Ag5[BW12]) [66] and [Zn4(BTC)24-O)(H2O)2]@Ag5[BW12] (denoted as Zn-BTC@Ag5[BW12]) [67] were obtained with a core–shell structure. Ag-BTC@Ag5[BW12] achieved a 94.4% MB removal [66] and Zn-BTC@Ag5[BW12] [67] yielded a 96.1% removal of MB under the same conditions compared to that of Mn-BTC@Ag5[BW12] (Table 1). Both catalysts demonstrated outstanding activity and stability (Table 1). The 3D host–guest POMOF ([Ag5(pz)6(H2O)4[BW12], denoted as [Ag5(pz)6][BW12]) was synthesized through hydrothermal reaction for MB removal [68]. Under UV irradiation, using 500 mg/L of [Ag5(pz)6][BW12] and 15 mg/L of MB, 93.2% of the MB was removed in 140 min (Table 1), and the removal efficiency was almost unchanged after five cycles [68]. Due to the charge transfer absorption occurring from oxygen to metals within the shorter wavelength range, the photoresponsive attributes of POMs predominantly manifest in the ultraviolet range (λ = 200−400 nm) [40]. This limitation has restricted their utility in solar-driven photocatalysis [69,70]. However, by coupling with visible light-absorbing materials, an avenue is established to enhance the visible light absorption of POMs [71,72,73]. [Hpip]2.5[BW12], formed through inserting BW12O405− clusters into the cavities of Hpip groups through heating reflux synthesis, removed 97.1% of MB (20 mg/L) within 24 min under visible light irradiation, which decreased to 90.0% after five cycles (Table 1) [74]. In addition to BW12O405−, the AsIII-capped Keggin arsenomolybdate with four V4+ substitutions ([As2IIIAsVMo8VIV4IVO40]5−) was also used as the template to produce the POMOF ([As2IIIAsVMo8VIV4IVO40]2[CuICu2II(pz)4]2, denoted as [Cu3(pz)4]2[As3Mo8V4]2; pz = pyrazine) by hydrothermal method [75]. [Cu3(pz)4]2[As3Mo8V4]2 achieved 97.8% removal efficiency within 120 min using 400 mg/L of [Cu3(pz)4]2[As3Mo8V4]2 and 10 mg/L of MB under UV irradiation (Table 1), and it reached 94.3% after five cycles [75]. Large channels constructed within {Cu-pz} MOFs decreased spatial site resistance, offering favorable sites for AsIII-capped Keggin arsenomolybdate accommodation [75]. In accordance with the intrinsic photocatalytic mechanism of POMs, photogenerated electrons migrated from the HOMO of oxygen to the LOMO of molybdenum and vanadium under UV irradiation. Concurrently, [As2IIIAsVMo8VIV4IVO40]5− transitioned into its excited state, leading to the generation of electron–hole pairs, OH, and O2•− for MB degradation.
An organic–inorganic hybrid material based on Mo8O26 (Cu2(L1)2(Mo8O26)0.5; L1:5-(pyrimidyl)tetrazolate) was synthesized via a hydrothermal reaction [76]. It achieved a significant 97.9% removal of MB (3.2 mg/L) within 7 min with 105 mg/L of Cu2(L1)2(Mo8O26)0.5, 200 mg/L of H2O2 (Table 1) [76]. After four cycles, there was only about a 1.9% decrease in removal efficiency (Table 1). Beyond the excitation of Cu2(L1)2(Mo8O26)0.5 that generated hole–electron pairs for active species formation (Equations (5)–(9)), MB also possessed the capability to reach an excited state (Equation (10)) and participated in a reaction with O2 to produce 1O2 (Equation (11)). Consequently, the subsequent decomposition of MB occurred under the influence of these active species (Equation (12)) [77]. With the same synthesis method as that of Cu2(L1)2(Mo8O26)0.5, 3D CuII hybrid materials based on Mo8O264−, termed Cu2(L2)3(Mo4O13)2 (with L2 indicating 2,6-(1,2,4-triazole-4-yl)pyridine), removed 99.1% of the MB (3.2 mg/L) in 90 min under UV irradiation (Table 1) [78]. And the performance remained stable without significant decreases after five cycles. This UV-induced irradiation initiated a charge transfer within the N-donor ligand, leading to the generation of OH and O2•− through interactions involving POM/organic ligands [79]. Ce2(BINDI)(Mo6O19) (H4BINDI: N,N′-bis(5-isophthalic acid)-1,4,5,8 -naphthalenediimide), obtained by heating reflux synthesis, removed 96.0% of MB (3.2 mg/L) after a 27 min visible irradiation period and the removal efficiency remained unchanged after 3 cycles (Table 1) [80]. Mo6O192− existed within the e-deficient NDIs plane, acting as a reservoir for electrons and facilitating charge transfer. Additionally, core–shell-shell Fe3O4/Ag/POMs, composed of Fe3O4/Ag/[Cu(PCA)2(H2O)]H2[Cu(PCA)2(P2Mo5O23)]·4H2O (abbreviated as Fe3O4/Ag/Cu2(PCA)4(P2Mo5), PCA: pyridine-2-carboxamide), which was synthesized via the sonochemical method, removed 98.7% of the MB within 26 min under visible irradiation (Table 1) [81]. It decreased only by about 1.2% after six repeated tests. Ag functioned as a photogenerated electron acceptor prompted the interaction between chemisorbed molecular oxygen and photogenerated electrons, consequently yielding the creation of O2•− [82]. This interaction notably contributed to the effective capture of photo-generated electrons.
Cu2(L1)2(Mo8O26)0.5 + UV → Cu2(L1)2(Mo8O26)0.5 (e + h+)
Cu2(L1)2(Mo8O26)0.5 (e + h+) + H2O → Cu2(L1)2(Mo8O26)0.5 (e) + OH + H+
Cu2(L1)2(Mo8O26)0.5 (e) + H2O2 + H+ → Cu2(L1)2(Mo8O26)0.5 + OH + H2O
Cu2(L1)2(Mo8O26)0.5 (e) + O2 → Cu2(L1)2(Mo8O26)0.5 + O2•−
Cu2(L1)2(Mo8O26)0.5 (e) + O2 + H+ → Cu2(L1)2(Mo8O26)0.5 + O2H
MB + UV → MB*
MB* + O2 → MB + 1O2
MB + 1O2, OH, O2H/O2•− → degradation intermediates → CO2 + H2O
Covalent organic frameworks (COFs) present another avenue for incorporating POMs into the framework to facilitate the photocatalytic activity of POMs. P@Ni-AndCOF and P@Cu-AndCOF were obtained through solvothermal synthesis using 4-functionalized porphyrins and Anderson-POM ([N(Bu)4]4[α-Mo8O26]), which eliminated 79.3% (P@Ni-AndCOF) and 89.0% (P@Cu-AndCOF) of the MB over a 660 min UV-Vis light irradiation period (Table 1) [83]. The heightened photocurrent response of P@Ni-AndCOF potentially accounted for its superior activity [84]. The pronounced electron-attracting ability of the Anderson-POM facilitated the transfer of excited electrons within P@M-AndCOF towards the Anderson-POM component, where the Anderson-POM fulfilled a pivotal role in electron storage during the photocatalysis process [37]. Subsequently, photogenerated electrons engaged with O2 to generate O2•−, while h+ interacted with H2O to give rise to OH, thereby facilitating the effective decomposition of MB.
Among the extensive array of POMs, a distinct subset is represented by the reduced forms, known as heteropoly blues, which are characterized by their heightened electron density and broader spectral absorption, encompassing both the visible and near-infrared regions [85]. An particular example is the “hourglass” species [M(P4Mo6O31)2] (abbreviated as [M(P4Mo6)]), which is formed by linking two identical [P4Mo6O31]12 units through bridging metal centers (M), thus featuring all reduced Mo atoms in a valence state of (+5) [86]. Based on this premise, a multifunctional photocatalyst of the “hourglass-type” POM ((H2bpp)2{[Na4(H2O)5][Co0.8Cd0.2(H2O)2][Cd[Mo6O12(OH)3(H2PO4)(HPO4)(PO4)2]2]}·2H2O, abbreviated as (Hbpp)2CoCd(P4Mo6)2) was obtained through a hydrothermal reaction for MB oxidation and Cr6+ reduction using visible light irradiation [85]. It achieved a 96.0% removal of MB (32 mg/L) and a 74.0% reduction in Cr6+ with 2000 mg/L catalyst over 180 min (Table 1), and the removal efficiency remained unchanged after five photocatalytic cycles. (Hbpp)2CoCd(P4Mo6)2 was irradiated to generate electrons and holes, which subsequently migrated to its surface and participated in redox reactions (Figure 2). Photogenerated electrons reduced Cr6+ to Cr3+, while photogenerated holes initiated the oxidation of MB. Notably, alongside h+, both OH and O2•− contributed to the oxidation of MB. The concerted reduction in Cr6+ and oxidation of MB in this system effectively decreased the recombination of photogenerated electron–hole pairs, thereby fostering the separation of photogenerated charge carriers and optimizing the utilization of active species [87]. In a subsequent study by Huili Guo et al. in 2022 [88], two distinctive 3D Fe2+ frameworks, namely {H(4,4′bipy)2[Fe4(PO4)(H2O)4Na6][Fe6(H2O)4][(Mo6O12)(HPO4)3(PO4) (OH)3]4·5H2O} (Fe4Fe6(P4Mo6)2) and {H3(C12H14N2)4[Fe4(PO4)(H2O)4Na4][Fe2(Mo6O12 (HPO4)3(PO4)(OH)3)4]·6H2O} (Fe4Fe2(P4Mo6)2), were synthesized through solvothermal synthesis using P4Mo6 units as the foundations. They achieved outstanding removal of 98.0% ((Fe4Fe6(P4Mo6)2) and 99.0% (Fe4Fe2(P4Mo6)2) for MB (19 mg/L), and 95.6% ((Fe4Fe6(P4Mo6)2) and 82.0% (Fe4Fe2(P4Mo6)2) for Cr6+ (57 mg/L) within 180 min using 1000 mg/L catalyst and 555 mg/L of H2O2 (Table 1). In addition, Fe4Fe6(P4Mo6)2) and (Fe4Fe2(P4Mo6)2 also demonstrated high recyclability with little decrease after ten cycles.
Table 1. Various procedures for photocatalytic degradation of MB with POM-based catalyst.
Table 1. Various procedures for photocatalytic degradation of MB with POM-based catalyst.
CatalystSynthesis MethodIrradiationCatalyst Dosage
(mg/L)
MB
Dosage
(mg/L)
pHTime
(min)
Removal
Efficiency
(%)
Ref.
1stnth
PMo12V3HydrothermalUV50010-6599.3~/5th[53]
PMo11V/g-C3N4DippingVis40010-12094.7-[56]
P2W18Sn3/Nd-TiO2One-Pot
Synthesis
UV-103591.0-[57]
Mn-BTC@Ag5[BW12]GrindingUV50015-14095.6~/5th[62]
Ag-BTC@Ag5[BW12]GrindingUV50015-14094.4~/5th[66]
Zn-BTC@Ag5[BW12]GrindingUV50015-14096.1~/5th[67]
[Ag5(pz)6][BW12]HydrothermalUV50015-14093.2~/5th[68]
[Hpip]2.5[BW12]Heating refluxVis80020-2497.190.0/5th[74]
[Cu3(pz)4]2[As3Mo8V4]2HydrothermalUV40010-12097.894.3/5th[75]
Cu2(L1)2(Mo8O26)0.5HydrothermalUV1053.26.8797.996.0/4th[76]
Cu2(L2)3(Mo4O13)2HydrothermalUV5003.2-9099.1~/5th[78]
Ce2(BINDI)(Mo6O19)Heating refluxVis3003.2-2796.0~/3rd[80]
Fe3O4/Ag/Cu2(PCA)4(P2Mo5)SonochemicalVis166.715-2698.797.5/6th[81]
P@Cu-AndCOF
P@Ni-AndCOF
SolvothermalUV-Vis10050-66089.0
79.3
-[83]
(Hbpp)2CoCd(P4Mo6)2HydrothermalVis2000326.818096.0~/5th[85]
Fe4Fe6(P4Mo6)2
Fe4Fe2(P4Mo6)2
SolvothermalVis100019-18098.0
99.0
~/10th[88]
UV: Ultraviolet irradiation; Vis: visible light irradiation; UV-Vis: ultraviolet- visible light irradiation. 1st: Dye removal efficiency with the catalyst’s first use for degradation; nth: dye removal efficiency after the catalyst has been recycled for n times; ~: dye removal efficiency was almost unchanged after n cycles of catalyst.

2.2. Rhodamine B (RhB)

Frequently employed as a photocatalyst, TiO2 possesses a relatively wide bandgap ranging from 3.0 to 3.2 eV [89,90,91]. Nevertheless, its photocatalytic efficiency can be significantly enhanced through the combination of POMs, which offer customizable electronic characteristics capable of tuning TiO2’s bandgap [92,93,94]. This strategic modulation amplifies the overall photocatalytic performance. K5CoW12/TiO2 (CoW12/TiO2, 5000 mg/L) synthesized by sol-gel/hydrothermal method achieved the complete removal of RhB (15 mg/L) within 30 min at pH = 5 under visible light irradiation (Table 2) [95]. And the performance remained unchanged after four cycles. Within this system, O2•− and photogenerated holes emerged as the primary active species driving RhB degradation [96,97]. Similarly, effective RhB removal under visible light conditions was achieved with a remarkable 98.0% removal within 40 min by employing the Fenton-like heterogeneous catalyst Co-H3PMo12O40/N-TiO2 (Co-PMo12/N-TiO2) (400 mg/L) in conjunction with the addition of 1110 mg/L of H2O2 (Table 2) [98], and Co-PMo12/N-TiO2, obtained through one-pot synthesis, was of great stability (Table 2). Beyond facilitating photoinduced valence electron transitions, POMs also functioned as electron acceptors, effectively curtailing the recombination of photo-generated holes and electrons (Figure 3) [98]. Upon excitation by visible light, valence electrons within N-TiO2 migrated to its CB (Equation (13)). These electrons subsequently underwent rapid transfer to Co-POM, causing it to transition into a reduced state (POM*) (Equation (14)) due to the robust hydrogen bonding between Co-PMo12 and N-TiO2 [77]. This phenomenon substantially suppressed the recombination of photo-generated electrons and holes, thereby facilitating the progression of reactions. During this process, photo-generated electrons reacted with oxygen, leading to the formation of O2•− (Equation (15)). Concurrently, POM* in its reduced state reacted with O2, yielding O2•− and reverting POM* to POM (Equation (16)), thus entering the Co2+/Co3+ cycle (Equations (17) and (18)) [99]. The catalytic effect of Co2+ results in the generation of Co3+ and OH from H2O2 (Equation (18)). This collaborative interplay between OH, holes, and O2•− culminates in the degradation of RhB (Figure 3).
Co-PMo12/N-TiO2 + hv → e + h+
PMo12 + e → PMo12*
O2 + e → O2•−
PMo12* + O2 → PMo12 + O2•−
Co3+ + PMo12* → Co2+ + PMo12
Co2+ + H2O2 → Co3+ +OH + OH
In addition to their photo-responsive attributes, POMs possess the capability to engage with other materials, facilitating the creation of interfaces that enable the manipulation of both electronic structure and the chemical environment at the interface [100]. Shengnan Cai et al. [101] employed Ag3PW12O40 (Ag3PW) in tandem with TiO2 through hydrothermal synthesis method to craft the Z-scheme Ag3PW12O40/TiO2 (Ag3PW12/TiO2) system. This configuration aimed to bolster charge carrier separation and amplify visible-light catalytic activity [101]. Ag3PW12/TiO2 achieved 95.0% removal of RhB (10 mg/L) within 120 min with Ag3PW12/TiO2 (1000 mg/L) under visible irradiation, and the removal efficiency was almost unchanged after four cycles (Table 2). The degradation was primarily governed by h+, accompanied by a minor contribution from O2•− (Figure 4). Under visible irradiation, photogenerated electrons migrated from CB of TiO2 to that of the metallic Ag due to the more negative CB potential of TiO2 in comparison to Ag’s Fermi level (Figure 4). Concurrently, holes within the VB of Ag3PW12 transferred to Ag+, engaging in binding with electrons due to Ag’s more positively positioned Fermi level [102]. Within the configuration of the Z-scheme Ag3PW12/TiO2, Ag played the pivotal role of a bridging agent [103,104,105]. This role effectively facilitated charge transfer through Ag3PW12, thereby promoting the separation of electrons and holes. Expanding beyond the realm of TiO2, the band structure alignment of Ag4SiW12O40 (Ag4SiW12) and Cs3PW12O40 (Cs3PW12) with ZnO is noteworthy [106]. This alignment enabled the creation of ZnO/Ag4SiW12 and ZnO/Cs3PW12 systems, which were synthesized through the precipitation method and achieved 92.3% (ZnO/Ag4SiW) and 72.7% (ZnO/Cs3PW12) of RhB removal within 60 min under simulated sunlight irradiation, whereas the performance of ZnO alone yielded 17.2% removal of RhB (Table 2) [106]. After recycling three times, the removal efficiency decreased to 84.8% (ZnO/Ag4SiW) and 65.3% (ZnO/Cs3PW12), respectively. The CB potential of ZnO was more negative than that of O2/O2•−, leading to the reaction of O2 with electrons on ZnO’s CB, generating O2•− [107,108]. Furthermore, VB potential surpassed that of OH/OH and H2O/OH, thereby fostering the formation of OH on the VB of Cs3PW12/AgPW. In this scenario, electrons on the CB of Cs3PW12/AgPW were directed to the CB of ZnO (Figure 5a,b).
POMOFs, anchored on SiW12O404− (SiW12), exhibit a distinctive potential for the photocatalytic degradation of dyes due to the pronounced negative charges and the reversible multi-electron transfer characteristics inherent to SiW12 [109]. Co2(3,3′-bpy)(3,5′-bpp)(4,3′-bpy)3[SiW12O40] ([Co2(bpy)3][SiW12]); 4,3′-bpy (4,3′-dipyridine); 3,5′-bpp (3,5′-bis(pyrid-4-yl)pyridine); and 3,3′-bpy (3,3′-bis(pyrid-4-yl) dipyridine)), synthesized by the hydrothermal method, achieved a 92.0% removal of RhB (50 mg/L) at 120 min under UV irradiation (Table 2) [109]. The removal efficiency did not decrease significantly after three cycles (Table 2). When irradiated, the SiW12 within [Co2(bpy)3][SiW12] was prompted into an excited state, thereby giving rise to the generation of holes and electrons through charge transfer from O to W in W-O-W ligands. These active species were pivotal for the ensuing RhB degradation. In addition, the hybrid complex, [Cu8Cl5(CPT)8 (H2O)4](HSiW12)(H2O)20(CH3CN)4}n (abbreviated as [Cu8(CPT)8](SiW12)(CH3CN)4; HCPT: 4-(4-carboxyphenyl)-1,2,4-triazole) accomplished 96.6% of RhB removal (10 mg/L) within 70 min by utilizing 3.4 × 104 mg/L of H2O2 under visible irradiation (Table 2) [110]. In contrast, the efficiency was only 46.4% with [Cu8(CPT)8](SiW12)(CH3CN)4 alone. [Cu8Cl5(CPT)8 (H2O)4], synthesized by the solvothermal method, maintained its original structure in comparison to the initial state. Moreover, the activity of [Cu8Cl5(CPT)8 (H2O)4] exhibited a negligible decrease after undergoing four cycles (Table 2). The introduction of H2O2, which is capable of accepting electrons, played a pivotal role in enhancing the degradation [111]. This was attributable to its capture of photogenerated electrons and the reduction in hole–electron pair recombination [112]. However, excessive H2O2 content can potentially quench the generated OH [113].
In parallel with SiW12, the POMOFs rooted in BW12 demonstrated comparable potential for the photocatalytic degradation of RhB. Specifically, core–shell configurations, such as Ag-BTC@Ag5[BW12] [66], Zn-BTC@Ag5[BW12] [67], and Mn-BTC@Ag5[BW12] [62], displayed noteworthy removal efficiencies of 90.6%, 91.1%, and 94.1%, respectively, for RhB (15 mg/L) within 140 min using 500 mg/L of catalyst under UV irradiation (Table 2). Additionally, [Cu(en)2(H2O)][{Cu(pdc)(en)}(Cu(en)2)(BW12)]·2H2O (abbreviated as [Cu3(en)]4(pdc)[BW12]) and [{CuI5(pz)6(H2O)4}(BW12)] (abbreviated as [CuI5(pz)6][BW12]), prepared by the hydrothermal method, achieved RhB removal efficiencies of 91.3% and 92.6%, respectively, when subjected to 150 min of UV light irradiation with 500 mg/L catalyst and 10 mg/L RhB (Table 2) [114]. Though the synthesized methods were different, Ag-BTC@Ag5[BW12] [66], Zn-BTC@Ag5[BW12] [67], Mn-BTC@Ag5[BW12] [62] [Cu3(en)]4(pdc)[BW12]) and [CuI5(pz)6][BW12] were of great stability (Table 2). Notably, [Cu3(pz)4]2[As3Mo8V4]2 accomplished a 97.4% RhB removal (10 mg/L) in 120 min using 400 mg/L of catalyst under UV irradiation (Table 2) [75]. And the removal efficiency of RhB was 92.7% after five cycles, with only a slight decrease of 4.7%.
{P2Mo5}-based POMOFs, exemplified by [H4′4-bipy)2][H2P2Mo5O23] [Cu(4′4-bipy)2]·18H2O (abbreviated as Cu(bipy)4(P2Mo5)), achieved 89.6% removal of RhB (30 mg/L) within 180 min through visible light irradiation (Table 2) [115]. Cu(bipy)4(P2Mo5), obtained through the hydrothermal reaction, exhibited excellent stability, and there was no significant decline in catalytic activity observed even after four cycles. Evidently, the presence of POMs-RhB complexes prior to degradation facilitated the interaction of active species with RhB during the photocatalytic process [116]. Intriguingly, aside from the active species generated, RhB can also undergo degradation through the excited POMs (Figure 6) [115]. Additionally, Cu2(L2)3(Mo4O13)2 removed 98.7% of RhB (4.79 mg/L) within 90 min under UV irradiation (Table 2) [78]. Moreover, P@Ni-AndCOF showcased a 71.3% removal of RhB (100 mg/L) within 660 min under UV-Vis light irradiation (Table 2), while P@Cu-AndCOF yielded an 83.0% removal of RhB (Table 2) [83].
Table 2. Various procedures for photocatalytic degradation of RhB with POM-based catalyst.
Table 2. Various procedures for photocatalytic degradation of RhB with POM-based catalyst.
CatalystSynthesis MethodIrradiationCatalyst Dosage
(mg/L)
RhB
Dosage
(mg/L)
pHTime
(min)
Removal Efficiency
(%)
Ref.
1stnth
CoW12/TiO2Sol-gel/hydrothermalVis500015530100~/4th[95]
Co-PMo12/N-TiO2One-pot synthesisVis4001074098.0~/4th[98]
Ag3PW12/TiO2HydrothermalVis100010-12095.0~/4th[101]
ZnO/Ag4SiW
ZnO/Cs3PW12
PrecipitationSimulated sunlight30050-6092.3
72.7
84.8/3rd
65.3/3rd
[106]
Co2(bpy)3][SiW12]HydrothermalUV40050-12092.0~/3rd[109]
[Cu8(CPT)8](SiW12)(CH3CN)4SolvothermalVis37510-7096.6~/4th[110]
Ag-BTC@Ag5[BW12]GrindingUV50015-14090.6~/5th[66]
Zn-BTC@Ag5[BW12]GrindingUV50015-14091.1~/5th[67]
Mn-BTC@Ag5[BW12]GrindingUV50015-14094.1~/5th[62]
[Cu3(en)]4(pdc)[BW12]
[CuI5(pz)6][BW12]
HydrothermalUV50010-15091.3
92.6
~/5th[114]
[Cu3(pz)4]2[As3Mo8V4]2HydrothermalUV40010-12097.492.7/5th[75]
Cu(bipy)4(P2Mo5)HydrothermalVis30030-18089.6~/4th[115]
Cu2(L2)3(Mo4O13)2HydrothermalUV5004.79-9098.7~/4th[78]
P@Ni-AndCOF
P@Cu-AndCOF
SolvothermalUV-Vis10050-66071.3
83.0
-[83]
Vis: Visible light irradiation; UV: ultraviolet irradiation; UV-Vis: ultraviolet- visible light irradiation. 1st: Dye removal efficiency with the catalyst’s first use for degradation; nth: dye removal efficiency after the catalyst has been recycled for n times; ~: dye removal efficiency was almost unchanged after n cycles of catalyst.

2.3. Methyl Orange (MO)

Keggin-type POMs, particularly H3PMo12O40 (PMo12), characterized by a narrow energy gap (Eg) of 2.4 eV, are well known for its exceptional stability and remarkable visible-light absorption attributes. Pt/PMo12/TiO2 composite nanofibers were fabricated via the electrospinning/calcination route and photoreduction for MO degradation [117,118,119]. Remarkably, it achieved an 88.1% removal of MO within 180 min at pH 1, employing 1000 mg/L of Pt/PMo12/TiO2 and 20 mg/L of MO (Table 3) [119]. The PMo12’s VB potential (3.13 eV) was notably more positive than that of TiO2 (2.91 eV), facilitating the migration of holes generated by PMo12 into TiO2’s domain (Figure 7a). In contrast, the photogenerated electrons did not react with O2 due to PMo12’s higher LUMO value (0.73 V vs. NHE) in comparison to O2/O2•− (−0.046 V vs. NHE). Consequently, photogenerated electrons produced by PMo12 combined with the plasmonic holes within Pt [120,121,122]. Subsequently, the plasma electrons within Pt were directly captured by absorbed O2 on the surface, resulting in the formation of O2•−. MO was degraded through the collaboration of OH and O2•−. Noble metal deposition with Bi further augmented the photocatalytic performance. Bi/PMo12-doped TiO2 (Bi/PMo12/TiO2, denoted as x = 10, 20, and 30), obtained through the electrospinning/calcination and hydrothermal methods, removed 92.5% of MO within 180 min at pH 1 under visible light irradiation, utilizing 1000 mg/L Bi/PMo12/TiO2 and 40 mg/L MO (Table 3) [123]. In comparison, single TiO2, PMo12-doped TiO2 (PMo12/TiO2), and Bi/TiO2 achieved removal rates of 14.6%, 18.2%, and 64.6%, respectively [123]. And both Pt/PMo12/TiO2 [119] and Bi/PMo12/TiO2 [123] were of great stability, with little decreases in activity observed even after five cycles (Table 3). Complete MO removal was realized within 30 min employing 1000 mg/L of Ag/PMo10V2O405−/TiO2 (Ag/PMo10V2/TiO2) and 20 mg/L of MO under visible light irradiation (Table 3) [124]. Ag/PMo10V2/TiO2 was synthesized by heating reflux and photoreduction [124]. Notably, an observed S-type heterojunction within Ag/PMo10V2/TiO2 facilitated the excitation of both PMo10V2 and PMo10V2/TiO2 by visible irradiation (Figure 7b). The photogenerated electrons from PMo10V2’s LUMO migrated to the VB of PMo10V2/TiO2, resulting in the subsequent combination with the holes present in the VB of PMo10V2/TiO2. Concurrently, photoelectrons generated on Ag/PMo10V2-TiO2 transported to the surface of Ag nanoparticles through Schottky junctions, engaging in interaction with O2, leading to the formation of O2•− (Figure 7b). Remarkably, a 98.4% removal efficiency of MO was achieved within just 5 min with Sandwich-type POM/TiO2 (P2W18Sn3/Nd-TiO2) at pH = 3 under UV irradiation, and the removal efficiency was 95.0% after five cycles (Table 3) [57]. MO was anionic, whereas TiO2’s pHpzc was 6.8, rendering it positively charged in acidic conditions (pH < 6.8). Thus, higher response rates were observed in acidic settings owing to the electrostatic attraction between TiO2 and MO [125]. The magnetic microsphere of Fe3O4@SiO2@[TiO2/H3PW12O40]10 (Fe3O4@SiO2@[TiO2/PW12]10) was synthesized at room temperature using a Layer-by-Layer method. It achieved a 83.9% removal of MO under UV irradiation at pH 2, utilizing 2000 mg/L Fe3O4@SiO2@[TiO2/PW12]10 and 10 mg/L MO (Table 3) [126]. In contrast, little degradation occurred with Fe3O4@SiO2 alone. The exceptional performance was attributed to the synergistic action of TiO2 and PW12. And Keggin-typed SiW10 exhibited characteristic absorption below 400 nm, and its absorption peak gradually broadened with increasing Co content [127]. Based on the remarkable UV-Vis absorption of Co2Co4(SiW10O37)2, [Co2Co4(SiW10O37)2/Fe2O3] (Co2Co4(SiW10)2/Fe2O3) achieved a 76.2% removal of MO under UV irradiation within 90 min at pH = 1 (Table 3) [128]. And the removal efficiency reached 69.4% after three cycles, exhibiting a decrease of 6.8%.
POMOFs originating from BW12, namely [Ag5(pz)6][BW12] [68], Ag-BTC@Ag5[BW12] [66] and Zn-BTC@Ag5[BW12] [67], demonstrated exceptional efficacy in the removal of MO at concentrations of 15 mg/L. When subjected to UV irradiation, they achieved removal efficiencies of 90.9%, 92.4%, and 95.2%, respectively, within 140 min, when utilizing a catalyst concentration of 500 mg/L and 15 mg/L of MB (Table 3). Exceptional stability was observed in these three catalysts, as the removal efficiency of MO remained unchanged even after undergoing five cycles (Table 3). Furthermore, [Cu3(pz)4]2[As3Mo8V4]2, built upon on [As2IIIAsVMo8VIV4IVO40]5−, demonstrated an impressive MO removal of 96.8% within 120 min under UV irradiation. And the removal efficiency remained at 91.4% after undergoing five cycles (Table 3) [75].
POMs feature extensive electron-donating conjugation systems, enabling the stabilization of Fe2+ and thus enhancing the functionality of the Fenton system [129,130]. Exploiting this attribute, the Sandwich-type phosphomolybdate [H4MoV6O15(PO4)4]8− was employed to create heterogeneous Fenton-like systems for MO removal [131]. The compound (H3O)3.5(H3DETA)3.5{FeII[H4MoV6O15(PO4)4]2} (abbreviated as (DETA)3.5Fe(P4Mo6)2, EDTA = diethylenetriamine), synthesized through hydrothermal reaction, exhibited remarkable performance by removal of 99.8% of the MO (20 mg/L) within 20 min with the addition of H2O2 (Table 3) [131]. In comparison, the removal efficiency was only 20.0% with H2O2 alone. The presence of metal–oxo clusters in POMs contributes to extensive electron donation, creating a positively charged environment around Fe2+ [129,130]. Consequently, the oxidation of Fe2+ to Fe3+ was significantly inhibited under both non-illuminated and visible light conditions [131]. Specifically, under UV irradiation, the charge transfer from O to Mo (i.e., Fe→O→Mo) reduced the electron density surrounding Fe2+, enabling Fe2+ to react with H2O2 to form Fenton reagents (Figure 8).
Table 3. Various procedures for photocatalytic degradation of MO with POM-based catalysts.
Table 3. Various procedures for photocatalytic degradation of MO with POM-based catalysts.
CatalystSynthesis MethodIrradiationCatalyst Dosage
(mg/L)
MO Dosage
(mg/L)
pHTime
(min)
Removal
Efficiency
(%)
Ref.
1stnth
Pt/PMo12/TiO2Electrospinning/calcination and photoreductionVis100020118088.1~/5th[119]
Bi/PMo12/TiO2Electrospinning/calcination and hydrothermalVis100040118092.5~/8th[123]
Ag/PMo10V2/TiO2Heating reflux
and photoreduction
Vis100020-30100-[124]
P2W18Sn3/Nd-TiO2One-Pot SynthesisUV-103598.495.0/5th[57]
Fe3O4@SiO2@[TiO2/PW12]10Layer-by-Layer methodUV200010210083.9-[126]
Co2Co4(SiW10)2/Fe2O3PrecipitationUV10001019076.269.4/3rd[128]
[Ag5(pz)6][BW12]HydrothermalUV50015-14090.9~/5th[68]
Ag-BTC@Ag5[BW12]GrindingUV50015-14092.4~/5th[66]
Zn-BTC@Ag5[BW12]GrindingUV50015-14095.2~/5th[67]
[Cu3(pz)4]2[As3Mo8V4]2HydrothermalUV40010-12096.891.4/5th[75]
(DETA)3.5Fe(P4Mo6)2HydrothermalUV6020-2099.8-[131]
Vis: Visible light irradiation; UV: ultraviolet irradiation; 1st: dye removal efficiency with the catalyst’s first use for degradation; nth: dye removal efficiency after the catalyst has been recycled for n times; ~: dye removal efficiency was almost unchanged after n cycles of catalyst.

3. Photodegradation Mechanism of Dyes and Enhancement Strategies for POM-Based Materials

3.1. Photodegradation Mechanism of Dyes

Metals within POMs predominantly exhibit a d0 electron configuration, leading to ligand-to-metal charge transfer upon exposure to light irradiation [40]. Moreover, for most POMs, a distinct energy gap emerges between the HOMO of oxygen and the LUMO of metals within metal–oxo clusters. As a result, the activation of most POMs necessitates ultraviolet or near-ultraviolet light irradiation [40]. For instance, in the case of PMo12V3, when photon energy matches or exceeds the bandgap, electrons transition from O2− (2p) orbitals to Mo6+ (5d) orbitals, inducing PMo12V3 to enter an excited state and generate electron–hole pairs [53]. The photogenerated electrons in CB gain high energy and mobility, while photogenerated holes in VB possess potent oxidative capabilities [132,133,134]. If the VB potential of POMs exceeds that of H2O/OH (2.7 V vs. NHE), these photogenerated holes would react with water, producing OH [106,135]. Alternatively, photogenerated electrons engage in reduction reactions, converting O2 into O2•−, facilitated by the more negative potential of CB of POMs compared to that of O2/O2 (−0.046 V vs. NHE) [55,107]. Subsequently, O2•− react with water to generate OH (Figure 9) [39,136]. These dynamic active species initiate the attack and breakdown of dyes, efficiently converting them into smaller, less harmful byproducts, such as carbon dioxide and water [31]. Alongside the generated active species, excited POMs can also contribute to the degradation of dyes [137,138].
The combination of POMs with specific materials presents an alternative for enhancing the light absorption capabilities and photocatalytic activity of POM-based catalysts. For example, the coupling of Anderson-POM ([N(Bu)4]4[α-Mo8O26]) with a visible light-absorbing porphyrin leads to the creation of the photoreactive COF catalyst (P@Ni-AndCOF), which demonstrates visible light catalysis activity [83]. Additionally, the energy band structures, interface characteristics, and charge migration properties can also be modified by integrating POMs with different materials. Notably, the fabrication of a Z-scheme Ag3PW12/TiO2 (Figure 4) and the establishment of an S-type heterojunction in Ag/PMo10V2/TiO2 (Figure 7b) [101,124]. Upon light excitation, both POMs and carriers generate electron–hole pairs [39,139]. The electrons originating from the CB of POMs migrate to VB of the partnering carriers and combine with its holes. Concurrently, holes in VB of POMs remain localized within the POMs structure, participating in oxidation processes or promoting the generation of OH. Alternatively, photogenerated electrons of the partnering carriers combine with POMs’ holes. Then, the photogenerated electrons on CB engage in the formation of O2•− [140].
Despite the removal of dyes in photocatalytic systems, the degradation pathways, and the specific roles of reactive oxygen species (ROS), such as h+, OH and O2•−, in their degradation are challenging to ascertain. These dyes have complex organic structures, making it difficult to analyze their degradation pathways as chemical bonds and interactions among functional groups can lead to various reactions. Additionally, degradation pathways are influenced by environmental conditions (such as temperature, pH, etc.), degradation methods, and catalysts. Variations in these factors can result in different degradation pathways, increasing the complexity of the research. Consequently, different researchers may propose different degradation mechanisms. For example, in the case of the degradation of Rhodamine B, Pengxiang Lei et al. [141] suggested that within the POM(Na3PW12O40)-resin/H2O2/visible light system, it underwent degradation through intermediates like N-ethyl-N′-ethyl-rhodamine, N, N-diethyl-N′-ethylrhodamine, and N, N-diethyl-rhodamine. While Zhentao Yu et al. [142] proposed a stepwise degradation mechanism involving the gradual loss of C2H5 units, facilitated by intermediates such as N, N′,N′-triethylrhodamine; N, N′-diethylrhodamine; and N-ethylrhodamine. The lack of consistent results not only complicates the determination of the exact degradation pathways but also hinders the understanding of the precise roles of ROS and POMs in the degradation process.
POMs, despite their diverse structures, typically possess small specific surface areas (1–10 m2·g−1) [33]. Consequently, specific surface area seems to have a minimal impact on their adsorption performance. For example, in the case of Ag5BW12 and PMo11V, under dark conditions, their removal efficiencies for MB were 3.8% (Ag5BW12) [62] and 8.0% (PMo11V) [56], respectively, with minimal variation observed. Nevertheless, when examining various POMs types, significant disparities in photocatalytic activity emerge. Notably, under visible light irradiation, Keggin-type Ag5[BW12O40] degraded 39.3% of the RhB within 140 min [62], while Dawson-type KNa-P8W48 accomplished a 91.0% removal of the RhB within 180 min [143]. Additionally, Keggin-type PMo12V3 achieved 99.3% MB removal after 65 min of UV irradiation [53], while Keggin-type PMo11V obtained 50.8% MB removal efficiency after 120 min of visible light exposure [56]. These findings underscore that the composition and structure of POMs are pivotal in governing their photocatalytic activity.

3.2. Enhancement Strategies for POM-Based Materials

By employing strategic material combinations and thoughtful interface design, it can not only to augment the specific surface area of POM-based materials but also to enhance the interaction between substrates and active sites. Additionally, effective electron–hole separation can be achieved, resulting in a significant improvement in photocatalytic efficiency. In accordance with the functional mechanism, the primary strategies for achieving this enhancement are outlined as follows:
(1)
Incorporation onto Suitable Carriers: The introduction of POMs into appropriate carriers, such as SiO2, provides a stable scaffold that enhances stability and activity [126,138,144];
(2)
Surface Modification: After immobilizing POMs molecules onto carrier surfaces, the introduction of light-absorbing groups or the adjustment of surface chemistry enhances the POMs’ absorption capacity, even within the visible light range [66,67,83];
(3)
Fabrication of Composite Materials: The combination of POMs with other optically active materials, such as semiconductor nanoparticles (TiO2 and ZnO), results in synergistic composite systems that elevate visible light absorption and electron transfer efficiency [95,98,106];
(4)
Introduction of Conjugated Structures: The incorporation of conjugated structures, such as benzene rings or carbazole moieties, into the POMs framework extends its light-absorption range, enabling broader visible light utilization [145];
(5)
Nano materialization: The transformation of POMs molecules into nanoparticles increases the surface area, thereby enhancing light absorption efficiency [119,124].
POM-based materials’ photocatalytic activity and stability depend on both compositions and synthesis methods. Common methods include the following:
(1)
Dipping: Simple but may result in uneven impregnation, especially for large materials.
(2)
Grinding: Easy but challenging to control particle size, leading to non-uniformity of materials.
(3)
Heating Reflux: Versatile for various reactions, offering precise temperature control but requiring reflux equipment and involving intricate procedures.
(4)
One-Pot Synthesis: Suitable for complex multi-component materials, saving time and resources by avoiding intermediate steps, but necessitates precise reaction control and may produce byproducts.
(5)
Hydrothermal and Solvothermal Methods: Ideal for synthesizing crystals, nanoparticles, and complex structures, with control over material size and shape. However, they typically involve high-temperature and high-pressure conditions.
The choice of the most suitable synthesis method depends on the desired material properties and the specific experimental conditions.

4. Application Potential of POM-Based Materials

Factors such as intricate synthesis, purity requirements, and material specificity may result in higher costs for POM-based photocatalysts compared to simpler counterparts. However, in situations where catalytic activity, selectivity, and long-term stability are pivotal considerations, the performance advantages and potential long-term benefits of POMs position them as the preferred choice for photocatalysts. These advantages can be assessed by researchers and industries to ascertain the cost-effectiveness of utilizing POM-based materials in particular applications.

5. Conclusions and Perspectives

The utilization of POM-based photocatalysis presents a promising avenue for addressing water contamination issues through effective dye degradation. The distinctive optical properties, diverse catalytic sites, and adaptable energy levels inherent to POMs offer a compelling platform for efficient solar-driven catalysis across a broad spectrum, including visible light. The presence of metal–oxygen clusters and vacant d orbitals within transition metals facilitates electron transitions from oxygen to metals upon light exposure, resulting in the generation of electron–hole pairs. These pairs serve in the direct oxidation of dyes or react with H2O/O2 to produce OH/O2•− and other oxidative species.
The synergy achieved through the coupling of POMs with other materials in heterojunctions enhances light absorption and photocatalytic efficiency. Notably, the adjustment of band structures and interface electronic states within heterojunctions facilitates improved electron and hole migration, thereby reducing charges recombination and bolstering the photocatalytic efficiency. Beyond the realm of heterojunctions, the integration of POMs with materials like TiO2 and organic-inorganic hybrid frameworks like MOFOF and COFs holds promise for enhancing light-absorption efficiency and catalytic activity. Strategic approaches such as carrier incorporation, surface modification, composite material fabrication, the introduction of conjugated structures, and nanomaterialization contribute to the amplification of light utilization and efficiency.
As the field progresses, future research necessitates more experimental investigations and the development of theoretical models. A thorough exploration into the generation and action of ROS is essential to uncover the precise mechanism of dye degradation. Furthermore, as the photocatalytic performance advances, additional research may be required to clarify the precise relationship between the composition, structure, and photocatalytic activity of POMs. Additionally, translating laboratory successes into practical applications becomes crucial. Scalable synthesis methods, material stability assessment under real conditions, and effective integration into water treatment systems are paramount. Efficient and practical POM-based water remediation demands interdisciplinary collaboration and a deep understanding of materials science, catalysis, and environmental engineering. By tackling challenges and seizing opportunities, POM-based photocatalysis can significantly influence the advancement of water purification technologies.

Author Contributions

Conceptualization, J.S. and D.Z.; methodology, D.Z., C.S., Y.W. and F.Z.; data curation, R.L.; writing—original draft preparation, R.L.; writing—review and editing, R.L. and D.Z.; visualization, R.L.; supervision, J.S., D.Z. and W.X.; project administration, D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (52000002) and Jilin Provincial Science and Technology Department (YDZJ202101ZYTS037).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Benkhaya, S.; M’rabet, S.; El Harfi, A. A review on classifications, recent synthesis and applications of textile dyes. Inorg. Chem. Commun. 2020, 115, 107891. [Google Scholar] [CrossRef]
  2. Ferreira, E.S.B.; Hulme, A.N.; McNab, H.; Quye, A. The natural constituents of historical textile dyes. Chem. Soc. Rev. 2004, 33, 329–336. [Google Scholar] [CrossRef]
  3. Raman, C.D.; Kanmani, S. Textile dye degradation using nano zero valent iron: A review. J. Environ. Manag. 2016, 177, 341–355. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, J.; Li, Q.; Li, W.; Li, Y.; Wang, G.; Li, A.; Li, H. Efficient removal of acid dyes using permanent magnetic resin and its preliminary investigation for advanced treatment of dyeing effluents. J. Clean. Prod. 2020, 251, 119694. [Google Scholar] [CrossRef]
  5. Cheng, J.; Wu, S.; Zhang, L.; Zhou, P.; Zhong, Y.; Xu, H.; Mao, Z. Oversized macroporous flower-like Cu9S8 used as an efficient peroxidase mimetic enzyme for the degradation of organic dyes. J. Phys. Chem. Solids 2022, 163, 110534. [Google Scholar] [CrossRef]
  6. Chang, J.S.; Lin, C.Y. Decolorization kinetics of a recombinant Escherichia coli strain harboring azo-dye-decolorizing determinants from Rhodococcus sp. Biotechnol. Lett. 2001, 23, 631–636. [Google Scholar] [CrossRef]
  7. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef]
  8. Didier de Vasconcelos, G.M.; Mulinari, J.; de Arruda Guelli Ulson de Souza, S.M.; Ulson de Souza, A.A.; de Oliveira, D.; de Andrade, C.J. Biodegradation of azo dye-containing wastewater by activated sludge: A critical review. World J. Microbiol. Biotechnol. 2021, 37, 101. [Google Scholar] [CrossRef]
  9. Chen, Y.; Sun, R.; Yan, W.; Wu, M.; Zhou, Y.; Gao, C. Antibacterial polyvinyl alcohol nanofiltration membrane incorporated with Cu(OH)2 nanowires for dye/salt wastewater treatment. Sci. Total Environ. 2022, 817, 152897. [Google Scholar] [CrossRef]
  10. Holkar, C.R.; Jadhav, A.J.; Pinjari, D.V.; Mahamuni, N.M.; Pandit, A.B. A critical review on textile wastewater treatments: Possible approaches. J. Environ. Manag. 2016, 182, 351–366. [Google Scholar] [CrossRef]
  11. Akpomie, K.G.; Conradie, J. Advances in application of cotton-based adsorbents for heavy metals trapping, surface modifications and future perspectives. Ecotoxicol. Environ. Saf. 2020, 201, 110825. [Google Scholar] [CrossRef]
  12. Ghosh, T.; Ullah, K.; Nikam, V.; Park, C.Y.; Meng, Z.D.; Oh, W.C. The characteristic study and sonocatalytic performance of CdSe–graphene as catalyst in the degradation of azo dyes in aqueous solution under dark conditions. Ultrason. Sonochem. 2013, 20, 768–776. [Google Scholar] [CrossRef] [PubMed]
  13. Haghighi, P.; Hekmati, M.; Ziyadi, H.; Ghasemi, E.; Esmaeili, D. Hibiscus sabdariffa extract modified magnetic polymer nanocomposite for azo dyes removal from aqueous samples. Mater. Chem. Phys. 2021, 267, 124608. [Google Scholar] [CrossRef]
  14. Chakravarthi, B.; Mathkala, V.; Palempalli, U.M.D. Degradation and detoxification of congo red azo dye by immobilized laccase of streptomyces sviceus. J. Pure Appl. Microbiol. 2021, 15, 864–876. [Google Scholar] [CrossRef]
  15. Li, Y.; Cao, P.; Wang, S.; Xu, X. Research on the treatment mechanism of anthraquinone dye wastewater by algal-bacterial symbiotic system. Bioresour. Technol. 2022, 347, 126691. [Google Scholar] [CrossRef]
  16. Murphree, S.S. Heterocyclic dyes: Preparation, properties, and applications. Prog. Heterocycl. Chem. 2011, 22, 21–58. [Google Scholar]
  17. Liang, L.; Cheng, L.; Zhang, Y.; Wang, Q.; Wu, Q.; Xue, Y.; Meng, X. Efficiency and mechanisms of rhodamine B degradation in Fenton-like systems based on zero-valent iron. RSC Adv. 2020, 10, 28509–28515. [Google Scholar] [CrossRef] [PubMed]
  18. Sriram, G.; Bendre, A.; Mariappan, E.; Altalhi, T.; Kigga, M.; Ching, Y.C.; Jung, H.Y.; Bhaduri, B.; Kurkuri, M. Recent trends in the application of metal-organic frameworks (MOFs) for the removal of toxic dyes and their removal mechanism—A review. Sustain. Mater. Technol. 2022, 31, e00378. [Google Scholar] [CrossRef]
  19. Gad, H.M.H.; El-Sayed, A.A. Activated carbon from agricultural by-products for the removal of rhodamine-B from aqueous solution. J. Hazard. Mater. 2009, 168, 1070–1081. [Google Scholar] [CrossRef] [PubMed]
  20. Lafi, R.; Hafiane, A. Removal of methyl orange (MO) from aqueous solution using cationic surfactants modified coffee waste (MCWs). J. Taiwan Inst. Chem. Eng. 2016, 58, 424–433. [Google Scholar] [CrossRef]
  21. Litefti, K.; Freire, M.S.; Stitou, M.; González-Álvarez, J. Adsorption of an anionic dye (congo red) from aqueous solutions by pine bark. Sci. Rep. 2019, 9, 16530. [Google Scholar] [CrossRef]
  22. Verma, N.; Yadav, S.; Marí, B.; Mittal, A.; Jindal, J. Synthesis and charcterization of coupled ZnO/SnO2 photocatalysts and their activity towards degradation of cibacron red dye. Trans. Indian Ceram. Soc. 2018, 77, 1–7. [Google Scholar] [CrossRef]
  23. Waghchaure, R.H.; Adole, V.A.; Jagdale, B.S. Photocatalytic degradation of methylene blue, rhodamine B, methyl orange and eriochrome black T dyes by modified ZnO nanocatalysts: A concise review. Inorg. Chem. Commun. 2022, 143, 109764. [Google Scholar] [CrossRef]
  24. Maruthanayagam, A.; Mani, P.; Kaliappan, K.; Chinnappan, S. In vitro and in silico studies on the removal of methyl orange from aqueous solution using Oedogonium subplagiostomum AP1. Water Air Soil Pollut. 2020, 231, 232. [Google Scholar] [CrossRef]
  25. Aljuaid, A.; Almehmadi, M.; Alsaiari, A.A.; Allahyani, M.; Abdulaziz, O.; Alsharif, A.; Alsaiari, J.A.; Saih, M.; Alotaibi, R.T.; Khan, I. g-C3N4 based photocatalyst for the efficient photodegradation of toxic methyl orange dye: Recent modifications and future perspectives. Molecules 2023, 28, 3199. [Google Scholar] [CrossRef] [PubMed]
  26. Lyu, H.; Gao, B.; He, F.; Zimmerman, A.R.; Ding, C.; Tang, J.C.; Crittenden, J.C. Experimental and modeling investigations of ball-milled biochar for the removal of aqueous methylene blue. Chem. Eng. J. 2018, 335, 110–119. [Google Scholar] [CrossRef]
  27. Wan, X.Y.; Zhan, Y.Q.; Long, Z.H.; Zeng, G.Y.; He, Y. Core@ double-shell structured magnetic halloysite nanotube nano-hybrid as efficient recyclable adsorbent for methylene blue removal. Chem. Eng. J. 2017, 330, 491–504. [Google Scholar] [CrossRef]
  28. Hu, X.S.; Liang, R.; Sun, G.X. Super-adsorbent hydrogel for removal of methylene blue dye from aqueous solution. J. Mater. Chem. A 2018, 6, 17612–17624. [Google Scholar] [CrossRef]
  29. Li, F.; Chen, Y.; Huang, H.; Cao, W.; Li, T. Removal of rhodamine B and Cr(VI) from aqueous solutions by a polyoxometalate adsorbent. Chem. Eng. Res. Des. 2015, 100, 192–202. [Google Scholar] [CrossRef]
  30. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2019, 27, 2522–2565. [Google Scholar] [CrossRef]
  31. Khalid, N.R.; Majid, A.; Tahir, M.B.; Niaz, N.A.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar] [CrossRef]
  32. Hasanpour, M.; Hatami, M. Photocatalytic performance of aerogels for organic dyes removal from wastewaters: Review study. J. Mol. Liq. 2020, 309, 113094. [Google Scholar] [CrossRef]
  33. D’Cruz, B.; Amin, M.O.; Al-Hetlani, E. Polyoxometalate-based materials for the removal of contaminants from wastewater: A review. Ind. Eng. Chem. Res. 2021, 60, 10960–10977. [Google Scholar] [CrossRef]
  34. Müller, A.; Roy, S. Oxomolybdates: From structures to functions in a new era of nanochemistry. In The Chemistry of Nanomaterials: Synthesis, Properties and Applications, 2nd ed.; Rao, C.N.R., Müller, A., Cheetham, A.K., Eds.; Wiley: Hoboken, NJ, USA, 2004; pp. 452–475. ISBN 9783527306862. [Google Scholar]
  35. D’Cruz, B.; Samuel, J.; Sreedhar, M.K.; George, L. Green synthesis of novel polyoxoanions of tungsten containing phosphorus as a heteroatom: Characterization, non-isothermal decomposition kinetics and photocatalytic activity. New J. Chem. 2014, 38, 5436–5444. [Google Scholar] [CrossRef]
  36. Hutin, M.; Rosnes, M.H.; Long, D.L.; Cronin, L. Polyoxometalates: Synthesis and structure–from building blocks to emergent materials. In Comprehensive Inorganic Chemistry II, 2nd ed.; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 241–269. ISBN 9780080965291. [Google Scholar]
  37. Gumerova, N.I.; Rompel, A. Synthesis, structures and applications of electron-rich polyoxometalates. Nat. Rev. Chem. 2018, 2, 0112. [Google Scholar] [CrossRef]
  38. Carraro, M.; Gross, S. Hybrid materials based on the embedding of organically modified transition metal oxoclusters or polyoxometalates into polymers for functional applications: A review. Materials 2014, 7, 3956–3989. [Google Scholar] [CrossRef]
  39. Lan, J.; Wang, Y.; Huang, B.; Xiao, Z.; Wu, P. Application of polyoxometalates in photocatalytic degradation of organic pollutants. Nanoscale Adv. 2021, 3, 4646–4658. [Google Scholar] [CrossRef]
  40. Suzuki, K.; Mizuno, N.; Yamaguchi, K. Polyoxometalate photocatalysis for liquid-phase selective organic functional group transformations. ACS Catal. 2018, 8, 10809–10825. [Google Scholar] [CrossRef]
  41. Stuckart, M.; Monakhov, K.Y. Polyoxometalate encapsulation into metal–organic frameworks: The way towards functional nanomaterials for water splitting. J. Mater. Chem. A 2018, 6, 17849–17853. [Google Scholar] [CrossRef]
  42. Lai, S.Y.; Ng, K.H.; Cheng, C.K.; Nur, H.; Nurhadi, M.; Arumugam, M. Photocatalytic remediation of organic waste over Keggin-based polyoxometalate materials: A review. Chemosphere 2021, 263, 128244. [Google Scholar] [CrossRef]
  43. Wang, H.; He, W.; Dong, X.; Wang, H.; Dong, F. In situ FT-IR investigation on the reaction mechanism of visible light photocatalytic NO oxidation with defective g-C3N4. Sci. Bull. 2018, 63, 117–125. [Google Scholar] [CrossRef]
  44. Wang, X.L.; Rong, X.; Lin, H.Y.; Liu, D.N.; Wang, X.; Liu, G.C.; Song, G. Metal ions induced various polymolybdate-based metal–organic complexes with a pyridyl-amide-carboxylate ligand: Synthesis, structures and selective separation of cationic dyes. Polyhedron 2017, 126, 92–99. [Google Scholar] [CrossRef]
  45. Liu, L.; Wang, B.; Lv, J.H.; Yu, K.; Wang, L.; Zhang, H.; Wang, S.; Zhou, B.B. One-step synthesis of two Wells–Dawson arsenotungstate hybrids via M–O–M bridges for efficient adsorption and selective separation of organic pollutants. CrystEngComm 2017, 19, 5653–5661. [Google Scholar] [CrossRef]
  46. Yi, F.Y.; Zhu, W.; Dang, S.; Li, J.P.; Wu, D.; Li, Y.H.; Sun, Z.M. Polyoxometalates-based heterometallic organic–inorganic hybrid materials for rapid adsorption and selective separation of methylene blue from aqueous solutions. Chem. Commun. 2015, 51, 3336–3339. [Google Scholar] [CrossRef]
  47. Wang, M.; Tan, G.; Dang, M.; Wang, Y.; Zhang, B.; Ren, H.; Lv, L.; Xia, A. Dual defects and build-in electric field mediated direct Z-scheme W18O49/g-C3N4−x heterojunction for photocatalytic NO removal and organic pollutant degradation. J. Colloid Interface Sci. 2021, 582, 212–226. [Google Scholar] [CrossRef]
  48. Fujimoto, S.; Cameron, J.M.; Wei, R.J.; Kastner, K.; Robinson, D.; Sans, V.; Newton, G.N.; Oshio, H. A simple approach to the visible-light photoactivation of molecular metal oxides. Inorg. Chem. 2017, 56, 12169–12177. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, Y.; Luo, F.; Liu, S.M.; Liu, S.X.; Lai, X.Y.; Li, X.H.; Lu, Y.; Li, Y.G.; Hu, C.W.; Shi, Z.; et al. Aminated graphene oxide impregnated with photocatalytic polyoxometalate for efficient adsorption of dye pollutants and its facile and complete photoregeneration. Small 2017, 13, 1603174. [Google Scholar] [CrossRef] [PubMed]
  50. Jiang, J.; Wang, X.; Liu, Y.; Ma, Y.H.; Li, T.R.; Li, Y.H.; Xie, T.F.; Dong, S.S. Photo-Fenton degradation of emerging pollutants over Fe-POM nanoparticle/porous and ultrathin g-C3N4 nanosheet with rich nitrogen defect: Degradation mechanism, pathways, and products toxicity assessment. Appl. Catal. B 2020, 278, 119349. [Google Scholar] [CrossRef]
  51. Li, D.D.; Ma, P.T.; Niu, J.Y.; Wang, J.P. Recent advances in transition-metal-containing Keggin-type polyoxometalate-based coordination polymers. Coord. Chem. Rev. 2019, 392, 49–80. [Google Scholar] [CrossRef]
  52. Srivani, A.; Venkateswara Rao, K.T.; Prasad, P.S.; Lingaiah, N. Role of vanadium in Keggin heteropoly molybdate supported on titania catalysts for oxidation reactions. J. Chem. Sci. 2014, 126, 467–472. [Google Scholar] [CrossRef]
  53. Li, F.R.; Ji, T.; Chen, W.L. A tri-vanadium-capped Keggin phosphomolybdate: Synthesis, characterization, photocatalytic and bifunctional electrocatalytic properties. Tungsten 2021, 4, 99–108. [Google Scholar] [CrossRef]
  54. Zhang, Y.M.; An, C.W.; Zhang, D.F.; Liu, T.; Yan, J.S.; Zhang, J. Photocatalytic activity of vanadium-substituted polyoxometalate doped magnetic carbon nitride towards antibiotics. Russ. J. Inorg. Chem. 2021, 66, 679–683. [Google Scholar] [CrossRef]
  55. Wang, J.; Wang, S. Reactive species in advanced oxidation processes: Formation, identification and reaction mechanism. Chem. Eng. J. 2020, 401, 126158. [Google Scholar] [CrossRef]
  56. Zhang, D.; Liu, T.; An, C.; Liu, H.; Wu, Q. Preparation of vanadium-substituted polyoxometalate doped carbon nitride hybrid materials POM/g-C3N4 and their photocatalytic oxidation performance. Mater. Lett. 2020, 262, 126954. [Google Scholar] [CrossRef]
  57. Pazhooh, P.; Khoshnavazi, R.; Bahrami, L.; Naseri, E. Synthesis and photocatalytic activity assessing of the TiO2 nanocomposites modified by some lanthanide ions and tin-derivative sandwich-type polyoxometalates. J. Iran. Chem. Soc. 2018, 15, 1775–1783. [Google Scholar] [CrossRef]
  58. Roy, S.; Vemuri, V.; Maiti, S.; Manoj, K.S.; Subbarao, U.; Peter, S.C. Two Keggin-based isostructural POMOF hybrids: Synthesis, crystal structure, and catalytic properties. Inorg. Chem. 2018, 57, 12078–12092. [Google Scholar] [CrossRef] [PubMed]
  59. Ruiz, V.; Blanco, C.; Raymundo-Piñero, E.; Khomenko, V.; Béguin, F.; Santamaría, R. Effects of thermal treatment of activated carbon on the electrochemical behaviour in supercapacitors. Electrochim. Acta 2007, 52, 4969–4973. [Google Scholar] [CrossRef]
  60. Wang, D.G.; Liang, Z.; Gao, S.; Qu, C.; Zou, R. Metal-organic framework-based materials for hybrid supercapacitor application. Coord. Chem. Rev. 2020, 404, 213093. [Google Scholar] [CrossRef]
  61. Mohamed, A.M.; Ramadan, M.; Ahmed, N.; ElNaga, A.O.A.; Alalawy, H.H.; Zaki, T.; Shaban, S.A.; Hassan, H.B.; Allam, N.K. Metal–Organic frameworks encapsulated with vanadium-substituted heteropoly acid for highly stable asymmetric supercapacitors. J. Energy Storage 2020, 28, 101292. [Google Scholar] [CrossRef]
  62. Shi, C.; Kang, N.; Wang, C.; Yu, K.; Lv, J.; Wang, C.; Zhou, B. An inorganic-organic hybrid nanomaterial with a core-shell structure constructed by using Mn-BTC and Ag5[BW12O40] for supercapacitors and photocatalytic dye degradation. Nanoscale Adv. 2022, 4, 4358–4365. [Google Scholar] [CrossRef]
  63. Yue, L.; Cao, Y.; Han, Y.; Li, Z.; Luo, X.; Liu, Y. Preparation of core-shell structured Fe3O4@Sn-MOF composite and photocatalytic performance. J. Alloys Compd. 2021, 870, 159339. [Google Scholar] [CrossRef]
  64. Shen, H.; Zhao, X.; Duan, L.; Liu, R.; Wu, H.; Hou, T.; Jiang, X.; Gao, H. Influence of interface combination of RGO-photosensitized SnO2 @RGO core-shell structures on their photocatalytic performance. Appl. Surf. Sci. 2017, 391, 627–634. [Google Scholar] [CrossRef]
  65. Zhang, W.; Zhang, Z.; Kwon, S.; Zhang, F.; Stephen, B.; Kim, K.K.; Jung, R.; Kwon, S.; Chung, K.; Yang, W. Photocatalytic improvement of Mn-adsorbed g-C3N4. Appl. Catal. B Environ. 2017, 206, 271–281. [Google Scholar] [CrossRef]
  66. Yu, L.; Ning, K.; Chunmei, W.; Kai, Y.; Jinghua, L.; Chunxiao, W.; Baibin, Z. A hybrid borotungstate-coated metal-organic framework with supercapacitance, photocatalytic dye degradation and H2O2 sensing properties. Dalton Trans. 2022, 51, 7613–7621. [Google Scholar] [CrossRef]
  67. Wang, L.; Kang, N.; Gong, L.; Wang, C.; Yu, K.; Wang, C.; Zhou, B. A novel core-shell structured hybrid composed of zinc homobenzotrizoate and silver borotungstate with supercapacitor and photocatalytic dye degradation performance. J. Energy Storage 2022, 46, 103873. [Google Scholar] [CrossRef]
  68. Song, Z.J.; Wang, L.Y.; Kang, N.; Yu, K.; Lv, J.; Zhou, B. 3D host-guest material of {Ag (pz)} modified {BW12O40} with supercapacitor, photocatalytic dye degradation and H2O2 sensing performances. J. Solid State Chem. 2023, 323, 124038. [Google Scholar] [CrossRef]
  69. Cameron, J.M.; Wales, D.J.; Newton, G.N. Shining a light on the photo-sensitisation of organic–inorganic hybrid polyoxometalates. Dalton. Trans. 2018, 47, 5120–5136. [Google Scholar] [CrossRef] [PubMed]
  70. Azcarate, I.; Ahmed, I.; Farha, R.; Goldmann, M.; Wang, X.; Xu, H.; Hasenknopf, B.; Lacôte, E.; Ruhlmann, L. Synthesis and characterization of conjugated Dawson-type polyoxometalate–porphyrin copolymers. Dalton Trans. 2013, 42, 12688–12698. [Google Scholar] [CrossRef]
  71. Paille, G.; Gomez-Mingot, M.; Roch-Marchal, C.; Lassalle-Kaiser, B.; Mialane, P.; Fontecave, M.; Mellot-Draznieks, C.; Dolbecq, A. A fully noble metal-free photosystem based on cobalt-polyoxometalates immobilized in a porphyrinic metal-organic framework for water oxidation. J. Am. Chem. Soc. 2018, 140, 3613–3618. [Google Scholar] [CrossRef]
  72. Iqbal, A.; Asif, H.M.; Zhou, Y.S.; Zhang, L.J.; Wang, T.; Shehzad, F.K.; Ren, X. From simplicity to complexity in grafting Dawson-type polyoxometalates on porphyrin, leading to the formation of new organic-inorganic hybrids for the investigation of third-order optical nonlinearities. Inorg. Chem. 2019, 58, 8763–8774. [Google Scholar] [CrossRef]
  73. Zhu, S.L.; Xu, X.; Ou, S.; Zhao, M.; He, W.L.; Wu, C.D. Assembly of a metalloporphyrin-polyoxometalate hybrid material for highly efficient activation of molecular oxygen. Inorg. Chem. 2016, 55, 7295–7300. [Google Scholar] [PubMed]
  74. Jamshidi, A.; Mohammadi Zonoz, F.; Wei, Y. A new Keggin-based organic-inorganic nanohybrid in the role of a dual-purpose catalyst. J. Chem. Sci. 2020, 132, 37. [Google Scholar] [CrossRef]
  75. Wang, D.; Zhao, Z.; Song, Y.; Su, Z.; Chen, J. The 3D POMOFs based two AsIII-capped Keggin arsenomolybdates with four VIV substituted: Synthesis, structures and properties. J. Solid State Chem. 2020, 291, 121639. [Google Scholar] [CrossRef]
  76. Jiang, F.; Liu, Q.Q.; Cui, Z.W.; Shi, S.; Long, J.; Wang, X.; Fei, B. A novel octamolybdate-based organic–inorganic hybrid as photo-Fenton-like catalyst for degradation of methylene blue. Appl. Organomet. Chem. 2023, 37, e6966. [Google Scholar] [CrossRef]
  77. Fei, B.L.; Deng, N.P.; Wang, J.H.; Liu, Q.B.; Long, J.Y.; Li, Y.G.; Mei, X. A heteropoly blue as environmental friendly material: An excellent heterogeneous Fenton-like catalyst and flocculent. J. Hazard. Mater. 2017, 340, 326–335. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, Y.Y.; An, J.D.; Wang, T.T.; Li, Y.; Ding, B. Hydrothermal assembly, structural diversity, and photocatalytic characterization of two polyoxometalates-based hybrid CuII and CuI coordination polymers with 2,6-(1,2,4-triazole -4-yl)pyridine. Inorg. Nano-Met. Chem. 2020, 51, 976–984. [Google Scholar] [CrossRef]
  79. Hao, Z.C.; Wang, S.C.; Yang, Y.J.; Cui, G.H. Syntheses, structural diversities and photocatalytic properties of three nickel (II) coordination polymers based semi-bis(benzimidazole) and aromatic dicarboxylic acid ligands. Polyhedron 2020, 181, 114466. [Google Scholar] [CrossRef]
  80. Liu, J.J.; Fu, J.J.; Liu, T.; Cheng, F. Photochromic polyoxometalate/naphthalenediimide hybrid structure with visble -light-driven dye degradation. J. Solid State Chem. 2022, 312, 123236. [Google Scholar] [CrossRef]
  81. Zhan, S.; Li, C.; Tian, H.; Ma, C.; Liu, H.; Luo, J.; Li, M. Synthesis, characterization and dye removal behavior of core–shell–shell Fe3O4/Ag/polyoxometalates ternary nanocomposites. Nanomaterials 2019, 9, 1255. [Google Scholar] [CrossRef]
  82. Wang, X.H.; Liu, H.L.; Zhang, W.X.; Cheng, W.Z.; Liu, X.; Li, X.M.; Wu, J.H. Synthesis and characterization of polymer-coated AgZnO nanoparticles with enhanced photocatalytic activity. RSC Adv. 2014, 4, 44011–44017. [Google Scholar] [CrossRef]
  83. Rani, S.; Tariq, M.; Bhatti, M.H.; Abdelmohsen, S.A.M.; Alanazi, M.M.; Khan, M.A.; Asif, H.M.; Nadeem, M.; Khan, R. A simplistic approach for the synthesis of covalent organic frameworks (COFs) comprising of tetrafunctionalized porphyrin and polyoxometalates to uncover catalytic applications. Opt. Mater. 2023, 138, 113672. [Google Scholar] [CrossRef]
  84. Liu, J.X.; Zhang, X.B.; Li, Y.L.; Huang, S.L.; Yang, G.Y. Polyoxometalate functionalized architectures. Coord. Chem. Rev. 2020, 414, 213260. [Google Scholar] [CrossRef]
  85. Tian, X.; Zhang, Y.; Ma, Y.; Zhao, Q.; Han, Z. Hourglass-type polyoxometalate-based crystalline materials as efficient cooperating photocatalysts for the reduction of Cr(VI) and oxidation of dyes. Catal. Sci. Technol. 2020, 10, 2593–2601. [Google Scholar] [CrossRef]
  86. Chang, W.J.; Jiang, Y.C.; Wang, S.L.; Lii, K.H. Hydrothermal synthesis of a three-dimensional organic-inorganic hybrid network formed by poly(oxomolybdophosphate) anions and nickel coordination cations. Inorg. Chem. 2006, 45, 6586–6588. [Google Scholar] [CrossRef]
  87. Yuan, Q.; Chen, L.; Xiong, M.; He, J.; Luo, S.L.; Au, C.T.; Yin, S.F. Cu2O/BiVO4 heterostructures: Synthesis and application in simultaneous photocatalytic oxidation of organic dyes and reduction of Cr(VI) under visible light. Chem. Eng. J. 2014, 255, 394–402. [Google Scholar] [CrossRef]
  88. Guo, H.L.; Xing, X.X.; Mao, S.X.; Feng, T.; Fan, Y.H.; Qin, Z.J.; Pang, J.Y.; Bai, Y.; Dang, D.B. Two three-dimensional Fe(II) frameworks based on P(4)Mo(6) tetrameric clusters exhibiting efficient visible-light photocatalytic properties for the degradation of Cr(VI) and methylene blue. Dalton Trans. 2022, 51, 18090–18098. [Google Scholar] [CrossRef] [PubMed]
  89. Nemiwal, M.; Zhang, T.C.; Kumar, D. Recent progress in g-C3N4, TiO2 and ZnO based photocatalysts for dye degradation: Strategies to improve photocatalytic activity. Sci. Total Environ. 2021, 767, 144896. [Google Scholar] [CrossRef]
  90. Koodali, R.T.; Zhao, D. Photocatalytic degradation of aqueous organic pollutants using titania supported periodic mesoporous silica. Energy Environ. Sci. 2010, 3, 608–614. [Google Scholar] [CrossRef]
  91. Xing, Z.P.; Zhang, J.Q.; Cui, J.Y.; Yin, J.W.; Zhao, T.Y.; Kuang, J.Y.; Xiu, Z.Y.; Wan, N.; Zhou, W. Recent advances in floating TiO2-based photocatalysts for environmental application. Appl. Catal. B Environ. 2018, 225, 452–467. [Google Scholar] [CrossRef]
  92. Jin, H.; Wu, Q.; Pang, W. Photocatalytic degradation of textile dye X-3B using polyoxometalate–TiO2 hybrid materials. J. Hazard. Mater. 2007, 141, 123–127. [Google Scholar] [CrossRef] [PubMed]
  93. Li, K.; Guo, Y.; Ma, F.; Li, H.; Chen, L.; Guo, Y. Design of ordered mesoporous H3PW12O40-titania materials and their photocatalytic activity to dye methyl orange degradation. Catal. Commun. 2010, 11, 839–843. [Google Scholar] [CrossRef]
  94. Khoshnavazi, R.; Fereydouni, S.; Bahrami, L. Enhanced photocatalytic activity of nanocomposites of TiO2 doped with Zr, Y or Ce polyoxometalates for degradation of methyl orange dye. Water Sci. Technol. 2016, 73, 1746–1755. [Google Scholar] [CrossRef] [PubMed]
  95. Mahmoodi, M.; Rafiee, E.; Eavani, S. Photocatalytic removal of toxic dyes, liquorice and tetracycline wastewaters by a mesoporous photocatalyst under irradiation of different lamps and sunlight. J. Environ. Manag. 2022, 313, 115023. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, Y.; Zhang, H.; Chen, L. Ultrasound enhanced catalytic ozonation of tetracycline in a rectangular air-lift reactor. Catal. Today 2011, 175, 283–292. [Google Scholar] [CrossRef]
  97. Mahmoodi, M.; Rafiee, E.; Eavani, S. Introducing of a novel polyoxometalate-based organic-inorganic hybrid: Insights into electochemical property-photoactivity relationship. J. Mater. Sci. Mater. Electron. 2021, 32, 1121–1138. [Google Scholar] [CrossRef]
  98. Li, S.; Zhang, J.; Cao, Y.; Yang, Y.; Xie, T.; Lin, Y. Visible light assisted heterogeneous photo-Fenton-like degradation of rhodamine B based on the Co-POM/N-TiO2 composites: Catalyst properties, photogenerated carrier transfer and degradation mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129248. [Google Scholar] [CrossRef]
  99. Fei, B.L.; Zhong, J.K.; Deng, N.P.; Wang, J.H.; Liu, Q.B.; Li, Y.G.; Mei, X. A novel 3D heteropoly blue type photo-Fenton-like catalyst and its ability to remove dye pollution. Chemosphere 2018, 197, 241–250. [Google Scholar] [CrossRef]
  100. Li, K.; Yang, X.; Guo, Y.; Ma, F.; Li, H.; Chen, L.; Guo, Y. Design of mesostructured H3PW12O40–titania materials with controllable structural orderings and pore geometries and their simulated sunlight photocatalytic activity towards diethyl phthalate degradation. Appl. Catal. B Environ. 2010, 99, 364–375. [Google Scholar] [CrossRef]
  101. Cai, S.; Shi, S.; Li, H.; Bai, Y.; Dang, D. Construction of self-sufficient Z-scheme Ag3PW12O40/TiO2 photocatalysts for the improved visible-light-driven photo-degradation of rhodamine B. Res. Chem. Intermed. 2018, 44, 7769–7788. [Google Scholar] [CrossRef]
  102. Chen, C.; Ma, W.; Zhao, J. Semiconductor-mediated photodegradation of pollutants under visible-light irradiation. Chem. Soc. Rev. 2010, 39, 4206–4219. [Google Scholar] [CrossRef]
  103. Ye, L.; Liu, J.; Gong, C.; Tian, L.; Peng, T.; Zan, L. Two different roles of metallic Ag on Ag/AgX/BiOX (X = Cl, Br) visible light photocatalysts: Surface plasmon resonance and Z-Scheme bridge. ACS Catal. 2012, 2, 1677–1683. [Google Scholar] [CrossRef]
  104. Yan, T.; Tian, J.; Guan, W.; Qiao, Z.; Li, W.; You, J.; Huang, B. Ultra-low loading of Ag3PO4 on hierarchical In2S3 microspheres to improve the photocatalytic performance: The cocatalytic effect of Ag and Ag3PO4. Appl. Catal. B Environ. 2017, 202, 84–94. [Google Scholar] [CrossRef]
  105. He, Y.; Zhang, L.; Teng, B.; Fan, M. New application of Z-scheme Ag3PO4/g-C3N4 composite in converting CO2 to fuel. Environ. Sci. Technol. 2015, 49, 649–656. [Google Scholar] [CrossRef] [PubMed]
  106. Guo, R.; Bai, L.; Dong, G.; Chai, D.; Lang, K.; Mou, Z.; Zhao, M. Construction of ZnO/Keggin polyoxometalate Nano-heterojunction catalyst for efficient removal of rhodamine B in aqueous solution. J. Inorg. Organomet. Polym. Mater. 2022, 32, 1599–1615. [Google Scholar] [CrossRef]
  107. Shi, H.F.; Yan, G.; Zhang, Y.; Tan, H.Q.; Zhou, W.Z.; Ma, Y.Y.; Li, Y.G.; Chen, W.; Wang, E. Ag/AgxH3–xPMo12O40 nanowires with enhanced visible-light-driven photocatalytic performance. ACS Appl. Mater. Interfaces 2016, 9, 422–430. [Google Scholar] [CrossRef]
  108. Yu, H.; Liu, R.; Wang, X.; Wang, P.; Yu, J. Enhanced visible-light photocatalytic activity of Bi2WO6 nanoparticles by Ag2O cocatalyst. Appl. Catal. B Environ. 2012, 111–112, 326–333. [Google Scholar] [CrossRef]
  109. Zhang, Z.; Gómez-García, C.J.; Wu, Q.; Xin, J.; Pang, H.; Ma, H.; Chai, D.; Li, S.; Zhao, C. Synthesis of a polyoxometalate-encapsulated metal–organic framework via in situ ligand transformation showing highly catalytic activity in both hydrogen evolution and dye degradation. Inorg. Chem. 2022, 61, 11830–11836. [Google Scholar] [CrossRef] [PubMed]
  110. Chen, D.M.; Zheng, Y.P.; Shi, D.Y.; Fang, S.M. An acid-base resistant polyoxometalate-based metal–organic framework constructed from {Cu4Cl}7+ and {Cu2 (CO2)4} clusters for photocatalytic degradation of organic dye. J. Solid State Chem. 2020, 287, 121384. [Google Scholar] [CrossRef]
  111. Rahmati, R.; Nayebi, B.; Ayati, B. Investigating the effect of hydrogen peroxide as an electron acceptor in increasing the capability of slurry photocatalytic process in dye removal. Water Sci. Technol. 2021, 83, 2414–2423. [Google Scholar] [CrossRef]
  112. Chen, Z.; Yao, D.; Chu, C.; Mao, S. Photocatalytic H2O2 production systems: Design strategies and environmental applications. Chem. Eng. J. 2023, 451, 138489. [Google Scholar] [CrossRef]
  113. Wang, L.; Li, B.; Dionysiou, D.D.; Chen, B.; Yang, J.; Li, J. Overlooked formation of H2O2 during the hydroxyl radical-scavenging process when using alcohols as scavengers. Environ. Sci. Technol. 2022, 56, 3386–3396. [Google Scholar] [CrossRef]
  114. Zhang, W.; Gong, L.; Du, N.; Wang, C.; Yu, K.; Wang, C.; Zhou, B. {BW12O40} hybrids modified by in situ synthesized rigid ligand with supercapacitance and photocatalytic properties. Inorg. Chem. 2021, 60, 16357–16369. [Google Scholar] [PubMed]
  115. Liang, Z.; Cheng, H.; Zhang, X.; Mao, Q. Two polyoxometalates based on {P2Mo5} catalysts: Synthesis, characterization, and photocatalytic degradation of RhB. J. Mol. Liq. 2023, 377, 121483. [Google Scholar] [CrossRef]
  116. Guo, Y.H.; Hu, C.W.; Jiang, S.C.; Guo, C.X.; Yang, Y.; Wang, E. Heterogeneous photodegradation of aqueous hydroxy butanedioic acid by microporous polyoxometalates. Appl. Catal. B Environ. 2002, 36, 9–17. [Google Scholar] [CrossRef]
  117. Shi, H.F.; Yu, Y.C.; Zhang, Y.; Feng, X.J.; Zhao, X.Y.; Tan, H.Q.; Khan, S.U.; Li, Y.G.; Wang, E. Polyoxometalate/TiO2/Ag composite nanofibers with enhanced photocatalytic performance under visible light. Appl. Catal. B Environ. 2018, 221, 280–289. [Google Scholar] [CrossRef]
  118. Zhao, X.; Zhang, Y.; Zhao, Y.; Tan, H.; Zhao, Z.; Shi, H.; Wang, E.; Li, Y.G. AgxH3−xPMo12O40/Ag nanorods/g-C3N4 1D/2D Z-scheme heterojunction for highly efficient visible-light photocatalysis. Dalton Trans. 2019, 48, 6484–6491. [Google Scholar] [CrossRef]
  119. Shi, H.; Zhao, T.; Zhang, Y.; Tan, H.; Shen, W.; Wang, W.; Li, Y.G.; Wang, E. Pt/POMs/TiO2 composite nanofibers with an enhanced visible-light photocatalytic performance for environmental remediation. Dalton Trans. 2019, 48, 13353–13359. [Google Scholar] [CrossRef]
  120. Yu, L.H.; Shao, Y.; Li, D.Z. Direct combination of hydrogen evolution from water and methane conversion in a photocatalytic system over Pt/TiO2. Appl. Catal. B Environ. 2017, 204, 216–223. [Google Scholar] [CrossRef]
  121. Gong, J.Y.; Imbault, A.; Farnood, R. The promoting role of bismuth for the enhanced photocatalytic oxidation of lignin on Pt-TiO2 under solar light illumination. Appl. Catal. B Environ. 2017, 204, 296–303. [Google Scholar] [CrossRef]
  122. Luo, B.F.; Xu, D.B.; Li, D.; Wu, G.L.; Wu, M.M.; Shi, W.D.; Chen, M. Fabrication of a Ag/Bi3TaO7 plasmonic photocatalyst with enhanced photocatalytic activity for degradation of tetracycline. ACS Appl. Mater. Interfaces 2015, 7, 17061–17069. [Google Scholar] [CrossRef]
  123. Shi, H.F.; Zhu, H.W.; Jin, T.; Chen, L.; Zhang, J.Y.; Qiao, K.Y.; Chen, Z. Construction of Bi/Polyoxometalate doped TiO2 composite with efficient visible-light photocatalytic performance: Mechanism insight, degradation pathway and toxicity evaluation. Appl. Surf. Sci. 2023, 615, 156310. [Google Scholar] [CrossRef]
  124. Zhao, X.; Wang, X.; Zhao, Y.; Sun, H.; Tan, H.; Qiu, T.; Zhao, Z.; Zhao, X.; Cheng, S.; Li, Y. Polyoxometalates-doped TiO2/Ag hybrid heterojunction: Removal of multiple pollutants and mechanism investigation. Environ. Sci. Nano 2021, 8, 3855–3864. [Google Scholar] [CrossRef]
  125. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative overview. RSC Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  126. Niu, P.; Wang, D.; Wang, A.; Liang, Y.; Wang, X. Fabrication of bifunctional TiO2/POM microspheres using a Layer-by-Layer method and photocatalytic activity for methyl orange degradation. J. Nanomater. 2018, 2018, 4212187. [Google Scholar] [CrossRef]
  127. Zhou, J.; Chen, W.; Sun, C.; Han, L.; Qin, C.; Chen, M.; Wang, X.; Wang, E.; Su, Z. Oxidative polyoxometalates modified graphitic carbon nitride for visible-light CO2 reduction. ACS Appl. Mater. Interfaces 2017, 9, 11689–11695. [Google Scholar] [CrossRef] [PubMed]
  128. Wei, B.; Zhang, L.; Guan, Y.L.; Fu, L.J.; Wang, H.; Chi, J.Y.; Jiang, Z.Q.; Wang, X.D.; Yu, H.H. Synthesis, characterisation and photocatalytic properties of visible light responsive Co2Co4(SiW10O37)2/Fe2O3 composites. Chem. Zvesti 2022, 76, 6997–7006. [Google Scholar] [CrossRef]
  129. Li, S.J.; Liu, S.X.; Li, C.C.; Ma, F.J.; Liang, D.D.; Zhang, W.; Tan, R.K.; Zhang, Y.Y.; Xu, L. Reactivity of polyoxoniobates in acidic solution: Controllable assembly and disassembly based on niobium-substituted germanotungstates. Chem. Eur. J. 2010, 16, 13435–13442. [Google Scholar] [CrossRef]
  130. Keita, B.; Liu, T.; Nadjo, L. Synthesis of remarkably stabilized metal nanostructures using polyoxometalates. J. Mater. Chem. 2009, 19, 19–33. [Google Scholar] [CrossRef]
  131. Chen, W.H.; Hu, Z.B.; Zhou, J.C.; Xiong, J.H.; Luo, J.S.; Zhang, H.Q.; Mi, J.X. Two new Sandwich-type phosphomolybdates: Thermal decomposition and photocatalytic degradation behavior of a UV-excited solid-phase Fenton catalyst. Eur. J. Inorg. Chem. 2019, 2019, 3015–3022. [Google Scholar] [CrossRef]
  132. Zou, J.P.; Chen, Y.; Zhu, M.; Wang, D.K.; Luo, X.B.; Luo, S.L. Semiconductor-based nanocomposites for photodegradation of organic pollutants. In Nanomaterials for the Removal of Pollutants and Resource Reutilization, 2nd ed.; Luo, X.B., Deng, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 25–28. ISBN 9780128148372. [Google Scholar]
  133. Feng, J.; Xiong, S.; Wang, Y. Atomic layer deposition of hybrid metal oxides on carbon nanotube membranes for photodegradation of dyes. Compos. Commun. 2019, 12, 39–46. [Google Scholar] [CrossRef]
  134. Chiu, Y.H.; Chang, T.F.M.; Chen, C.Y.; Sone, M.; Hsu, Y.J. Mechanistic insights into photodegradation of organic dyes using heterostructure photocatalysts. Catalysts 2019, 9, 430. [Google Scholar] [CrossRef]
  135. Nosaka, Y.; Nosaka, A. Understanding hydroxyl radical (OH) generation processes in photocatalysis. ACS Energy Lett. 2016, 1, 356–359. [Google Scholar] [CrossRef]
  136. Wang, J.L.; Xu, L.J. Advanced oxidation processes for wastewater treatment: Formation of hydroxyl radical and application. Crit. Rev. Environ. Sci. Technol. 2012, 42, 251–325. [Google Scholar] [CrossRef]
  137. Ong, B.C.; Lim, H.K.; Tay, C.Y.; Lim, T.T.; Dong, Z.L. Polyoxometalates for bifunctional applications: Catalytic dye degradation and anticancer activity. Chemosphere 2022, 286, 131869. [Google Scholar] [CrossRef] [PubMed]
  138. Valipour, A.; Rezvani, M.A.; Alina Asli, M.; Oveisi, M.; Mahmoodi, N.M. Bi-amino surface functionalized polyoxometalate nanocomposite as an environmentally friendly catalyst: Synthesis and dye degradation. Water Sci. Technol. 2017, 75, 2381–2389. [Google Scholar]
  139. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  140. Alharthi, F.A.; AlFawaz, A.; Ahmad, N. Photocatalytic degradation of anionic dye using well-crystalline SnWO4 polyoxometalate. Phys. Scr. 2022, 97, 085813. [Google Scholar] [CrossRef]
  141. Lei, P.; Chen, C.; Yang, J.; Wanhong, M.; Zhao, J.; Zang, L. Degradation of dye pollutants by immobilized polyoxometalate with H2O2 under visible-light irradiation. Environ. Sci. Technol. 2005, 39, 8466–8474. [Google Scholar] [CrossRef]
  142. Yu, Z.T.; Liao, Z.L.; Jiang, Y.S. Li, G.H.; Chen, J.S. Water-insoluble Ag–U–Organic assemblies with photocatalytic activity. Chem. Eur. J. 2005, 11, 2642–2650. [Google Scholar] [CrossRef]
  143. Yu, X.S.; Cui, H.J.; Wang, Q.Z.; Li, J.S.; Su, F.; Zhang, L.C.; Sang, X.J.; Zhu, Z.M. Construction and visible-light photocatalytic performance of carboxyethyltin/transition metal–functionalized wheel-like tungstophosphates. Appl. Organomet. Chem. 2020, 34, e5720. [Google Scholar] [CrossRef]
  144. Ai, L.; Zhang, D.; Wang, Q.; Yan, J.; Wu, Q. Photocatalytic degradation of textile dye X-3B using TiW11Ti/SiO2 hybrid materials. Catal. Commun. 2019, 126, 10–14. [Google Scholar] [CrossRef]
  145. Zhang, H.Y.; Liu, L.; Wang, H.J.; Sun, J.W. Asymmetrical modification of Keggin polyoxometalates by sextuple Ag–N coordination polymeric chains: Synthesis, structure and selective separation of cationic dyes. J. Solid State Chem. 2021, 296, 121986. [Google Scholar] [CrossRef]
Figure 1. The degradation mechanism of MB with Mn-BTC@Ag5[BW12] ((1) represented the action of light on the photocatalyst; (2) represented the reaction of water with h+ in the photocatalyst to produce OH; (3) represented the reaction of O2 with e in the photocatalyst to produce O2•−; (4) represents the generated OH acting on the substrate). Adapted with permission from Ref. [62].
Figure 1. The degradation mechanism of MB with Mn-BTC@Ag5[BW12] ((1) represented the action of light on the photocatalyst; (2) represented the reaction of water with h+ in the photocatalyst to produce OH; (3) represented the reaction of O2 with e in the photocatalyst to produce O2•−; (4) represents the generated OH acting on the substrate). Adapted with permission from Ref. [62].
Ijms 24 15244 g001
Figure 2. The possible mechanism of the simultaneous MB oxidation and Cr(VI) reduction with (Hbpp)2CoCd(P4Mo6)2 upon visible light irradiation. Adapted with permission from Ref. [85].
Figure 2. The possible mechanism of the simultaneous MB oxidation and Cr(VI) reduction with (Hbpp)2CoCd(P4Mo6)2 upon visible light irradiation. Adapted with permission from Ref. [85].
Ijms 24 15244 g002
Figure 3. The mechanism diagram of RhB degradation by Co-PMo12/N-TiO2 (POM * meant the reduced state POM). Adapted with permission from Ref. [98].
Figure 3. The mechanism diagram of RhB degradation by Co-PMo12/N-TiO2 (POM * meant the reduced state POM). Adapted with permission from Ref. [98].
Ijms 24 15244 g003
Figure 4. The photocatalytic degradation mechanism of RhB with Z-system photocatalyst Ag3PW12/TiO2. Adapted with permission from Ref. [101].
Figure 4. The photocatalytic degradation mechanism of RhB with Z-system photocatalyst Ag3PW12/TiO2. Adapted with permission from Ref. [101].
Ijms 24 15244 g004
Figure 5. Possible photocatalytic mechanism of (a) ZnO/Ag4SiW12 and (b) ZnO/Cs3PW12 under simulated sunlight irradiation. Adapted with permission from Ref. [106].
Figure 5. Possible photocatalytic mechanism of (a) ZnO/Ag4SiW12 and (b) ZnO/Cs3PW12 under simulated sunlight irradiation. Adapted with permission from Ref. [106].
Ijms 24 15244 g005
Figure 6. The degradation path of RhB (* meant excited state). Adapted with permission from Ref. [115].
Figure 6. The degradation path of RhB (* meant excited state). Adapted with permission from Ref. [115].
Ijms 24 15244 g006
Figure 7. The possible photocatalytic degradation mechanism of MO with (a) Pt/PMo12/TiO2 [119] and (b) Ag/PMo10V2/TiO2 [124]. Adapted with permissions from Refs. [119,124].
Figure 7. The possible photocatalytic degradation mechanism of MO with (a) Pt/PMo12/TiO2 [119] and (b) Ag/PMo10V2/TiO2 [124]. Adapted with permissions from Refs. [119,124].
Ijms 24 15244 g007
Figure 8. The possible photocatalytic degradation mechanism MO of with the heterogeneous Fenton-like catalyst of (DETA)3.5Fe(P4Mo6)2 (a meant (DETA)3.5Fe(P4Mo6)2; b meant excited state (DETA)3.5Fe(P4Mo6)2). Adapted with permission from Ref. [131].
Figure 8. The possible photocatalytic degradation mechanism MO of with the heterogeneous Fenton-like catalyst of (DETA)3.5Fe(P4Mo6)2 (a meant (DETA)3.5Fe(P4Mo6)2; b meant excited state (DETA)3.5Fe(P4Mo6)2). Adapted with permission from Ref. [131].
Ijms 24 15244 g008
Figure 9. The photocatalytic mechanism of organic substrate with POMs. Adapted with permission from Ref. [39].
Figure 9. The photocatalytic mechanism of organic substrate with POMs. Adapted with permission from Ref. [39].
Ijms 24 15244 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, R.; Wang, Y.; Zeng, F.; Si, C.; Zhang, D.; Xu, W.; Shi, J. Advances in Polyoxometalates as Electron Mediators for Photocatalytic Dye Degradation. Int. J. Mol. Sci. 2023, 24, 15244. https://doi.org/10.3390/ijms242015244

AMA Style

Li R, Wang Y, Zeng F, Si C, Zhang D, Xu W, Shi J. Advances in Polyoxometalates as Electron Mediators for Photocatalytic Dye Degradation. International Journal of Molecular Sciences. 2023; 24(20):15244. https://doi.org/10.3390/ijms242015244

Chicago/Turabian Style

Li, Ruyue, Yaqi Wang, Fei Zeng, Cuiqing Si, Dan Zhang, Wenbiao Xu, and Junyou Shi. 2023. "Advances in Polyoxometalates as Electron Mediators for Photocatalytic Dye Degradation" International Journal of Molecular Sciences 24, no. 20: 15244. https://doi.org/10.3390/ijms242015244

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

Article Metrics

Back to TopTop