Skip to main content
ACS Omega logoLink to ACS Omega
. 2021 Oct 26;6(44):29291–29324. doi: 10.1021/acsomega.1c04018

Reticular-Chemistry-Inspired Supramolecule Design as a Tool to Achieve Efficient Photocatalysts for CO2 Reduction

Bilal Masood Pirzada †,*, Arif Hassan Dar , M Nasiruzzaman Shaikh §, Ahsanulhaq Qurashi †,*
PMCID: PMC8581999  PMID: 34778605

Abstract

graphic file with name ao1c04018_0039.jpg

Photocatalytic CO2 reduction into C1 products is one of the most trending research subjects of current times as sustainable energy generation is the utmost need of the hour. In this review, we have tried to comprehensively summarize the potential of supramolecule-based photocatalysts for CO2 reduction into C1 compounds. At the outset, we have thrown light on the inert nature of gaseous CO2 and the various challenges researchers are facing in its reduction. The evolution of photocatalysts used for CO2 reduction, from heterogeneous catalysis to supramolecule-based molecular catalysis, and subsequent semiconductor–supramolecule hybrid catalysis has been thoroughly discussed. Since CO2 is thermodynamically a very stable molecule, a huge reduction potential is required to undergo its one- or multielectron reduction. For this reason, various supramolecule photocatalysts were designed involving a photosensitizer unit and a catalyst unit connected by a linker. Later on, solid semiconductor support was also introduced in this supramolecule system to achieve enhanced durability, structural compactness, enhanced charge mobility, and extra overpotential for CO2 reduction. Reticular chemistry is seen to play a pivotal role as it allows bringing all of the positive features together from various components of this hybrid semiconductor–supramolecule photocatalyst system. Thus, here in this review, we have discussed the selection and role of various components, viz. the photosensitizer component, the catalyst component, the linker, the semiconductor support, the anchoring ligands, and the peripheral ligands for the design of highly performing CO2 reduction photocatalysts. The selection and role of various sacrificial electron donors have also been highlighted. This review is aimed to help researchers reach an understanding that may translate into the development of excellent CO2 reduction photocatalysts that are operational under visible light and possess superior activity, efficiency, and selectivity.

1. Introduction

The incessant demand for energy, unabated deterioration of fossil fuel reserves, and global warming are three serious challenges mankind is facing at present.1 It is believed that the continuous release of CO2 into the atmosphere through anthropogenic activities, generally by burning of fossil fuels, affects the global temperature with drastic consequences.2 In this direction, the sequestration and subsequent chemical processing of CO2 have become a global focus so as to reduce its emission and accumulation. Various methods have been tested by researchers so far, which include: (1) sequestration and storage of CO2 under the earth and ocean;3,4 (2) chemical processing of CO2, and (3) biological fixation of CO2.5,6

However, there are various limitations associated with the above-mentioned procedures concerning viable applications. Geological dumping is not considered an efficient method as it may lead to wastage of resources due to leakages of CO2 gas. CO2 dissolved in oceans may acidify the waters, which will affect the marine ecosystem. Further, the mineralization of CO2 could be very costly and a potential hazard to the environment.3,4 Biological CO2 fixation is a feasible method, but the uneven distribution of the global forest area and deforestation are limiting factors.

Hence, the current focus is mainly on the research related to the chemical processing of CO2. However, the thermodynamic stability and limited reactivity of CO2 make its sequestration and chemical processing a daunting task. CO2 has a very high standard Gibbs free energy of −394.39 kJ mol–1,4 indicating that it is a kind of inert molecule. Thus, activating this molecule requires overcoming the kinetic inertia and the thermodynamics energy barrier. Scientists have applied high temperature and7,8 high pressure9,10 and employed various catalysts,11,12 including coordinating activation,13,14 acid–base synergism-based activation,15,16 photoelectrochemical activation,1720 enzymatic activation,21 and plasma-based activation.22 However, these procedures did not produce viable results. More recently, photocatalytic CO2 reduction is being foreseen as the most effective method for CO2 reduction due to the easy availability of sunlight, ambient reaction conditions, and simpler processing.4,23,24 However, there are various obstacles to the development of optimized systems for photocatalytic reduction of CO2 into C1 compounds, which will be discussed below in detail.

1.1. Bottlenecks

Photocatalytic reactions are determined in terms of three phenomena that actually control the economic viability of the reaction, i.e., efficiency, activity, and selectivity. The interplay among these three plays a pivotal role in photocatalytic CO2 reduction into value-added carbon products (Figure 1). The efficiency depends on the catalyst’s optical and electronic properties, and it determines the energy cost of the reaction. While optical properties could be optimized by band-gap engineering, charge carrier mobility is an intrinsic material property. The optical property determines the quantum efficiency, while the electronic property influences the faradic efficiency. The activity is a function of the concentration of active sites, and it governs the turnover and yield of the photocatalytic reaction. Thus, the photocatalyst morphology, dimension, and surface area are pivotal to achieve the desired activity. The productivity and purity of a product are governed by selectivity, which is the function of the catalyst’s chemical nature. The selective binding efficiency and the redox properties of the photocatalyst are key in determining the selectivity of the desired product. Albeit a huge number of photocatalysts have been developed by researchers, the central bottleneck is that all three phenomena could not be combined together in one material. Nature has evolved perfect catalysts to fix CO2 with remarkable activity and selectivity. However, the efficiency is limited since a vast source of light is applied to reduce a small concentration of CO2 in the atmosphere at a particular time. Hence, a lot of energy goes unused.

Figure 1.

Figure 1

Interplay of activity, efficiency, and selectivity of photocatalysts for reduction of CO2 into C1 compounds.

Metal and metal oxide heterogeneous catalysts are the most primitive class of photocatalyst materials synthesized.25,26 These materials exhibit enhanced efficiency and activity subject to optimizing their morphological, optical, and electronic properties.27,28 However, using heterogeneous inorganic catalysts, it is very difficult to control all three phenomena at the molecular level rationally.29 As a step forward, homogeneous molecular catalysts came to the scene. These catalysts can be tailored toward enhanced selectivity and better activity by tailoring the organic ligands. However, the activity could be a limitation.26,30 The main causes of the limited activity could be the catalyst deactivation and solubility issues of the catalyst. Furthermore, molecular catalysts require multiple components, which hinder the activity due to the randomness issues in the solution. Hence, if we could combine the multiple components with the desired features in an ordered manner, a thought arises that they can perform synergistically.31 This prospect of achieving all three aspects is made possible by exploring reticular chemistry.26 Hence, researchers are trying to combine different types of components to design a photocatalyst with synergistic properties. In this direction, molecular catalysts have gained greater attention due to the wide scope of molecular engineering potential in them.26

1.2. Why Supramolecules?

One of the most serious limitations of photocatalytic CO2 reduction is that it requires a high reduction potential to activate CO2 molecules. The photogenerated electrons should have an extra potential (overpotential) to carry on the efficient photoreduction of CO2.32 The primary activation of CO2 to CO2 requires a standard potential of −1.90V vs NHE at pH 7, which is very high (eq 1).33,34 The required potentials (pH = 7) to achieve various solar fuels from CO2 reduction are presented in eqs 1–8.24,32,35

1.2. 1

However, multiproton- and/or multielectron-mediated reduction of CO2 requires a smaller potential30,34,36 by overcoming high thermodynamic thresholds, which originate from a single-electron CO2 reduction to form CO2. It can be inferred here that methane and methanol formation from CO2 reduction requires a smaller reduction potential. However, these reactions require more electrons. Hence, these reactions are kinetically controlled where methane and methanol production is much difficult with respect to carbon monoxide, formic acid, and formaldehyde, which require a smaller number of electrons.37 The reactions that require 2–8 electrons and protons for the reduction reactions are very much difficult to occur.32

Inorganic photocatalysts were not found to be effective for multielectron processes due to their complicated electrochemical surface and lower charge mobility. Hence, it is imperative to explore molecular and supramolecular photocatalysts with a superior visible-light absorption coefficient, generating sufficient electrons, and have a high overpotential, which could improve the reaction efficiency.38 Supramolecules are seen as highly attractive photocatalysts as they exhibit good selectivity, activity, and efficiency. More importantly, the modifications in the catalyst structure can be easily engineered to suit the desired reaction through synthetic means.39

2. Results and Discussions

2.1. Evolution of Supramolecule Photocatalysts

In view of the above introduction, the transition from molecular catalysis to development of supramolecule photocatalysts is discussed below.

2.1.1. Two-Component Systems

The most primitive molecular systems as photocatalysts for reduction of CO2 are rhenium(I)-based molecular systems.1,4042 Lehn and co-workers (1983) presented the first report regarding the fabrication of a Re-based complex—fac-Re(bpy)(CO)3X (bpy = 2,2′-bipyridine; X = Cl or Br)—and investigated its photocatalysis for CO2 reduction in a mixed solution of triethanolamine (TEOA) and N,N-dimethylformamide (DMF).43,44 They obtained a good selectivity and efficiency for CO production (ΦCO = 0.14; X = Cl). However, these complexes exhibit stability issues, resulting in smaller turnover numbers (TONCO = 27; X = Cl) and small CO production, and they also possess a weak visible-light absorption coefficient. To address these limitations, mechanistic investigations were performed using laser flash photolysis techniques.4547 It was observed that the lowest excited state of the rhenium(I) complex, i.e., the triplet metal-to-ligand-charge-transfer state (3MLCT), undergoes reductive quenching by TEOA without undergoing oxidative quenching by CO2. This leads to the generation of one electron-reduced species (OERS) of the rhenium(I) complex. Since the removal of the Cl ligand associated with the OERS is a delayed process, a 17-electron species is being observed. Fujita et al. reported that [Re(4,4′-Me2bpy)(CO)3] can take up CO2 despite its reduced and coordinatively unsaturated nature. However, this reaction was found to be very slow as the extra electron got mainly accumulated on the bpy ligand moiety, represented as ReI(4,4′- Me2bpy•–)(CO)3. This electronic structure must be more feasible with respect to Re0(4,4′ Me2bpy)(CO)3. Hence, it was inferred that the slow formation of the former might be the reason for the smaller activity of rhenium(I) complexes for the photocatalytic CO2 reduction.48 Ishitani and co-workers reported that using only one rhenium(I) complex inhibits the activity. They cite the reason that the OERS generated in the process must involve two opposite characteristics: one is the stability for behaving as an electron donor, and another is the instability for producing the “CO2 adduct”.

To address these limitations, Ishitani and co-workers proposed the two-component-based supramolecular photocatalytic systems. These photochemical systems comprise a redox photosensitizer complex that actually facilitates the one-electron transfer and a catalyst component that modulates the one-electron transfer to the multielectron reduction of CO2, requiring a smaller negative reduction potential. Ishitani and co-workers used the fac-[Re{4,4′-(MeO)2bpy}-(CO)3{P(OEt)3}]+ complex as the photosensitizer component and fac-[Re(bpy)(CO)3(MeCN)]+ was taken as the catalyst component. The OERS generated by the photosensitizer complex was found to be highly durable as its P(OEt)3 ligands exhibited an enhanced π-accepting property and high reduction power owing to the presence of electron-releasing methoxy groups. They found that a ratio of 25:1 photosensitizer to catalyst exhibited a significant CO2 reduction efficiency with respect to the single-component systems. A quantum efficiency of ΦCO = 0.59 was obtained while irradiating at λex = 365 nm.49 A more efficient ring-shaped rhenium(I) trinuclear complex was developed by these researchers (Figure 2), in which the bidentate phosphine ligands were used to connect between Re components. This rhenium ring exhibits good properties as a redox photosensitizer in terms of enhanced visible-light absorption, lifetime of the excited state, and stability of the OERS. These properties were attributed to the effective π–π interaction among the phenyl moieties on phosphine ligands and the diimine.1

Figure 2.

Figure 2

Ring-shaped rhenium(I) trinuclear complex where the constituent complexes are connected by bidentate phosphine ligands; reprinted with permission from ref (1). Copyright [2015] [American Chemical Society].

Facilitating an effective electron transfer between the photosensitizer component and the catalyst component requires collisions among the OERS of the photosensitizer unit and the catalyst unit during the process.50,51 To address this limitation, researchers thought of a bridging ligand that could connect the photosensitizer component to the catalyst component. This idea led to the supramolecular photocatalyst systems, where different functions are performed by different component units in one molecule.52 These supramolecular photocatalysts are considered more feasible systems as the OERS species of the photosensitizer and the catalyst component do not require to collide as being bridged to each other.1

2.2. Selection of Photosensitizers

The selection of photosensitizers in supramolecule photocatalysts is a very important issue. Two modes of reaction mechanisms are possible for the photoreduction of CO2. One is the reductive quenching (RQ) and the other is an oxidative quenching (OQ) mechanism. The OQ mechanism involves the transfer of an electron from the excited state of the photosensitizer (*PS) to the catalyst unit producing an OEOS pertaining to the photosensitizer component (PS+) and an OERS on the catalyst component (Cat). Subsequently, the sacrificial electron donor (D) donates an electron to the PS+. On the other hand, in the case of the RQ mechanism, a sacrificial electron donor first reduces the excited photosensitizer *PS, producing an OERS (PS), which then transfers the electron to the catalyst unit where the reduction of CO2 finally takes place.

2.2. 9

The OQ mechanism has the disadvantage that there is the possibility of a simultaneous fast intramolecular backward electron transfer where the oxidized photosensitizer unit may readily take the electron back. Therefore, the RQ mechanism has been responsible for the enhanced activity in various high-performance supramolecular photocatalysts.53 The photosensitizers should have a larger light absorption coefficient with respect to the reductant and the catalyst component for efficient CO2 reduction. The photosensitizer component should also have a larger emission lifetime, which provides ample time for an efficient reductive quenching process. The photosensitizer in its excited state should have a good oxidation capability so as to accept an electron from the reductant, and the OERS produced after the electron transfer must have an enhanced stability.54 For efficient electron transfer, the reduction potential of PS* should be comparable to or higher than the oxidation potential of the sacrificial donor/reductant.54,55

In view of the above discussion, researchers have investigated several classes of photosensitizers to optimize the activity, efficiency, and selectivity. Here, we will mention a few of the most commonly used photosensitizers. The selection of photosensitizer complexes and the challenges thereof have been discussed in Table 1.

Table 1. List of Common Photosensitizers that Have Been Employed with Various Catalyst Components for Photocatalytic CO2 Reductionk.

2.2.

2.2.

Solvents: a = Acetonitrile.

b = DMF.

c = Dichloromethane.

d = Methanol.

e = Toluene.

f = Acetonitrile/dioxane (1:1 (v/v)).

g = Tetrahydrofuran.

h = Benzene.

i = Ep (irreversible wave).

j

P-P = Diphenylphosphinoethane.

k

The table has been reproduced with permission from ref (54). Copyright [2015] [Elsevier. B. V.].

2.2.1. Ru(II) Diimine Complexes

Photosensitizer complexes based on ruthenium are the most widely used in photocatalytic CO2 reduction reactions. These types of complexes, viz. [RuII(NN)3]n+, containing a diimine ligand like 2,2′-bipyridine (bpy), exhibit powerful visible-light absorption owing to the singlet metal-to-ligand charge transfer (1MLCT) band of absorption.56,57 On excitation of the photosensitizer, a triplet excited-state 3MLCT is produced involving the rapid and efficient intersystem crossing.58 These Ru(II) complexes are photochemically stable and have a relatively longer lifetime of 3MLCT. However, rapid ligand substitution may occur under acidic pH. Albeit various [RuII(NN)3]n+ kinds of complexes produce stable OERS when in the dark, it may undergo decomposition on photoexcitation, producing [Ru(NN)2L2]m+ types of complexes, where L may be the solvent containing Cl or CO.55,5961

2.2.2. Re(I) Carbonyl Diimine Complexes

Like Ru(II) complexes, rhenium-based photosensitizer complexes exhibit relatively enhanced visible-light absorption, longer excited-state lifetime (τem = 5.4 μs), and enhanced OERS stability owing to the effective π–π overlap between the diimine and phosphine ligands containing the phenyl moieties. Among the various complexes of this kind, fac-[Re(NN)(CO)3(PR3)]+ type of complexes has been considered as superior photosensitizers.49,6264 The OERS generated with these photosensitizers possesses a very high quantum yield (fac-[Re{4,4′-(MeO)2bpy}(CO)3{P(OEt)3}]+ exhibited ΦOERS = 1.6 when triethanolamine was used as the reductant).49 Similarly, complexes like [Re(NN)(CO)2(PAr3)2]+ have been reported as suitable photosensitizers owing to their longer-wavelength 1MLCT absorption band, unlike complexes with nonaromatic ligands.6568 This leads to enhanced emission lifetimes with enhanced oxidation capability of the 3MLCT excited states.

2.2.3. Os(II) Diimine Complexes

In view of the various limitations of Ru-based photosensitizers, like limited photostability and a narrow band of absorption, researchers are looking for alternative photosensitizer complexes. Os(II) diimine complexes, viz. [OsII(NN)3]n+, possess enhanced visible-light absorption up to longer wavelengths (∼700 nm).69,70 This property originates as Os-based photosensitizer complexes do not allow a singlet to triplet (3MLCT) absorption band around 500–700 nm due to the larger heavy-atom effect. Although these complexes exhibit higher negative reduction potentials in their excited states, the shorter emission lifetimes pose a limitation requiring strong reductants in the photocatalytic CO2 reduction processes.24,54

2.2.4. Ir(III) Complexes

Cyclometalated Ir(III) complexes were also explored as photosensitizers for CO2 reduction photocatalysts although they exhibit poor absorption in the lower wavelength area of the visible-light spectrum. One example is the fac-[Ir(ppy)3] complex, where ppy = 2-phenylpyridine,50,7176 which was exploited owing to its good photostability and the enhanced oxidation power possessed by its excited states. These properties exhibited by the [Ir(ppy)2(bpy)]+ and Ir(ppy)3 complexes were attributed to their large ligand field stabilization by ppy ligands due to the significant σ-donor ability, which leads to the high photochemical stability and a long lifetime of the excited states in these complexes.7779 For [Ir(CN)2(NN)]+, where CN = cyclometalated ligands, viz. ppy, the HOMOs comprise the d orbitals associated with the Ir atom and the π-orbitals are located on the cyclometalated ligands. The π* orbitals of the NN ligands comprise the LUMOs of these types of Ir complexes. The LUMO obtained from the π* orbitals of the CN ligands results in larger lifetimes in their excited states and hence the reduction power.54 However, the relatively weak absorption in visible light is a limitation. Kuramochi and co-workers79 used a 1-phenylisoquinoline (piq)-based Ir complex, [Ir(piq)2(dmb)]+, where dmb = 4,4′-dimethyl-2,2′-bipyridine (Figure 3 right).80 This complex exhibits a relatively larger absorption at wavelengths higher than 450 nm and also shows a longer lifetime of the excited states with respect to that of [Ru(dmb)3]2+.24

Figure 3.

Figure 3

Structures and abbreviations of the iridium(III) complexes; reprinted with permission from ref (79). Copyright [2016] [American Chemical Society].

Chang and co-workers (2013) demonstrated that an Ir-based photosensitizer could introduce a high selectivity and superior activity in nickel complexes containing N-heterocyclic carbene-amine as ligands for the conversion of CO2 to CO and reached a TON of up to 98 000.50,81

2.2.5. Chlorophylls and Metal Porphyrins

Chlorophylls82,83 and metal porphyrins8487 are the two large classes of photosensitizers that include the naturally occurring chlorophylls. Such complexes exhibit two strong visible-light absorption bands. The first is the Soret band, which involves the transition to the S2 excited state, ε(Soret band) > 105 M–1 cm–1, while the Q-band involves the transition to the S1 excited state, ε(Q band) > 104 M–1 cm–1. PdII-based porphyrin complexes are characterized by enhanced excited-state lifetimes owing to the rapid intersystem crossing and hence act as efficient photosensitizers like Ru diimine complexes.86 Although Zn2+-based porphyrins exhibit quite short emission lifetimes, the internal conversion taking place from the S2 excited state to the S1 excited state is a delayed process,88 which enhances their reduction power. Hence, Zn-porphyrins are also under wide focus as photosensitizers for photocatalytic CO2 reduction.

2.2.6. Organic Photosensitizers

Various organic compounds that contain superior visible-light absorption and also exhibit relatively long lifetimes of their excited states were also developed as photosensitizers. For example, phenazine (Phen) has a very stable triplet excited state, and triethylamine (TEA) can reductively quench it to give Phen•–.89 The as-generated Phen•– may react with a proton to produce PhenH, which can eventually provide a hydride toward the catalyst component, such as a Co-cyclam-based complex.90 Likewise, oligo(p-phenylenes) (OPPs) can yield OPP•– on getting reductively quenched by TEA.91 OPP•– exhibits a very strong reduction power (E1/2 = −2.45 V vs SCE), which is very favorable for the photoreduction of CO2.54

A copper purpurin-based complex, which exhibits a reduction potential more negative by 540 mV than its organic dye moiety, was reported by Yuan and co-workers.92 They employed a copper purpurin photosensitizer along with an iron porphyrin complex as the catalyst component. They achieved a remarkable TON of over 16 100 CO2 to CO conversion with high selectivity (95%) using 1,3-dimethyl-2-phenyl-2,3-dihydro-1H-benzo[d]imidazole as the sacrificial reductant. This is the best performance reported for noble-metal-free homogeneous systems.92

Yamashita and co-workers93 developed a photocatalyst for CO2 reduction by intercalating the [Re(bpy)(CO)3Cl] complex into the layers of Zr(HPO4)2·H2O (a-ZrP), and rhodamine B was used as the photosensitizer, as shown in Figure 4. Rhodamine B (RhB) is a stable and very fluorescent molecule and has a strong absorption in the visible range of the solar spectrum.94 Hence, it can be used as a promising photosensitizer owing to the ease of electron transfer from its LUMO to the catalyst component.

Figure 4.

Figure 4

(a) Interlayer structures of a-ZrP (Zr(HPO4)2·H2O) complexes, (b) rhodamine B molecule, and (c) [Re(bpy)(CO)3Cl] complex. Reprinted with permission from ref (93). Copyright [2015] [Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim].

The attractive performance of RhB as a metal-free photosensitizer has opened the doors for investigating various organic-molecule-based photosensitizers for photocatalytic CO2 reduction systems.93

2.3. Supramolecules for CO2 Reduction

Ishitani and co-workers demonstrated the photocatalytic CO2 reduction properties of supramolecule photocatalysts on the basis of the following traits: (1) selectivity (Γ); (2) quantum yield (Φ); (3) turnover number (TON); and (4) turnover frequency (TOF).53 In view of the various developments in terms of the above traits, different classes of supramolecular photocatalysts were designed for CO2 reduction.

The various supramolecule photocatalyst systems for CO2 reduction have been discussed next.

2.3.1. Ru(II)–Ni(II)- and Ru(II)–Co(III)-Based Systems

Ruthenium-based supramolecule photocatalysts with non-noble metals were very scarce until the 1990s. Kimura and co-workers (1992) reported the first such bimetallic system for photocatalytic CO2 reduction.102 This supramolecule photocatalyst consisted of a [Ru(phen)3]2+ component as the photosensitizer, where phen = 1,10-phenanthroline, and a [Ni(cyclam)]2+ complex as the catalyst component, where cyclam = 1,4,8,11-tetraazacyclotetradecane. In the presence of ascorbic acid as a reductant, the photocatalyst showed good selectivity (ΓCO = 72%) for CO generation with respect to H2 evolution. Tamaki and Ishitani (2017)53 summarized the photocatalytic activities exhibited by these photocatalysts, as shown in Table 2. When mixed systems were used, the activity was far less (Table 2: entry 2; ΓCO = 36%). However, S1 (Chart 1) is not considered an efficient photocatalyst as its TON for CO generation is below unity. Also, while using a pyridinium cation as a bridging unit between the photosensitizer unit and the catalyst unit (Chart 1; S2), improvement in the photocatalytic performance was again not achieved (Table 2: entry 3; TONCO < 1), and the TON = 2 was even much smaller than obtained from the mixed systems involving the [Ru(bpy)3]2+ moiety, where bpy = 2,2′-bipyridine and the [Ni(6-((N-benzylpyridin-4-yl)methyl)-1,4,8,11-tetraazacyclotetradecane)]3+ complex (Table 2: entry 4).9597 Ni-based supramolecule complexes are hence not considered good photocatalysts for CO2 reduction owing to their weak absorption in the higher wavelength visible spectrum and decreased excited-state lifetimes of the ruthenium photosensitizer component.99103 The various chemical structures and designations of the Ru(II)–Ni(II) supramolecules have been presented in the chart below.

Table 2. Summary of the Photocatalytic Performance by Ru(II)–Ni(II)- and Ru(II)–Co(III)-Based Systemsa.
entry photocatalyst reductant product formed selectivity (%) TON reference
1 S1 ascorbic acid CO 72 <1 (102)
2 [Ru(Phen)3]2+ + [Ni(cyclam)]2+ ascorbic acid CO 36 <1 (102)
3 S2 ascorbic acid CO 11 <1 (95)
4 [Ru(bpy)3]2+ + Ni moiety (as in S2) ascorbic acid CO 09 2 (95)
5 S3 TEOA CO 50 2 (105)
6 [Ru(bpy)3]2+ + [Ni(bpy)3]2+ TEOA CO 79 2 (105)
7 S4 TEOA CO 73 3 (105)
8 S5 TEOA CO 79 5 (105)
9 [Ru(bpy)3]2+ + [Co(bpy)3]3+ TEOA CO 35 9 (105)
a

Reproduced with permission from ref (53). Copyright [2017] [American Chemical Society].

Chart 1. Ru(II)–Ni(II) and Ru(II)–Co(III)-Based Supramolecule Complexes: Various Chemical Structures and Designations.

Chart 1

Komatsuzaki and co-workers (1999) also reported the synthesis of supramolecule photocatalysts, where the [Ru(bpy)2(phen)]2+ component was taken as a photosensitizer and a Co(III) or Ni(II) polypyridyl component was employed as a catalyst component.104 Such bimetallic complexes generated both CO and H2 when irradiated under 400–750 nm using TEOA as a sacrificial electron donor. RuNi3 (Table 2: entry 5) generated an equal amount of CO and H2 with the turnover number TONCO = TONH2 = 2. On the contrary, the mixed system involving [Ru(bpy)3]2+ as a photosensitizer and [Ni(bpy)3]2+ as the catalyst component selectively evolves CO (Table 2: entry 6; TONCO = 2; TONH2 < 1).53,54

Further, the bimetallic supramolecules based on Ru(II)–Co(III), viz. S4 (Table 2: entry 7; TONCO = 3; TONH2 = 1) and S5 (Table 2: entry 8; TONCO = 5; TONH2 = 1), exhibited a smaller activity with respect to a mixed system involving their constituents, viz. the [Ru(bpy)3]2+ complex and the [Co(bpy)3]3+ complex (Table 2: entry 9; TONCO = 9; TONH2 = 16).53

Wang and co-workers (2017)106 reported new dinuclear Ru–Co complexes for CO2 photoreduction (Chart 2).44 The Ru–Co complexes (Chart 2; S6–S10) are considered the most efficient non-noble-metal-based supramolecular systems for photocatalytic CO2 reduction into both CO and H2 reported so far. The complexes designated S6, S8, and S10 in Chart 2 actually exhibited good results pertaining to their TOFs and stability.106,107

Chart 2. Dinuclear Complexes of Ru and Co.

Chart 2

Apart from the above systems, Fabry and co-workers (2020) reported the Ru(II)–Mn(I) supramolecular photocatalysts for the predominant production of HCOOH by photocatalytic reduction of CO2. [MnBr(CO)3(BL)] was taken as the catalyst component, and [Ru(dmb)2(BL)]2+ was employed as the photosensitizer, where dmb = 4,4′-dimethyl-2,2′-bipyridine and BL is the bridging ligand. An equimolar ratio of the redox photosensitizer and catalyst components generated a supramolecule, which predominantly produced HCOOH, while the corresponding mixed system of individual units had a far lower activity.108

2.3.2. Ru(II)–Re(I) Systems

Ishitani and co-workers reported various Ru(II)–Re(I)-based complexes for photocatalytic CO2 reduction, which are presented in Chart 3. The photocatalytic activities achieved by such systems have been presented in Table 3.53 The foremost successful Ru–Re supramolecule photocatalyst reported by Ishitani and co-workers is S12.52 This complex comprised a [Ru(NN)3]2+-based photosensitizer component and a fac-Re(NN)(CO)3Cl complex as a catalyst component.43,44 The two units are bridged by a −CH2CH(OH)CH2– chain through two 4-methyl-bpy moieties. S12 exhibited high stability, selectivity, and efficiency (Table 3: entry 10; ΦCO = 0.12; TONCO = 170) for photocatalytic CO2 reduction when visible light was used in the presence of a sacrificial electron donor called 1-benzyl-1,4-dihydronicotinamide (BNAH). These results were much better against a 1:1 mixture of the constituent mononuclear complexes (Table 3: entry 11; ΦCO = 0.062; TONCO = 101). Contrary to this, S13 (Table 3: entry 12) and S14 (Table 3: entry 13), involving bpy or ((CF3)2bpy), i.e., 4,4-bis(trifluoromethyl)-2,2′-bipyridine on the Ru component as the peripheral ligand in place of 4dmb, exhibit smaller photocatalytic performance with respect to S12 and even smaller when compared with the mixed systems involving [Ru(4dmb)3]2+ and fac-Re(4dmb)(CO)3Cl.

Chart 3. Various Chemical Structures and Designations of Ru(II)–Re(I)-Based Supramolecule Complexes.

Chart 3

Table 3. Summary of the Photocatalytic Performance by Various Ru(II)–Re(I)-Based Supramolecules Involving One or More Photosensitizers and/or Catalyst Componentsa.
entry photocatalyst reductant product selectivity (%) TON reference
10 S12 BNAH CO 96 170 (52)
11 [Ru(4dmb)3]2+ + fac-Re(4dmb)(CO)3Cl BNAH CO/HCOOH changes from CO to HCOOH 101 (52, 79)
12 S13 BNAH CO   50 (52)
13 S14 BNAH CO   03 (52)
14 S15 BNAH CO   14 (52)
15 S16 BNAH CO   28 (52)
16 S17 BNAH CO   240 (52)
17 S18 BNAH CO 97 232 (111)
18 S19 BNAH CO 96 97 (111)
19 S20 BNAH CO 97 180 (79, 112)
20 S21 BNAH CO   120 (112)
21 S22 BNAH CO   120 (112)
22 S23 BNAH CO 94 207 (98)
23   BIH CO >99 3029 (113)
24 S24 BNAH CO 91 144 (98)
25 S25 BNAH CO 73 27 (98)
26 S26 BNAH CO 98 233 (53)
27   BIH CO >99 2915 (53)
28 S27 BNAH CO 95 253 (114)
29 S28 BNAH CO     (114)
30   BIH CO >99 >1000 (115)
31 S29 BNAH CO   123 (116)
32 S30 BNAH CO   204 (116)
33 S31 sodium ascorbate HCOOH 81 25 (117)
34 S32 BI(CO2H)H CO 81 130 (118)
35 S33 BNAH CO   315 (109)
36 S34 BNAH CO   50 (109)
37 S35 BNAH CO   110 (110)
38 S36 BNAH CO   190 (110)
39 S37 BNAH CO 69 283 (119)
40 S38 BNAH CO 75 313 (119)
a

Reproduced with permission from ref (53). Copyright [2017] [American Chemical Society].

The lower performance was attributed to the fact that the intramolecular electron transfer from the OERS of the Ru component to the Re component was ineffective due to its endergonic nature (e.g., S14 has E1/2 red (Ru) = −1.23 V and E1/2 red (Re) = −1.76 V vs Ag/AgNO3). However, S12 involved the reduction of Ru and Re components at almost equivalent potential (E1/2 red = −1.77 V) using one electron.

Various photocatalysts involving more than one photosensitizer and/or catalyst components have been introduced by Ishitani et al. (Table 3: S17),52 Rieger et al. (Table 3: S33),109 and Furue et al. (Table 3: S35, S36).110 The supramolecule photocatalysts containing more than one catalyst unit, viz. S17 (Table 3: entry 16; TONCO = 240), S33 (Table 3: entry 35; TONCO = 315), and S36 (Table 3: entry 38; TONCO = 190), presented a better photocatalytic performance with respect to S12 (Table 3: entry 10; TONCO = 170). On the contrary, S35, containing two photosensitizer components and one catalyst component, exhibited a lower performance (Table 3: entry 37; TONCO = 110) against S12.

Ishitani and co-workers investigated that the OERS of the photosensitizer, [Ru(4dmb)3]2+, undergoes photochemical degradation, while the Re(I)-based catalyst component is quite stable and retains its original chemical structure when applied for a photocatalytic reduction reaction (Scheme 1).120

Scheme 1. Solvent-Based Deactivation of the [Ru(4dmb)3]2+ Photosensitizer on Accepting the Electron from the Sacrificial Reductant.

Scheme 1

It is interesting to observe that the product obtained on decomposition, viz. [Ru-(4dmb)2(solvent)2]2+, acts as a catalyst component and undergoes a CO2 reduction reaction along with the [Ru(4dmb)3]2+ photosensitizer component to evolve HCOOH as the major product.53,79

Stanley and co-workers (2021) tried to address the stability issue by embedding these molecular systems in a metal−organic framework (MOF). They chose a benchmark catalyst component, fac-ReBr(CO)3(4,4′-dcbpy) (dcbpy = dicarboxy-2,2′-bipyridine), and photosensitizer Ru(bpy)2(5,5′-dcbpy)Cl2 (bpy = 2,2′-bipyridine) and entrapped them inside the framework of an inexpensive and nontoxic MOF, MIL-101-NH2(Al), via noncovalent host–guest interactions. The photocatalyst exhibited enhanced stability for photocatalytic CO2 reduction, and selective evolution of CO was achieved for 1.5–40 h, which was a much better performance with respect to its free molecular component systems.121

Cancelliere and co-workers (2020)122 developed a novel tris-chelating polypyridine ligand, which was employed as a bridging ligand for producing different supramolecular photocatalysts. The central bridging ligand is a phenylene ring containing three ethylene chains at its 1, 3, and 5 positions, which connect to the 2,2′-bipyridine moieties of the Ru photosensitizer or Re catalyst units (Chart 4). As such, various combinations of photosensitizer and catalyst components were taken to develop various supramolecule photocatalysts. These trinuclear systems could selectively produce CO (up to 97%) by photocatalytic reduction of CO2 with good efficiency while BIH was used as a reductant. They exhibited excellent durability with the largest TONs for CO2 reduction employing these supramolecular photocatalysts in homogeneous solutions.122

Chart 4. Various Chemical Structures and Designations of Ru(II)–Re(I)-Based Supramolecule Complexes Having a Central Phenylene Ring Bridging Ligand Containing Three Ethylene Chains at its 1, 3, and 5 Positions, which Are Connected to the 2,2′-Bipyridine Moieties of the Ru Photosensitizer or Re Catalyst Units122.

Chart 4

Ishitani and co-workers (2016) reported new Ru(II)–Re(I)-based supramolecule photocatalysts involving cyclic bridging ligands (Chart 5). It was expected that significant electronic interaction between the Ru and Re components would be acquired as the two diimine moieties are connected by two ethylene chains. However, the number of bridging ligands and the extended conjugation were found to determine the extent of electronic interaction.

Chart 5. Structure and Abbreviations of Various Bridging Ligands and a Representative Ru(II)–Re(I)-Based Supramolecule Photocatalyst Based on These Ligands.

Chart 5

Ohkubo and co-workers reported the synthesis of two new Ru(II)–Re(I)-based supramolecular photocatalysts, viz. [Ru(BL1)Re(CO)3Cl]2+ and [Ru(BL1)Re(CO)2{P(p-F-C6H4)3}2]3+, and their photocatalytic activities were compared with respect to the BL2 containing supramolecular photocatalysts, viz. [Ru(BL2)Re(CO)3Cl]2+ and [Ru(BL2) Re(CO)2{P(p-F-C6H4)3}2]3+. It was found that CO was the main product of CO2 reduction.116

A new biscarbonyl-complex-based supramolecule photocatalyst was reported by Tamaki and Ishitani et al., involving cis,trans-[Re(NN)(CO)2(PR3)2]+ (where R = p-Fph, Ph, OEt) as a catalyst component. The authors suggest that these supramolecules with tri(p-fluorophenyl)phosphine ligands, Ru–Re(FPh), exhibited high efficiency, selectivity, and durability (Chart 6).65,66,98

Chart 6. Ru(II)–Re(I)-Based Supramolecule Complexes with Various Substituents on the Ligands.

Chart 6

2.3.3. Ir(III)–Re(I)-Based Systems

The representative structure and designation of the Ir(III)–Re(I)-based complex are presented in Chart 7.

Chart 7. Representative Structure of the Ir(III)–Re(I)-Based Complex.

Chart 7

When two 1-phenylisoquinoline (piq) ligands are used as a bridge between the Ir(III)-based photosensitizer component and the Re-based catalyst component, an enhanced light absorption below 560 nm was obtained with better emission lifetime (τem = 3.0 μs). CO was the main product obtained by it on reduction of CO2 using BNAH as a sacrificial electron donor with selectivity ΓCO > 99%.79 This photocatalyst also exhibited superior stability and efficiency (ΦCO = 0.21, TONCO = 130).53

The photosensitizer can also play a dual role in some cases. It can act as a light-harvesting unit and also a catalytic site for CO2 reduction. The complexes of the kind, viz. fac-[Re(bpy)(CO)3Br]43 and [Ir(tpy)(bpy)Cl]123 (where tpy= 2,2′:6′,2″-terpyridine), were developed by the Lehn group and the Ishitani group, respectively. The supramolecules so obtained have been explored for the selective photocatalytic CO2 reduction to CO achieving TONs close to 50. Subsequently, Beller and co-workers81 used [Ir(ppy)2(bpy)]PF6, where ppy = 2-(pyridine-2-yl)benzene-1-ide, as the photosensitizer to develop a new photocatalyst for the selective photocatalytic reduction of CO2 to obtain HCOOH.124128 These supramolecule photocatalyst systems exhibit good visible-light absorption and long lifetimes of their excited states.81,129,130

2.3.4. Os(II)–Re(I) Systems

Os(II)-based complexes, viz. [Os(NN)3]2+, act as efficient photosensitizers owing to the strong visible-light absorption up to λabs < 730 nm, and the reducing power of their OERS was found to be equivalent to that of [Ru(NN)3]2+-based photosensitizers. For example, a photocatalyst of this kind was developed when the [(5dmb)2Os(4dmb)]2+ complex (where 5dmb = 5,5′-dimethyl-2,2′- bipyridine) was bridged to a cis,trans-[Re(NN)- (CO)2{P(p-X-C6H4)3}2]+ catalyst component (where X = F, Cl) through an ethylene chain. This photocatalyst was found to selectively produce CO when exploited in a photocatalytic CO2 reduction reaction, using BIH as a sacrificial reductant and irradiation at λex > 620 nm (Chart 8).69

Chart 8. Os(II)–Re(I)-Based Supramolecule Complexes with Various Substituents on Their Ligands.

Chart 8

It has been reported that the photocatalytic performance gets influenced by the introduction of phosphine ligands in the Re component, and addition of Cl on the phosphine moiety as in S48 showed a better photocatalytic performance (ΦCO = 0.12, TONCO = 1138, TOFCO = 3.3 min–1) as compared to that of S47, which contains the F substituent in the phosphine moiety (ΦCO = 0.10, TONCO = 762, TOFCO = 1.6 min–1). The relatively high photocatalytic activity was also observed with S49, where the catalyst component contains fac-[Re(NN)(CO)3(PPh3)]+(TONCO = 973).53

Ishitani and co-workers131 reported for the first time a trinuclear complex based on the Os(II)–Re(I)–Ru(II) supramolecular photocatalyst (Chart 9). These complexes show superior absorption in a wide range of visible light and act as stable photocatalysts for CO2 reduction. These supramolecule photocatalysts showed the highest TON for CO generation (TONCO = 4347) among the various Re(I) catalyst component-based supramolecular photocatalysts.131

Chart 9. Trinuclear Os(II)–Re(I)–Ru(II) Supramolecular Complexes and Constituent Fragments.

Chart 9

2.3.5. Binuclear Ru(II)-Based Systems

The various chemical structures and designations of binuclear Ru(II) supramolecule complexes are presented in Chart 10. Tamaki and Ishitani (2017) summarized their photocatalytic reduction performances as given in Table 4.53 Tanaka and co-workers originally reported that [Ru(NN)2(CO)2]2+-based supramolecule photocatalysts generate HCOOH with good selectivity under basic pH.132134

Chart 10. Ru(II)–Ru(II)-Based Photocatalyst Systems.

Chart 10

Table 4. Summary of Performance by Ru(II)–Ru(II) Systemsa.
entry photocatalyst reductant product selectivity (%) TON reference
41 S52 BNAH HCOOH 90 315 (135)
42 S53 BNAH HCOOH 91 562 (135)
43 S53 MeO-BNAH HCOOH 89 671 (135)
44 S53 BI(OH)H HCOOH 87 2766 (136)
45 S53 BIH HCOOH 72 641 (136)
46 S54 BNAH HCOOH 77 353 (135)
47 S55 BNAH HCOOH 70 234 (135)
48 S56 BNAH HCOOH 91 337 (53)
49 S57 BNAH CO 98 40 (53)
a

Reproduced with permission from ref (53). Copyright [2017] [American Chemical Society].

Various other Ru-complex-based systems, viz. [Ru(4dmb)3m(BL)m]2+, where m = 1, 2, 3 and BL = 1,2-bis(4′- methyl-[2,2′]bipyridin-4-yl)-ethane, act as photosensitizers, which bridge to cis-[Ru(4dmb)2–n(BL)n(CO)2]2+ (n = 1, 2) catalyst components through a nonconjugated bridging ligand. The as-developed photocatalysts could successfully reduce CO2 to HCOOH when irradiated in the presence of BNAH.135 The ratio of photosensitizer to catalyst components influences the photocatalytic activity. For example, S53 (Table 4: entry 42; ΓHCOOH = 91%) and S52 (Table 4: entry 41; ΓHCOOH = 90%) exhibited better selectivity in the photocatalytic reduction reaction, whereas S54 (Table 4: entry 46; ΓHCOOH = 77%) and S55 (Table 4: entry 47; ΓHCOOH = 70%) deactivated readily due to modification in the catalyst components.137,138

When a higher number of catalyst components are used in one photocatalyst, viz. S54 and S55, it facilitates the degradation of the catalyst components as the photosensitizer is unable to transfer sufficient electrons to it. However, when the ratio of photosensitizers is increased, viz. S52 and S53, it facilitates CO2 reduction and suppresses deactivation. Interestingly, the Ru(II)–Re(I)-based supramolecular photocatalysts exhibited higher durability when multiple catalyst components were used.52,109,110

Various multinuclear Ru-based complexes containing cis,trans-Ru(BL)(CO)2Cl2139,140 (S56 and S57) have been reported. Various ratios of photosensitizer to the catalyst component were developed. S56, having the ratio of 1:1, majorly evolved HCOOH (Table 4: entry 48; TONHCOOH = 337; TONCO = 22, TONH2 = 12), whereas S57 containing three catalyst components (1:3 ratio) selectively evolved CO (Table 4: entry 49; TONHCOOH < 1, TONCO = 40, TONH2 < 1). The selectivity of S57 toward CO formation and not HCOOH was attributed to the polymerization of the catalyst components during the photocatalytic reaction.141144 The generation of CO instead was attributed to the Ru oligomer involving the Ru photosensitizer component generated from S57 (Scheme 2).53

Scheme 2. Polymerization of the Catalyst Components during the Photocatalytic Reaction.

Scheme 2

2.3.6. Porphyrin–Re(I) and Ru(II)-Based Systems

There are fewer reports regarding dinuclear supramolecule complexes based on a porphyrin component and a Re(I) bipyridine tricarbonyl component.84,86,87,145147 In most of these complexes, an aliphatic amide bond connects the bpy ligand of the Re(I) component to one phenyl group in the meso-position of the porphyrin component.84,86,87 Many other groups have reported the use of a phenanthroline moiety to combine a porphyrin component so as to facilitate the interaction between two different metal centers;148150 however, their photocatalytic activity was not investigated. Tamaki and Ishtani (2017) summarized the photocatalytic performance by porphyrin–Re(I) systems as shown in Table 5.53 A Zn(II)-based porphyrin photosensitizer (S58) was developed by Inoue et al., which could photocatalytically reduce CO2 to CO using visible light for irradiation at λex = 428 nm, and triethylamine (TEA) was used as the sacrificial electron donor (Table 5: entry 50; ΦCO = 0.0064).84 Perutz and co-workers reported the synthesis of a supramolecule photocatalyst (S59) involving a Pd(II) tetraphenylporphyrin-based photosensitizer and a fac-[Re(NN)(CO)3(3-picoline)]+ derivative-based catalyst component, and the two are connected through an amide linkage.86 This photocatalyst assembly also reduced CO2 to CO when visible-light irradiation was used in the presence of TEA as the sacrificial electron donor (Table 5: entry 51; TONCO = 2). However, their photocatalytic activity was found to be lower than the mixed system containing its constituent components (Table 5: entry 52; TONCO = 3). When Zn(II) tetraphenylporphyrin derivative was used as the photosensitizer unit in place of Pd, again, a lower photocatalytic efficiency was achieved with respect to their mixed system (Table 5: S60; entry 53; TONCO = 14; S61; entry 54; TONCO = 32 and mixed system: entry 56; TONCO = 103).87 The photocatalytic efficiency could be enhanced by introducing a methylene group on the −NHCO– linkage and the Re component (Table 5: S62; entry 55; TONCO = 332) owing to their larger reducing power with respect to S60 and S61 (S62: E1/2 red = −1.68 V vs Fc/Fc+; S60: E1/2 red = −1.44 V; S61: E1/2 red = −1.42 V).53,82,151 More recently, Kuramochi and co-workers (2020) developed a dyad, ZnP-phen = Re, where a zinc-based porphyrin photosensitizer has been coupled with a fac-Re(phen)(CO)3Br catalyst component via its meso-position. This system exhibited an excellent durability, achieved a healthy TONCO of 1300, and selectivity for CO > 99.9%, which was attributed to the strong interaction between the porphyrin and the Re component where the electron reservoir on the porphyrin readily transfers to the Re component and causes hydrogenation of the porphyrin structure.152 The various structures and designations of porphyrin–Re(I)-based supramolecule photocatalysts, along with their photocatalytic performances, are presented in Chart 11 and Table 5, respectively.

Table 5. Summary of Performances by Porphyrin–Re(I)-Based Supramolecule Photocatalystsa.
entry photocatalyst reductant product ϕproduct TON reference
50 S58 TEA CO 0.0064   (84)
51 S59 TEA CO   2 (86)
52 Pd(POR) + fac-[Re(bpy)(CO)3(3-picolene)]+ TEA CO   3 (86)
53 S60 TEOA CO   14 (87)
54 S61 TEOA CO   32 (87)
55 S62 TEOA CO   332 (151)
56 Zn(POR) + fac-[Re(bpy)(CO)3(3-picolene)]+ TEOA CO   103 (87)
57 S63 TEA CO   13 (153)
58 S64 TEA CO   <1 (153)
59 S65 TEA CO   <1 (153)
60 S66 TEA CO   <1 (153)
61 S67 TEA CO   <1 (153)
62 S68 TEA CO   <1 (153)
63 S69 BIH CO   18 (82)
a

Reproduced with permission from ref (53). Copyright [2017] [American Chemical Society].

Chart 11. Various Structures and Designations of Porphyrin–Re(I)-Based Supramolecule Complexes.

Chart 11

Various mononuclear porphyrin metal complexes, viz. M-TPP, containing a redox-active metal such as iron or cobalt, have also been reported to undergo a photocatalytic reduction of CO2.154157 Fujita et al. and Neta et al. reported the syntheses of FeCl-TPP and Co-TPP. They evolved CO as the reduction product with TONCO = 80 for Co-TPP and TONCO = 65 for FeCl-TPP; both were irradiated at λ > 320 nm for 200 h.156,157

Schwalbe and co-workers153 reported the development of M-Por and M-Por-Ru compounds, where M = H2, Cu, Pd, Co, and FeCl evolved from the phenanthroline extended tetramesityl porphyrin ligand (H2-Por). Here, the Ru-based component is the [Ru(tbbpy)2]2+ complex, where tbbpy = 4,4′-di-tert-butyl-2,2′-bipyridine.158,159 Among the various developed supramolecules, FeCl-1 and FeCl-1-Ru exhibited the best photocatalytic performance, even higher than the earlier known photocatalysts like Co-TPP and FeCl-TPP. They, for the first time, developed a new type of M-1-Re complex with Re = Re(CO)3Cl, where a Re(I) tricarbonyl component is connected through a delocalized π-system linker (Chart 12).153

Chart 12. Structures of M-Por and M-Por-Re Compounds where M = H2, Cu, Pd, Co, Zn, and Fe.

Chart 12

Takeda and co-workers (2016)160 reported the design of a novel, durable, and efficient photocatalyst for CO2 reduction based on some nonprecious-metal complexes. They developed a CuI diimine complex as the photosensitizer and an FeII diimine moiety as the catalyst component (Chart 13).

Chart 13. Various Structures and Designations of CuI Photosensitizers (a) and FeII-Based Catalyst Component (b).

Chart 13

Thus, various reports on photocatalytic CO2 reduction by supramolecule complexes based on Co,161165 Ni,50,166,167 Fe,160,168,169 and Mn170172 have been reported in the literature. However, the overall efficiency and selectivity of most of these photocatalytic systems are not satisfactory.

Leung and co-workers (2012) have reported that a Co-based complex [Co(qpy)(OH2)2]2+ (where qpy = 2,2′:6′,2″:6″,2‴-quaterpyridine) can prove to be a superior photocatalyst for water oxidation and CO2 reduction.173 Guo et al. (2016) also supported such reports. Interestingly, these photocatalysts enhance their photocatalytic performance when a cheap organic dye purpurin is used as a photosensitizer (Chart 14).174

Chart 14. Cobalt-Based Porphyrin-Type Catalyst and Its Fe-Based Analogue.

Chart 14

2.3.7. Polyoxometalate-Based Supramolecules

The polyoxometalates (POMs) are combined with specific organic- or inorganic-type molecules through noncovalent interactions to develop a new class of supramolecules, which can be exploited for photocatalytic CO2 reduction reactions. These systems comprise POMs along with metal–organic frameworks, cationic metal complexes, organic molecular networks, or an assembly of various POM units. The POM-based supramolecular hybrids require compatibility in the dimensions of the constituent components so as to deliver the desired activity and stability.175

Das and co-workers (2018) developed a complex polyoxometalate Na48[HXMo368O1032(H2O)240(So4)48] ca. 1000 H2O designated as Mo368, which was exploited for the photocatalytic reduction of CO2 to HCOOH. They achieved a high TON of 27 666 with a TOF of 4611 h–1 and a quantum efficiency of 0.6%. This photocatalyst oxidizes the H2O in the medium to provide the electrons for the reduction of CO2. The superior activity was attributed to various Mo subunits in the cluster, and no external photosensitizer or reductant was used, which makes this reaction environmentally friendly.176

Huang and co-workers (2020) developed two structurally analogous porphyrin-POM coordination frameworks, where [PMo8VMo4VIO35(OH)5Zn4]2[Zn-TCPP][2H2O]·xGuest (NNU-13) and [PMo8VMo4VIO35(OH)5Zn4]2[Zn-TCPP][2H2O]·yGuest (NNU-14) are assembled with porphyrin {ϵ-PMo8VMo4VIO40Zn4}, designated as (Zn-ϵ-Keggin).177 Here, PMo12 is surrounded by four ZnII ion nodes and a photosensitive tetrakis(4-carboxylphenyl) porphyrin derivative (H2TCPP) behaving as bridging ligands. These systems showed good stability in aqueous solutions for photocatalytic CO2 reduction in the absence of an external photosensitizer or a cocatalyst. These systems showed a high selectivity for CH4 formation (96.6% for NNU-13 and 96.2% for NNU-14) on reduction of CO2. The superior photocatalytic activity was attributed to the strong reduction potential of the Zn-ϵ-Keggin moiety and enhanced charge-transfer properties of the TCPP linker.178,179

Haviv and co-workers (2017) demonstrated the photoreductive capability of an acidic polyoxometalate, H5PW2VW10O40, linked to a dirhenium molecular complex. This system exhibited a superior performance for the selective production of CO from photocatalytic reduction of CO2.180

Researchers envisioned that by combining a chiral organic moiety and a POM in a MOF, the chemical conversion of CO2 can be performed.181,182 Han and coworkers introduced Keggin-type [ZnW12O40]6– anions, NH2-bipyridine, zinc(II) ions, pyrrolidine-2-yl-imidazole, NH2-functionalized bridging links, and asymmetric organocatalytic groups within a single MOF. They revealed that the Lewis acid zinc centers could synergistically interact with CO2 to undergo its assimilation.183

Another MOF-entrapped POM, TBA5[P2Mo16VMo8VIO71(OH)9Zn8(L)4] (NNU-29), was reported by Li and co-workers (2019).184 These systems exhibited photocatalytic CO2 reduction for the production of HCOO (220 μmol g–1 h–1) and selectivity of 97.9% in an aqueous medium after 16 h.

Some other POM-based hybrid catalyst systems such as Na6[Co(H2O)2(H2tib)]2{Co[Mo6O15(HPO4)4]2}·5H2O (1), Na3[Co(H2O)3][Co2(bib)] (H2bib)2.5{HCo[Mo6O14(OH)(HPO4)4]2}·4H2O (2), and (H2bpp)3 {HNa[Mo6O12(OH)3(HPO4)3(H2PO4)]2}·6H2O (3) have been reported by Du and co-workers (2020).185 Here, bib = 1,4-bis(1-H-imidazol-4-yl) benzene, tib = 1,3,5-tris(1-imidazolyl) benzene, and bpp = 1,3-bi(4-pyridyl) propane. These POMs were integrated with the [Ru(bpy)3]Cl2 complex to develop novel photocatalytic systems, which exhibited good photocatalytic activity for CO2 reduction reactions. The first two systems (1 and 2) yielded CO (322.7 and 281.81 nmol) with selectivities of 96.3 and 96.4%, respectively. The superior performance was attributed to the fact that {P4Mo6}-based POMs could readily oxidize TEOA and Ru(bpy) acted as the catalytic component for the reduction of CO2 (Figure 5).185

Figure 5.

Figure 5

Possible regulation mechanism for photocatalytic CO2 reduction by 1 or 2/Ru(bpy). Reprinted with permission from ref (185). Copyright [2020] [Elsevier B.V.].

In view of the above discussions, POM-based supramolecular assemblies have found new dimensions for the photocatalytic assimilation of CO2. Albeit these systems are quite complex, extensive research can unravel their vast potential for reductive reactions.

2.4. Selection of the Bridging Ligand

The bridging ligand has an important role in the overall performance of the supramolecule photocatalysts.63 It has been generally found that the introduction of conjugation in the bridging ligand is detrimental to the photocatalytic performance of a supramolecule photocatalyst.186 This difference should occur for the reason that there must be a difference in their E1/2 red.53 The various supramolecule photocatalysts with different bridging ligands are presented in Chart 15. The UV–vis spectra of photocatalysts S78 and S81 are presented in Figure 6, which also differ only in terms of the bridging ligand.

Chart 15. Multinuclear Ru(II)–Re(I)-Based Supramolecule Photocatalysts with Different Bridging Ligands52a.

Chart 15

a The Chart has been reproduced with permission from ref (1). Copyright [2015] [American Chemical Society].

Figure 6.

Figure 6

UV–visible spectra of various metal complexes dissolved in acetonitrile: (A) S88 (dashed line), S89 (dot-dashed line), and S78 (solid line); (B) S90 (dot-dashed line) and S81 (solid line). Reprinted with permission from ref (52). Copyright [2005] [American Chemical Society].

As is evident from Figure 6, the complex S78 has a very identical spectrum to that of its mixed constituents (Figure 6a). However, the complex S81 has a bit different spectrum from that of the mixed system from its constituents (Figure 6b). It follows that in the case of complex S81, there is an effective interaction between the two diimine moieties in the bridging ligand due to conjugation. It was found that both complexes act as photocatalysts in the CO2 reduction reaction. However, unexpectedly, complex S81 had a much lower activity than S78 because of conjugation in the bridging ligand. Complex S78 exhibited TONCO = 170, which was much higher than that of S81. They found that complex S78 had a reduction potential of −1.10 V, which was very much lower than that of complex S78.

They revealed that the presence of conjugation in the bridging ligand actually deepens the π*-orbital energy in the diimine moiety of the Re component, which lowers its reduction potential.1

Koike and Ishitani187 for the first time employed time-resolved infrared (TRIR) spectroscopy to three Ru(II)–Re(I)-based supramolecules, viz. RuRe(X), where X = FPh, Ph, and OEt (Chart 16), in the presence and absence of a reductant BNAH (Scheme 3).

Chart 16. Binuclear Complexes Based on Ru(II)–Re(I) with Varying Substituents on the Ligands and Structures of BNAH and BIH Reductants.

Chart 16

Scheme 3. Oxidative and Reductive Quenching Pathways for the Photochemical Generation of OERS from the Ru(II)–Re(I)-Based System.

Scheme 3

They found that while using a saturated linker between them, there is a rapid intramolecular electron transfer between the excited states and OERS of the Ru component to the Re-based catalyst component (kET > 2 × 107 s–1).53 This observation discouraged the use of conjugated linkers in supramolecule photocatalysts.

Kato and co-workers (2015) also reported the development of supramolecule photocatalysts based on Ru(II)–Re(I) assemblies connected through a saturated chain, viz. Ru(CH2CH2)Re.52 They achieved a good quantum yield and turnover number for CO generation, 0.15 and 207, respectively, when BNAH was used as a reductant.98 This was the best-performing photocatalyst of the time.113 They also introduced −CH2OCH2– or −CH2SCH2– as bridging linkers. The −CH2OCH2– chain containing the supramolecule photocatalyst exhibited the highest activity when BNAH was used as the reductant.114 Cheung and co-workers (2019) added the amide groups on the Ru- and Re-based photosensitizer and catalyst components, respectively, to ensure hydrogen bonding between the two moieties.188 They revealed that hydrogen bonding enhanced the performance of the supramolecule photocatalyst for CO2 reduction with respect to the individual components. This photocatalyst exhibited almost thrice better turnover number (TONCO = 100 ± 4) and quantum yield (ΦCO = 23.3 ± 0.8%) for CO production with respect to the unsubstituted Ru and Re components (TONCO = 28 ± 4 and ΦCO = 7 ± 1%) in acetonitrile (MeCN) using BIH as the reductant. The better performance was attributed to the enhanced intramolecular charge transfer in the supramolecule system and the prevention of dimerization in the individual components.188

Wang and co-workers (2021) investigated the role of coordinative interactions between the photosensitizer and the catalyst components.189 They revealed that the coordinative interaction between the iridium-based photosensitizer and the cobalt phthalocyanine-based catalyst component potentially enhanced the electron transfer and resulted in superior quantum efficiency (10.2 ± 0.5% at 450 nm) for CO production. It exhibited almost 100% selectivity for CO, and a turnover number of 391 ± 7 was achieved, which is 4 times better than a comparative system with no coordinate interactions.189

From the above discussion, it can be concluded that there are a few technicalities in choosing a linker. (1) The electron on the excited photosensitizer component should be located on the bridging ligand so as to ensure effective electron mobility to the catalyst component. (2) A saturated or nonconjugated linker must be tried.53 (3) The peripheral ligands of the photosensitizer component should have equal or higher π* orbital energy with respect to that of the bridging ligand so as to ensure smooth electron mobility from the photosensitizer OERS to the catalyst component. (4) The bridging ligand must have sufficient π* orbital energy to reinforce CO2 reduction.54

2.5. Length of the Alkyl Chain

The photocatalytic activity of supramolecule photocatalysts for CO2 reduction is well influenced by the alkyl chain length in the bridging ligand.112 The Ru(II)–Re(I) supramolecule complex achieved the highest photocatalytic performance when an ethylene chain (−CH2CH2−) was used as the alkyl chain in the linker. On increasing the carbon number of the alkyl chain, the photocatalytic performance diminished. The weaker transfer of electrons along Ru and Re components on using the ethylene chain as the linker modulated the oxidation potential of the Ru component. This rapidly quenches the 3MLCT excited state involving the sacrificial donor BNAH. As such, the reduction potential of the Re component went much more toward the negative—Ered = −1.72 V vs Ag/AgNO3—so as to undergo a photocatalytic reduction reaction (Chart 17).1,98

Chart 17. Ru(II)–Re(I) Supramolecule Complex with Different Alkyl Chain Lengths.

Chart 17

Yamazaki and co-workers (2019) investigated the rate constants of intramolecular charge transfer between OERS of the Ru moiety and the Re catalyst unit in Ru–Re supramolecular photocatalysts containing different bridging ligands.190 They revealed that the rate of charge transfer is the function of the alkyl chain length in the bridging ligand and the number of chains. A linear relationship was obtained between the logarithm of the rate constants and the chain length of the bridging ligands. The gradual slope demonstrates a small decay coefficient factor (β) of 0.74 Å–1. However, the introduction of two ethylene chains as bridging ligands results in an enhanced electron transfer with respect to the complex containing only one ethylene chain. It was found that the electron transfer in lower chain bridging ligands was even faster than the reductive quenching process happening on the photocatalyst, thus avoiding the rate-determining step and being of great importance for a higher activity.190

2.6. Peripheral Ligand

Ishitani and co-workers (2015)54 studied the role of peripheral anionic ligands in Re(I)-based complexes for the photocatalytic reduction of CO2 using fac-Re(bpy)(CO)3X, where X = Cl, CN, and SCN.49 They found that the complex with the SCN ligand exhibited a better performance (ΦCO = 0.20 and TONCO = 30) with respect to that of the Cl-containing complex, and the CN-containing complex was quite ineffective for CO2 reduction. The zero activity of the CN-containing complex is because the OERS does not release the CN ligand. However, the OERS of the SCN-containing complex was much higher than that of the Cl-containing complex, which provided a better electron source for photocatalytic CO2 in the SCN-containing complex.1

Triethylphosphite was also introduced as a peripheral ligand (S98) in place of Cl, which enhanced the photocatalytic performance (ΦCO = 0.16, TONCO = 232). On the other hand, use of a pyridine ligand (S99) diminished the performance of the photocatalyst (TONCO = 97).111 It was revealed that (S98) got rapidly modified into a complex possessing the −OC(O)OC2H4N(C2H4OH)2 moiety as a ligand, generated by the reaction between deprotonated TEOA at the Re center and CO2.111,115,191 This assembly was actually responsible for the photocatalytic reduction of CO2. In (S99), the free pyridine in the medium facilitated the deterioration of the photocatalyst (Charts 18 and 19).53

Chart 18. Ru(II)–Re(I)-Based Supramolecule Photocatalysts with Different Peripheral Ligands.

Chart 18

Chart 19. Representative Structure of Ru(II)–Re(I)-Based Supramolecular Photocatalyst.

Chart 19

Gholamkhaas and co-workers (2005) reported the influence of −CF3, −H, and −CH3 as substituents on the peripheral ligands for the photocatalytic reduction of CO2.52 The photocatalytic performances of the different possible photocatalysts have been summarized by Ishitani and co-workers (2015), as shown in Table 6.54

Table 6. Summary of the Photocatalytic Performance by Ru(II)–Re(I)-Based Supramolecules Containing Different Substituents on the Peripheral Ligandsa.

entry Complex R X L L′ n TON ϕCO
64 Ru(CF3)–X–ReCl CF3 CH2CH(OH)CH2 CO Cl 2 3  
65 Ru(H)–X–ReCl H CH2CH(OH)CH2 CO Cl 2 50  
66 Ru(Me)–X–ReCl CH3 CH2CH(OH)CH2 CO Cl 2 170 0.12
a

Reproduced with permission from ref (54). Copyright [2015] [Elsevier. B. V.].

They found that Ru(Me)–X–ReCl (entry 66) exhibited much better TONCO and ΦCO in comparison to (Ru(CF3)–X–ReCl) and (Ru(H)–X–ReCl). This phenomenon was attributed to the better reduction potential in OERS of the former than the latter, governed by the nature of the substituent.1,54,111 Morimoto and co-workers (2013) demonstrated the role of triethanolamine (TEOA) as the peripheral ligand. They added TEOA to a rhenium(I) tricarbonyl diimine complex containing N,N-dimethylformamide (DMF) as one ligand. The DMF ligand gets replaced by TEOA, which then captures a CO2 molecule. This helps assimilate even the lower concentrations of CO2, as in air.191 Similarly, Kamogawa and co-workers (2021) reported the role of the triethanolamine ligand on the Re-based catalyst unit in a Ru-C2-Re-based supramolecule photocatalyst. They demonstrated that the triethanolamine ligand on the catalyst moiety reacts with CO2 to produce a carbonate ester complex, which then subsequently gets reduced to CO. This phenomenon was found beneficial in the sense that even lower concentrations of CO2 (10%) could get reduced efficiently.192

2.7. Selection of the Semiconductor Component

There are various visible-light-responsive semiconductor photocatalysts exploited for photocatalytic oxygen evolution from H2O.193196 However, these photocatalysts have very low activity and selectivity for photocatalytic CO2 reduction.197199 Researchers suggest if they could hybridize a supramolecule photocatalyst with a semiconductor component containing a strong oxidation power, a new and efficient photocatalyst for CO2 reduction can be achieved with water as the reductant. The semiconductor–supramolecule hybrid system can be advantageous as the photosensitizer and catalyst components of the supramolecule can be brought in proximity to each other on the semiconductor substrate. In these systems, the excited electrons in the conduction band of the semiconductor component can pump electrons on the photosensitizer unit of the supramolecular component. The photosensitizer component will eventually transfer the electrons to the catalyst component with a strong reduction power. The schematic representation of the whole process is provided in Figure 7.

Figure 7.

Figure 7

Schematic representation showing the electron transport in a semiconductor–supramolecule hybrid for CO2 reduction.

2.7.1. Metal Oxides

Various metal oxide-based semiconductors, possessing a higher conduction band potential, have been employed to integrate with the supramolecular photocatalysts through the anchoring groups. Various researchers processed these metal oxide supports by doping with metallic and nonmetallic atoms to optimize the light absorption and reduction potential characteristics. Some reports demonstrate the use of a heterostructure formed between a metal oxide and some other organic or inorganic species, which enhances the charge-transfer properties and mitigates charge recombination. Sato and co-workers used N-doped Ta2O5, a p-type semiconductor (ECB = −1.3 V, EVB = 2.7 V, EDonor-Level = 1.1 V vs NHE), as the substrate for the supramolecule involving [Ru{4,4′-(COOH)2bpy}2(CO)2]2+ (Ep red = −0.8 V vs NHE) as the photosensitizer.200 Using TEOA as the sacrificial electron donor, HCOOH was again obtained as the major product with TONHCOOH = 89 and selectivity more than 75%, along with H2 and CO. A carbon nitride (C3N4)-based semiconductor–supramolecule hybrid was also developed involving cis,trans-Ru{4,4′-(PO3H2)2bpy}(CO)2Cl2 (Ep red = −1.4 V vs Ag/AgNO3). Using TEOA as the sacrificial reductant, TONHCOOH > 1000 and apparent ΦHCOOH = 5.7% were observed using irradiation at λex = 400 nm.201203

TiO2 was used as a semiconductor support involving a dye-based photosensitizer and a Re(I)-type complex as the catalyst component so as to develop an effective photocatalyst for CO2 reduction (TONCO > 435).204

A NiO electrode was used to develop a photocathode that combines with the Zn(II) porphyrin–Re(I) component through carboxyl anchor groups to develop a CO2 reduction photocatalyst.205 Ru(II)–Re(I)-based supramolecules have also been hybridized with the NiO electrode through methylphosphonate groups (Figure 8), and TONCO = 32 was achieved.205207 Similarly, a p-type InP photocathode was combined with the Ru(II) polymer to develop a CO2 reduction photocatalyst. This photocathode hybrid photocatalyst was also combined with a TiO2 photoanode containing Pt particles as cocatalysts. HCOOH and O2 with 0.03% of the solar conversion efficiency were achieved.208 The photocatalytic efficiency was enhanced when a reduced SrTiO3 photoanode was used in place of TiO2 (0.14%).206 Amorphous silicon–germanium and Ru polymer catalyst and IrOx were brought together to achieve a better solar-energy conversion efficiency (4.6%).209

Figure 8.

Figure 8

NiO-Ru(II)–Re(I) hybrid supramolecule photocatalyst anchored by methylphosphonate groups; reprinted with permission from ref (210). Copyright [2019] [American Chemical Society].

Faustino and co-workers (2018) reported the use of hexaniobate nanoscrolls (KxH(4–x)Nb6O17) in place of TiO2 as a semiconductor substrate. The hexaniobates are bestowed with a better conduction band reduction potential (ECB0 ≈ −0.75 V vs NHE)211,212 with respect to the TiO2 substrate (ECB0 ≈ −0.3 V vs NHE),212 which enhances their reducing capability. Further, their scrolled lamellar morphology can introduce better electron transfer by facilitating a vectorial charge transfer along with the individual layers. This phenomenon is seen to suppress the ohmic losses occurring along the grain boundaries.213 This semiconductor was made hybrid with two different Re(I) complexes. The insights obtained from the results highlight its role and that of the anchor group on the photocatalytic performance of immobilized Re(I) complexes.

Wang and co-workers (2012) investigated various supramolecule photocatalysts based on porphyrins of Zn(II), Cu(II), and Co(II) as photosensitizers. They integrated the Ru(II) polypyridyl/complexes with these porphyrins and were supported by modified TiO2 nanotubes. They demonstrated the role of metal-ion coordination and peripheral ligands in the porphyrins on the efficiency of the photosensitization when they were impregnated on a TiO2 nanotube surface. It was revealed that the Zn(II) porphyrin-based supramolecule complex showed the best efficiency for CO2 reduction to methanol in an aqueous medium.

Kumagai and co-workers developed a Ru(II)–Re(I)/CuGaO2 photocathode where CuGaO2 works as a p-type semiconductor electrode and the Ru(II)–Re(I) complex acts as a photocatalyst for CO2 reduction.214,215 The synthesized Ru(II)–Re(I)/CuGaO2 photocatalysts provided an onset potential for CO2 reduction 0.4 V more positive with respect to that of the Ru(II)–Re(I)/NiO electrode. Ru(II)–Re(I)/CuGaO2 and a CoOx/TaON photoanode were used to develop a photoelectrochemical cell that could behave as a CO2 reduction photocatalyst, and water was taken as a reductant.216

Saito and co-workers (2020) demonstrated the photocatalytic reduction of CO2 to CO using a Ru–Re-based supramolecule catalyst supported on insulating Al2O3 through phosphate linkers. They found that the Al2O3 support actually brought the photosensitizer and catalyst component in proximity, hence enhancing the durability of the resultant supramolecule through covalent bridging of the molecular units. The higher density of the individual components reduces the TOF as the sacrificial reductant is hindered from approaching the catalyst.217

2.7.2. Metal Oxynitrides

Metal oxynitrides have a good visible-light absorption characteristic due to their small band gap and have a favorable conduction band potential for reduction reactions. Researchers have tried to optimize these properties by doping of metal ions or heterostructure formation with metal oxynitrides.

Sahara and Ishitani (2015) reported a Ag/TaON-based semiconductor–supramolecule hybrid where [Ru(4,4′-Me2bpy)(NN){4,4′-(H2O3PCH2)2bpy}]2+ is the photosensitizer and the cis,trans-Ru(NN)-(CO)2Cl2 moiety is the catalyst component. Methylphosphonate groups were used as anchors to bring the connection between the Ag/TaON component and the supramolecule component.218 This hybrid catalyst was found to produce HCOOH as the major product along with CO, HCHO, and H2. The supramolecule photocatalyst without Ag/TaON did not reduce the CO2 in methanol. Similarly, Ag/TaON was not able to reduce CO2 without the supramolecule component but evolves H2.1 When CaTaO2N was used instead of TaON to produce the hybrid photocatalyst, high selectivity for HCOOH production (>99%) was achieved in the dimethylacetamide–triethanolamine mixed solution.219 Hence, metal oxynitrides were found to have more selectivity for the formation of HCOOH from CO2.

2.7.3. Metal Chalcogenides

Metal chalcogenides, viz. metal sulfides and metal selenides, have also been employed for making hybrids with supramolecule-based CO2-reduction photocatalysts.220222 Various metal chalcogenides, which were used for HER reactions, such as CdS, Ni-doped ZnS, (AgIn)0.22Zn1.56S2, and (CuGa)0.8Zn0.4S2, have now been employed as semiconductor supports for complex catalysts such as trans(Cl)-[Ru{4,4′-X2-2,2′-bipyridine}(CO)2Cl2] (X = H, CH3, COOH, PO3H2, or CH2PO3H2), for visible-light-driven CO2 reduction to obtain HCOOH using acetonitrile as the solvent and TEOA as the reductant.223 Some chalcogenide quantum dots have also been used with supramolecules in aqueous mediums.220,222 For example, CdS quantum dots were combined with nickel terpyridine complex photocatalysts, which exhibited photocatalytic reduction of CO2 to produce CO in 0.1 M aqueous TEOA solution, where a TONCO of ≈20 and a good selectivity of ≈90% for CO production were obtained.220,224 ZnSe quantum dots combined with a Ni(cyclam) catalyst through a phosphonate anchoring group showed a good CO2 reduction activity to produce CO, having TONCO higher than 120.221 When these ZnSe quantum dots were surface-modified with 2-(dimethylamino)ethanethiol, the selectivity for CO was enhanced while suppressing the HER reaction. A TON of 280 was obtained with 33% selectivity for CO production after visible-light irradiation for 20h. Another hybrid catalyst was developed when CuInS2/ZnS (core/shell) quantum dots combined with meso-tetraphenylporphyrin iron(III) chloride to undergo photocatalytic reduction of CO2 to CO. A TON higher than 50 (on 40 h irradiation) and selectivity of 84% for CO were obtained in dimethyl sulfoxide (DMSO) with traces of H2O.225 When a negatively charged CuInS2/ZnS quantum dot colloidal solution was combined with oppositely charged trimethylamino-functionalized iron tetraphenylporphyrin complexes, the TON for photoreduction of CO2 to CO was enhanced to 450 (on 30 h irradiation) and a selectivity of 99% was achieved in an aqueous solution of 15 mM TEOA and 5 mM KCl.222

From the above discussion, the metal chalcogenide support can be exploited for enhancing the selectivity for CO production from photocatalytic CO2 reduction.226

2.7.4. Carbon-Based Materials

Various metal-free semiconductor materials have been used as supports for supramolecule photocatalysts, viz. graphene, carbon nitride, carbon nanotubes, carbon nanodots, etc. These materials have been found to be potential candidates due to their low band gap, high visible-light absorption, and enhanced charge transfer, and their functionalized surface can efficiently integrate with the supramolecule moiety via the anchoring groups, leading to superior charge transfer throughout the system.

Wada and co-workers (2017) developed a hybrid photocatalyst involving a binuclear Ru(II)–Re(I) complex and mesoporous graphitic carbon nitride (mpg-C3N4). This photocatalyst could selectively generate CO by CO2 reduction.227 When the Ru(II)–Re(I) moiety is immobilized on the TiO2/NS-C3N4 composite (NS = nanosheet), a better photocatalyst is achieved as the TiO2/NS-C3N4 substrate provides a Z-scheme mechanism. This was possible because the conduction band maximum of TiO2 is more negative with respect to the reduction potential attained by the excited state of the Ru unit in Ru(II)–Re(I)228 and hence readily transfers the electrons to the photosensitizer component. Hence, the Ru(II)–Re(I)/TiO2/NS-C3N4 hybrid photocatalyst showed an enhanced performance four times better than Ru(II)–Re(I)/NS-C3N4, selectively producing CO with TONCO = 73.229 Chen and co-workers (2020) investigated the fabrication of a lanthanum–nitrogen charge-transfer bridge on g-C3N4 as the active site for photocatalytic CO2 reduction. They revealed that the bridge between La–N was responsible for the higher CO production (92 μmol g1 h–1) with an attractive CO selectivity of 80.3%, which was superior to most of the g-C3N4-based photocatalysts. This performance was attributed to the formation of a charge-transfer channel between the La–N atoms through p–d orbital hybridization.230 Thus, engineering with C3N4-based materials has been explored to enhance the selectivity for CO production in photocatalytic CO2 reduction reaction.

2.7.5. Silica-Based Materials

Periodic mesoporous organosilica (PMO) has been widely used as a light-harvesting (LH) material. The silica framework with various organic groups introduced into it acts as a light-absorbing antenna. These materials have also been hybridized with metal complexes so as to function as photocatalysts.228,231 As a first example, PMOs containing biphenyl groups in their framework were used to immobilize the Re(I) complex so as to achieve a photocatalyst.231 Another photocatalyst used PMO where acridone or methylacridone moieties have been introduced in the framework, and a Ru(II)–Re(I) supramolecule complex is anchored with it through methylphosphonate groups (Figure 9).228 These acridone groups harvest light energy, which can transfer to the Ru photosensitizer component of the supramolecular photocatalyst and then reduce CO2 to selectively produce CO while using BIH as a sacrificial donor (TONCO = 635). The methylacridone hybrid system exhibited a photocatalytic capability almost tenfold larger than that of the same supramolecule system immobilized on an ordinary mesoporous silica gel (MCM-41).

Figure 9.

Figure 9

PMO having acridone groups in its framework and the Ru(II)–Re(I) supramolecule photocatalyst immobilized on it. Reprinted with permission from ref (54). Copyright [2015] [Elsevier Ltd].

Various other hybrid systems involving Re(I) tricarbonyl complexes immobilized over zeolites,232,233 silica nanoparticles,234 mesoporous silica,232,235,236 MXenes,237 and clays238 were also applied for photocatalytic CO2 reduction. Metal–organic frameworks (MOFs) and other porous coordination polymers have also been blended with Ru(II)–Re(I) supramolecule photocatalysts.239241 More recently, perovskite materials are also considered as efficient supports for supramolecule photocatalysts for CO2 reduction reactions. Sheng and co-workers (2020) demonstrated the performance of bismuth-based stable perovskite quantum dots, Cs3Bi2X9 (where X = Cl, Br, I), for photocatalytic reduction of CO2 to selectively produce CO. These materials showed a superior performance for CO2 conversion and produced 134.76 μmol g–1 CO with a good selectivity of 98.7% under solar light.242

2.8. Role of Anchoring Groups

The anchoring groups must have appropriate terminal functionality through which they can attach to the semiconductor support on one side and to the photosensitizer on the other side. Various linkers having terminal −COOH, −PO3H2, −SO3H, pyridine, −OH, −SH, etc. have been investigated. For example, mesoporous graphitic carbon nitride (mpg-C3N4) was coupled with trans(Cl)-[Ru{4,4′-X2-2,2′-bipyridine}(CO)2Cl2]-based complexes via X = COOH, PO3H2, CH2PO3H2, H, or CH3.243,244 The electron-donating groups (such as CH3, OCH3, or CH2PO3H2) in the linker were found to enhance Ered of the complex, while the electron-withdrawing groups (such as COOH or PO3H2) reduce the Ered. Thus, we can tune and optimize the reduction potential of the semiconductor–supramolecule hybrid catalyst by changing the functionality of the anchoring linkers. This way, the selectivity for a desired product can also be optimized. A phosphonate anchoring group on the [Ru(X2bpy)2(CO)2]2+ complex has been found to be more selective for formate (HCOOH) production than an anchoring group with carboxylate.245 While using methylphosphonate (CH2PO3H2) as the anchoring group for coupling Ru–Re complexes with the mpg-C3N4 semiconductor, the selectivity for HCOOH production can be enhanced in the photocatalytic CO2 reduction.226

Ishitani and co-workers (2015) investigated the role of phosphonate groups used to anchor a Ru-based complex with carbon nitride (C3N4).246 It was observed that the anchoring group having no methylene spacers connecting between the phosphonic acid groups and bpy showed a better photocatalytic performance. This hybrid system achieved TONHCOOH = 1000 with a quantum yield of 5.7% when irradiated at λ = 400 nm (Chart 20).246

Chart 20. Ru(II)–Ru(II) Supramolecule Photocatalyst with Methylphosphonate Groups as Anchoring Groups to Attach a Solid Substrate.

Chart 20

Kang and co-workers developed a hybrid system in which a [ReL(bpy)(CO)3] (where L = 3-picoline or Br) complex was immobilized on the TiO2 substrate through phosphonate groups.247 They revealed that the [ReL(CO)3] unit (L = 3-picoline, Br) is coordinately bonded to a bpy ligand through phosphonate anchor groups along with the 4- and 4′-positions, firmly bound to TiO2. This heterogeneous photocatalyst system (TONCO = 52) outperformed the homogeneous molecular photocatalysts, viz. [ReCl(bpy)(CO)3] (TONCO = 30). Thus, the selection and modification of the anchoring ligands significantly determine the activity and product selectivity in photocatalytic CO2 reduction.

2.9. Role of Reductants

Most of the photosensitizers based on transition-metal complexes require a sacrificial donor to undergo reductive quenching where an electron is taken by the OERS of the photosensitizer from the reductant/sacrificial donor. Hence, the selection of the reductant has a great influence on the photocatalytic activities of the photocatalyst systems. The reductive quenching process gives rise to OERS of the photosensitizer (PS•–) and the OEOS of the sacrificial donor (D•+). The durability of D•+ highly influences the performance of the photocatalyst as the back-electron transfer from PS•– to D•+ is detrimental to the photocatalytic CO2 reduction reaction. Hence, D•+ should readily get degraded via deprotonation so as to mitigate its oxidation capability. The final oxidation products of the sacrificial donor must not react with the PS*, PS•–, or Cat•– species so as to suppress the photocatalytic activity.

From the above discussion, various sacrificial electron donors for supramolecular photocatalysts for CO2 reduction have been employed, viz. BIH, BNAH, ascorbic acid, TEOA, isopropanol, Na2S, and Na2SO3.248 The following classes of sacrificial reductants are being extensively used for photocatalytic CO2 reduction using semiconductor–supramolecule hybrid systems.

2.9.1. NAD(P)H Model-Based Reductants

1-Benzyl-1,4-dihydronicotinamide (BNAH) is the main NAD(P)H-type compound. It has been applied in various photocatalytic reduction reactions as the sacrificial electron donor for CO2 reduction. These types of compounds exhibit a higher reducing power with respect to triethylamine (TEA; Eoxp = 0.96 V vs SCE) and triethanolamine (TEOA; Eoxp = 0.80 V).46 BNAH is reported to have a high reduction potential—E°(BNAH/BNAH·+) = 0.57 V249—which facilitates the reduction of the excited [Ru(4dmb)3]2+-type photosensitizers. Thus, BNAH can actually undergo reductive quenching of the excited photosensitizer component of supramolecule photocatalysts. A higher quenching fraction (ηq) causes the value of quantum efficiency to be higher and vice versa. A better sacrificial donor, viz. 1-(4-methoxybenzyl)-1,4-dihydronicotinamide (MeO-BNAH, E°ox = 0.50 V),249 enhances the efficacy of the reductive quenching process and eventually improves the photocatalytic performance.135

It has been investigated that the OEOS of BNAH (BNAH•+) deprotonates quite easily to give BNA under basic conditions.135 On the other hand, TEOA does not undergo quenching of the excited state but can efficiently behave as a base so as to take a proton from BNAH•+ present in the reaction medium. Two types of reactions can occur with BNA generated in the reactions; one is the formation BNA dimers, and in another process, BNA loses an electron to the medium and gets converted to BNA+. Hence, BNAH can behave as a single-electron donor or a two-electron donor. The BNA dimers pose a problem for photocatalytic reduction reactions since they are stronger electron donors (E°ox = 0.26 V vs SCE)250 with respect to BNAH. Hence, the quenching process of the dimers competes with that of BNAH. The OEOS values of the dimers are stable, and hence, the back-electron transfer preferentially occurs from the reduced photosensitizer to the OEOS of the dimer.98,135 This reductive quenching and subsequent backward electron transfer are detrimental to the efficacy of photocatalytic reactions involving BNA dimers (Scheme 4).

Scheme 4. Mechanism of Sacrificial Reduction by BNAH and BIH Reductants.

Scheme 4

2.9.2. Dihydrobenzoimidazole-Based Reductants

1,3-Dimethyl-2-phenyl-2,3-dihydro-1H-benzoimidazole (BIH) and BI(OH)H were found to be possessing a much better reducing power (BIH: E1/2 ox = 0.33 V; BI(OH)H: E1/2 ox = 0.31 V vs SCE) with respect to that of BNAH.251255 BIH undergoes oxidation quite differently from that of BNAH. It has been investigated that BIH gives two electrons on a photon excitation of the photocatalyst sequentially through electron transfer, deprotonation, and subsequent electron transfer. Since the OEOS and a deprotonated moiety of BIH (BI) possess a very high reducing power (Ep ox = −2.06 V vs Fc+/Fc),251 it easily donates an electron to the catalyst component in the photocatalytic CO2 reduction reaction. When BIH is used as a sacrificial donor in photocatalytic reactions, TEOA can also be used as a base so as to enhance the photocatalytic performance of supramolecular photocatalyst systems. BIH can also act as a base along with being a sacrificial agent. However, the protonated BIH (BIHH+) is not capable of reductive quenching and exhibits a lower activity when TEOA is not used.

BI(OH)H also acts as a better reductant than BNAH.136 BI(OH)H also donates two electrons. In a photocatalytic CO2 reduction reaction, a BI(OH)H molecule gives two electrons and two protons, which are taken by CO2 so as to obtain BI(OH)H (BI(O)+) and HCOOH. Hence, BI(OH)H has better selectivity for HCOOH production than BIH. Thus, BNAH can provide one electron and one proton, BIH supplies two electrons and one proton, while BI(OH)H offers two electrons and two protons.

3. Conclusions

In this work, a comprehensive review has been provided on supramolecule catalyst systems for photoreduction of CO2. The review has focused on the various challenges associated with CO2 reduction and the ways forward. The evolution of various photocatalyst systems for CO2 reduction has been thoroughly discussed. It can be understood from the literature that supramolecule photocatalysts have opened new ways of thinking and strategy in this field. Supramolecules have an advantage in that we can engineer with the variety of its constituent moieties so as get the desired performance in terms of activity, efficiency, or selectivity. The main motive of researchers is the efficient electron transfer from the photosensitizer unit to the catalyst unit and, also, the enhanced extra overpotential acquired by the catalyst component. The back-electron transfer from the OERS of the photosensitizer is also an issue. Researchers suggest the introduction of hybrid semiconductor–supramolecule photocatalysts. The semiconductor support can pump the electrons on the photosensitizer, facilitating multiple electron transfer to the catalyst component with a high overpotential. The semiconductor support can hence indirectly suppress the back-electron transfer on the photosensitizer. This review throws light on the selection of various components of the hybrid semiconductor–supramolecule photocatalyst. Researchers suggest that the semiconductor component must have a low band gap and visible-light active with good charge-transfer capabilities. It is also important that the conduction band potential of the semiconductor support must be higher than that of the photosensitizer unit so as to allow the electron transfer. The photosensitizer must have a strong visible-light absorption and should readily go into the excited state. The excited state must readily get the electron from the sacrificial reductant so as to undergo reductive quenching. The OERS of the PS obtained on reductive quenching should have a strong reducing power and should readily transfer the electron on the catalyst component. This is possible when the OERS has a reduction potential equal to or higher than the catalyst unit. The oxidized sacrificial donor should decompose through a process so as to suppress the back-electron transfer. This can be done by the deprotonation process through which it can lose its oxidation power. It is to be mentioned that the oxidation product of the sacrificial donor must not interfere with the progress of the photocatalytic reaction. The linker between the photosensitizer and the catalyst unit is also an important factor that determines the activity, efficiency, and selectivity of the CO2 reduction products. Generally, aliphatic linkers without conjugation were found to show better performance. Some bridging linkers were also used to achieve better selectivity for a product. The length of the linker was found to influence the performance, and it was recommended to use smaller-chain-length linkers. Similarly, the anchoring groups used to attach the supramolecule on the semiconductor substrate must have a good affinity with the substrate. They should augment the charge transfer from the semiconductor to the photosensitizer unit. Finally, the catalyst unit has the main role in terms of performance and selectivity. The distribution of active sites, the overpotential, and the durability of the catalyst unit determine the quantum efficiency and turnover frequency of a product. The various peripheral ligands attached to the catalyst component are crucial for inculcating such properties in the catalyst unit.

To dream about the practical utility of artificial photosynthesis, a lot is yet to be done in terms of a deeper understanding of the subject. The fundamental mechanisms involved in photocatalytic CO2 reduction must be clearer and fool-proof so that efficient photocatalysts can be designed.

Acknowledgments

The authors acknowledge the support by the Khalifa University of Science and Technology to A.Q. vide Award No. FSU-2020-01.

The authors declare no competing financial interest.

References

  1. Sahara G.; Ishitani O. Efficient Photocatalysts for CO2 Reduction. Inorg. Chem. 2015, 54, 5096–5104. 10.1021/ic502675a. [DOI] [PubMed] [Google Scholar]
  2. Nocera D. G. Living healthy on a dying planet. Chem. Soc. Rev. 2009, 38, 13–15. 10.1039/B820660K. [DOI] [PubMed] [Google Scholar]
  3. Li L.; Zhao N.; Wei W.; Sun Y. A review of research progress on CO2 capture, storage, and utilization in Chinese Academy of Sciences. Fuel 2013, 108, 112–130. 10.1016/j.fuel.2011.08.022. [DOI] [Google Scholar]
  4. Su T.-m.; Qin Z.-z.; Ji H.-b.; Jiang Y.-x.; Huang G. Recent advances in the photocatalytic reduction of carbon dioxide. Environ. Chem. Lett. 2016, 14, 99–112. 10.1007/s10311-015-0528-0. [DOI] [Google Scholar]
  5. Guadalupe-Medina V.; Wisselink H. W.; Luttik M. A.; de Hulster E.; Daran J.-M.; Pronk J. T.; van Maris A. J. Carbon dioxide fixation by Calvin-Cycle enzymes improves ethanol yield in yeast. Biotechnol. Biofuels 2013, 6, 125 10.1186/1754-6834-6-125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Yu J.; Dow A.; Pingali S. The energy efficiency of carbon dioxide fixation by a hydrogen-oxidizing bacterium. Int. J. Hydrogen Energy 2013, 38, 8683–8690. 10.1016/j.ijhydene.2013.04.153. [DOI] [Google Scholar]
  7. Vaiano V.; Sannino D.; Ciambelli P. Steam reduction of CO2 on Pd/TiO2 catalysts: a comparison between thermal and photocatalytic reactions. Photochem. Photobiol. Sci. 2015, 14, 550–555. 10.1039/C4PP00252K. [DOI] [PubMed] [Google Scholar]
  8. Schwartzenberg K. C.; Hamilton J. W. J.; Lucid A. K.; Weitz E.; Notestein J.; Nolan M.; Byrne J. A.; Gray K. A. Multifunctional photo/thermal catalysts for the reduction of carbon dioxide. Catal. Today 2017, 280, 65–73. 10.1016/j.cattod.2016.06.002. [DOI] [Google Scholar]
  9. Hirota K.; Tryk D. A.; Yamamoto T.; Hashimoto K.; Okawa M.; Fujishima A. Photoelectrochemical Reduction of CO2 in a High-Pressure CO2 + Methanol Medium at p-Type Semiconductor Electrodes. J. Phys. Chem. B 1998, 102, 9834–9843. 10.1021/jp9822945. [DOI] [Google Scholar]
  10. Hara K.; Tsuneto A.; Kudo A.; Sakata T. Electrochemical Reduction of CO 2 on a Cu Electrode under High Pressure: Factors that Determine the Product Selectivity. J. Electrochem. Soc. 1994, 141, 2097–2103. 10.1149/1.2055067. [DOI] [Google Scholar]
  11. Drees M.; Cokoja M.; Kühn F. E. Recycling CO2? Computational Considerations of the Activation of CO2 with Homogeneous Transition Metal Catalysts. ChemCatChem 2012, 4, 1703–1712. 10.1002/cctc.201200145. [DOI] [Google Scholar]
  12. Zhao G.; Huang X.; Wang X.; Wang X. Progress in catalyst exploration for heterogeneous CO2 reduction and utilization: a critical review. J. Mater. Chem. A 2017, 5, 21625–21649. 10.1039/C7TA07290B. [DOI] [Google Scholar]
  13. Huff C. A.; Kampf J. W.; Sanford M. S. Role of a Noninnocent Pincer Ligand in the Activation of CO2 at (PNN)Ru(H)(CO). Organometallics 2012, 31, 4643–4645. 10.1021/om300403b. [DOI] [Google Scholar]
  14. Vogt M.; Gargir M.; Iron M. A.; Diskin-Posner Y.; Ben-David Y.; Milstein D. A New Mode of Activation of CO2 by Metal–Ligand Cooperation with Reversible C–C and M–O Bond Formation at Ambient Temperature. Chem. – Eur. J. 2012, 18, 9194–9197. 10.1002/chem.201201730. [DOI] [PubMed] [Google Scholar]
  15. Ashley A. E.; Thompson A. L.; O’Hare D. Non-Metal-Mediated Homogeneous Hydrogenation of CO2 to CH3OH. Angew. Chem., Int. Ed. 2009, 48, 9839–9843. 10.1002/anie.200905466. [DOI] [PubMed] [Google Scholar]
  16. Mömming C. M.; Otten E.; Kehr G.; Fröhlich R.; Grimme S.; Stephan D. W.; Erker G. Reversible Metal-Free Carbon Dioxide Binding by Frustrated Lewis Pairs. Angew. Chem., Int. Ed. 2009, 48, 6643–6646. 10.1002/anie.200901636. [DOI] [PubMed] [Google Scholar]
  17. Li P.; Zhang J.; Wang H.; Jing H.; Xu J.; Sui X.; Hu H.; Yin H. The photoelectric catalytic reduction of CO2 to methanol on CdSeTe NSs/TiO2 NTs. Catal. Sci. Technol. 2014, 4, 1070–1077. 10.1039/C3CY00978E. [DOI] [Google Scholar]
  18. Munir A.; Joya K. S.; Ul Haq T.; Babar N. U. A.; Hussain S. Z.; Qurashi A.; Ullah N.; Hussain I. Metal nanoclusters: new paradigm in catalysis for water splitting, solar and chemical energy conversion. ChemSusChem 2019, 12, 1517–1548. 10.1002/cssc.201802069. [DOI] [PubMed] [Google Scholar]
  19. Hardwick T.; Ahmed N. C–H Functionalization via Electrophotocatalysis and Photoelectrochemistry: Complementary Synthetic Approach. ACS Sustainable Chem. Eng. 2021, 9, 4324–4340. 10.1021/acssuschemeng.0c08434. [DOI] [Google Scholar]
  20. Long Ngo H.; Kumar Mishra D.; Mishra V.; Chien Truong C. Recent advances in the synthesis of heterocycles and pharmaceuticals from the photo/electrochemical fixation of carbon dioxide. Chem. Eng. Sci. 2021, 229, 116142 10.1016/j.ces.2020.116142. [DOI] [Google Scholar]
  21. Shi J.; Jiang Y.; Jiang Z.; Wang X.; Wang X.; Zhang S.; Han P.; Yang C. Enzymatic conversion of carbon dioxide. Chem. Soc. Rev. 2015, 44, 5981–6000. 10.1039/C5CS00182J. [DOI] [PubMed] [Google Scholar]
  22. Mistry H.; Varela A. S.; Bonifacio C. S.; Zegkinoglou I.; Sinev I.; Choi Y.-W.; Kisslinger K.; Stach E. A.; Yang J. C.; Strasser P.; Cuenya B. R. Highly selective plasma-activated copper catalysts for carbon dioxide reduction to ethylene. Nat. Commun. 2016, 7, 12123 10.1038/ncomms12123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Li K.; Peng B.; Peng T. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. ACS Catal. 2016, 6, 7485–7527. 10.1021/acscatal.6b02089. [DOI] [Google Scholar]
  24. Kuramochi Y.; Ishitani O.; Ishida H. Reaction mechanisms of catalytic photochemical CO2 reduction using Re(I) and Ru(II) complexes. Coord. Chem. Rev. 2018, 373, 333–356. 10.1016/j.ccr.2017.11.023. [DOI] [Google Scholar]
  25. White J. L.; Baruch M. F.; Pander J. E.; Hu Y.; Fortmeyer I. C.; Park J. E.; Zhang T.; Liao K.; Gu J.; Yan Y.; Shaw T. W.; Abelev E.; Bocarsly A. B. Light-Driven Heterogeneous Reduction of Carbon Dioxide: Photocatalysts and Photoelectrodes. Chem. Rev. 2015, 115, 12888–12935. 10.1021/acs.chemrev.5b00370. [DOI] [PubMed] [Google Scholar]
  26. Diercks C. S.; Liu Y.; Cordova K. E.; Yaghi O. M. The role of reticular chemistry in the design of CO2 reduction catalysts. Nat. Mater. 2018, 17, 301–307. 10.1038/s41563-018-0033-5. [DOI] [PubMed] [Google Scholar]
  27. Chen Y.; Li C. W.; Kanan M. W. Aqueous CO2 Reduction at Very Low Overpotential on Oxide-Derived Au Nanoparticles. J. Am. Chem. Soc. 2012, 134, 19969–19972. 10.1021/ja309317u. [DOI] [PubMed] [Google Scholar]
  28. Kim D.; Resasco J.; Yu Y.; Asiri A. M.; Yang P. Synergistic geometric and electronic effects for electrochemical reduction of carbon dioxide using gold–copper bimetallic nanoparticles. Nat. Commun. 2014, 5, 4948 10.1038/ncomms5948. [DOI] [PubMed] [Google Scholar]
  29. Cao Z.; Kim D.; Hong D.; Yu Y.; Xu J.; Lin S.; Wen X.; Nichols E. M.; Jeong K.; Reimer J. A.; Yang P.; Chang C. J. A Molecular Surface Functionalization Approach to Tuning Nanoparticle Electrocatalysts for Carbon Dioxide Reduction. J. Am. Chem. Soc. 2016, 138, 8120–8125. 10.1021/jacs.6b02878. [DOI] [PubMed] [Google Scholar]
  30. Morris A. J.; Meyer G. J.; Fujita E. Molecular Approaches to the Photocatalytic Reduction of Carbon Dioxide for Solar Fuels. Acc. Chem. Res. 2009, 42, 1983–1994. 10.1021/ar9001679. [DOI] [PubMed] [Google Scholar]
  31. Rajeshwar K.; Thomas A.; Janáky C. Photocatalytic Activity of Inorganic Semiconductor Surfaces: Myths, Hype, and Reality. J. Phys. Chem. Lett. 2015, 6, 139–147. 10.1021/jz502586p. [DOI] [PubMed] [Google Scholar]
  32. Li K.; An X.; Park K. H.; Khraisheh M.; Tang J. A critical review of CO2 photoconversion: Catalysts and reactors. Catal. Today 2014, 224, 3–12. 10.1016/j.cattod.2013.12.006. [DOI] [Google Scholar]
  33. Schwarz H. A.; Dodson R. W. Reduction potentials of CO2- and the alcohol radicals. J. Phys. Chem. A 1989, 93, 409–414. 10.1021/j100338a079. [DOI] [Google Scholar]
  34. Wang W.-H.; Himeda Y.; Muckerman J. T.; Manbeck G. F.; Fujita E. CO2 Hydrogenation to Formate and Methanol as an Alternative to Photo- and Electrochemical CO2 Reduction. Chem. Rev. 2015, 115, 12936–12973. 10.1021/acs.chemrev.5b00197. [DOI] [PubMed] [Google Scholar]
  35. Kubacka A.; Fernández-García M.; Colón G. Advanced Nanoarchitectures for Solar Photocatalytic Applications. Chem. Rev. 2012, 112, 1555–1614. 10.1021/cr100454n. [DOI] [PubMed] [Google Scholar]
  36. Benson E. E.; Kubiak C. P.; Sathrum A. J.; Smieja J. M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. 10.1039/B804323J. [DOI] [PubMed] [Google Scholar]
  37. Navalón S.; Dhakshinamoorthy A.; Álvaro M.; Garcia H. Photocatalytic CO2 Reduction using Non-Titanium Metal Oxides and Sulfides. ChemSusChem 2013, 6, 562–577. 10.1002/cssc.201200670. [DOI] [PubMed] [Google Scholar]
  38. Liao P.; Carter E. A. New concepts and modeling strategies to design and evaluate photo-electro-catalysts based on transition metal oxides. Chem. Soc. Rev. 2013, 42, 2401–2422. 10.1039/C2CS35267B. [DOI] [PubMed] [Google Scholar]
  39. Sampson M. D.; Kubiak C. P. Manganese Electrocatalysts with Bulky Bipyridine Ligands: Utilizing Lewis Acids To Promote Carbon Dioxide Reduction at Low Overpotentials. J. Am. Chem. Soc. 2016, 138, 1386–1393. 10.1021/jacs.5b12215. [DOI] [PubMed] [Google Scholar]
  40. Kumar B.; Smieja J. M.; Kubiak C. P. Photoreduction of CO2 on p-type Silicon Using Re(bipy-But)(CO)3Cl: Photovoltages Exceeding 600 mV for the Selective Reduction of CO2 to CO. J. Phys. Chem. C 2010, 114, 14220–14223. 10.1021/jp105171b. [DOI] [Google Scholar]
  41. Kumar B.; Smieja J. M.; Sasayama A. F.; Kubiak C. P. Tunable, light-assisted co-generation of CO and H2 from CO2 and H2O by Re(bipy-tbu)(CO)3Cl and p-Si in non-aqueous medium. Chem. Commun. 2012, 48, 272–274. 10.1039/C1CC16024A. [DOI] [PubMed] [Google Scholar]
  42. Bruckmeier C.; Lehenmeier M. W.; Reithmeier R.; Rieger B.; Herranz J.; Kavakli C. Binuclear rhenium(i) complexes for the photocatalytic reduction of CO2. Dalton Trans. 2012, 41, 5026–5037. 10.1039/c2dt30273j. [DOI] [PubMed] [Google Scholar]
  43. Hawecker J.; Lehn J.-M.; Ziessel R. Efficient photochemical reduction of CO2 to CO by visible light irradiation of systems containing Re(bipy)(CO)3X or Ru(bipy)32+–Co2+ combinations as homogeneous catalysts. J. Chem. Soc., Chem. Commun. 1983, 536–538. 10.1039/C39830000536. [DOI] [Google Scholar]
  44. Hawecker J.; Lehn J.-M.; Ziessel R. Photochemical and Electrochemical Reduction of Carbon Dioxide to Carbon Monoxide Mediated by (2,2′-Bipyridine)tricarbonylchlororhenium(I) and Related Complexes as Homogeneous Catalysts. Helv. Chim. Acta 1986, 69, 1990–2012. 10.1002/hlca.19860690824. [DOI] [Google Scholar]
  45. Kutal C.; Weber M. A.; Ferraudi G.; Geiger D. A mechanistic investigation of the photoinduced reduction of carbon dioxide mediated by tricarbonylbromo(2,2′-bipyridine)rhenium(I). Organometallics 1985, 4, 2161–2166. 10.1021/om00131a016. [DOI] [Google Scholar]
  46. Kalyanasundaram K. Luminescence and redox reactions of the metal-to-ligand charge-transfer excited state of tricarbonylchloro-(polypyridyl)rhenium(I) complexes. J. Chem. Soc., Faraday Trans. 2 1986, 82, 2401–2415. 10.1039/f29868202401. [DOI] [Google Scholar]
  47. Kutal C.; Corbin A. J.; Ferraudi G. Further studies of the photoinduced reduction of carbon dioxide mediated by tricarbonylbromo(2,2′-bipyridine)rhenium(I). Organometallics 1987, 6, 553–557. 10.1021/om00146a020. [DOI] [Google Scholar]
  48. Hayashi Y.; Kita S.; Brunschwig B. S.; Fujita E. Involvement of a Binuclear Species with the Re–C(O)O–Re Moiety in CO2 Reduction Catalyzed by Tricarbonyl Rhenium(I) Complexes with Diimine Ligands: Strikingly Slow Formation of the Re–Re and Re–C(O)O–Re Species from Re(dmb)(CO)3S (dmb = 4,4′-Dimethyl-2,2′-bipyridine, S = Solvent). J. Am. Chem. Soc. 2003, 125, 11976–11987. 10.1021/ja035960a. [DOI] [PubMed] [Google Scholar]
  49. Takeda H.; Koike K.; Inoue H.; Ishitani O. Development of an Efficient Photocatalytic System for CO2 Reduction Using Rhenium(I) Complexes Based on Mechanistic Studies. J. Am. Chem. Soc. 2008, 130, 2023–2031. 10.1021/ja077752e. [DOI] [PubMed] [Google Scholar]
  50. Thoi V. S.; Kornienko N.; Margarit C. G.; Yang P.; Chang C. J. Visible-Light Photoredox Catalysis: Selective Reduction of Carbon Dioxide to Carbon Monoxide by a Nickel N-Heterocyclic Carbene–Isoquinoline Complex. J. Am. Chem. Soc. 2013, 135, 14413–14424. 10.1021/ja4074003. [DOI] [PubMed] [Google Scholar]
  51. Kuramochi Y.; Kamiya M.; Ishida H. Photocatalytic CO2 Reduction in N,N-Dimethylacetamide/Water as an Alternative Solvent System. Inorg. Chem. 2014, 53, 3326–3332. 10.1021/ic500050q. [DOI] [PubMed] [Google Scholar]
  52. Gholamkhass B.; Mametsuka H.; Koike K.; Tanabe T.; Furue M.; Ishitani O. Architecture of Supramolecular Metal Complexes for Photocatalytic CO2 Reduction: Ruthenium–Rhenium Bi- and Tetranuclear Complexes. Inorg. Chem. 2005, 44, 2326–2336. 10.1021/ic048779r. [DOI] [PubMed] [Google Scholar]
  53. Tamaki Y.; Ishitani O. Supramolecular Photocatalysts for the Reduction of CO2. ACS Catal. 2017, 7, 3394–3409. 10.1021/acscatal.7b00440. [DOI] [Google Scholar]
  54. Yamazaki Y.; Takeda H.; Ishitani O. Photocatalytic reduction of CO2 using metal complexes. J. Photochem. Photobiol., C 2015, 25, 106–137. 10.1016/j.jphotochemrev.2015.09.001. [DOI] [Google Scholar]
  55. Allen G. H.; White R. P.; Rillema D. P.; Meyer T. J. Synthetic control of excited-state properties. Tris-chelate complexes containing the ligands 2,2′-bipyrazine, 2,2′-bipyridine, and 2,2′-bipyrimidine. J. Am. Chem. Soc. 1984, 106, 2613–2620. 10.1021/ja00321a020. [DOI] [Google Scholar]
  56. Juris A.; Balzani V.; Barigelletti F.; Campagna S.; Belser P.; von Zelewsky A. Ru(II) polypyridine complexes: photophysics, photochemistry, eletrochemistry, and chemiluminescence. Coord. Chem. Rev. 1988, 84, 85–277. 10.1016/0010-8545(88)80032-8. [DOI] [Google Scholar]
  57. Ito A.; Meyer T. J. The Golden Rule. Application for fun and profit in electron transfer, energy transfer, and excited-state decay. Phys. Chem. Chem. Phys. 2012, 14, 13731–13745. 10.1039/c2cp41658a. [DOI] [PubMed] [Google Scholar]
  58. Bhasikuttan A. C.; Suzuki M.; Nakashima S.; Okada T. Ultrafast Fluorescence Detection in Tris(2,2′-bipyridine)ruthenium(II) Complex in Solution: Relaxation Dynamics Involving Higher Excited States. J. Am. Chem. Soc. 2002, 124, 8398–8405. 10.1021/ja026135h. [DOI] [PubMed] [Google Scholar]
  59. Van Houten J.; Watts R. J. Photochemistry of tris(2,2′-bipyridyl)ruthenium(II) in aqueous solutions. Inorg. Chem. 1978, 17, 3381–3385. 10.1021/ic50190a016. [DOI] [Google Scholar]
  60. Durham B.; Caspar J. V.; Nagle J. K.; Meyer T. J. Photochemistry of tris(2,2′-bipyridine)ruthenium(2+) ion. J. Am. Chem. Soc. 1982, 104, 4803–4810. 10.1021/ja00382a012. [DOI] [Google Scholar]
  61. Kalyanasundaram K. Photophysics, photochemistry and solar energy conversion with tris(bipyridyl)ruthenium(II) and its analogues. Coord. Chem. Rev. 1982, 46, 159–244. 10.1016/0010-8545(82)85003-0. [DOI] [Google Scholar]
  62. Hori H.; Johnson F. P. A.; Koike K.; Ishitani O.; Ibusuki T. Efficient photocatalytic CO2 reduction using [Re(bpy) (CO)3{P(OEt)3}]+. J. Photochem. Photobiol., A 1996, 96, 171–174. 10.1016/1010-6030(95)04298-9. [DOI] [Google Scholar]
  63. Koike K.; Hori H.; Ishizuka M.; Westwell J. R.; Takeuchi K.; Ibusuki T.; Enjouji K.; Konno H.; Sakamoto K.; Ishitani O. Key Process of the Photocatalytic Reduction of CO2 Using [Re(4,4′-X2-bipyridine)(CO)3PR3]+ (X = CH3, H, CF3; PR3 = Phosphorus Ligands): Dark Reaction of the One-Electron-Reduced Complexes with CO2. Organometallics 1997, 16, 5724–5729. 10.1021/om970608p. [DOI] [Google Scholar]
  64. Tsubaki H.; Tohyama S.; Koike K.; Saitoh H.; Ishitani O. Effect of intramolecular π–π and CH−π interactions between ligands on structure, electrochemical and spectroscopic properties of fac-[Re(bpy)(CO)3(PR3)]+ (bpy = 2,2′-bipyridine; PR3 = trialkyl or triarylphosphines). Dalton Trans. 2005, 385–395. 10.1039/B407947G. [DOI] [PubMed] [Google Scholar]
  65. Tsubaki H.; Sekine A.; Ohashi Y.; Koike K.; Takeda H.; Ishitani O. Control of Photochemical, Photophysical, Electrochemical, and Photocatalytic Properties of Rhenium(I) Complexes Using Intramolecular Weak Interactions between Ligands. J. Am. Chem. Soc. 2005, 127, 15544–15555. 10.1021/ja053814u. [DOI] [PubMed] [Google Scholar]
  66. Tsubaki H.; Sugawara A.; Takeda H.; Gholamkhass B.; Koike K.; Ishitani O. Photocatalytic reduction of CO2 using cis,trans-[Re(dmbpy)(CO)2(PR3)(PR′3)]+ (dmbpy=4,4′-dimethyl-2,2′-bipyridine). Res. Chem. Intermed. 2007, 33, 37–48. 10.1163/156856707779160771. [DOI] [Google Scholar]
  67. Morimoto T.; Nishiura C.; Tanaka M.; Rohacova J.; Nakagawa Y.; Funada Y.; Koike K.; Yamamoto Y.; Shishido S.; Kojima T.; Saeki T.; Ozeki T.; Ishitani O. Ring-Shaped Re(I) Multinuclear Complexes with Unique Photofunctional Properties. J. Am. Chem. Soc. 2013, 135, 13266–13269. 10.1021/ja406144h. [DOI] [PubMed] [Google Scholar]
  68. Asatani T.; Nakagawa Y.; Funada Y.; Sawa S.; Takeda H.; Morimoto T.; Koike K.; Ishitani O. Ring-Shaped Rhenium(I) Multinuclear Complexes: Improved Synthesis and Photoinduced Multielectron Accumulation. Inorg. Chem. 2014, 53, 7170–7180. 10.1021/ic501196q. [DOI] [PubMed] [Google Scholar]
  69. Tamaki Y.; Koike K.; Morimoto T.; Yamazaki Y.; Ishitani O. Red-Light-Driven Photocatalytic Reduction of CO2 using Os(II)–Re(I) Supramolecular Complexes. Inorg. Chem. 2013, 52, 11902–11909. 10.1021/ic4015543. [DOI] [PubMed] [Google Scholar]
  70. Kober E. M.; Caspar J. V.; Lumpkin R. S.; Meyer T. J. Application of the energy gap law to excited-state decay of osmium(II)-polypyridine complexes: calculation of relative nonradiative decay rates from emission spectral profiles. J. Phys. Chem. A 1986, 90, 3722–3734. 10.1021/j100407a046. [DOI] [Google Scholar]
  71. Hay P. J. Theoretical Studies of the Ground and Excited Electronic States in Cyclometalated Phenylpyridine Ir(III) Complexes Using Density Functional Theory. J. Phys. Chem. A 2002, 106, 1634–1641. 10.1021/jp013949w. [DOI] [Google Scholar]
  72. Kapturkiewicz A.; Angulo G. Extremely efficient electrochemiluminescence systems based on tris(2-phenylpyridine)iridium(iii). Dalton Trans. 2003, 3907–3913. 10.1039/b308964a. [DOI] [Google Scholar]
  73. Garces F. O.; King K. A.; Watts R. J. Synthesis, structure, electrochemistry, and photophysics of methyl-substituted phenylpyridine ortho-metalated iridium(III) complexes. Inorg. Chem. 1988, 27, 3464–3471. 10.1021/ic00293a008. [DOI] [Google Scholar]
  74. Colombo M. G.; Brunold T. C.; Riedener T.; Guedel H. U.; Fortsch M.; Buergi H.-B. Facial tris cyclometalated rhodium(3+) and iridium(3+) complexes: their synthesis, structure, and optical spectroscopic properties. Inorg. Chem. 1994, 33, 545–550. 10.1021/ic00081a024. [DOI] [Google Scholar]
  75. King K. A.; Spellane P. J.; Watts R. J. Excited-state properties of a triply ortho-metalated iridium(III) complex. J. Am. Chem. Soc. 1985, 107, 1431–1432. 10.1021/ja00291a064. [DOI] [Google Scholar]
  76. Gärtner F.; Cozzula D.; Losse S.; Boddien A.; Anilkumar G.; Junge H.; Schulz T.; Marquet N.; Spannenberg A.; Gladiali S.; Beller M. Synthesis, Characterisation and Application of Iridium(III) Photosensitisers for Catalytic Water Reduction. Chem. – Eur. J. 2011, 17, 6998–7006. 10.1002/chem.201100235. [DOI] [PubMed] [Google Scholar]
  77. Sajoto T.; Djurovich P. I.; Tamayo A. B.; Oxgaard J.; Goddard W. A.; Thompson M. E. Temperature Dependence of Blue Phosphorescent Cyclometalated Ir(III) Complexes. J. Am. Chem. Soc. 2009, 131, 9813–9822. 10.1021/ja903317w. [DOI] [PubMed] [Google Scholar]
  78. Yang C.-H.; Cheng Y.-M.; Chi Y.; Hsu C.-J.; Fang F.-C.; Wong K.-T.; Chou P.-T.; Chang C.-H.; Tsai M.-H.; Wu C.-C. Blue-Emitting Heteroleptic Iridium(III) Complexes Suitable for High-Efficiency Phosphorescent OLEDs. Angew. Chem., Int. Ed. 2007, 46, 2418–2421. 10.1002/anie.200604733. [DOI] [PubMed] [Google Scholar]
  79. Kuramochi Y.; Ishitani O. Iridium(III) 1-Phenylisoquinoline Complexes as a Photosensitizer for Photocatalytic CO2 Reduction: A Mixed System with a Re(I) Catalyst and a Supramolecular Photocatalyst. Inorg. Chem. 2016, 55, 5702–5709. 10.1021/acs.inorgchem.6b00777. [DOI] [PubMed] [Google Scholar]
  80. Li Z.; Cui P.; Wang C.; Kilina S.; Sun W. Nonlinear Absorbing Cationic Bipyridyl Iridium(III) Complexes Bearing Cyclometalating Ligands with Different Degrees of π-Conjugation: Synthesis, Photophysics, and Reverse Saturable Absorption. J. Phys. Chem. C 2014, 118, 28764–28775. 10.1021/jp5073457. [DOI] [Google Scholar]
  81. Rosas-Hernández A.; Junge H.; Beller M. Photochemical Reduction of Carbon Dioxide to Formic Acid using Ruthenium(II)-Based Catalysts and Visible Light. ChemCatChem 2015, 7, 3316–3321. 10.1002/cctc.201500494. [DOI] [Google Scholar]
  82. Kitagawa Y.; Takeda H.; Ohashi K.; Asatani T.; Kosumi D.; Hashimoto H.; Ishitani O.; Tamiaki H. Photochemical Reduction of CO2 with Red Light Using Synthetic Chlorophyll–Rhenium Bipyridine Dyad. Chem. Lett. 2014, 43, 1383–1385. 10.1246/cl.140439. [DOI] [Google Scholar]
  83. Kitagawa Y.; Ogasawara S.; Kosumi D.; Hashimoto H.; Tamiaki H. Photophysical Properties of Chlorophyll Derivatives Linked with Rhenium Bipyridine Complexes. Bull. Chem. Soc. Jpn. 2015, 88, 346–351. 10.1246/bcsj.20140326. [DOI] [Google Scholar]
  84. Kiyosawa K.; Shiraishi N.; Shimada T.; Masui D.; Tachibana H.; Takagi S.; Ishitani O.; Tryk D. A.; Inoue H. Electron Transfer from the Porphyrin S2 State in a Zinc Porphyrin-Rhenium Bipyridyl Dyad having Carbon Dioxide Reduction Activity. J. Phys. Chem. C 2009, 113, 11667–11673. 10.1021/jp901548y. [DOI] [Google Scholar]
  85. Kou Y.; Nakatani S.; Sunagawa G.; Tachikawa Y.; Masui D.; Shimada T.; Takagi S.; Tryk D. A.; Nabetani Y.; Tachibana H.; Inoue H. Visible light-induced reduction of carbon dioxide sensitized by a porphyrin–rhenium dyad metal complex on p-type semiconducting NiO as the reduction terminal end of an artificial photosynthetic system. J. Catal. 2014, 310, 57–66. 10.1016/j.jcat.2013.03.025. [DOI] [Google Scholar]
  86. Schneider J.; Vuong K. Q.; Calladine J. A.; Sun X.-Z.; Whitwood A. C.; George M. W.; Perutz R. N. Photochemistry and Photophysics of a Pd(II) Metalloporphyrin: Re(I) Tricarbonyl Bipyridine Molecular Dyad and its Activity Toward the Photoreduction of CO2 to CO. Inorg. Chem. 2011, 50, 11877–11889. 10.1021/ic200243y. [DOI] [PubMed] [Google Scholar]
  87. Windle C. D.; Câmpian M. V.; Duhme-Klair A.-K.; Gibson E. A.; Perutz R. N.; Schneider J. CO2 photoreduction with long-wavelength light: dyads and monomers of zinc porphyrin and rhenium bipyridine. Chem. Commun. 2012, 48, 8189–8191. 10.1039/c2cc33308b. [DOI] [PubMed] [Google Scholar]
  88. Tokumaru K. Photochemical and photophysical behaviour of porphyrins and phthalocyanines irradiated with violet or ultraviolet light. J. Porphyrins Phthalocyanines 2001, 5, 77–86. . [DOI] [Google Scholar]
  89. Davis G. A.; Gresser J. D.; Carapellucci P. A. Photoreduction of nitrogen heterocycles. I. Photoreduction phenazine: evidence for singlet-state reactivity. J. Am. Chem. Soc. 1971, 93, 2179–2182. 10.1021/ja00738a014. [DOI] [Google Scholar]
  90. Ogata T.; Yamamoto Y.; Wada Y.; Murakoshi K.; Kusaba M.; Nakashima N.; Ishida A.; Takamuku S.; Yanagida S. Phenazine-Photosensitized Reduction of CO2 Mediated by a Cobalt-Cyclam Complex through Electron and Hydrogen Transfer. J. Phys. Chem. A 1995, 99, 11916–11922. 10.1021/j100031a020. [DOI] [Google Scholar]
  91. Matsuoka S.; Kohzuki T.; Pac C.; Ishida A.; Takamuku S.; Kusaba M.; Nakashima N.; Yanagida S. Photocatalysis of oligo(p-phenylenes): photochemical reduction of carbon dioxide with triethylamine. J. Phys. Chem. A 1992, 96, 4437–4442. 10.1021/j100190a057. [DOI] [Google Scholar]
  92. Yuan H.; Cheng B.; Lei J.; Jiang L.; Han Z. Promoting photocatalytic CO2 reduction with a molecular copper purpurin chromophore. Nat. Commun. 2021, 12, 1835 10.1038/s41467-021-21923-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Wen M.; Mori K.; Kuwahara Y.; Yamashita H. Visible-Light-Responsive Carbon Dioxide Reduction System: Rhenium Complex Intercalated into a Zirconium Phosphate Layered Matrix. ChemCatChem 2015, 7, 3519–3525. 10.1002/cctc.201500480. [DOI] [Google Scholar]
  94. Wen M.; Mori K.; Kamegawa T.; Yamashita H. Amine-functionalized MIL-101(Cr) with imbedded platinum nanoparticles as a durable photocatalyst for hydrogen production from water. Chem. Commun. 2014, 50, 11645–11648. 10.1039/C4CC02994A. [DOI] [PubMed] [Google Scholar]
  95. Kimura E.; Wada S.; Shionoya M.; Okazaki Y. New Series of Multifunctionalized Nickel(II)-Cyclam (Cyclam = 1,4,8,11-Tetraazacyclotetradecane) Complexes. Application to the Photoreduction of Carbon Dioxide. Inorg. Chem. 1994, 33, 770–778. 10.1021/ic00082a025. [DOI] [Google Scholar]
  96. Suzuki K.; Kobayashi A.; Kaneko S.; Takehira K.; Yoshihara T.; Ishida H.; Shiina Y.; Oishi S.; Tobita S. Reevaluation of absolute luminescence quantum yields of standard solutions using a spectrometer with an integrating sphere and a back-thinned CCD detector. Phys. Chem. Chem. Phys. 2009, 11, 9850–9860. 10.1039/b912178a. [DOI] [PubMed] [Google Scholar]
  97. Caspar J. V.; Meyer T. J. Photochemistry of tris(2,2′-bipyridine)ruthenium(2+) ion (Ru(bpy)32+). Solvent effects. J. Am. Chem. Soc. 1983, 105, 5583–5590. 10.1021/ja00355a009. [DOI] [Google Scholar]
  98. Tamaki Y.; Watanabe K.; Koike K.; Inoue H.; Morimoto T.; Ishitani O. Development of highly efficient supramolecular CO2 reduction photocatalysts with high turnover frequency and durability. Faraday Discuss. 2012, 155, 115–127. 10.1039/C1FD00091H. [DOI] [PubMed] [Google Scholar]
  99. Kuramochi Y.; Ishitani O. An Ir(III) Complex Photosensitizer With Strong Visible Light Absorption for Photocatalytic CO2 Reduction. Front. Chem. 2019, 7, 259. 10.3389/fchem.2019.00259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Bailey D. N.; Roe D. K.; Hercules D. M. Photoreduction of phenazine in acidic methanol. J. Am. Chem. Soc. 1968, 90, 6291–6297. 10.1021/ja01025a007. [DOI] [Google Scholar]
  101. Matsuoka S.; Fujii H.; Yamada T.; Pac C.; Ishida A.; Takamuku S.; Kusaba M.; Nakashima N.; Yanagida S. Photocatalysis of oligo(p-phenylenes): photoreductive production of hydrogen and ethanol in aqueous triethylamine. J. Phys. Chem. A 1991, 95, 5802–5808. 10.1021/j100168a018. [DOI] [Google Scholar]
  102. Kimura E.; Bu X.; Shionoya M.; Wada S.; Maruyama S. A new nickel(II) cyclam (cyclam = 1,4,8,11-tetraazacyclotetradecane) complex covalently attached to tris(1,10-phenanthroline)ruthenium(2+). A new candidate for the catalytic photoreduction of carbon dioxide. Inorg. Chem. 1992, 31, 4542–4546. 10.1021/ic00048a020. [DOI] [Google Scholar]
  103. Fujita E.; Milder S. J.; Brunschwig B. S. Photophysical properties of covalently attached tris(bipyridine)ruthenium(2+) and Mcyclam2+ (M = nickel, H2) complexes. Inorg. Chem. 1992, 31, 2079–2085. 10.1021/ic00037a019. [DOI] [Google Scholar]
  104. Komatsuzaki N.; Himeda Y.; Hirose T.; Sugihara H.; Kasuga K. Synthesis and Photochemical Properties of Ruthenium–Cobalt and Ruthenium–Nickel Dinuclear Complexes. Bull. Chem. Soc. Jpn. 1999, 72, 725–731. 10.1246/bcsj.72.725. [DOI] [Google Scholar]
  105. Komatsuzaki N.; Himeda Y.; Hirose T.; Sugihara H.; Kasuga K. Synthesis and Photochemical Properties of Ruthenium–Cobalt and Ruthenium–Nickel Dinuclear Complexes. Bull. Chem. Soc. Jpn. 1999, 72, 725–731. 10.1246/bcsj.72.725. [DOI] [Google Scholar]
  106. Wang X.; Goudy V.; Genesio G.; Maynadié J.; Meyer D.; Fontecave M. Ruthenium–cobalt dinuclear complexes as photocatalysts for CO2 reduction. Chem. Commun. 2017, 53, 5040–5043. 10.1039/C6CC09941F. [DOI] [PubMed] [Google Scholar]
  107. Steck E. A.; Day A. R. Reactions of Phenanthraquinone and Retenequinone with Aldehydes and Ammonium Acetate in Acetic Acid Solution1. J. Am. Chem. Soc. 1943, 65, 452–456. 10.1021/ja01243a043. [DOI] [Google Scholar]
  108. Fabry D. C.; Koizumi H.; Ghosh D.; Yamazaki Y.; Takeda H.; Tamaki Y.; Ishitani O. A Ru(II)–Mn(I) Supramolecular Photocatalyst for CO2 Reduction. Organometallics 2020, 39, 1511–1518. 10.1021/acs.organomet.9b00755. [DOI] [Google Scholar]
  109. Meister S.; Reithmeier R. O.; Ogrodnik A.; Rieger B. Bridging Efficiency within Multinuclear Homogeneous Catalysts in the Photocatalytic Reduction of Carbon Dioxide. ChemCatChem 2015, 7, 3562–3569. 10.1002/cctc.201500674. [DOI] [Google Scholar]
  110. Bian Z.-Y.; Sumi K.; Furue M.; Sato S.; Koike K.; Ishitani O. Synthesis and properties of a novel tripodal bipyridyl ligand tb-carbinol and its Ru(II)–Re(I) trimetallic complexes: investigation of multimetallic artificial systems for photocatalytic CO2 reduction. Dalton Trans. 2009, 983–993. 10.1039/B814340D. [DOI] [PubMed] [Google Scholar]
  111. Sato S.; Koike K.; Inoue H.; Ishitani O. Highly efficient supramolecular photocatalysts for CO2 reduction using visible light. Photochem. Photobiol. Sci. 2007, 6, 454–461. 10.1039/B613419J. [DOI] [PubMed] [Google Scholar]
  112. Koike K.; Naito S.; Sato S.; Tamaki Y.; Ishitani O. Architecture of supramolecular metal complexes for photocatalytic CO2 reduction: III: Effects of length of alkyl chain connecting photosensitizer to catalyst. J. Photochem. Photobiol., A 2009, 207, 109–114. 10.1016/j.jphotochem.2008.12.014. [DOI] [Google Scholar]
  113. Tamaki Y.; Koike K.; Morimoto T.; Ishitani O. Substantial improvement in the efficiency and durability of a photocatalyst for carbon dioxide reduction using a benzoimidazole derivative as an electron donor. J. Catal. 2013, 304, 22–28. 10.1016/j.jcat.2013.04.002. [DOI] [Google Scholar]
  114. Kato E.; Takeda H.; Koike K.; Ohkubo K.; Ishitani O. Ru(ii)–Re(i) binuclear photocatalysts connected by −CH2XCH2– (X = O, S, CH2) for CO2 reduction. Chem. Sci. 2015, 6, 3003–3012. 10.1039/C4SC03710C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Nakajima T.; Tamaki Y.; Ueno K.; Kato E.; Nishikawa T.; Ohkubo K.; Yamazaki Y.; Morimoto T.; Ishitani O. Photocatalytic Reduction of Low Concentration of CO2. J. Am. Chem. Soc. 2016, 138, 13818–13821. 10.1021/jacs.6b08824. [DOI] [PubMed] [Google Scholar]
  116. Ohkubo K.; Yamazaki Y.; Nakashima T.; Tamaki Y.; Koike K.; Ishitani O. Photocatalyses of Ru(II)–Re(I) binuclear complexes connected through two ethylene chains for CO2 reduction. J. Catal. 2016, 343, 278–289. 10.1016/j.jcat.2015.12.025. [DOI] [Google Scholar]
  117. Nakada A.; Koike K.; Nakashima T.; Morimoto T.; Ishitani O. Photocatalytic CO2 Reduction to Formic Acid Using a Ru(II)–Re(I) Supramolecular Complex in an Aqueous Solution. Inorg. Chem. 2015, 54, 1800–1807. 10.1021/ic502707t. [DOI] [PubMed] [Google Scholar]
  118. Nakada A.; Koike K.; Maeda K.; Ishitani O. Highly efficient visible-light-driven CO2 reduction to CO using a Ru(ii)–Re(i) supramolecular photocatalyst in an aqueous solution. Green Chem. 2016, 18, 139–143. 10.1039/C5GC01720C. [DOI] [Google Scholar]
  119. Yamazaki Y.; Umemoto A.; Ishitani O. Photochemical Hydrogenation of π-Conjugated Bridging Ligands in Photofunctional Multinuclear Complexes. Inorg. Chem. 2016, 55, 11110–11124. 10.1021/acs.inorgchem.6b01736. [DOI] [PubMed] [Google Scholar]
  120. Tamaki Y.; Imori D.; Morimoto T.; Koike K.; Ishitani O. High catalytic abilities of binuclear rhenium(i) complexes in the photochemical reduction of CO2 with a ruthenium(ii) photosensitiser. Dalton Trans. 2016, 45, 14668–14677. 10.1039/C6DT00996D. [DOI] [PubMed] [Google Scholar]
  121. Stanley P. M.; Thomas C.; Thyrhaug E.; Urstoeger A.; Schuster M.; Hauer J.; Rieger B.; Warnan J.; Fischer R. A. Entrapped Molecular Photocatalyst and Photosensitizer in Metal–Organic Framework Nanoreactors for Enhanced Solar CO2 Reduction. ACS Catal. 2021, 11, 871–882. 10.1021/acscatal.0c04673. [DOI] [Google Scholar]
  122. Cancelliere A. M.; Puntoriero F.; Serroni S.; Campagna S.; Tamaki Y.; Saito D.; Ishitani O. Efficient trinuclear Ru(ii)–Re(i) supramolecular photocatalysts for CO2 reduction based on a new tris-chelating bridging ligand built around a central aromatic ring. Chem. Sci. 2020, 11, 1556–1563. 10.1039/C9SC04532E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Sato S.; Morikawa T.; Kajino T.; Ishitani O. A Highly Efficient Mononuclear Iridium Complex Photocatalyst for CO2 Reduction under Visible Light. Angew. Chem., Int. Ed. 2013, 52, 988–992. 10.1002/anie.201206137. [DOI] [PubMed] [Google Scholar]
  124. Yuan Y.-J.; Yu Z.-T.; Chen X.-Y.; Zhang J.-Y.; Zou Z.-G. Visible-Light-Driven H2 Generation from Water and CO2 Conversion by Using a Zwitterionic Cyclometalated Iridium(III) Complex. Chem. – Eur. J. 2011, 17, 12891–12895. 10.1002/chem.201102147. [DOI] [PubMed] [Google Scholar]
  125. Gärtner F.; Sundararaju B.; Surkus A.-E.; Boddien A.; Loges B.; Junge H.; Dixneuf P. H.; Beller M. Light-Driven Hydrogen Generation: Efficient Iron-Based Water Reduction Catalysts. Angew. Chem., Int. Ed. 2009, 48, 9962–9965. 10.1002/anie.200905115. [DOI] [PubMed] [Google Scholar]
  126. Curtin P. N.; Tinker L. L.; Burgess C. M.; Cline E. D.; Bernhard S. Structure–Activity Correlations Among Iridium(III) Photosensitizers in a Robust Water-Reducing System. Inorg. Chem. 2009, 48, 10498–10506. 10.1021/ic9007763. [DOI] [PubMed] [Google Scholar]
  127. Tinker L. L.; McDaniel N. D.; Curtin P. N.; Smith C. K.; Ireland M. J.; Bernhard S. Visible Light Induced Catalytic Water Reduction without an Electron Relay. Chem. – Eur. J. 2007, 13, 8726–8732. 10.1002/chem.200700480. [DOI] [PubMed] [Google Scholar]
  128. Goldsmith J. I.; Hudson W. R.; Lowry M. S.; Anderson T. H.; Bernhard S. Discovery and High-Throughput Screening of Heteroleptic Iridium Complexes for Photoinduced Hydrogen Production. J. Am. Chem. Soc. 2005, 127, 7502–7510. 10.1021/ja0427101. [DOI] [PubMed] [Google Scholar]
  129. Dovletoglou A.; Adeyemi S. A.; Meyer T. J. Coordination and Redox Chemistry of Substituted-Polypyridyl Complexes of Ruthenium. Inorg. Chem. 1996, 35, 4120–4127. 10.1021/ic9512587. [DOI] [PubMed] [Google Scholar]
  130. Colombo M. G.; Hauser A.; Guedel H. U. Evidence for strong mixing between the LC and MLCT excited states in bis(2-phenylpyridinato-C2,N′)(2,2′-bipyridine)iridium(III). Inorg. Chem. 1993, 32, 3088–3092. 10.1021/ic00066a020. [DOI] [Google Scholar]
  131. Yamazaki Y.; Ishitani O. Synthesis of Os(ii)–Re(i)–Ru(ii) hetero-trinuclear complexes and their photophysical properties and photocatalytic abilities. Chem. Sci. 2018, 9, 1031–1041. 10.1039/C7SC04162D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Ishida H.; Tanaka K.; Tanaka T. THE ELECTROCHEMICAL REDUCTION OF CO2 CATALYZED BY RUTHENIUM CARBONYL COMPLEXES. Chem. Lett. 1985, 14, 405–406. 10.1246/cl.1985.405. [DOI] [Google Scholar]
  133. Ishida H.; Tanaka K.; Tanaka T. Electrochemical CO2 reduction catalyzed by ruthenium complexes [Ru(bpy)2(CO)2]2+ and [Ru(bpy)2(CO)Cl]+. Effect of pH on the formation of CO and HCOO. Organometallics 1987, 6, 181–186. 10.1021/om00144a033. [DOI] [Google Scholar]
  134. Ishida H.; Fujiki K.; Ohba T.; Ohkubo K.; Tanaka K.; Terada T.; Tanaka T. Ligand effects of ruthenium 2,2′-bipyridine and 1,10-phenanthroline complexes on the electrochemical reduction of CO2. J. Chem. Soc., Dalton Trans. 1990, 2155–2160. 10.1039/DT9900002155. [DOI] [Google Scholar]
  135. Tamaki Y.; Morimoto T.; Koike K.; Ishitani O. Photocatalytic CO<sub>2</sub> reduction with high turnover frequency and selectivity of formic acid formation using Ru(II) multinuclear complexes. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 15673. 10.1073/pnas.1118336109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Tamaki Y.; Koike K.; Ishitani O. Highly efficient, selective, and durable photocatalytic system for CO2 reduction to formic acid. Chem. Sci. 2015, 6, 7213–7221. 10.1039/C5SC02018B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Chardon-Noblat S.; Collomb-Dunand-Sauthier M. N.; Deronzier A.; Ziessel R.; Zsoldos D. Formation of Polymeric [{Ru0(bpy)(CO)2}n] Films by Electrochemical Reduction of [Ru(bpy)2(CO)2](PF6)2: Its Implication in CO2 Electrocatalytic Reduction. Inorg. Chem. 1994, 33, 4410–4412. 10.1021/ic00097a034. [DOI] [Google Scholar]
  138. Nagao H.; Mizukawa T.; Tanaka K. Carbon-Carbon Bond Formation in the Electrochemical Reduction of Carbon Dioxide Catalyzed by a Ruthenium Complex. Inorg. Chem. 1994, 33, 3415–3420. 10.1021/ic00093a033. [DOI] [Google Scholar]
  139. Lehn J.-M.; Ziessel R. Photochemical reduction of carbon dioxide to formate catalyzed by 2,2t́-bipyridine- or 1,10-phenanthroline-ruthenium(II) complexes. J. Organomet. Chem. 1990, 382, 157–173. 10.1016/0022-328X(90)85224-M. [DOI] [Google Scholar]
  140. Kuramochi Y.; Itabashi J.; Fukaya K.; Enomoto A.; Yoshida M.; Ishida H. Unexpected effect of catalyst concentration on photochemical CO2 reduction by trans(Cl)–Ru(bpy)(CO)2Cl2: new mechanistic insight into the CO/HCOO– selectivity. Chem. Sci. 2015, 6, 3063–3074. 10.1039/C5SC00199D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Collomb-Dunand-Sauthier M.-N.; Deronzier A.; Ziessel R. Electrocatalytic Reduction of Carbon Dioxide with Mono(bipyridine)carbonylruthenium Complexes in Solution or as Polymeric Thin Films. Inorg. Chem. 1994, 33, 2961–2967. 10.1021/ic00091a040. [DOI] [Google Scholar]
  142. Chardon-Noblat S.; Deronzier A.; Ziessel R.; Zsoldos D. Selective Synthesis and Electrochemical Behavior of trans(Cl)- and cis(Cl)-[Ru(bpy)(CO)2Cl2] Complexes (bpy = 2,2′-Bipyridine). Comparative Studies of Their Electrocatalytic Activity toward the Reduction of Carbon Dioxide. Inorg. Chem. 1997, 36, 5384–5389. 10.1021/ic9701975. [DOI] [Google Scholar]
  143. Chardon-Noblat S.; Deronzier A.; Ziessel R.; Zsoldos D. Electroreduction of CO2 catalyzed by polymeric [Ru(bpy)(CO)2]n films in aqueous media: Parameters influencing the reaction selectivity. J. Electroanal. Chem. 1998, 444, 253–260. 10.1016/S0022-0728(97)00584-6. [DOI] [Google Scholar]
  144. Caix-Cecillon C.; Chardon-Noblat S.; Deronzier A.; Haukka M.; Pakkanen T. A.; Ziessel R.; Zsoldos D. Electrochemical formation and spectroelectrochemical characterization of organometallic [Ru(L)(CO)2]n polymers; L=disubstituted-2,2′-bipyridine. J. Electroanal. Chem. 1999, 466, 187–196. 10.1016/S0022-0728(99)00143-6. [DOI] [Google Scholar]
  145. Gabrielsson A.; Hartl F.; Zhang H.; Lindsay Smith J. R.; Towrie M.; Vlček A.; Perutz R. N. Ultrafast Charge Separation in a Photoreactive Rhenium-Appended Porphyrin Assembly Monitored by Picosecond Transient Infrared Spectroscopy. J. Am. Chem. Soc. 2006, 128, 4253–4266. 10.1021/ja0539802. [DOI] [PubMed] [Google Scholar]
  146. Gabrielsson A.; Lindsay Smith J. R.; Perutz R. N. Remote site photosubstitution in metalloporphyrin–rhenium tricarbonylbipyridine assemblies: photo-reactions of molecules with very short lived excited states. Dalton Trans. 2008, 4259–4269. 10.1039/b806267f. [DOI] [PubMed] [Google Scholar]
  147. Casanova M.; Zangrando E.; Iengo E.; Alessio E.; Indelli M. T.; Scandola F.; Orlandi M. Structural and Photophysical Characterization of Multichromophoric Pyridylporphyrin-Rhenium(I) Conjugates. Inorg. Chem. 2008, 47, 10407–10418. 10.1021/ic800971e. [DOI] [PubMed] [Google Scholar]
  148. Crossley M. J.; Burn P. L.; Langford S. J.; Prashar J. K. Porphyrins with appended phenanthroline units: a means by which porphyrin π-systems can be connected to an external redox centre. J. Chem. Soc., Chem. Commun. 1995, 1921–1923. 10.1039/C39950001921. [DOI] [Google Scholar]
  149. Vannelli T. A.; Karpishin T. B. Neocuproine-Extended Porphyrin Coordination Complexes. 2. Spectroscopic Properties of the Metalloporphyrin Derivatives and Investigations into the HOMO Ordering. Inorg. Chem. 2000, 39, 340–347. 10.1021/ic990914q. [DOI] [PubMed] [Google Scholar]
  150. Vannelli T. A.; Karpishin T. B. Neocuproine-Extended Porphyrin Coordination Complexes. Inorg. Chem. 1999, 38, 2246–2247. 10.1021/ic9812914. [DOI] [PubMed] [Google Scholar]
  151. Windle C. D.; George M. W.; Perutz R. N.; Summers P. A.; Sun X. Z.; Whitwood A. C. Comparison of rhenium–porphyrin dyads for CO2 photoreduction: photocatalytic studies and charge separation dynamics studied by time-resolved IR spectroscopy. Chem. Sci. 2015, 6, 6847–6864. 10.1039/C5SC02099A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Kuramochi Y.; Fujisawa Y.; Satake A. Photocatalytic CO2 Reduction Mediated by Electron Transfer via the Excited Triplet State of Zn(II) Porphyrin. J. Am. Chem. Soc. 2020, 142, 705–709. 10.1021/jacs.9b12712. [DOI] [PubMed] [Google Scholar]
  153. Matlachowski C.; Braun B.; Tschierlei S.; Schwalbe M. Photochemical CO2 Reduction Catalyzed by Phenanthroline Extended Tetramesityl Porphyrin Complexes Linked with a Rhenium(I) Tricarbonyl Unit. Inorg. Chem. 2015, 54, 10351–10360. 10.1021/acs.inorgchem.5b01717. [DOI] [PubMed] [Google Scholar]
  154. Bhugun I.; Lexa D.; Savéant J.-M. Homogeneous Catalysis of Electrochemical Hydrogen Evolution by Iron(0) Porphyrins. J. Am. Chem. Soc. 1996, 118, 3982–3983. 10.1021/ja954326x. [DOI] [Google Scholar]
  155. Goodwin J. A.; Kurtikyan T. S. Electrocatalytic reactions of dioxygen and nitric oxide with reduced (nitrosyl) cobalt porphyrins — cyclic voltammetry and computational chemistry. J. Porphyrins Phthalocyanines 2011, 15, 99–105. 10.1142/S1088424611003033. [DOI] [Google Scholar]
  156. Behar D.; Dhanasekaran T.; Neta P.; Hosten C. M.; Ejeh D.; Hambright P.; Fujita E. Cobalt Porphyrin Catalyzed Reduction of CO2. Radiation Chemical, Photochemical, and Electrochemical Studies. J. Phys. Chem. A 1998, 102, 2870–2877. 10.1021/jp9807017. [DOI] [Google Scholar]
  157. Grodkowski J.; Behar D.; Neta P.; Hambright P. Iron Porphyrin-Catalyzed Reduction of CO2. Photochemical and Radiation Chemical Studies. J. Phys. Chem. A 1997, 101, 248–254. 10.1021/jp9628139. [DOI] [Google Scholar]
  158. Matlachowski C.; Schwalbe M. Synthesis and characterization of mono- and dinuclear phenanthroline-extended tetramesitylporphyrin complexes as well as UV-Vis and EPR studies on their one-electron reduced species. Dalton Trans. 2013, 42, 3490–3503. 10.1039/C2DT32196C. [DOI] [PubMed] [Google Scholar]
  159. Matlachowski C.; Schwalbe M. Photochemical CO2-reduction catalyzed by mono- and dinuclear phenanthroline-extended tetramesityl porphyrin complexes. Dalton Trans. 2015, 44, 6480–6489. 10.1039/C4DT03846K. [DOI] [PubMed] [Google Scholar]
  160. Takeda H.; Ohashi K.; Sekine A.; Ishitani O. Photocatalytic CO2 Reduction Using Cu(I) Photosensitizers with a Fe(II) Catalyst. J. Am. Chem. Soc. 2016, 138, 4354–4357. 10.1021/jacs.6b01970. [DOI] [PubMed] [Google Scholar]
  161. Lehn J.-M.; Ziessel R. Photochemical generation of carbon monoxide and hydrogen by reduction of carbon dioxide and water under visible light irradiation. Proc. Natl. Acad. Sci. U.S.A. 1982, 79, 701. 10.1073/pnas.79.2.701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Matsuoka S.; Yamamoto K.; Ogata T.; Kusaba M.; Nakashima N.; Fujita E.; Yanagida S. Efficient and selective electron mediation of cobalt complexes with cyclam and related macrocycles in the p-terphenyl-catalyzed photoreduction of carbon dioxide. J. Am. Chem. Soc. 1993, 115, 601–609. 10.1021/ja00055a032. [DOI] [Google Scholar]
  163. Varma S.; Castillo C. E.; Stoll T.; Fortage J.; Blackman A. G.; Molton F.; Deronzier A.; Collomb M.-N. Efficient photocatalytic hydrogen production in water using a cobalt(iii) tetraaza-macrocyclic catalyst: electrochemical generation of the low-valent Co(i) species and its reactivity toward proton reduction. Phys. Chem. Chem. Phys. 2013, 15, 17544–17552. 10.1039/c3cp52641k. [DOI] [PubMed] [Google Scholar]
  164. Chan S. L.-F.; Lam T. L.; Yang C.; Yan S.-C.; Cheng N. M. A robust and efficient cobalt molecular catalyst for CO2 reduction. Chem. Commun. 2015, 51, 7799–7801. 10.1039/C5CC00566C. [DOI] [PubMed] [Google Scholar]
  165. Chen L.; Guo Z.; Wei X.-G.; Gallenkamp C.; Bonin J.; Anxolabéhère-Mallart E.; Lau K.-C.; Lau T.-C.; Robert M. Molecular Catalysis of the Electrochemical and Photochemical Reduction of CO2 with Earth-Abundant Metal Complexes. Selective Production of CO vs HCOOH by Switching of the Metal Center. J. Am. Chem. Soc. 2015, 137, 10918–10921. 10.1021/jacs.5b06535. [DOI] [PubMed] [Google Scholar]
  166. Tinnemans A. H. A.; Koster T. P. M.; Thewissen D. H. M. W.; Mackor A. Tetraaza-macrocyclic cobalt(II) and nickel(II) complexes as electron-transfer agents in the photo(electro)chemical and electrochemical reduction of carbon dioxide. Recl. Trav. Chim. Pays-Bas 1984, 103, 288–295. 10.1002/recl.19841031004. [DOI] [Google Scholar]
  167. Beley M.; Collin J. P.; Ruppert R.; Sauvage J. P. Electrocatalytic reduction of carbon dioxide by nickel cyclam2+ in water: study of the factors affecting the efficiency and the selectivity of the process. J. Am. Chem. Soc. 1986, 108, 7461–7467. 10.1021/ja00284a003. [DOI] [PubMed] [Google Scholar]
  168. Bonin J.; Chaussemier M.; Robert M.; Routier M. Homogeneous Photocatalytic Reduction of CO2 to CO Using Iron(0) Porphyrin Catalysts: Mechanism and Intrinsic Limitations. ChemCatChem 2014, 6, 3200–3207. 10.1002/cctc.201402515. [DOI] [Google Scholar]
  169. Alsabeh P. G.; Rosas-Hernández A.; Barsch E.; Junge H.; Ludwig R.; Beller M. Iron-catalyzed photoreduction of carbon dioxide to synthesis gas. Catal. Sci. Technol. 2016, 6, 3623–3630. 10.1039/C5CY01129A. [DOI] [Google Scholar]
  170. Takeda H.; Koizumi H.; Okamoto K.; Ishitani O. Photocatalytic CO2 reduction using a Mn complex as a catalyst. Chem. Commun. 2014, 50, 1491–1493. 10.1039/C3CC48122K. [DOI] [PubMed] [Google Scholar]
  171. Torralba-Peñalver E.; Luo Y.; Compain J.-D.; Chardon-Noblat S.; Fabre B. Selective Catalytic Electroreduction of CO2 at Silicon Nanowires (SiNWs) Photocathodes Using Non-Noble Metal-Based Manganese Carbonyl Bipyridyl Molecular Catalysts in Solution and Grafted onto SiNWs. ACS Catal. 2015, 5, 6138–6147. 10.1021/acscatal.5b01546. [DOI] [Google Scholar]
  172. Cheung P. L.; Machan C. W.; Malkhasian A. Y. S.; Agarwal J.; Kubiak C. P. Photocatalytic Reduction of Carbon Dioxide to CO and HCO2H Using fac-Mn(CN)(bpy)(CO)3. Inorg. Chem. 2016, 55, 3192–3198. 10.1021/acs.inorgchem.6b00379. [DOI] [PubMed] [Google Scholar]
  173. Leung C.-F.; Ng S.-M.; Ko C.-C.; Man W.-L.; Wu J.; Chen L.; Lau T.-C. A cobalt(ii) quaterpyridine complex as a visible light-driven catalyst for both water oxidation and reduction. Energy Environ. Sci. 2012, 5, 7903–7907. 10.1039/c2ee21840b. [DOI] [Google Scholar]
  174. Guo Z.; Cheng S.; Cometto C.; Anxolabéhère-Mallart E.; Ng S.-M.; Ko C.-C.; Liu G.; Chen L.; Robert M.; Lau T.-C. Highly Efficient and Selective Photocatalytic CO2 Reduction by Iron and Cobalt Quaterpyridine Complexes. J. Am. Chem. Soc. 2016, 138, 9413–9416. 10.1021/jacs.6b06002. [DOI] [PubMed] [Google Scholar]
  175. Xie S.-L.; Liu J.; Dong L.-Z.; Li S.-L.; Lan Y.-Q.; Su Z.-M. Hetero-metallic active sites coupled with strongly reductive polyoxometalate for selective photocatalytic CO2-to-CH4 conversion in water. Chem. Sci. 2019, 10, 185–190. 10.1039/C8SC03471K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Das S.; Balaraju T.; Barman S.; Sreejith S. S.; Pochamoni R.; Roy S. A Molecular CO2 Reduction Catalyst Based on Giant Polyoxometalate {Mo368}. Front. Chem. 2018, 6, 514. 10.3389/fchem.2018.00514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Qin J.-S.; Du D.-Y.; Guan W.; Bo X.-J.; Li Y.-F.; Guo L.-P.; Su Z.-M.; Wang Y.-Y.; Lan Y.-Q.; Zhou H.-C. Ultrastable Polymolybdate-Based Metal–Organic Frameworks as Highly Active Electrocatalysts for Hydrogen Generation from Water. J. Am. Chem. Soc. 2015, 137, 7169–7177. 10.1021/jacs.5b02688. [DOI] [PubMed] [Google Scholar]
  178. Huang Q.; Liu J.; Feng L.; Wang Q.; Guan W.; Dong L.-Z.; Zhang L.; Yan L.-K.; Lan Y.-Q.; Zhou H.-C. Multielectron transportation of polyoxometalate-grafted metalloporphyrin coordination frameworks for selective CO2-to-CH4 photoconversion. Nat. Sci. Rev. 2020, 7, 53–63. 10.1093/nsr/nwz096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Fang Y.; Ma Y.; Zheng M.; Yang P.; Asiri A. M.; Wang X. Metal–organic frameworks for solar energy conversion by photoredox catalysis. Coord. Chem. Rev. 2018, 373, 83–115. 10.1016/j.ccr.2017.09.013. [DOI] [Google Scholar]
  180. Haviv E.; Shimon L. J. W.; Neumann R. Photochemical Reduction of CO2 with Visible Light Using a Polyoxometalate as Photoreductant. Chem. – Eur. J. 2017, 23, 92–95. 10.1002/chem.201605084. [DOI] [PubMed] [Google Scholar]
  181. Gao W.-Y.; Chen Y.; Niu Y.; Williams K.; Cash L.; Perez P. J.; Wojtas L.; Cai J.; Chen Y.-S.; Ma S. Crystal Engineering of an nbo Topology Metal–Organic Framework for Chemical Fixation of CO2 under Ambient Conditions. Angew. Chem., Int. Ed. 2014, 53, 2615–2619. 10.1002/anie.201309778. [DOI] [PubMed] [Google Scholar]
  182. Beyzavi H.; Klet R. C.; Tussupbayev S.; Borycz J.; Vermeulen N. A.; Cramer C. J.; Stoddart J. F.; Hupp J. T.; Farha O. K. A Hafnium-Based Metal–Organic Framework as an Efficient and Multifunctional Catalyst for Facile CO2 Fixation and Regioselective and Enantioretentive Epoxide Activation. J. Am. Chem. Soc. 2014, 136, 15861–15864. 10.1021/ja508626n. [DOI] [PubMed] [Google Scholar]
  183. Han Q.; Qi B.; Ren W.; He C.; Niu J.; Duan C. Polyoxometalate-based homochiral metal-organic frameworks for tandem asymmetric transformation of cyclic carbonates from olefins. Nat. Commun. 2015, 6, 10007 10.1038/ncomms10007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Li X.-X.; Liu J.; Zhang L.; Dong L.-Z.; Xin Z.-F.; Li S.-L.; Huang-Fu X.-Q.; Huang K.; Lan Y.-Q. Hydrophobic Polyoxometalate-Based Metal-Organic Framework for Efficient CO2 Photoconversion. ACS Appl. Mater. Interfaces 2019, 11, 25790–25795. 10.1021/acsami.9b03861. [DOI] [PubMed] [Google Scholar]
  185. Du J.; Ma Y.; Xin X.; Na H.; Zhao Y.; Tan H.; Han Z.; Li Y.; Kang Z. Reduced polyoxometalates and bipyridine ruthenium complex forming a tunable photocatalytic system for high efficient CO2 reduction. Chem. Eng. J. 2020, 398, 125518 10.1016/j.cej.2020.125518. [DOI] [Google Scholar]
  186. Bian Z.-Y.; Wang H.; Fu W.-F.; Li L.; Ding A.-Z. Two bifunctional RuII/ReI photocatalysts for CO2 reduction: A spectroscopic, photocatalytic, and computational study. Polyhedron 2012, 32, 78–85. 10.1016/j.poly.2011.08.037. [DOI] [Google Scholar]
  187. Koike K.; Grills D. C.; Tamaki Y.; Fujita E.; Okubo K.; Yamazaki Y.; Saigo M.; Mukuta T.; Onda K.; Ishitani O. Investigation of excited state, reductive quenching, and intramolecular electron transfer of Ru(ii)–Re(i) supramolecular photocatalysts for CO2 reduction using time-resolved IR measurements. Chem. Sci. 2018, 9, 2961–2974. 10.1039/C7SC05338J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Cheung P. L.; Kapper S. C.; Zeng T.; Thompson M. E.; Kubiak C. P. Improving Photocatalysis for the Reduction of CO2 through Non-covalent Supramolecular Assembly. J. Am. Chem. Soc. 2019, 141, 14961–14965. 10.1021/jacs.9b07067. [DOI] [PubMed] [Google Scholar]
  189. Wang J.-W.; Jiang L.; Huang H.-H.; Han Z.; Ouyang G. Rapid electron transfer via dynamic coordinative interaction boosts quantum efficiency for photocatalytic CO2 reduction. Nat. Commun. 2021, 12, 4276 10.1038/s41467-021-24647-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Yamazaki Y.; Ohkubo K.; Saito D.; Yatsu T.; Tamaki Y.; Tanaka Si.; Koike K.; Onda K.; Ishitani O. Kinetics and Mechanism of Intramolecular Electron Transfer in Ru(II)–Re(I) Supramolecular CO2–Reduction Photocatalysts: Effects of Bridging Ligands. Inorg. Chem. 2019, 58, 11480–11492. 10.1021/acs.inorgchem.9b01256. [DOI] [PubMed] [Google Scholar]
  191. Morimoto T.; Nakajima T.; Sawa S.; Nakanishi R.; Imori D.; Ishitani O. CO2 Capture by a Rhenium(I) Complex with the Aid of Triethanolamine. J. Am. Chem. Soc. 2013, 135, 16825–16828. 10.1021/ja409271s. [DOI] [PubMed] [Google Scholar]
  192. Kamogawa K.; Shimoda Y.; Miyata K.; Onda K.; Yamazaki Y.; Tamaki Y.; Ishitani O. Mechanistic study of photocatalytic CO2 reduction using a Ru(ii)–Re(i) supramolecular photocatalyst. Chem. Sci. 2021, 12, 9682–9693. 10.1039/D1SC02213J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Hitoki G.; Takata T.; Kondo J. N.; Hara M.; Kobayashi H.; Domen K. An oxynitride, TaON, as an efficient water oxidation photocatalyst under visible light irradiation (λ ≤ 500 nm). Chem. Commun. 2002, 1698–1699. 10.1039/B202393H. [DOI] [PubMed] [Google Scholar]
  194. Zhang F.; Yamakata A.; Maeda K.; Moriya Y.; Takata T.; Kubota J.; Teshima K.; Oishi S.; Domen K. Cobalt-Modified Porous Single-Crystalline LaTiO2N for Highly Efficient Water Oxidation under Visible Light. J. Am. Chem. Soc. 2012, 134, 8348–8351. 10.1021/ja301726c. [DOI] [PubMed] [Google Scholar]
  195. Kudo A.; Omori K.; Kato H. A Novel Aqueous Process for Preparation of Crystal Form-Controlled and Highly Crystalline BiVO4 Powder from Layered Vanadates at Room Temperature and Its Photocatalytic and Photophysical Properties. J. Am. Chem. Soc. 1999, 121, 11459–11467. 10.1021/ja992541y. [DOI] [Google Scholar]
  196. Ma S. S. K.; Maeda K.; Abe R.; Domen K. Visible-light-driven nonsacrificial water oxidation over tungsten trioxide powder modified with two different cocatalysts. Energy Environ. Sci. 2012, 5, 8390–8397. 10.1039/c2ee21801a. [DOI] [Google Scholar]
  197. Yui T.; Kan A.; Saitoh C.; Koike K.; Ibusuki T.; Ishitani O. Photochemical Reduction of CO2 Using TiO2: Effects of Organic Adsorbates on TiO2 and Deposition of Pd onto TiO2. ACS Appl. Mater. Interfaces 2011, 3, 2594–2600. 10.1021/am200425y. [DOI] [PubMed] [Google Scholar]
  198. Iizuka K.; Wato T.; Miseki Y.; Saito K.; Kudo A. Photocatalytic Reduction of Carbon Dioxide over Ag Cocatalyst-Loaded ALa4Ti4O15 (A = Ca, Sr, and Ba) Using Water as a Reducing Reagent. J. Am. Chem. Soc. 2011, 133, 20863–20868. 10.1021/ja207586e. [DOI] [PubMed] [Google Scholar]
  199. Teramura K.; Iguchi S.; Mizuno Y.; Shishido T.; Tanaka T. Photocatalytic Conversion of CO2 in Water over Layered Double Hydroxides. Angew. Chem., Int. Ed. 2012, 51, 8008–8011. 10.1002/anie.201201847. [DOI] [PubMed] [Google Scholar]
  200. Sato S.; Morikawa T.; Saeki S.; Kajino T.; Motohiro T. Visible-Light-Induced Selective CO2 Reduction Utilizing a Ruthenium Complex Electrocatalyst Linked to a p-Type Nitrogen-Doped Ta2O5 Semiconductor. Angew. Chem., Int. Ed. 2010, 49, 5101–5105. 10.1002/anie.201000613. [DOI] [PubMed] [Google Scholar]
  201. Maeda K.; Sekizawa K.; Ishitani O. A polymeric-semiconductor–metal-complex hybrid photocatalyst for visible-light CO2 reduction. Chem. Commun. 2013, 49, 10127–10129. 10.1039/c3cc45532g. [DOI] [PubMed] [Google Scholar]
  202. Maeda K.; Kuriki R.; Zhang M.; Wang X.; Ishitani O. The effect of the pore-wall structure of carbon nitride on photocatalytic CO2 reduction under visible light. J. Mater. Chem. A 2014, 2, 15146–15151. 10.1039/C4TA03128H. [DOI] [Google Scholar]
  203. Kuriki R.; Sekizawa K.; Ishitani O.; Maeda K. Visible-Light-Driven CO2 Reduction with Carbon Nitride: Enhancing the Activity of Ruthenium Catalysts. Angew. Chem., Int. Ed. 2015, 54, 2406–2409. 10.1002/anie.201411170. [DOI] [PubMed] [Google Scholar]
  204. Ha E.-G.; Chang J.-A.; Byun S.-M.; Pac C.; Jang D.-M.; Park J.; Kang S. O. High-turnover visible-light photoreduction of CO2 by a Re(i) complex stabilized on dye-sensitized TiO2. Chem. Commun. 2014, 50, 4462–4464. 10.1039/C3CC49744E. [DOI] [PubMed] [Google Scholar]
  205. Sahara G.; Abe R.; Higashi M.; Morikawa T.; Maeda K.; Ueda Y.; Ishitani O. Photoelectrochemical CO2 reduction using a Ru(ii)–Re(i) multinuclear metal complex on a p-type semiconducting NiO electrode. Chem. Commun. 2015, 51, 10722–10725. 10.1039/C5CC02403J. [DOI] [PubMed] [Google Scholar]
  206. Arai T.; Sato S.; Kajino T.; Morikawa T. Solar CO2 reduction using H2O by a semiconductor/metal-complex hybrid photocatalyst: enhanced efficiency and demonstration of a wireless system using SrTiO3 photoanodes. Energy Environ. Sci. 2013, 6, 1274–1282. 10.1039/c3ee24317f. [DOI] [Google Scholar]
  207. Sato S.; Arai T.; Morikawa T. Toward Solar-Driven Photocatalytic CO2 Reduction Using Water as an Electron Donor. Inorg. Chem. 2015, 54, 5105–5113. 10.1021/ic502766g. [DOI] [PubMed] [Google Scholar]
  208. Sato S.; Arai T.; Morikawa T.; Uemura K.; Suzuki T. M.; Tanaka H.; Kajino T. Selective CO2 Conversion to Formate Conjugated with H2O Oxidation Utilizing Semiconductor/Complex Hybrid Photocatalysts. J. Am. Chem. Soc. 2011, 133, 15240–15243. 10.1021/ja204881d. [DOI] [PubMed] [Google Scholar]
  209. Arai T.; Sato S.; Morikawa T. A monolithic device for CO2 photoreduction to generate liquid organic substances in a single-compartment reactor. Energy Environ. Sci. 2015, 8, 1998–2002. 10.1039/C5EE01314C. [DOI] [Google Scholar]
  210. Kamata R.; Kumagai H.; Yamazaki Y.; Sahara G.; Ishitani O. Photoelectrochemical CO2 Reduction Using a Ru(II)–Re(I) Supramolecular Photocatalyst Connected to a Vinyl Polymer on a NiO Electrode. ACS Appl. Mater. Interfaces 2019, 11, 5632–5641. 10.1021/acsami.8b05495. [DOI] [PubMed] [Google Scholar]
  211. Townsend T. K.; Sabio E. M.; Browning N. D.; Osterloh F. E. Improved Niobate Nanoscroll Photocatalysts for Partial Water Splitting. ChemSusChem 2011, 185–190. 10.1002/cssc.201000377. [DOI] [PubMed] [Google Scholar]
  212. Kim Y. I.; Atherton S. J.; Brigham E. S.; Mallouk T. E. Sensitized layered metal oxide semiconductor particles for photochemical hydrogen evolution from nonsacrificial electron donors. J. Phys. Chem. A 1993, 97, 11802–11810. 10.1021/j100147a038. [DOI] [Google Scholar]
  213. Butler S. Z.; Hollen S. M.; Cao L.; Cui Y.; Gupta J. A.; Gutiérrez H. R.; Heinz T. F.; Hong S. S.; Huang J.; Ismach A. F.; Johnston-Halperin E.; Kuno M.; Plashnitsa V. V.; Robinson R. D.; Ruoff R. S.; Salahuddin S.; Shan J.; Shi L.; Spencer M. G.; Terrones M.; Windl W.; Goldberger J. E. Progress, Challenges, and Opportunities in Two-Dimensional Materials Beyond Graphene. ACS Nano 2013, 7, 2898–2926. 10.1021/nn400280c. [DOI] [PubMed] [Google Scholar]
  214. Kawazoe H.; Yasukawa M.; Hyodo H.; Kurita M.; Yanagi H.; Hosono H. P-type electrical conduction in transparent thin films of CuAlO2. Nature 1997, 389, 939–942. 10.1038/40087. [DOI] [Google Scholar]
  215. Ueda K.; Hase T.; Yanagi H.; Kawazoe H.; Hosono H.; Ohta H.; Orita M.; Hirano M. Epitaxial growth of transparent p-type conducting CuGaO2 thin films on sapphire (001) substrates by pulsed laser deposition. J. Appl. Phys. 2001, 89, 1790–1793. 10.1063/1.1337587. [DOI] [Google Scholar]
  216. Kumagai H.; Sahara G.; Maeda K.; Higashi M.; Abe R.; Ishitani O. Hybrid photocathode consisting of a CuGaO2 p-type semiconductor and a Ru(ii)–Re(i) supramolecular photocatalyst: non-biased visible-light-driven CO2 reduction with water oxidation. Chem. Sci. 2017, 8, 4242–4249. 10.1039/C7SC00940B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Saito D.; Yamazaki Y.; Tamaki Y.; Ishitani O. Photocatalysis of a Dinuclear Ru(II)–Re(I) Complex for CO2 Reduction on a Solid Surface. J. Am. Chem. Soc. 2020, 142, 19249–19258. 10.1021/jacs.0c09170. [DOI] [PubMed] [Google Scholar]
  218. Sekizawa K.; Maeda K.; Domen K.; Koike K.; Ishitani O. Artificial Z-Scheme Constructed with a Supramolecular Metal Complex and Semiconductor for the Photocatalytic Reduction of CO2. J. Am. Chem. Soc. 2013, 135, 4596–4599. 10.1021/ja311541a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Yoshitomi F.; Sekizawa K.; Maeda K.; Ishitani O. Selective Formic Acid Production via CO2 Reduction with Visible Light Using a Hybrid of a Perovskite Tantalum Oxynitride and a Binuclear Ruthenium(II) Complex. ACS Appl. Mater. Interfaces 2015, 7, 13092–13097. 10.1021/acsami.5b03509. [DOI] [PubMed] [Google Scholar]
  220. Kuehnel M. F.; Orchard K. L.; Dalle K. E.; Reisner E. Selective photocatalytic CO2 reduction in water through anchoring of a molecular Ni catalyst on CdS nanocrystals. J. Am. Chem. Soc. 2017, 139, 7217–7223. 10.1021/jacs.7b00369. [DOI] [PubMed] [Google Scholar]
  221. Kuehnel M. F.; Sahm C. D.; Neri G.; Lee J. R.; Orchard K. L.; Cowan A. J.; Reisner E. ZnSe quantum dots modified with a Ni (cyclam) catalyst for efficient visible-light driven CO2 reduction in water. Chem. Sci. 2018, 9, 2501–2509. 10.1039/C7SC04429A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Lian S.; Kodaimati M. S.; Weiss E. A. Photocatalytically active superstructures of quantum dots and iron porphyrins for reduction of CO2 to CO in water. ACS Nano 2018, 12, 568–575. 10.1021/acsnano.7b07377. [DOI] [PubMed] [Google Scholar]
  223. Suzuki T. M.; Takayama T.; Sato S.; Iwase A.; Kudo A.; Morikawa T. Enhancement of CO2 reduction activity under visible light irradiation over Zn-based metal sulfides by combination with Ru-complex catalysts. Appl. Catal., B 2018, 224, 572–578. 10.1016/j.apcatb.2017.10.053. [DOI] [Google Scholar]
  224. Ellis A. B.; Kaiser S. W.; Bolts J. M.; Wrighton M. S. Study of n-type semiconducting cadmium chalcogenide-based photoelectrochemical cells employing polychalcogenide electrolytes. J. Am. Chem. Soc. 1977, 99, 2839–2848. 10.1021/ja00451a001. [DOI] [Google Scholar]
  225. Lian S.; Kodaimati M. S.; Dolzhnikov D. S.; Calzada R.; Weiss E. A. Powering a CO2 reduction catalyst with visible light through multiple sub-picosecond electron transfers from a quantum dot. J. Am. Chem. Soc. 2017, 139, 8931–8938. 10.1021/jacs.7b03134. [DOI] [PubMed] [Google Scholar]
  226. Maeda K. Metal-Complex/Semiconductor Hybrid Photocatalysts and Photoelectrodes for CO2 Reduction Driven by Visible Light. Adv. Mater. 2019, 31, 1808205 10.1002/adma.201808205. [DOI] [PubMed] [Google Scholar]
  227. Wada K.; Eguchi M.; Ishitani O.; Maeda K. Activation of the Carbon Nitride Surface by Silica in a CO-Evolving Hybrid Photocatalyst. ChemSusChem 2017, 10, 287–295. 10.1002/cssc.201600661. [DOI] [PubMed] [Google Scholar]
  228. Ueda Y.; Takeda H.; Yui T.; Koike K.; Goto Y.; Inagaki S.; Ishitani O. A Visible-Light Harvesting System for CO2 Reduction Using a RuII–ReI Photocatalyst Adsorbed in Mesoporous Organosilica. ChemSusChem 2015, 8, 439–442. 10.1002/cssc.201403194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Wada K.; Ranasinghe C. S. K.; Kuriki R.; Yamakata A.; Ishitani O.; Maeda K. Interfacial Manipulation by Rutile TiO2 Nanoparticles to Boost CO2 Reduction into CO on a Metal-Complex/Semiconductor Hybrid Photocatalyst. ACS Appl. Mater. Interfaces 2017, 9, 23869–23877. 10.1021/acsami.7b07484. [DOI] [PubMed] [Google Scholar]
  230. Chen P.; Lei B.; Dong X.; Wang H.; Sheng J.; Cui W.; Li J.; Sun Y.; Wang Z.; Dong F. Rare-Earth Single-Atom La–N Charge-Transfer Bridge on Carbon Nitride for Highly Efficient and Selective Photocatalytic CO2 Reduction. ACS Nano 2020, 14, 15841–15852. 10.1021/acsnano.0c07083. [DOI] [PubMed] [Google Scholar]
  231. Takeda H.; Ohashi M.; Tani T.; Ishitani O.; Inagaki S. Enhanced Photocatalysis of Rhenium(I) Complex by Light-Harvesting Periodic Mesoporous Organosilica. Inorg. Chem. 2010, 49, 4554–4559. 10.1021/ic1000914. [DOI] [PubMed] [Google Scholar]
  232. Sung-Suh H. M.; Kim D. S.; Lee C. W.; Park S.-E. Photoinduced activation of CO2 by rhenium complexes encapsulated in molecular sieves. Appl. Organomet. Chem. 2000, 14, 826–830. . [DOI] [Google Scholar]
  233. Dubois K. D.; Petushkov A.; Garcia Cardona E.; Larsen S. C.; Li G. Adsorption and Photochemical Properties of a Molecular CO2 Reduction Catalyst in Hierarchical Mesoporous ZSM-5: An In Situ FTIR Study. J. Phys. Chem. Lett. 2012, 3, 486–492. 10.1021/jz201701y. [DOI] [PubMed] [Google Scholar]
  234. Liu C.; Dubois K. D.; Louis M. E.; Vorushilov A. S.; Li G. Photocatalytic CO2 Reduction and Surface Immobilization of a Tricarbonyl Re(I) Compound Modified with Amide Groups. ACS Catal. 2013, 3, 655–662. 10.1021/cs300796e. [DOI] [Google Scholar]
  235. Hwang J.-S.; Kim D. S.; Lee C. W.; Park S.-E. Photocatalytic activation of CO2 under visible light by Rhenium complex encapsulated in molecular sieves. Korean J. Chem. Eng. 2001, 18, 919–923. 10.1007/BF02705619. [DOI] [Google Scholar]
  236. Dubois K. D.; He H.; Liu C.; Vorushilov A. S.; Li G. Covalent attachment of a molecular CO2-reduction photocatalyst to mesoporous silica. J. Mol. Catal. A: Chem. 2012, 363–364, 208–213. 10.1016/j.molcata.2012.06.011. [DOI] [Google Scholar]
  237. Bhat A.; Anwer S.; Bhat K. S.; Mohideen M. I. H.; Liao K.; Qurashi A. Prospects challenges and stability of 2D MXenes for clean energy conversion and storage applications. npj 2D Mater. Appl. 2021, 5, 61 10.1038/s41699-021-00239-8. [DOI] [Google Scholar]
  238. He H.; Liu C.; Louis M. E.; Li G. Infrared studies of a hybrid CO2-reduction photocatalyst consisting of a molecular Re(I) complex grafted on Kaolin. J. Mol. Catal. A: Chem. 2014, 395, 145–150. 10.1016/j.molcata.2014.08.020. [DOI] [Google Scholar]
  239. Wang C.; Xie Z.; deKrafft K. E.; Lin W. Doping Metal–Organic Frameworks for Water Oxidation, Carbon Dioxide Reduction, and Organic Photocatalysis. J. Am. Chem. Soc. 2011, 133, 13445–13454. 10.1021/ja203564w. [DOI] [PubMed] [Google Scholar]
  240. Sun D.; Gao Y.; Fu J.; Zeng X.; Chen Z.; Li Z. Construction of a supported Ru complex on bifunctional MOF-253 for photocatalytic CO2 reduction under visible light. Chem. Commun. 2015, 51, 2645–2648. 10.1039/C4CC09797A. [DOI] [PubMed] [Google Scholar]
  241. Li L.; Zhang S.; Xu L.; Wang J.; Shi L.-X.; Chen Z.-N.; Hong M.; Luo J. Effective visible-light driven CO2 photoreduction via a promising bifunctional iridium coordination polymer. Chem. Sci. 2014, 5, 3808–3813. 10.1039/C4SC00940A. [DOI] [Google Scholar]
  242. Sheng J.; He Y.; Li J.; Yuan C.; Huang H.; Wang S.; Sun Y.; Wang Z.; Dong F. Identification of Halogen-Associated Active Sites on Bismuth-Based Perovskite Quantum Dots for Efficient and Selective CO2-to-CO Photoreduction. ACS Nano 2020, 14, 13103–13114. 10.1021/acsnano.0c04659. [DOI] [PubMed] [Google Scholar]
  243. Kuriki R.; Sekizawa K.; Ishitani O.; Maeda K. Visible-light-driven CO2 reduction with carbon nitride: enhancing the activity of ruthenium catalysts. Angew. Chem., Int. Ed. 2015, 54, 2406–2409. 10.1002/anie.201411170. [DOI] [PubMed] [Google Scholar]
  244. Maeda K.; Kuriki R.; Ishitani O. Photocatalytic activity of carbon nitride modified with a Ruthenium (II) complex having carboxylic-or phosphonic acid anchoring groups for visible-light CO2 reduction. Chem. Lett. 2016, 45, 182–184. 10.1246/cl.151061. [DOI] [Google Scholar]
  245. Suzuki T. M.; Tanaka H.; Morikawa T.; Iwaki M.; Sato S.; Saeki S.; Inoue M.; Kajino T.; Motohiro T. Direct assembly synthesis of metal complex–semiconductor hybrid photocatalysts anchored by phosphonate for highly efficient CO 2 reduction. Chem. Commun. 2011, 47, 8673–8675. 10.1039/c1cc12491a. [DOI] [PubMed] [Google Scholar]
  246. Kuriki R.; Sekizawa K.; Ishitani O.; Maeda K. Visible-light-driven CO2 reduction with carbon nitride: enhancing the activity of ruthenium catalysts. Angew. Chem., Int. Ed. 2015, 54, 2406–2409. 10.1002/anie.201411170. [DOI] [PubMed] [Google Scholar]
  247. Windle C. D.; Pastor E.; Reynal A.; Whitwood A. C.; Vaynzof Y.; Durrant J. R.; Perutz R. N.; Reisner E. Improving the Photocatalytic Reduction of CO2 to CO through Immobilisation of a Molecular Re Catalyst on TiO2. Chem. – Eur. J. 2015, 21, 3746–3754. 10.1002/chem.201405041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Liu J.; Ren Y.-Y.; Wu J.; Xia W.; Deng B.-Y.; Wang F. Hybrid artificial photosynthetic systems constructed using quantum dots and molecular catalysts for solar fuel production: development and advances. J. Mater. Chem. A 2021, 9, 19346–19368. 10.1039/d1ta02673a. [DOI] [Google Scholar]
  249. Fukuzumi S.; Koumitsu S.; Hironaka K.; Tanaka T. Energetic comparison between photoinduced electron-transfer reactions from NADH model compounds to organic and inorganic oxidants and hydride-transfer reactions from NADH model compounds to p-benzoquinone derivatives. J. Am. Chem. Soc. 1987, 109, 305–316. 10.1021/ja00236a003. [DOI] [Google Scholar]
  250. Patz M.; Kuwahara Y.; Suenobu T.; Fukuzumi S. Oxidation Mechanism of NAD Dimer Model Compounds. Chem. Lett. 1997, 26, 567–568. 10.1246/cl.1997.567. [DOI] [Google Scholar]
  251. Zhu X.-Q.; Zhang M.-T.; Yu A.; Wang C.-H.; Cheng J.-P. Hydride, Hydrogen Atom, Proton, and Electron Transfer Driving Forces of Various Five-Membered Heterocyclic Organic Hydrides and Their Reaction Intermediates in Acetonitrile. J. Am. Chem. Soc. 2008, 130, 2501–2516. 10.1021/ja075523m. [DOI] [PubMed] [Google Scholar]
  252. Hasegawa E.; Seida T.; Chiba N.; Takahashi T.; Ikeda H. Contrastive Photoreduction Pathways of Benzophenones Governed by Regiospecific Deprotonation of Imidazoline Radical Cations and Additive Effects. J. Org. Chem. 2005, 70, 9632–9635. 10.1021/jo0514220. [DOI] [PubMed] [Google Scholar]
  253. Hasegawa E.; Takizawa S.; Seida T.; Yamaguchi A.; Yamaguchi N.; Chiba N.; Takahashi T.; Ikeda H.; Akiyama K. Photoinduced electron-transfer systems consisting of electron-donating pyrenes or anthracenes and benzimidazolines for reductive transformation of carbonyl compounds. Tetrahedron 2006, 62, 6581–6588. 10.1016/j.tet.2006.03.061. [DOI] [Google Scholar]
  254. Hasegawa E.; Tosaka E.; Yoneoka A.; Tamura Y.; Takizawa S.-y.; Tomura M.; Yamashita Y. Photoinduced electron-transfer reaction of α-bromomethyl-substituted benzocyclic β-keto esters with amines: selective reaction pathways depending on the nature of the amine radical cations. Res. Chem. Intermed. 2013, 39, 247–267. 10.1007/s11164-012-0646-2. [DOI] [Google Scholar]
  255. Hasegawa E.; Ohta T.; Tsuji S.; Mori K.; Uchida K.; Miura T.; Ikoma T.; Tayama E.; Iwamoto H.; Takizawa S.-y.; Murata S. Aryl-substituted dimethylbenzimidazolines as effective reductants of photoinduced electron transfer reactions. Tetrahedron 2015, 71, 5494–5505. 10.1016/j.tet.2015.06.071. [DOI] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES