Open Access Article
Christian
Cariño
a,
Anna
Proust
a,
Geoffroy
Guillemot
a,
Ludivine
K/Bidi
a,
Sébastien
Blanchard
a,
Elizabeth
A. Gibson
b and
Guillaume
Izzet
*a
aSorbonne Université, CNRS, Institut Parisien de Chimie Moléculaire, IPCM, F-75005 Paris, France. E-mail: guillaume.izzet@sorbonne-universite.fr
bEnergy Materials Laboratory, Chemistry, School of Natural and Environmental Sciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
First published on 2nd October 2025
Artificial photosynthesis faces the challenge of developing visible-light-driven strategies for converting and storing solar energy in the form of fuels and high-value chemicals. In such an approach, selective fuel production often depends on the accumulation of multiple electrons at a catalytic site. However, this process is constrained by the rapid recombination of photogenerated charges and the inherently slow kinetics of multi-electron catalytic reactions, which hinder efficient charge buildup and utilization. Polyoxometalates (POMs), a tunable class of nanoscale metal oxides, have emerged as promising multi-electron acceptors due to their redox versatility and stability. Their electron storage capabilities make them attractive as both reservoirs and catalysts. In most cases, their UV-limited absorption necessitates pairing of the POM with visible-light-absorbing antennas. Advances in photosensitized POM derivatives—via electrostatic assembly, covalent bonding, or band-gap engineering—are herein detailed. Covalent hybrids, in particular, allow precise control over electron transfer. Still, a detailed understanding of photoinduced electron transfer kinetics remains limited. This perspective article explores the potential applications of POMs in solar fuel generation, emphasizing the need for kinetic insight to design efficient, visible-light-driven photocatalysts and photoelectrochemical devices.
In this context, polyoxometalates (POMs), a class of discrete nanosized metal oxides, are attracting attention owing to their fascinating redox properties and synthetic versatility.12,13 As they are composed of metal ions in their highest oxidation state, they can act as multi-electron acceptors to collect, store then deliver electrons14 to an external catalyst. Furthermore, reduced forms of POMs also may have appealing reactivity15–18 notably for the hydrogen evolution reaction (HER) and the reductive activation of molecular oxygen (ORR).19 Consequently, POMs can be integrated in artificial photosynthetic devices having dual role of electron storage site and catalyst. Finally, POMs also have excellent intrinsic photocatalytic properties,20 that is yet often limited to the UV part of the solar spectrum.21 As a consequence their implementation into artificial photosynthetic devices often requires their prior association with a visible-range antenna. This has been achieved either by electrostatic interaction with a cationic photosensitizer22 or by direct functionalization of the POM framework.23,24
Organic–inorganic POM-based hybrids, hereafter named hybrid POMs,25–31 represent versatile models for artificial photosynthesis, although not without challenges. So far there are only few existing reports of charge photoaccumulation in hybrid POMs. Furthermore, despite numerous reports on the photosensitization of POMs, limited quantitative information on the kinetics of photo-induced electron transfers (separation, accumulation, recombination) are available, while these are of high importance in order to correlate the timescale of the photophysical processes with that of catalysis. Indeed, a challenging key-step in the use of POMs for photocatalysis is ensuring the efficient closure of the photocatalytic cycle, specifically the rate-limiting re-oxidation of the reduced POM. It is thus important to rationalize the kinetics of photoinduced electron transfer with POMs, and take into account the driving force dependence which is rarely estimated or even considered.
This perspective article will hence describe the opportunities and challenges in the use of POM derivatives as multi-electron acceptors for solar energy conversion, especially for reductive fuel formation with a particular focus on the critical role of the kinetics of electron transfer in facilitating these processes. While several transition-metal substituted POMs have been reported to display remarkable activity in the oxygen evolution reaction (OER) by water oxidation,16 this aspect will not be covered. The photocatalytic properties of POMs in the UV region are also not considered, since we are looking towards applications in solar-driven chemistry.
Among the different chromophores, the most commonly used were Ru(II)-polypyridyl complexes, following pioneering works of Balzani and coll. with Keggin-type polyoxotungstates ([SiW11O39Co(H2O)]6− and [PW11O39Mn(H2O)]6−).32 The association of these chromophores with Wells–Dawson-type POMs (e.g. [S2Mo18O62]4− and [S2W18O62]4−) was then further thoroughly studied by the groups of Keyes and Bond.33–35 Other polypyridyl complexes have been used, notably a cyclometalated Ir(III) complex applied to photoreduce multinuclear nickel substituted POMs in the presence of sacrificial electron donors (SEDs) to produce hydrogen.36,37 More recently Ishitani and coll. used a tetranuclear ring-shaped Re(I) complex that formed very strong 1
:
1 adducts with Keggin-type [XW12O40]4− (X = Si and Ge) owing to the full charge compensation of the POM.29
Cationic porphyrins have also been applied for the photosensitization of POMs particularly by Fukuzumi and coll. and Ruhlmann and coll. who respectively used Keggin-38 and Wells–Dawson-type POMs.39 In the case of a monocationic porphyrin 1
:
1 POM–porphyrins complexes were observed with rather important association constant (7.5 × 102 M−1).38 When using tetracationic porphyrin and highly charged tetracobalt-substituted POMs, the formation of a 1
:
4 POM
:
porphyrin complex was reported with very high association constant (estimated to be 7 × 1030 M−4). Although slightly counter intuitive, the photosensitization of Keggin- and Wells–Dawson-type POMs was also achieved by an anionic electron rich [Mo6I8Cl6]2− cluster despite both species being negatively charged.40,41
The electrostatic photosensitization of POMs can also be achieved by incorporating POMs into photoactive MOFs, and form photocatalytic materials with high activity in reduction (HER, CO2 reduction)42–44 or oxidation (OER). Quantum dots-sensitized photocathodes were also associated with a tetra-nickel substituted POM, as HER catalyst.45 Finally positively charged carbon dots were associated with POMs and the resulting materials showed some activity in the photoreduction of silver ions.46
![]() | ||
| Fig. 1 Selected examples of hybrid POM–PS reported in the literature with different POM-based platforms: (a) Keggin Co(II)-porphyrin: [SiW11O39Co]6−-(C46H35N5),48 (b) Dawson V3 triol-porphyrin: [P2W15V3O62{C49H34N5OZn}]6−,49 (c) Keggin germane-Ru(II): [PW11O39{GeC44H29N6Ru}]2−,50 (d) Dawson tin-Ir(III): [P2W17O61{SnC36H23O2N3Ir}]7−,26,51 (e) Anderson triol-Ir(III): [MnMo6O24{C38H33IrN5}2]−,52 (f) Keggin silyl-bodipy: [PW11O39{O(SiC31H30N2BF2)2}]3−,53,54 (g) imido hexamolybdate: [Mo6O18N{diblock copolymer}]2−,49 (h) Dawson phosphonate–Ru(II): [P2W17O61{(PO)C48H59N6Ru}2]2−,28 (i) capsule Keggin-diarylethene: [(PMo9O31)2{C25H17F6N2S2}3]6−.55 Color code: blue octahedra, {WO6} or {MoO6}; pink octahedra {MnO6}, orange octahedra, {VO6}; green tetrahedra {PO4}, and yellow tetrahedra {SiO4}. | ||
Early developments in the photosensitization of POMs include the coordination between a pyridyl-functionalized hybrid POM and Ru(II) porphyrin photosensitizer,56 or between a TMS-POM and a photosensitizer containing a pyridyl ligand.48 This straightforward strategy has nevertheless rather similar drawbacks as the electrostatic approach, since dissociation between the POM and the PS may occur in diluted solutions, which makes the photophysical analysis complex. Following the coordination bond approach, Yamaguchi, Suzuki and coll. reported the synthesis of a new type of stable capsule-like hybrids that could be achieved by the straightforward reaction between a trivacant lacunary phosphomolybdate [A-α-PMo9O34]9− with organic moieties displaying pyridyl ligands.57 Stable dimers and tetramers could be selectively synthesized by reaction with ditopic and tetratopic ligands (i.e. porphyrin). Such strategy was recently followed by the group of Ritchie that inserted in the capsule diarylethene pyridyl units.55
The group of Bonchio was among the first ones to covalently graft a visible range antenna (i.e. a fullerene derivative) to a lacunary decatungstosilicate through an organosilyl linker for application in organic photocatalysis.58 The group of Peng then reported the covalent grafting of photoconductive poly(phenylene ethynylene) onto imido functionalized hexamolybdates (i.e., [Mo5O18M
NR]2− and [Mo4O17(M
NR)2]2−),59,60 both using direct and post-functionalization synthetic approaches. They further used the same hybrid platforms to graft photoactive diblock copolymers.61,62 They processed these hybrid polymers as thin film for photovoltaic application. Odobel, Mayer, and coll. reported then a series of hybrid POMs based on monovacant Wells–Dawson [α2-P2W17O61]10− using organosilane and organophosphonate functionalizations. They successfully developed a post-functionalization strategy based on the Huisgen reaction with a hybrid platform displaying either azido or terminal alkyne function. In most cases, the flexible aliphatic spacer connecting the POM to the photoactive antennae (perylene mono-imide63 and Zn(II) porphyrin25,64) significantly complicated the photophysical investigation owing to the presence of various conformers.
Our group further developed a series of organosilane and organotin derivatives of monovacant Keggin- and Wells–Dawson-type POMs. Through a post-functionalization approach, the POM-based platform containing one or two iodoaryl moieties were grafted to various photoactive antennae (Ru(II)65 and Ir(III) polypyridyl/cyclometalated31,51,66 complexes, pyrene,65,67 bodipy,53,54,68 porphyrins,69,70 bis-thiophenethienothiophene71) using Sonogashira reactions. In this case a same PS could be grafted onto different POM-based platforms displaying distinct electron accepting properties in order to evaluate the effect of the thermodynamic parameters on the kinetics of photoinduced electron transfer (see part 2). Furthermore, rigid systems with controlled geometry could be obtained. Hasenknopf, Ruhlmann and coll. developed POM–PS hybrids using the functionalization of Anderson–Evans ([MnMo6O24]9−), Lindqvist hexavanadate ([V6O19]8−) and polyoxovanadotungstate Wells–Dawson ([P2W15V3O62]9−) with triol-based ligands. Porphyrin,49,72–74 Ru(II) complexes,75 and Re(I)76 complexes could then be covalently grafted onto these POM-based platforms. The Anderson–Evans platforms ([MnMo6O24]9−, [FeMo6O24]9−, [CoMo6O24]9−) were also functionalized by Streb, Rau and coll. with Ir(III) cyclometalated photosensitizers52,77 and the photophysical and photocatalytic properties of the resulting POM–Ir hybrids were thoroughly investigated by the group of Dietzek.78–80 More recently the same consortium grafted Ru(II) tris-pyridyl complexes onto the monovacant [α2-P2W17O61]10− through an organophosphonate functionalization.28 Schubert and coll. also grafted Ru(II) bis-terpyridine onto a organogermane derivative of Keggin-type POM81 using a post-functionalization synthetic route with a Sonogashira coupling reaction.50,82,83
![]() | ||
| Fig. 2 Selected examples of low band gap POM derivatives. (a) [P4W35O124{Re(CO)3}2]16−,86 (b) [CoIIW12O40]6−,87 (c) [P2W15V3O62{C16H18N3Cu(C2H3O2)}]5−,30 (d) [P2W17O61{(PO)C6H5}2]6−,88 color code: blue octahedra, {WO6}; pink tetrahedron {CoO4}, orange octahedra, {VO6}; and green tetrahedra {PO4}. | ||
The first case, corresponding to LUMO engineering, is exemplified by polyoxovanadates and mixed V/W POMs displaying LMCT from oxo-centered orbitals to the V-based LUMOs tailing off in the visible range. This is the case of the bismuth capped dodecavanadates [X{Bi(dmso)3}2V12O33]− (X = Cl or Br) characterized by a visible absorption up to ∼570 nm with enhanced extinction coefficients compared to the decavanadates ascribed to the presence of bismuth.89
The second category illustrates HOMO engineering and comprises mostly TMS-POMs displaying long-lived metal-to-POM charge transfer (MPCT) arising from a transition involving the inserted TM-based HOMO to purely POM-based LUMOs (Fig. 2). Pioneering examples have been given by Hill and coll. with one to three Re(I) carbonyl fragments {Re(CO)3}+ grafted to a monovacant [P2W17O61]10−.86,90 These compounds are dark red and the MPCT nature of the absorption has been confirmed by computational modelling. Cobalt-containing Keggin POMs ([SiW11O39CoII(H2O)]6−, [CoIIW12O40]6−) were also investigated as these compounds displayed appealing photophysical properties.87 Similarly, a tetranuclear Ce(III)-containing silicotungstate [{Ce(H2O)}2{Ce(CH3CN)}2(μ4-O)(γ-SiW10O36)2] was reported to display visible light photoredox catalytic activity.91 Many other visible-light responsive TMS-POMs have been described and used in photo-assisted catalysis but no data about the dynamics of their excited state are available.84
This strategy has also been applied with hybrid POMs. Our group has recently shown that the mixed V/W POM [P2V3W15O62]9−, once functionalized with a triol-based dipyridylamine-Cu complex, could be photoexcited with visible light and was able to store successively up to 3 electrons.30 In a series of papers, Newton, Oshio and coll. have shown that engineering the HOMO–LUMO gap of hybrid polyoxotungstates via the organic group allowed for visible-light sensitization of Wells–Dawson POMs.27,88,92 Using arylphosphonate or arylarsonate functionalization, they have shown a mixing of the organic and POM orbitals both in the HOMO and the LUMO. The energy of each can be tuned by playing on the electronic properties of the substituent on the aryl ring with an increase photosensitivity to visible light with organoarsonate derivatives.21,93
As regards to POM chemistry, an electron decoupling between the POM and the organic moieties is most often reported. This is the case for most Keggin and Wells–Dawson hybrid POMs, including organosilane, organotin and organogermane derivatives.97 In these systems, density functional theory (DFT) calculations showed that the frontier molecular orbitals are either located on the organic moieties (HOMO) or the POM (LUMO, which mainly involves dxy orbitals), the 14-group atom acting as a node in the molecular orbitals (Fig. 3).51,54,98,99 To some extent, this is also the case for Anderson-type POMs,52 where the LUMO is localized on the central heterometal. In these cases, the POM–PS hybrids showed absorption/emission dominated by the PS in the visible range, indicating negligible electronic interaction with the POM unit in the ground state. The modification of the organic moieties does not influence the redox properties of the POM (unless in case of charged organic PS, in which the redox potentials are affected by electrostatic interactions between the charged-subunits).65 Similarly, for polyoxovanadotungstate Dawson [P2W15V3O62]9− with triol-based ligands, the HOMO was found to be localized on the organic moieties and the LUMO was localized on the V3 cap.30,76 Note that in the case of hexavanadates, some electronic interaction between the POM core and the triol-based ligand has been observed.100,101
![]() | ||
| Fig. 3 (Left) Frontier molecular orbitals of a redox coupled organophosphonate hybrid. Reproduced from ref. 92 with permission from American Chemical Society, copyright 2017 . (Right) Frontier molecular orbitals of a redox decoupled organosilane Dawson-bodipy derivative. Reproduced from ref. 54 with permission from John Wiley and Sons, copyright 2021. | ||
Other examples in which organic functionalization modified the redox properties of the POM, to some extent, are also reported in the literature. The first case is the imido functionalized hexamolybates, as developed by the groups of Peng,59,60 Wei102 and Fielden.103 In this case a DFT study showed that the Mo
N functionalization participate both to the HOMO (mostly localized on the PS) and the LUMO.104 Another example is the insertion of an amide bond on a polyoxovanadotungstate Dawson [P2W15V3O62]9−.105 Here, it was found that the modification of the organic moieties could tune the first reduction wave of the POM by up to 50 mV. Finally, the redox properties of organophosphonate92,106 and organoarsonate21 derivatives of POMs were found to differ even more according to the nature of the organic moieties grafted to the POM. The involvement of the P and As atoms and the mixing between the organic and inorganic components in the frontier molecular orbitals was evidenced by time dependant-DFT calculations21,92 (Fig. 3). Consequently, in these hybrid POMs, the electronic interactions between the organic part and the POM unit are responsible of the visible absorption.
| Reference/label in Fig. 6 | POM | PS | solvent | E° (PS+/PS)/V vs. SCE | E° (POM/POM + 1 e)/V vs. SCE | −ΔG0CS/eV | τ CS/ns | −ΔG0CR/eV | τ CR/ns | |
|---|---|---|---|---|---|---|---|---|---|---|
| a Redox potential extrapolated to SCE reference. | ||||||||||
| Non covalent | 32 | [SiW11O39Co(OH2)]6− | [(Rubpy)3]2+ | H2O/LiCl | −0.88 | 0.20 | 12 ns | Not reported | ||
| 32 | [PW11O39Mn(OH2)]6− | [(Rubpy)3]2+ | H2O/LiCl | −1.02 | 0.34 | <2 ns | Not reported | |||
| 32 | [SiW11O39Co(OH2)]6− | [(Rubpy)2(4,4′-Cl2bpy)]2+ | H2O/LiCl | −0.88 | 0.07 | 250 | Not reported | |||
| 32 | [PW11O39Mn(OH2)]6− | [(Rubpy)3]2+ | H2O/LiCl | −1.02 | 0.21 | <5 ns | Not reported | |||
| 38 | [PW12O40]3− | Tin(IV)-porphyrin | PhCN | 0.69 | −0.48 | 0.68 | 0.2 | Very fast | ||
| 38 | [PW12O40]3− | Tin(IV)-porphyrin | PhCN | 0.86 | −0.50 | 0.48 | CS much slower than the ISC (∼0.1 ns) | |||
| 40 | [PW12O40]3− 5 equiv. | [Mo6I8Cl6]2− | MeCN | ∼1.00a | ∼−0.17a | 0.6 | 23 × 103 | 1.17 | 350 × 103 | |
| 40 | [PW18O62]6− 5 equiv. | [Mo6I8Cl6]2− | MeCN | ∼1.00a | ∼−0.57a | 0.2 | 42 × 103 | 1.57 | 620 × 103 | |
| 40 | [BW12O40]5− 5 equiv. | [Mo6I8Cl6]2− | MeCN | ∼1.00a | ∼ −1.07a | −0.3 | No CS formed | |||
| 107 | [Mo6O19]2− | Ferrocenyl cation | Sol. state (alum.) | Charge transfer band at 550 nm | ≪25 × 10−3 | ∼25 × 10−3 | ||||
| 108 | [Mo6O19]2− | Pyrene cation | Cryst. mat. | Charge transfer band at 500–600 nm | ≪25 × 10−3 | 56 × 10−3 | ||||
| 108 | [Mo6O19]2− | Pyrene cation | Cryst. mat. | Charge transfer band at 500–600 nm | ≪25 × 10−3 | 0.20 | ||||
| Coord. Bond | 48 | [CoSiW11O39]6− | Porphyrin | C2H4Cl2 | Not given | 60 × 10−3 | Not given | |||
| TMS-POM | 86 | [P4W35O124{Re(CO)3}2]16− | CH2Cl2 | Charge transfer band up to 600 nm | 35 × 10−6 | 1.4 × 10−3 | ||||
| 87 | [SiW11O39CoII(H2O)]6− | H2O | 1.3 × 10−3 | |||||||
| 87 | [CoIIW12O40]6− | H2O | 0.42 | |||||||
| 87 | [CoIIW12O40]6− | MeCN | 1.7 | |||||||
| Covalent | 63 | [P2W17O61{O(SiPS)2}]6− | Perylene monoimide | DMF | 1.00a | −0.84a | 0.34 | 40% fluor. quench. diff. CS rates (ps–ns) | 1.84 | Faster than CS |
| 63 | [P2W17O61{POPS}2]6− | Perylene monoimide | DMF | 1.00a | Not given, estim. to −0.46 V | 0.87 | 97% fluor. quench. ∼70 × 10−3 | ∼1.33 | 0.16 | |
| 25 | [P2W17O61{O(SiPS)2}]6− | Zn(II)-porphyrin | DMF | 0.70 | −0.68 | S1 0.68 | τ 1 ∼ 0.50 | 1.38 | Faster than CS | |
| T1 0.18 | τ 2 ∼ 3.8 | |||||||||
| 25 | [P2W17O61{POPS}2]6− | Zn(II)-porphyrin | DMF | 0.72 | −0.18 | S1 1.16 | τ 1 = 0.1 | 0.9 | Faster than CS | |
| T1 0.66 | τ 2 = 22 | |||||||||
| 25/P | [P2W17O61{POPS}2]6− | Zn(II)-porphyrin | DMF | 0.71 | −0.30 | S1 1.05 | 22 | 1.10 | Faster than CS | |
| T1 0.55 | ||||||||||
| 51/M | [PW11O39{O(SiPS)2}]3− | Cyclometalated Ir(III) | DMF | 1.05 | −0.38 | 1.00 | 3.9 × 10−3 | 1.43 | 1.5 | |
| 51/F | [P2W17O61{O(SiPS)2}]6− | Cyclometalated Ir(III) | DMF | 1.03 | −0.73 | 0.67 | 0.5 | 1.76 | 27 | |
| Covalent | 51/C | [PW11O39{(SnPS)}]4− | Cyclometalated Ir(III) | DMF | 1.05 | −1.09 | 0.29 | 46 | 2.14 | 180 |
| 51/B | [P2W17O61{(SnPS}]7− | Cyclometalated Ir(III) | DMF | 1.01 | −1.28 | 0.14 | 82 | 2.29 | 480 | |
| 31/O | [PMo11O39{SnPS)}]4− | Cyclometalated Ir(III) | DMF | 1.29 (irr) | −0.51 | 0.30 | 67 | |||
| 53/D | [PMo11O39{(SnPS)}]4− | Bodipy | CH2Cl2 | 1.02 | −0.57 | 0.56 | 0.18 | 1.55 | 0.52 (evol. into PS triplet state) | |
| 53 | [PW11O39{(SnPS)}]4− | Bodipy | CH2Cl2 | 0.98 | −1.11 | 0.03 | No CS formed | |||
| 68/J | [PMo11O39{(SnPS)}]4− | Bodipy | MeCN | ∼1.02 | −0.99 | 0.86 | 2.2 | 1.48 | 16 | |
| 68 | [PW11O39{(SnPS)}]4− | Bodipy | MeCN | ∼1.02 | −0.48 | 0.35 | No CS formed | |||
| 54/N | [PW11O39{O(SiPS)2}]3− | Bodipy | MeCN | 1.01 | −0.33 | 1.02 | 38 × 10−3 | 1.32 | 1.3 | |
| 53 and 54/L | [PW11O39{O(SiPS)2}]3− | Bodipy | CH2Cl2 | 1.02 | −0.45 | 0.94 | 54 × 10−3 | 1.39 | 4.8 | |
| 54/K | [PW11O39{O(SiPS)2}]3− | Bodipy | DMF | 1.03 | −0.36 | 0.95 | 2.0 | 1.39 | 5.7 | |
| 54/I | [P2W17O61{O(SiPS)2}]6− | Bodipy | MeCN | 0.99 | −0.57 | 0.80 | 0.38 | 1.54 | 3.1 | |
| 54/G | [P2W17O61{O(SiPS)2}]6− | Bodipy | CH2Cl2 | 1.00 | −0.71 | 0.70 | 0.34 | 1.63 | 13 | |
| 54/E | [P2W17O61{O(SiPS)2}]6− | Bodipy | DMF | 1.03 | −0.69 | 0.62 | 2.4 | 1.70 | 10 | |
| 50 and 82/A | [PW11O39{(GePS)}]4− | Ru(II) bis terpyridine | DMSO | 1.22 | −0.99 | 0.11 | 32 | 2.21 | 470 | |
| 28/H | [P2W17O61{POPS}2]2− | Ru(II) tris bipyridine | DMF | 1.10a | −0.21a | ∼0.7 | 12 × 10−3 | 1.31 | 0.09 | |
| 49/Q | [P2V3W15O59{(OCH2)3PS}]− | Zn(II)-porphyrin | DMF | 0.81 | −0.03a | S2 1.97 | 2.3 × 10−3 | 0.84 | >1 ns | |
| S1 1.13 | 62 × 10−3 | |||||||||
| 49 | [MnMo6O18{(OCH2)3PS}]− | Zn(II)-porphyrin | DMF | 0.86 | −1.17a | S2 0.93 | No CS formed | |||
| S1 0.09 | ||||||||||
| 52 and 79 | [MnMo6O18{(OCH2)3PS}]− | Cyclometalated Ir(III) | DMF | 1.37a | −0.83a | Hot state ≫0.13 | 0.5 × 10−3 | 2.20 | 0.29 | |
| 3MLCT 0.13 | — | |||||||||
| 52 and 79 | [FeMo6O18{(OCH2)3PS}]− | Cyclometalated Ir(III) | DMF | 1.37a | −0.95a | Hot state ≫0.01 | 0.5 × 10−3 | 2.32 | 0.43 | |
| 3MLCT 0.01 | — | |||||||||
| 52 and 79 | [CoMo6O18{(OCH2)3PS}]− | Cyclometalated Ir(III) | DMF | 1.37a | −0.84a | Hot state ≫0.12 | 0.5 × 10−3 | 2.21 | 0.54 | |
| 3MLCT 0.12 | — | |||||||||
| 80 and 109 | [MnMo6O18{(OCH2)3PS}]− | Cyclometalated Ir(III) | DMF | 1.33a | −0.76a | 0.24 | 0.18 | 2.09 | <0.18 | |
In Sections 3 and 4 we will mainly focus on POM–Ir(III)51 and POM–bodipy54 hybrids as case studies. Indeed, for these hybrid POMs, the kinetics of both monoelectronic charge separation/recombination but also their abilities to perform charge photoaccumulation26,110 are reported and quantified. These systems will provide insightful information regarding the advantages and limitations of the development of POM–PS for artificial photosynthesis.
In few cases, where the POM-based system was an electrostatic adduct33,107,108 or a TMS-POM,86 the authors observed new optical features, putatively assigned to a charge transfer (CT) transition, as the absorption spectrum of the whole system differed from the sum of the individual subunits. In one case, the authors indicated that such transition led to a luminescent CT state.33 The unambiguous characterization of charge-separated states by TAS of multimolecular POM–PS assemblies was only seldomly reported.40,107,108 In the case of a ferrocenyl/[Mo6O19]2− adduct, TAS of the material in the solid-state indicated that the resulting charge-separated (CS) state displayed very short lifetime (in the ps timescale) because of the close proximity between the POM and the electron donating unit. A similar feature was observed for a series of crystalline materials associating cationic polyaromatic compounds (anthracene, pyrene) to Lindqvist-type ([W6O19]2−, [Mo6O19]2−) and Keggin-type POMs ([SiW12O40]4−, [SiMo12O40]4−).108 In this instructive study, the authors correlated the distance between the POM and the polyaromatic moieties to the coloration of the compound in the solid state and to the kinetic of charge recombination. It was inferred that the shorter the POM–polyaromatic compound distance, the darker the colour of the crystal and the faster the decay of the CT state. Using a similar POM ([Mo6O19]2−)/cationic pyrene combination, two polymorphic crystalline materials were isolated, each one featuring different POM–pyrene distances, coloration (orange/yellow, red) and kinetics of CR (50/200 ps).
In the case of the association of Keggin- and Wells–Dawson-type POMs to anionic electron rich [Mo6I8Cl6]2− cluster, charge-separated states with very long lifetime were observed (hundreds of μs, Table 1).40 The long-lived character of the CS state is probably due to the low affinity between both polyanionic species in solution. The authors demonstrated that for this specific case of POM–PS, the cluster quenching could be correlated with the charge density of the POM and not its redox property.
In TMS-POMs, the systems often displayed very short lifetime again probably owing to the close proximity between the POM and the transition metal.86,107 For instance, excitation at 400 nm of [P4W35O124{Re(CO)3}2]16− or [P2W17O61{Re(CO)3}3{ORb(H2O)}(μ3-OH)]9− showed the instant formation of CS states with very short lifetime (below 10 ps) as evidenced by TAS.86,90 Longer-lived CT excited states have been described for Co-containing Keggin POMs, with a clear effect of the substitution site, either at the tetrahedral site or at a pseudo-octahedral addendum position.87 The dynamics of the excited states of [CoIIW12O40]6− and [SiW11O39CoII(H2O)]6− have been monitored after excitation at 400 nm. Both species exhibited a bi-exponential decay, with a first excited state decaying in a few hundreds of fs and a second longer-lived intermediate assigned to a CT state. The longer lifetime observed in [CoIIW12O40]6− is consistent with the ability of the central Co to undergo structural distortion and to display poor orbital overlap between the tetrahedral site and the reduced tungstate framework. A putative transient valence-trap with the excited electron localized on a single tungsten among the twelve was also proposed to explain the slower CR (τ = 420 ps in H2O, τ = 1700 ps in MeCN) of excited [CoIIW12O40]6− compared to that of excited [SiW11O39CoII(H2O)]6− (τ = 1.3 ps in H2O). Later investigation of the heterobimetallic TMS-POMs [CoIIW11O39{(MxOHy)}](12−x−y)− (MOHy = VIVO, CrIII(OH2), MnII(OH2), FeIII(OH2), CoII(OH2), NiII(OH2), CuII(OH2), ZnII(OH2)), did not show any improvement of the lifetime (sub-300 ps in aqueous media), probably due to the increase of the anion charge and ion-pairing.111 Excited state with even longer lifetime (τ = 20 ns) could be observed in the case of [X{Bi(dmso)3}2V12O33]−,89 in which the Bi(III) centre is not involved in the excited state solely localized on the dodecavanatade framework.
![]() | ||
| Fig. 4 Comparison of transient absorption spectra (thin black solid line) at indicated delay time and simulated spectra (green dashed line) of charge-separated states in for POM–Ir(III) hybrids: (a) Keggin-organosilane, (b) Keggin-organotin, (c) Dawson-organosilane, (d) Dawson-organotin. The simulated spectra are a sum of the spectrum of the mono-reduced POM (thick black solid line) and the oxidized iridium references (red dashed line). Reproduced from ref. 51 with permission from the Royal Society of Chemistry, copyright 2013. | ||
Among the developed POM–PS systems, the TAS study of those based on porphyrin were difficult to investigate owing to the complex relaxation dynamics (S2–S1, ISC) involved in these systems. Consequently, the persistence at long timescale of the signal of the porphyrin triplet state often hindered the observation of the CS state. Nevertheless, several studies with organic dyes or organometallic complexes allowed the kinetics of both CS and CR to be estimated in these POM–PS assemblies.
In the non-adiabatic limit (i.e. when the hybridization between D and A is limited), the rate constant for an electron transfer process between a donor (D) and acceptor (A) is described by the Marcus theory,112,113 and the driving force of the CS (ΔG0CS) and CR (ΔG0CR) are estimated using the Rehm–Weller equations.114,115
| ΔG0CS = Eox(D) − Ered(A) − E00 − wel |
| ΔG0CR = Ered(A) − Eox(D) + wel |
ν the electronic frequency depends, (among other parameters) on the overlap between the electronic wavefunction of the donor and acceptor units, λ is the reorganization energy and ΔG0ET is the free energy of the electron transfer. E00 is the energy of PS*. The work term for electrostatic interactions wel is often neglected in polar organic solvent. For polyanionic species, the exact evaluation of the reorganization energy is particularly challenging as it requires the exact distances between all charged species to be known, including the counter ions associated with the POMs. As a first approximation, it can be considered that the associated counter ions screen the charges of the POM and that the POM/counter ions behave as a global neutral species. Using such an approximation, when the distance between the POM and the PS is ca. 2 nm, the work term is calculated to be 0.08 eV in CH2Cl2 and 0.02 eV in MeCN and DMF, which probably underestimates its actual value.54 The redox and photophysical properties of all reported systems are reported in Table 1.
Our group reported the photophysical properties of a series of four POM–Ir(III) covalent assemblies (Fig. 4).51 The same cyclometalated Ir(III) complex was grafted on four hybrid platforms displaying distinct electron accepting properties. In these molecular systems in which the POM and the PS were electronically decoupled, we found that the CS rate increased with ∣ΔG0CS∣ (ranging from 0.14 eV to 1.00 eV) while the CR rate decreased with ∣ΔG0CR∣. This suggested, assuming a similar reorganization, that the CS process occurred in the Marcus normal region (∣ΔG0CS∣ < λ) while the more exergonic CR were located in the Marcus inverted region (∣ΔG0CR∣ > λ). In this specific case, it was also found that because of the heteroleptic character of the Ir(III) complex, the photoinduced electron transfer was favoured for the CS and delayed for the CR (Fig. 5), as the overlap between the electronic wavefunctions of the donor and acceptor groups differed in the Ir(III) excited state (donor: reduced picolinate ligand; acceptor: POM) and the CS state (donor: reduced POM; acceptor, oxidized Ir-phenylpyridine). The directionality in the photoinduced electron transfer is indeed a key parameter to obtain efficient charge photoaccumulation (see Section 4).
![]() | ||
| Fig. 5 Directionality in the photoinduced electron transfers of the POM–Ir(III) assembly [P2W17O61{O(SiC36H23O2N3Ir)2}]6−, conferred by the heteroleptic character of the Ir(III) complex. The red and blue cloud respectively represent the localization of the positive and negative charges in the different excited states.51 | ||
More recently a series of POM–bodipy hybrids were developed to avoid the use of noble metals. A similar trend (i.e. the CS and CR rates respectively increased and decreased with the driving force) was observed on these hybrids.53,54,68 In a work that combined experimental and theoretical approaches,54 the reorganization energy was estimated to be ca. 0.94–1.24 eV according to the nature of the POM or the solvent. Actually, the difference in reorganization energy was found to be ca. 0.05 eV smaller for the Wells–Dawson-type hybrids compared to the Keggin ones. This system was further investigated to evaluate the effect of the POM counter ion and solvent on the kinetics of photoinduced electron transfer (vide infra).
Other reports using TAS include Keggin- and Wells–Dawson-type POM connected to Ru(II) polypyridyl complexes,28,50 and Dawson-type POMs coupled to porphyrins.25,49 Very interestingly, when plotting the kinetics of CS and CR (when the charge-separated state was observed) in the logarithmic scale (kET = 1/τ with τ values listed in Table 1) according to the driving force of the electron transfer for all these hybrid systems, a correlation can be observed (Fig. 6). The slower CS processes are associated with small driving force (below 0.5 eV), while fast CS processes (below 100 ps) are only observed when the driving force is ca. 1 eV. Conversely systems with slow CR (above 10 ns) were found to display high driving force. In other words, systems that displayed high-energy charge-separated state (i.e., when POMs have low electron accepting properties) are slow for both CS and CR, while those based on POMs with good electron accepting properties, are fast for both electron transfer processes. The deviation from the observed trend is probably due to the difference in the chemical bridge between the POM and the PS (distance, conjugation, decoupling of the electronic state, directionality, etc.) and also the difference in solvent (vide infra) that has a significant effect on the photophysical properties in these systems.
![]() | ||
| Fig. 6 Correlation between kinetics of CS (blue circles) and CR (red circles) and the thermodynamic driving force (−ΔG0) in covalent POM–PS assemblies based on Keggin and Wells–Dawson structures according to Table 1. The alphabetic labels correspond to the same compound (see Table 1). Note that for points labelled O, P, Q the kinetics of CR is not reported in the literature. | ||
The photophysical properties of Anderson-type POM-based hybrid are more complex. This may arise from the fact that while in most reduced POMs the added electron is delocalized onto the whole POM framework, in Anderson–Evans, the first reduction corresponds to a process localized on the central metal (e.g. Mn(III), Co(III), Fe(III)). Consequently, the intrinsic electron delocalization of reduced forms of POMs is not present in mono-reduced Anderson systems, resulting in very unusual behaviour of the photosensitized hybrids. In some cases, no quenching of the excited PS (porphyrin,49 bodipy116,117) was observed owing to the rather negative potential of the reduced POM. By contrast with Ir(III) cyclometalated chromophores some partial, if not total, quenching of the luminescence was observed depending on the organic spacer between the POM and the PS. With an imine spacer, the reminiscence of the 3MLCT state with the same lifetime as the parent chromophore, suggested that this relaxed excited state is not responsible of the luminescence quenching. Indeed, with this relaxed excited state the driving force for the CS is very low (less than 0.15 eV). The authors attributed the partial formation of a charge separated state with a very fast process (ca. 0.5 ps) from a hot 3MLCTppy state79 even though the energy of hot excited state was estimated by other authors to be rather close to that of the relaxed 3MLCTppy state.118 The resulting POM–Ir charge-separated state displayed very short-lived character (<1 ns) while its energy is very high. More recently, the authors also evaluated the effect of the modification of the organic linker.80 Using a shorter organic tether between the Ir(III) photosensitizer and the POM, they observed an almost total quenching of the Ir(III) luminescence that can be attributed to a more efficient electron transfer from the photo-excited PS to the POM. From femtosecond TAS, the authors concluded that the CS was operative from the relaxed 3MLCTppy state with a kinetic of 180 ps. Yet with this system, the charge-separated state could not be observed, indicating that it would recombine at a faster pace.
Finally, the photophysical properties of an elegant POM-based molecular capsule consisting of diarylethene sandwiched by two trivacant lacunary phosphomolybdate [A-α-PMo9O34]9− was recently reported.55 This system features interesting photophysical behaviour as the diarylethene can either act as an electron donor or as a photochromic unit. Furthermore, the formation of triplet species was also proposed. In this article, the authors propose the formation of a charge separated state with an intriguing lifetime exceeding 200 μs, which exceed by far the values of all previously mentioned systems.
From the comparison of all the covalent POM–PS systems, it is found that those based on larger POMs (Keggin and Wells–Dawson structures) display the longest charge-separated state. This outlines the importance of the electron delocalization of the reduced POMs in the kinetics of photoinduced electron transfers with POM–PS hybrids. The comparison of two different PS (cyclometalated Ir(III) complex and bodipy) can be made on similar hybrid platforms and in the same solvent.51,54 It can be clearly found that the τCR/τCS ratio is considerably higher for the organometallic PS (τCR/τCS in the range of 50–380) as it displays directional electron transfer (Fig. 5), a feature that is absent in the bodipy dye (τCR/τCS in the of range 3–4). For these POM–PS systems, only fast photoinduced electron transfer (below 100 ps) could be observed when the driving force for the charge separation was important (ca. 1 eV), which implies that (i) some energy loss is necessary to ensure efficient electron transfer, (ii) only high-energy excited state PS and hybrid POMs with good electron accepting properties can lead to rapid photoinduced charge separation.
To illustrate this, organosilane hybrid derivatives incorporating Wells–Dawson and Keggin POMs were compared in CH2Cl2, DMF and MeCN (with tetrabutylammonium (TBA) hexafluorophosphate electrolyte), and the first two reductions found to be shifted by ca. 115–160 mV for these hybrids.54 Changing the electrolyte also altered the redox behaviour. In DMF, switching from TBAPF6 to tetraethylammonium (TEA) perchlorate and tetramethylammonium (TMA) tetrafluoroborate shifts the Dawson-based hybrid's first reduction by 230 mV. Molecular dynamics simulations revealed closer cation–POM interactions with smaller cations (TMA, TEA) than with bulky TBA. Solvent effects varied but the observed trends were consistent with coordination numbers and interaction energies, with TBA having weaker binding to the POM (especially in MeCN) but stronger solvent affinity (especially in DMF). DFT calculations showed that the LUMO energy of the hybrids depended on solvent polarity and number of POM-contacting counterions, with stronger ion pairing (for the smaller cations) leading to a lower LUMO. This sensitivity adds complexity to photophysical behaviour, influencing both the reorganization energy (λ ∼0.94–0.99 eV in CH2Cl2 to 1.19–1.24 eV in MeCN due to differences in polarity for the Keggin- and Wells–Dawson-bodipy) and thermodynamics of photoinduced electron transfer. In CH2Cl2 and MeCN, bodipy fluorescence was similarly quenched (∼85%) due to efficient CS. The similar CS rates in CH2Cl2 and MeCN suggest that CH2Cl2 compensates for its lower driving force with a lower λ. By contrast it was found that the charge-separated state lasts 3–4 times longer in CH2Cl2 than in MeCN. This is because, in CH2Cl2, CR has lower reorganization energy and a higher driving force. Since recombination occurs in the Marcus inverted region, both factors raise the activation energy, extending the excited state lifetime. In DMF, both excited-state decay and CS were surprisingly slow though DMF's redox potentials and λ values fall between those of CH2Cl2 and MeCN. In this case, it was proposed that the high viscosity of DMF and the strong TBA solvation likely hinder rapid counterion rearrangement, slowing electron transfer. Simulations confirm that in DMF, TBA cations are largely in the bulk, unlike in CH2Cl2, where they remain near the POM. This suggests that slow ion exchange in DMF acts as a gating mechanism for ET. The counterions also influenced the kinetics. In DMF, the TMA salt of the Dawson-bodipy showed faster decay than TEA or TBA analogues. In MeCN, TMA accelerated CS the most. In CH2Cl2, a mixed TMA/TBA salt displayed extremely fast charge injection (∼4 ps), but the result may be confounded by aggregation. Although solubility was limited, this hybrid displayed both ultrafast CS and long-lived charge-separated states, making it the most efficient system studied.
Photoinduced electron transfer in Keggin hybrids can also be modulated by protonation, shown with POM–bodipy and POM–Ir(III) organotin hybrids.31,68 The findings are consistent with proton-coupled electron transfer (PCET), where the interplay of protonation and electron transfer governs the CS and CR processes. In the polyoxomolybdate-based POM–bodipy hybrid, photoinduced CS occurred spontaneously in solution, while in the polyoxotungstate analogue, electron transfer was inefficient due to a poorer electron accepting property of the latter.68 A similar feature was later observed with POM–Ir(III) hybrids.31 Protonation (in this case via addition of trifluoroacetic acid (TFA)), shifts the redox potential of the POM to more positive values, enhancing the thermodynamic driving force for light-induced electron transfer (Fig. 7). In the polyoxotungstate–bodipy hybrid, TFA enables partial CS, but the driving force remained insufficient for high efficiency, and CR in tens of ns led to the formation of bodipy triplet states due to tungsten's heavy atom effect. In contrast, for the polyoxomolybdate analogue, TFA drastically accelerated CS (from 2.2 ns in the absence of acid down to 91 ps), and CR slowed (up to 1.3 μs), plateauing at 500 equivalents of TFA. A polyoxomolybdate–Ir(III) hybrid showed a similar trend albeit the increase in the electron transfer upon the addition of an acid source was less pronounced (CS accelerated from 60 ns in the absence of proton down to 26 ns in the presence of 250 equiv. TFA).31 In these hybrids, charge-separation enhancement with acid likely occurs via sequential proton transfer and electron transfer, while charge-recombination appears to involve a concerted or sequential mechanism influenced by POM basicity. Weak interactions between H+ and oxidized POMs complicate full protonation at low acid concentrations. A detailed theoretical investigation of the intricate mechanism of the protonation step in the excited state would be of high interest to account for these experimental findings.
![]() | ||
| Fig. 7 (Top) Energy level diagram of the polyoxotungstate- (blue) and poloxomolybdate- (orange) bodipy hybrids showing the effect of the presence of TFA in MeCN. (Down) Cyclic voltammograms of the POM–bodipy hybrids (black curves, top panel) 1 mM in MeCN containing 0.1 M TBAPF6 solutions and after the addition of 30 (red), 60 (orange), 100 (green), 200 (purple) and 500 (blue) equiv. of trifluoroacetic acid (TFA); working electrode: glassy carbon; reference electrode: SCE. Reproduced and adapted from ref. 68 with permission from the Chinese Chemical Society (CCS) Peking University (PKU), and the Royal Society of Chemistry, copyright 2021. | ||
Two main strategies have been followed when developing molecular systems able for charge photoaccumulation.125–127 The most prevalent one relies on the use of sacrificial electron donors (SEDs), in large excess, to efficiently reduce the charge-separated state, thereby minimizing charge recombination and promoting the formation of the targeted photoproducts. This method, as demonstrated in the seminal work of MacDonnell and co-workers, enabled the accumulation of more than two electrons within a single molecular acceptor.128 The second approach, which has not been yet applied to POM systems, involves the application of intense irradiation to excite simultaneously multiple photosensitizers that are spatially arranged around a central multielectron acceptor and peripheral single-electron donors. Similarly, the use of a pump–pump probe setup (also not explored with POMs systems) allows for the sequential excitation of a unique chromophore generating a doubly excited state.129,130 The challenges when dealing with charge photo-accumulation is that, once partially reduced, the electron reservoir can potentially promote new energy-wasting pathways that compete with the electron accumulation process. A simplified illustration of the charge accumulation pathway for a POM–PS dyad in the presence of a SED is depicted in Fig. 8. For sake of simplicity, here we consider the SED as an ideal single electron donor (which is rarely the case). Furthermore, we do not consider the reductive quenching of the excited chromophore by the SED but only the oxidation of the SED by PS+ in the charge-separated state. Ideally, the accumulation of two electrons in the POM proceeds through two light-induced charge separation steps. While the first reduction usually occurs efficiently, the second reduction is slower as reduced POMs, (which behave as electron donor and exhibits visible-light absorption) may undesirably quench the excited PS through reverse charge transfer (blue arrow in Fig. 8) or energy transfer (green arrow in Fig. 8). In general, the design requires a PS with sufficient reducing power to drive several electron transfers to the POM acceptor. In addition, it should also favour a faster accumulative pathway compared to the energy-wasting pathways and the spontaneous reoxidation of the POM, to ensure high yield of the accumulated reducing equivalents.
Our group has reported light-driven charge accumulation using the Wells–Dawson-type POM–heteroleptic Ir(III) assembly described above.26 In this hybrid, the excited Ir(III) PS has sufficient reducing power (E(PS+/PS*) = −1.36 V vs. SCE) to drive the two-electron reduction of the Dawson polyoxotungstate unit (ΔG01 ≈ −0.67 eV and ΔG02 ≈ −0.22 eV for the first and second POM reduction). Irradiation of the Dawson hybrid under visible light (λ > 400 nm) in DMF in the presence of triethylamine (TEA) as SED led to the stepwise reduction of the POM unit. Rapid formation of the one-electron reduced species (1000 W Xe lamp, τ1/2 = 43 s, τ1/2 being the time constant, ϕ = 10.5%) was followed by a slower formation of the two-electron reduced species (τ1/2 = 270 s, ϕ = 2.3%). It was shown that adding acetic acid (AcOH) considerably accelerated the rate of the formation of the second reduction (τ1/2 = 60 s) by promoting PCET for which POMs have strong propensity. The system was also able to produce hydrogen, albeit at low pace, achieving a turnover frequency (TOF) of ca. 0.25 h−1 and a turnover number (TON) of up to 41 under irradiation for 7 days. Expanding our work on POM–Ir(III) systems, we replaced the picolinate ligand with a less acid labile imidazo-1,10-phenanthroline scaffold and very recently reported the accumulation of up to three electrons in the polyoxomolybdate unit when irradiated in DMF in the presence of TEA and TFA.31 Meanwhile, no electron was stored in the polyoxotungstate counterpart in which the POM reduction potential is more negative and CS remains ineffective even with acid present. Overall, within these POM–Ir(III) hybrids, the SED is not directly involved in reducing the excited Ir(III)* PS. Instead, the hybrid first undergoes a fast CS, that is, an oxidative quenching of the excited PS by the POM, prior to the regeneration of the PS by the SED.
While often overlooked, the role of the SED is also crucial in the development of charge accumulating systems131,132 and this was addressed in our very recent work on the POM–bodipy hybrid [P2W17O61{O(SiC31H30N2BF2)2}]6−.110 Several SEDs such as xanthate, triethanolamine, TEA were tested. Yet only TEA (1 M in CH3CN) promoted the two-electron photo-accumulation in this noble-metal-free hybrid, which completed within 10 min when irradiated with visible light (300 W Xe lamp, λ > 385 nm) under argon atmosphere. Evaluation of the kinetic profile reveals that the first reduction of the POM is efficient and even improved by acid (ϕ = 11% to 23%). By contrast, the formation of the two-electron reduced POM is much slower (ϕ = 0.6% in the absence of acid), more complex and highly dependent on the strength and quantity of the acid (ϕ = 5.2% in the presence of 20 mM TFA, Fig. 9). The aid of theoretical calculations elucidated an intricate mechanism of the formation of the two-electron reduced species, which is influenced by PCET and acid-promoted POM dismutation. Another pathway involves thermal reduction by the TEA(–H)˙ radical, bypassing the photon absorption path, and overall rendering TEA a two-electron, one-proton donor. This study underscores the importance of considering the potential non-innocent role of the SED byproducts and also taking into account how the SED affects the redox properties of the hybrid POM given that it is often used in large excess (or even as co-solvent).
![]() | ||
| Fig. 9 (Left) Molecular representation of the Dawson–bodipy hybrid. (Middle) Evolution of the absorbance spectra upon irradiation of the hybrid (0.2 mM) in MeCN containing 1 M TEA without acid with 20 mM TFA. (Right) Kinetics of electron accumulation of the hybrid (0.2 mM) in 1 M TEA in MeCN at varying amount of AcOH or TFA monitored at λabs = 690 nm. Reproduced and adapted from ref. 107 with permission from the Chinese Chemical Society (CCS) Peking University (PKU), and the Royal Society of Chemistry, copyright 2025.110 | ||
Charge accumulation was also demonstrated by an unconventional example consisting of a Cu(II)-dipyridylamine complex grafted onto the {VO6}3 cap of a mixed Dawson POM, [P2V3W15O62]9−.30 Here, three electrons were successively stored photochemically (one through Cu(II)/Cu(I) reduction, and two through V(V)/V(IV)-centred reductions), although the electron donor remains unidentified. Interestingly, the hybrid was useful in photocatalytically mediating CF3 radical generation via reduction of Togni reagent II, opening avenues to synthetic applications of charge photoaccumulation. By far, non-covalent POM–PS systems exhibiting charge accumulation has been rare. Ishitani and coll., very recently reported a hybrid from the 1
:
1 ion pairing between a cationic ring-shaped Re(I) tetranuclear complex and a Keggin-type POM, ([XW12O40]4− (X = Si, Ge)29 electron in the ring). Despite the remarkable number of electrons stored, only one or two with sufficient reducing power were supplied to an external catalyst and drive the CO2 reduction reaction (CO2RR). Due to the complex composition of most charge accumulation systems, numerous competing pathways that diminish the yield of accumulated reducing equivalents can take place. This may result from side reactions with SED by-products or the inherent reactivity of reduced POMs to traces of oxidants, especially molecular oxygen or oxidizing species formed during the photoirradiation, as we have observed with the POM–bodipy system.110
The quenching of the excited PS by reverse electron transfer from the reduced POM (also called ‘ping-pong’ mechanism)64,78 is also a limiting issue given that it is a thermodynamically favourable process (Fig. 8). One way to circumvent this is by designing a POM–PS hybrid that possesses rigid molecular bridges and that promotes directional electron transfer (such as in the initial POM–Ir(III) system).26 Establishing design criteria that improves the stability of photoreduced POM–PS systems and understanding the mechanism of re-oxidation in the dark are some aspects worthy of attention to maximize charge accumulation.
The comparison of the POM–PS compounds displaying charge photoaccumulation properties with other non-POM based systems, in the presence of SEDs, is tedious as the quantum yields of the photoaccumulation are rarely reported.128,133 In several cases, the efficiency of the overall process were low, probably owing to the presence of aforementioned energy-wasting pathways.130,134 Yet in one case, a remarkable quantum yield of 66% was reported for the 2e− photoreduction of a Ru(II) complex featuring a fused dipyridophenazine–pyridoquinolinone acceptor ligand.135 The high efficiency of such system probably relies on the fact that the electron acceptor is involved in the PS* MLCT state that readily react with the SED, a feature that cannot be present in a POM–PS system. Charge photoaccumulation, in the absence of SED, has rarely been reported, and never with a POM acceptor. In most systems featuring a central multielectron acceptor and peripheral PSs, very fast charge separation (below 50 ps) with a low driving force (below 0.5 eV) is observed.136,137 This characteristic, however, appears challenging to achieve upon electron transfer to POMs owing to the high reorganization energy associated with their polyanionic nature, including solvation effects and counter cations reorganization. Another inspiring strategy, developed by Odobel, Hammarström and coll., relies on the grafting of the molecular system to TiO2 nanomaterial.129 Using such an approach, they reported the quantitative formation of a doubly charge separated-state with the oxide materials providing directionality for electron transfer and serving as an electron reservoir.
We have briefly mentioned in the previous examples how the accumulated electrons in the POM can be “released” to drive the reduction of small molecules. Compared to most photocatalytic systems, the most advantageous practical feature of POM–PS is the possibility to temporally decouple photochemical electron storage and thermal re-oxidation in the dark which enables on-demand solar fuel production, like HER. The group of Cronin and Symes has spurred the application of the decoupling strategy through their pioneering works where they used POMs as mediators to temporally and spatially separate H2 and O2 generation in redox flow batteries and water electrolyzers.138–141 More recently Li, Chen and coll. pursued this approach by incorporating BiVO4 in the system to construct a photo-electrocatalytic system able to decouple H2 and O2 evolution in solar-driven water splitting.142 This concept was also adapted in photochemical POM–PS systems by Streb, Rau, Dietzek and coll. who prepared a hybrid comprising a Ru(II) poylpyridyl complex PS linked to a Dawson polyoxotungstate via a phosphonate group.28 Quantitative accumulation of two electrons was observed after 6 min of irradiation by a monochromatic LED (λmax = 470 nm) in the presence of SEDs such as sodium ascorbate (NaAsc, ϕ = 0.25%) or triethanolamine (TEOA, ϕ = 0.05%). Addition of H2SO4 (>10
000 eq.) in the dark triggered instant H2 generation (up to ∼40% yield) and complete re-oxidation of the Ru(II) POM–PS. Remarkably, they were able to store the reduced POM in the dark under inert atmosphere for a long period (t1/2 >24 h) and further showed acid-induced H2 release after 2.5 h storage albeit at a reduced yield (33%) due to partial re-oxidation from an unknown background process.
Since the area of POM–PS photoaccumulation systems is still under development, there is insufficient understanding of reactivity once reducing equivalents are accumulated in the POM, although most systems have been geared towards HER. To understand the “release” of the stored electrons, we can gain insights from studies of the behavior of reduced POMs (i.e., reduced via electrochemical or chemical means) exhibiting electron discharge upon the addition of an organic acid.
Guillemot and coll., recently described the study of electron charge/discharge on hybrid Keggin-type polyoxotungstates featuring tertbutylsilanol function coordinated to oxovanadium (V
O).143 While electrochemical reduction was facilitated in the presence of TFA by virtue of PCET, no hydrogen production was observed due to the weak reducing potential of the electrochemically generated reduced POM species. Chemical reduction using sodium naphthalenide in THF, in contrast, allowed the isolation and characterization of the different reduced forms of the parent POM. One equivalent of the reductant leads to the reduction of the V(V) center to V(IV), while the second equivalent reduces the polyoxotungstic framework leading to a bireduced species possessing decoupled d1-V(IV) and d1-W(V). Addition of lutidinium triflate as proton source to the latter resulted in the release of H2 and the formation of a monoreduced, monoprotonated species. The same chemical reduction carried out on the same POM without the oxovanadium confirmed that hydrogen evolution takes place at the reduced polyoxotungstic framework. In this system, it was proposed that proton-driven disproportionation of the bireduced species enables the release of hydrogen.
More recently, Matson and coll., also reported spontaneous HER by release of electrons from a chemically mono-reduced [PW12O40]3− in the presence of stoichiometric amounts of a strong organic acid, diphenylammonium (H2NPh2+).18 However, as the monoreduced, monoprotonated species is unstable, it cannot be isolated and characterized. They argued that the ability of this POM to produce H2 is related to the bond dissociation free energy (BDFE) of the O–H bond on the reduced-protonated species which they estimated from the pKa and the standard redox potential, E°, of [PW12O40]3− as described by the Bordwell equation (see below), where Cg is a constant associated to proton reduction in the given solvent.144–146 The study confirms that the BDFE(O–H) of monoreduced, monoprotonated [PW12O40]3− is comparable to that of H2 BDFE(H–H) in acetonitrile (approx. 50 kcal mol−1) which implies that hydrogen release is thermodynamically favored.
| BDFE (X–H) = 1.37pKa + 23.06E° + Cg |
Extending this BDFE-guided study, the same team look at the effect of substituting a W- or P-atom in [PW12O40]3− by a V(V) resulting to [PVW11O40]4− and [VW12O40]3−, respectively.147 The presence of a vanadium dopant positively shifted the first redox potential in both complexes. The presence of V atom at the surface in [PVW11O40]4− also increases its basicity and therefore its affinity for hydrogen atoms. The combined factors ultimately gave rise to higher BDFE (O–H): 68 kcal mol−1 in [PVW11O40]4−vs. 48 kcal mol−1 in [PW12O40]3−. Conversely, the presence of an internal V atom dopant in [VW12O40]3− does not affect the surface basicity compared to [PW12O40]3−. As a result, HER is less favorable in both cases. The same pattern is observed for vanadium-substituted Lindqvist-type polyoxotungstate, leading to a stable mono reduced monoprotonated species with BDFE(O–H) of 64 kcal mol−1.148 Overall, this shows that BDFE is a useful metric for predicting a POM's ability to facilitate HER or H-atom transfer, and can guide the design of more effective POMs that can be adapted in hybrid systems.
Although self-evident, it is often overlooked that the reactivity of reduced POMs is ultimately governed not only by the number, but also importantly the reducing strength of the stored electrons. It should be pointed out that while protons facilitate charge accumulation through PCET, this also simultaneously diminishes the reducing power of the stored electrons. Thus, a finely tuned balance of conditions is necessary. While the preceding examples are not photochemical, they highlight the advantage of spatial and/or temporal decoupling between POM reduction and subsequent proton reduction which is a strategy that can preserve the reducing power while promoting HER or the reduction of other substrates.
Other than being catalysts themselves, POMs can also take the role of redox mediators or electron shuttles between the photoactive materials and the relevant catalyst. When the development of photo-redox catalysis comes to the interfacing with transparent (semi)-conductive oxide electrodes as in Photo Electrochemical Cells (PECs) and Dye Sensitized Photo Electrochemical Cells (DSPECs), the redox mediator can turn out to be decisive to spatially extend the CS, retard the electron–hole recombination and eventually improve the catalytic efficiency.151 Note that the lifetime gain is however at the expense of a loss of the driving force to reduce the catalyst.152 Indeed, photophysical studies on POM–PS hybrids have shown that electron transfers onto/from POMs were significantly slow when the driving force was low, which could be a limitation in the use of these species as redox active shuttle. Mimicking the Z-scheme of natural photosynthesis, Abe and coll. have used [SiW11MoO40]4−/5−and [SiW11O39MnIII(H2O)]5−/6− couples as effective redox mediators in a visible light-driven multi-photon water splitting system.153 The electron shuttle role is akin to that of POMs as charge transport layers in optoelectronic devices, where they improve the charge injection at the electrodes.154 POMs have indeed been diversely applied to improve the performance of dye-sensitized solar cells,155 either as redox mediators, photosensitizers, interface modification and catalysts for the counter electrode. A study by Gibson, Fielden and coll. showed how Lindqvist polyoxometalate hybrids, co-assembled with a photosensitizer on the surface of NiO photocathodes, could enhance the photovoltage (VOC) by up to 140%.156 This was caused by CT from the photoreduced dye to the POM electron acceptor and a positive shift in the NiO valence band edge. Charge lifetime and transient IR measurements confirmed that POMs retard both recombination and electron transfer processes. The approach can lead to significant efficiency gains particularly in systems with poorly matched dye/redox mediator combinations. These improvements, however, were offset by reductions in short-circuit current density (JSC). Nonetheless, the approach, if well engineered, could enable regeneration of the PS where there is slow charge-extraction by the redox mediator or catalyst and prevent photodegradation and/or suit applications in photocapacitors.
Biological systems possess self-repairing capabilities that help restore or extend their functions (e.g. movement, motility, chemical activity). In natural photosynthesis, for instance, the oxygen-evolving catalyst regenerates approximately every 30 minutes under solar irradiation.11 In contrast, human-made materials, and particularly those exposed to harsh conditions such as solar flux, degrade over time due to damage induced under operation conditions. Until now, artificial photosynthesis has mostly focused on the performance and robustness of the device. Yet, considering that artificial systems maintained under illumination in aerobic conditions can operate without self-protection is reckless. While the concept of self-healing in artificial photosynthetic device is still in its infancy, supramolecular chemistry with POMs provides tools allowing the elaboration of such materials. For instance, POMs have been integrated in few self-healing gels.157–159 They also behave as covalently dynamic species in water due to the reversible nature of their structure.160,161 Such a property was leveraged by Hill, Weinstock and coll. to develop all-inorganic equilibrium systems capable of regenerating the catalytically active species.162 Similarly, self-repairing properties of hybrid POMs were also reported in case of a reversible anchorage of the organic moieties onto the POM framework.163 Considering the dynamic nature of the covalent bonds of POMs and the various modes of interactions at work in related materials,164,165 POMs appear as promising building-blocks for the elaboration of future generation of self-repairing materials for energy applications. Supramolecular chemistry can also be used to control the molecular organization between the polyanionic POM and the organic PS. Typically, controlling molecular organization in photoactive thin films is crucial for enhancing optoelectronic performance.166,167 For example, the self-assembly of electron donor and acceptor units into highly ordered architectures offers ideal percolation pathways for charge carriers through well-defined nanosegregated D and A domains.168,169 Furthermore, experimental investigations have shown that D–A systems that assemble into extended architectures can enhance the charge transport properties and considerably prolong the lifetimes of photogenerated excitons.170 However, creating nanostructured arrays with regularly alternating D and A regions remains challenging due to entropic constraints and the natural tendency of D and A units to stack via electronic complementarity, hindering charge transport.170 A promising approach is to design D–A systems with chemically distinct components to promote effective nanosegregation. As a proof-of-concept, we developed a photoactive mesogenic hybrid POM.71 In the solid state, the compound, displays a multi-lamellar organization in which double-layers of POMs and the mesogenic organic antenna alternate regularly. Preliminary studies on a next generation of photoactive mesogenic systems showed that such materials display in the solid-state charge-separated states with very long-lived character.171 Considering the importance of CT processes, such self-assembled materials displaying large D–A interfaces and directional hole/electron transporting pathways could have important impact to further guide the design of optoelectronics materials.
Charge photo-accumulation potentially affords more powerful reductants and could favour multi-electronic reduction processes. However, its scope is much more limited than electrochemically driven electron storage and may be hindered by the reverse electron transfer from the reduced POM to the excited photosensitizer.78,110 The use of chromophores displaying polarized excited state, such as in heteroleptic metal complexes51 or in push–pull organic dyes172–174 can provide directionality to the photoinduced electron transfer, and hence preclude the undesired reverse electron transfer. While electron photoaccumulation has been reported several times on POM–PS the mechanism is not fully understood, especially as these studies required the use of SED, which can be non-innocent in the electron accumulation mechanism. Important effort should be devoted now to perform reversible multi-photon induced charge accumulation in the absence of SEDs.127 To this end, hybrids associating POMs to multiple photosensitizers are potentially prone to display such multiple charge-separated states using intense laser excitation energy, as observed by pioneering works of Wasielewski and coll.136 Typically as many POMs feature bielectronic reduction waves in the presence of an acid source, potential inversion may favour the second reduction of the central core and may lead to long-lived multiple charge separated states, as reported by Wenger and coll.175 While some information is now accessible on monoelectronic photoinduced electron transfers in POM–PS hybrids, it seems of prime importance to evaluate the kinetics of CS/CR of multiple excited states, and study the effect of the POM protonation in these systems. Understanding the recombination mechanisms of multicharged POM–PS species is of critical importance for carrying out multielectron photoredox catalysis. Another way of performing electron photoaccumulation in the absence of a SED would be to graft the POM–PS species onto a semiconductor (NiO) that would act as an electron pool, inspired by the system developed by Odobel, Hammaström and coll.129 However, in such a device, protonation is also expected to influence the semiconductor Fermi level by modifying the band edges, which in turn could impact charge transfer at the electrode/PS interface. For NiO, a relatively large driving force for charge separation is required (>0.6 eV), whereas for TiO2 only ca. 0.1–0.2 eV is sufficient.176 The local environment must also be taken into account (e.g., screening effects, supporting electrolyte, salts, and solvent). In the case of polyanionic species such as POMs, where redox processes are coupled to cation motion and local dielectric, charge stabilization is hindered if ions cannot reorganize rapidly. In solid films or at electrode surfaces, limited ion mobility or the presence of unfavourable counterions could therefore restrict charge accumulation.177
The implementation of POM-based photocathodes is thus highly desired, the photoactive materials could be either a molecular photosensitizer or the semi-conducting material (e.g. Cu2O, CIGS). Dye-sensitized photocathodes are a promising approach for converting sunlight into chemical fuels such as H2 or CO.178–180 They are often paired with photoanodes (e.g., BiVO4 or dye-sensitized TiO2) in tandem PECs. In a solar cell, a reversible redox mediator completes the circuit by shuttling charge between the two photoelectrodes.176,181 In fuel forming PECs, the photoanode oxidizes water, while the photocathode reduces protons to hydrogen or CO2 to fuels, enabling overall solar-driven water splitting or CO2 reduction. The fuel forming photocathodes typically consist of a p-type semiconductor (commonly nanostructured NiO) coated with a light-absorbing dye and a molecular catalyst. Suitable dye features have been recently reviewed.182–184 Device performance is tuned by adjusting dye structure, catalyst properties, and the nanostructure of the semiconductor to balance light absorption, CS, and catalytic activity. Despite much interest, challenges arise due to fast recombination between the reduced dye and oxidized NiO, which limits efficiency. Strategies such as using redox mediators, optimizing dye/catalyst arrangement, and developing new materials are being pursued to address this. As a proof of principle, we demonstrated that Keggin-type POM hybrids co-grafted with a push–pull dye onto nano-ITO cathodes have an amplification effect on the photocurrent response.179 By anchoring POMs with carboxylic or diazonium groups onto the electrode surface, we achieved controlled POM loading and improved stability against leaching. Remarkably, even at low loading, the POM–COOH hybrid increased the photocurrent response by up to 25 times. It was found that the anchoring group not only stabilized the POM on the electrode but also influenced the complex interplay of electron transfer processes, ultimately boosting photoelectrochemical efficiency. Besides, we observed that the POM presence increased the endurance of the photocathode. Similarly, in the solution state, it was shown that [Ru(bpy)3]2+ exhibited marked photostability when associated with a Wells–Dawson type POM.33
To date, covalently bound POM–PS hybrids have not been applied directly in photocathodes, despite many of the photosensitizers used having been integrated in solar cells or fuel forming PECs.185 Bodipy dyes,186 for example, can be anchored to metal oxide surfaces (such as TiO2 or NiO) either via traditional carboxylate groups, directly through the boron center or chelating nitrogens, resulting in robust attachment and efficient charge injection.187 The natural next step would be to develop POM-based photocathodes and, by coupling the dye with suitable molecular catalysts, to see whether the kinetics can be modified, using the molecular systems described above. The flexibility of POM–bodipy hybrids would allow optimization of energy level alignment at the NiO and catalyst interfaces for efficient charge injection and transfer, which is key for photocathode performance. The length, position and conjugation of the spacer between the photosensitizer and anchoring group can be tuned to compromise between fast injection and slow recombination.
Challenges for the integration of hybrid POMs into efficient photocathodes may arise in ensuring good solubility, limiting aggregation on the surface, and good electronic communication through the molecules to the metal oxide. This will be even more challenging when considering the use of aqueous electrolyte (and hence evaluate the stability of the resulting assemblies according to the pH of the solution) and the integration of the catalyst. It will be necessary to test the impacts of structural modifications, electronic coupling and spatial separation between the components, and the environment (electrolyte ions, solvent and pH, for example) on light absorption, photophysics, robustness and catalytic turnover. Therefore, a multidisciplinary approach is needed. However, POM–PS hybrids represent robust, versatile and powerful platforms for advancing the efficiency and functionality of DSPECs for both solar electricity and solar fuel production. Note that a significant step toward practical devices independent of dye design was made by Ryu and coll., who reported a rare example of a POM-based PEC for bias-free solar water splitting. It uses visible-light-active Cu2O and BiVO4 electrodes, functionalized with TMS-POM catalysts and assembled via the Layer-by-Layer technique.188
Finally, the use of reduced POMs photosensitivity has emerged as an additional asset for visible-light conversion. Following a series of reports on the visible-light driven reduction of CO2 to CO over an association of chemically- or electrochemically-reduced H3PW12O40 and rhenium(I) pyridyl–carbonyl complexes and supported by computational studies, the multiple role of the POM has been deciphered and fully exploited by Neumann and coll. A dispersion of graphitic carbon nitride g-C3N4 loaded with the POM and a rhenium catalyst was irradiated with a white LED for a Z-scheme like biphotonic excitation. Photocatalytic hydrocarbon dehydrogenation at the carbon nitride (blue light) was coupled to an electron transfer to the POM and to a subsequent photo-induced electron transfer from the POM to the Re(I) catalyst, by excitation of the intervalence charge transfer band (IVCT) of the bi-reduced POM (red light).189 In a related example, HER triggered by visible-light induced electron transfer from a reduced form of the Wells–Dawson [P2W18O62]6− to a cobalt catalyst was described by Duan.190 This still largely unexplored photo-sensitization of POMs, which has also found recent applications in photothermal therapy191 or photothermal catalysis,192,193 provides a new tool to increase the POM reducing power, and creates a need for insights in the photophysics of reduced POMs.
The POM family offers a full set of tuneable physicochemical properties to devise POM-based photocathodes and DSPECs with improved efficiency and durability for solar energy conversion and multi-electronic processes. One challenge is to make the electron transfer cascade timescales compatible with the fast kinetics of photocathodes, especially those using NiO as a p-type semi-conductor. Furthermore, a robust trapping layer incorporating POMs could be of interest for the photoprotection of molecular photoelectrodes where photodegradation is often attributed to a slow charge extraction, leading to the presence of instable radicals in the materials.194 Integration of POM-based materials into real device configuration will thus rely on a compromise between synthetic complexity and photophysical performance.
| This journal is © The Royal Society of Chemistry 2025 |