You are currently viewing a new version of our website. To view the old version click .
Catalysts
  • Review
  • Open Access

2 November 2025

Visible-Light-Driven CO2 Photoreduction Using Ruthenium (II) Complexes: Mechanisms, Hybrid Systems and Recent Advances

and
Department of Chemistry, College of Science, Engineering, and Technology, Florida Science Campus, University of South Africa, Johannesburg 1709, South Africa
*
Authors to whom correspondence should be addressed.
This article belongs to the Special Issue Heterogeneous Catalysis for Sustainable Conversion of Biomass, Carbon Dioxide and Plastic Waste into Fuels and Chemicals

Abstract

The photocatalytic reduction of carbon dioxide (CO2) into energy-dense fuels using visible light provides a sustainable approach for solar-to-chemical energy transformation. Among the diverse metal molecular systems developed, ruthenium (II) (Ru(II)) complexes have emerged as promising catalysts due to their superior redox properties, strong visible light absorption, and customizable ligand structures. This review explores recent advances in Ru(II)-catalyzed CO2 photoreduction, with particular attention given to catalyst design strategies, mechanistic pathways, and system integration methodologies. Key configurations, including photosensitizer/catalyst (PS/Cat) mixed systems, covalently bonded dyads, and hybrid/supramolecular frameworks, are evaluated in terms of efficiency, turnover numbers (TON), and selectivity. A critical analysis of challenges such as competing H2 generation, inefficient charge transfer, and limited long-term stability is presented. Emerging trends toward the use of pincer ligands, transition metal integration, and self-photosensitizing frameworks are discussed as potential approaches for improving efficiency. Overall, this review offers insights into the structural and mechanistic features driving CO2 photoreduction and provides perspectives for the rational design of next-generation Ru-based photocatalytic systems for efficient solar CO2 conversion and the photocatalytic reduction of carbon dioxide (CO2) into energy-dense fuels using visible light.

1. Introduction

The need for sustainable energy solutions that address climate change and energy demand has increased in tandem with the rise in atmospheric CO2 levels. One possible method for converting solar energy into chemical fuels is photocatalytic CO2 reduction with visible light. However, due to its strong C=O bonds (750 kJ/mol), high reduction potential, and large highest occupied molecular orbital–lowest unoccupied molecular orbital (HOMO–LUMO) gap (13.7 eV), CO2 is thermodynamically and kinetically stable, hindering its reduction []. Nonetheless, Ru complexes have garnered significant interest as both photosensitizers (PS) and catalysts (Cat) because of their advantageous photochemical and redox characteristics. To overcome the drawbacks of traditional PS/Cat mixed systems, integrated assemblies such as dyads and hybrid/supramolecular composites have been developed to enhance charge transfer and catalyst stability. Additionally, recent advancements, including self-photosensitizing complexes, pincer ligands, and transition metal (TM) cocatalysts, have demonstrated improved selectivity and turnover numbers (TON) under milder, more sustainable conditions.
While extensive research exists on heterogeneous photocatalysts, such as Ru-doped titanates [] and broader TiO2 derivatives [], these studies focus on morphology, band-gap reduction, and pollutant removal, with only indirect effects on charge transfer and CO2 photoreduction selectivity []. This review adopts an architecture-focused approach, examining Ru(II) complexes across PS/Cat systems, dyads, and hybrid/supramolecular frameworks, with benchmarking analyses (TON, Apparent Quantum Yield (AQY), selectivity (S). By dissecting competing CO and HCOO cycles, this review links mechanistic pathways to product selectivity. Furthermore, this review evaluates the influence of donors, solvents, and ligands, and highlights superficial results from photolysis. Emerging strategies, including Ru–pincer/self-photosensitizing dyads and Z-scheme-like aqueous systems, are emphasized to guide the development of durable, selective photocatalysts for efficient solar-to-chemical energy conversion.

2. Mechanism of Photocatalytic CO2 Reduction

The photoreduction of CO2 is predominantly a surface process that utilizes protons (H+) and photo-generated charge carriers (e/h+) as shown in Figure 1.
Figure 1. Representative illustration of photocatalytic CO2 reduction [].
Figure 1 shows that the key stages in CO2 photoreduction include light absorption, CO2 adsorption, surface-redox reactions, electron-hole (e/h+) pair formation, and product desorption. Another essential process component is the separation and transfer of charge. However, the one-electron reduction of CO2 generates a CO2 radical anion (CO2•−) (Equation (1)), which has a high reduction potential (−1.9 V vs. SHE in water at pH 7) [], making the process thermodynamically unfavorable. Alternatively, multi-electron reduction via proton-coupled electron transfer (Equations (2)–(6)) can yield stable, energy-dense products at relatively lower reduction potentials [,,].
CO2 + e → CO2•− (Eo = −1.9)
CO2 + 2H+ + 2e → COOH (Eo = −0.61 V)
CO2 + 2H+ + 2e → CO + H2O (Eo = −0.53 V)
CO2 + 4H+ + 4e → HCHO + H2O (Eo = −0.48 V)
CO2 + 6H+ + 6e → CH3OH + H2O (Eo = −0.38 V)
CO2 + 8H+ + 8e → CH4 + 2H2O (Eo = −0.24)
However, the kinetic barrier generally increases with additional electrons and protons participating in the reaction []. Furthermore, the multistep photocatalytic process frequently leads to low efficiency, particularly in the transport and separation of photoinduced charge carriers. Therefore, a catalyst capable of promoting multi-electron and multi-proton transfers is necessary to facilitate efficient turnovers. For example, the band gap energy of the photocatalyst must meet the thermodynamic requirements for CO2 photoreduction to facilitate electron transfer to the CO2 molecules adsorbed on its surface []. Additionally, the valence band (VB) edge must be positioned at a more negative potential than the standard reduction potential (E red) of CO2 to ensure the thermodynamic feasibility of its conversion to other products []. Importantly, the distribution of reduction products is influenced by the photocatalyst’s band edge position, which controls the build-up of electrons [].

3. Photocatalytic CO2 Reduction Using Metal Molecular Catalysts

As much as the highlighted limitations of CO2 hinder photocatalytic performance, significant progress has been realized in the development of more effective photocatalysts with superior qualities. Among them, visible light-active molecular catalysts have shown attractive results. The most common molecular catalysts include Re(I) (Re(bpy)(CO)3Cl), Ru(II) (Ru(bpy)32+), Ir(III) ([Ir(ppy)3]), Ni(II) (Ni(cyclam)), Co(II) (Co(bpy)32+), and Mn(I) (Mn(bpy)(CO)3Br) complexes (bpy: 2,2′-bipyridine; ppy: 2-phenylpyridine). These systems absorb light and transition to an excited state capable of facilitating redox reactions []. The advantages of molecular catalysts include their molecular-scale design and adjustable organic ligands, which allow precise control over functionality (e.g., photophysical, electrochemical, and catalytic properties), leading to enhanced light absorption, catalytic activity, and selectivity []. However, like other systems, metal complexes have drawbacks such as solubility issues, catalyst deactivation, poor recyclability, scalability challenges, high costs of precious metals, and the need for multiple components in solution, which can result in inefficient and random interactions [].
A typical photocatalytic CO2 reduction system involving molecular components consists of three parts: a redox PS for harvesting light energy, a sacrificial electron donor (SD) that provides electrons for the reduction process, and a Cat that serves as the active site for CO2 transformation. The photoreduction process can proceed via two main mechanisms: oxidative quenching and reductive quenching. In reductive quenching, the SD transfers an electron to the excited state of the PS (PS*), forming a reduced species (PS•−), commonly referred to as the one-electron-reduced species (OERS) []. The PS•− then donates an electron to the catalyst, facilitating CO2 reduction. Through successive iterations of this process, depending on the number of electrons required for the desired product, CO2 can be converted into chemical fuels. Noteworthy, in some cases, PS can directly transfer an electron to the catalyst, although this pathway is less common [,]. Figure 2 illustrates a typical CO2 reductive quenching system.
Figure 2. Mechanism of reductive quenching for photocatalytic CO2 reduction [].
On the other hand, although feasible, the oxidative quenching mechanism, wherein PS directly donates an electron to the Cat, is less common. This may be due to the highly negative Ered required for CO2 conversion, which necessitates that the OERS possess greater reducing power than the corresponding PS []. The formation of strong and relatively stable oxidants, such as PS•+, in the reaction medium is unfavorable for CO2 reduction, as these species can hinder efficient electron transfer to the catalyst [].
The initial electron transfer from SD to PS* (Figure 2) is thermodynamically favorable when the Eored of PS* is comparable to or more positive than the oxidation potential (Eoox) of SD. The Eored of PS* can be calculated using Equation (7) [,]:
E(PS*/PS•−) = E(PS/PS•−) + E0-0
The equation implies that the oxidation state of the catalyst is governed by the redox potential of the PS*/PS•− couple. The efficiency of electron transfer can be evaluated using the quenching rate constant (kq) and quenching efficiency (ηq). The kq is determined from Stern–Volmer plots and the lifetime of PS*, while ηq, representing the fraction of PS* that forms PS•−, is calculated using Equation (8) []
η q = 1 1 + K q ͳ × e l e c t r o n   d o n o r
A higher quenching fraction (ηq) leads to an increase in quantum efficiency, and vice versa. Other critical process parameters for measuring the performance of CO2 photoreduction include turnover number (TON), Apparent Quantum Yield (AQY) and selectivity (S):
T O N = p r o d u c t m o l c a t a l y s t m o l
TON > 1 indicates the feasibility of the CO2 reduction process, as well as high catalyst durability.
A Q Y % = n u m b e r   o f   r e a c t e d   e l e c t r o n s   o r   m o l e c u l e s n u m b e r   o f   i n c i d e n t   p h o t o n s × 100
A higher AQY suggests a more effective use of light in driving the CO2 photoreduction process.
S e l e c t i v i t y S % = n u m b e r   o f   r e a c t e d   e l e c t r o n s s u m   o f   a l l   r e a c t e d   e l e c t r o n s × 100

3.1. Electron Donors

Electron donors release a proton in the oxidized electron donor state (OEOS), which is crucial for suppressing back-electron transfer from the OERS of the PS to the OEOS of the SD [,]. Commonly used SDs include aliphatic amines such as triethylamine (TEA) and triethanolamine (TEOA), ascorbate (AscH), and NADH model compounds such as 1-benzyl-1,4-dihydronicotinamide (BNAH) and dihydro-1H-benzo[d]imidazole derivatives (BIH, BI(OH)H). These SDs have often been employed in photochemical CO2 reduction reactions. Figure 3 illustrates the mechanism of sacrificial reduction by BNAH and BIH in the presence of TEOA.
Figure 3. Schematic illustration of sacrificial reduction by BIH and BNAH in the presence of TEOA [].
TEOA is a commonly used SD for reductive quenching of PS*, functioning as a Brønsted base in the presence of stronger electron donors such as BNAH and BIH. While TEOA does not directly quench PS*, it can abstract a proton from BNAH•+ or BIH•+ in the reaction medium []. BNAH, with a stronger reducing potential Eo(BNAH/BNAH+) = 0.57 V vs. SCE in acetonitrile (MeCN), is more effective than TEA (EP = 0.96 V) and TEOA (EP = 0.80 V) at reducing PS* []. Furthermore, BNAH can efficiently undergo reductive quenching, and its reducing power is retained in aqueous media, unlike aliphatic amines, whose protonation diminishes their effectiveness [].
BIH and BI(OH)H show a stronger reducing tendency than BNAH with E1/2(ox) values of 0.33 V and 0.31 V vs. SCE, respectively []. BIH follows a unique two-electron oxidation via first electron transfer, followed by deprotonation and a second electron transfer []. The deprotonated intermediate (BI) exhibits negative oxidation potential, facilitating efficient electron transfer in CO2 photoreduction. BIH also shows weak basicity, although its protonated form (BIH2+) is inactive in reductive quenching and less effective without TEOA []. On the other hand, BI(OH)H donates two electrons and two protons, forming BI(O)+ and selectively converting CO2 to formic acid (HCOOH). This dual electron–proton transfer increases its effectiveness over BIH. In contrast, BNAH donates only one electron and one proton. These mechanistic differences strongly affect the donors’ reactivity and selectivity in CO2 photo conversion. Importantly, an ideal SD enhances reductive quenching efficiency, thereby improving overall photocatalytic performance.

3.2. Photosensitizers

Generally, the PS should enable electron transfer from the SD to the Cat upon photoexcitation. Thus, key requirements for an effective PS include (i) strong absorption at the excitation wavelength, particularly in the visible range, to efficiently harvest solar energy and suppress absorption by coexisting species such as the SD and Cat; (ii) a long-lived excited state to promote efficient reductive quenching; (iii) a high oxidative potential in the PS to readily accept an electron from the donor; and (iv) high photochemical stability of the PS•− along with efficient intersystem crossing []. Phosphorescent metal complexes, particularly those based on Ru, iridium (Ir), and osmium (Os) (Figure 4), are widely employed as photosensitizers in CO2 reduction systems. The popularity of these compounds stems from favorable redox properties and excellent photosensitizing performance, attributed to their metal-to-ligand charge transfer (MLCT) excited states, which typically exhibit lifetimes in the nanosecond to microsecond range [].
Figure 4. Phosphorescent metal complexes as photosensitizers in CO2 photo reduction [].
Ru provides more flexibility than Ir and Os due to its well-developed ligand chemistry (such as pincer, terpyridine, and bipyridine ligands). This allows for effective redox modulation and selective product distribution (CO or HCOO) []. Additionally, Ru presents less toxicity and volatility problems commonly associated with osmium (e.g., OsO4) []. These features make Ru an ideal substrate for developing selective and scalable, CO2 photoreduction catalysts. It is therefore not surprising that [Ru(bpy)3]2+ and its derivatives were among the first photosensitizers employed in CO2 reduction systems [].
While the OERS of most Ru(II)n+ complexes are generally stable in the dark, photoexcitation can occasionally induce their decomposition, leading to the formation of [Ru(bpy)2X2]ᵐ+-type complexes, where X may be a solvent molecule, CO, or Cl []. These decomposition products can themselves act as photocatalysts for CO2 reduction, enabling the reaction to proceed even in the absence of an added catalyst, which can affect product selectivity []. In the presence of water, CO2 can be converted to formate (HCOO) [].

3.3. Catalytic Roles of [Ru(bpy)2(CO)2]2+ Complexes

In recent years, [Ru(bpy)2(CO)2]2+ (Ru(CO))-type complexes and their derivatives have been frequently employed as catalysts in CO2 photoreduction. Table 1 displays some of the common Ru(II) complexes used as catalysts in CO2 reduction.
Table 1. Common Ru complexes and their properties [].
Similarly to [Ru(bpy)3]2+, Ru(CO) complexes exhibit strong visible-light absorption, high CO2-binding affinity, enhanced product selectivity, and multiple accessible redox states that support multielectron reduction processes []. These Ru(II) complexes also demonstrate excellent durability and preferentially reduce CO2 over protons, even in aqueous media []. The presence of labile halide ligands enables CO2 coordination via halogen replacement, forming M-C bonds that promote selective CO2 binding to the metal center []. Additionally, Ru(CO)-type complexes can serve dual roles, as both PSs and Cat in some systems. This was demonstrated by Tanaka et al. and the Meyer group in their pioneering work using Ru(CO) to convert CO2 into carbon monoxide (CO) and formate in MeCN with TEOA as the SD under visible light []. The authors went on to propose a possible mechanism for the photoreduction of CO2 into CO and HCOO, as shown in Figure 5.
Figure 5. Schematic representation of the (a) CO cycle and (b) HCOO cycle for CO2 reduction [].
The CO formation cycle (Figure 5a) involves a two-electron reduction, where the Ru(II) complex, [Ru(CO)2]2+ accepts two electrons, releases CO, and forms an unstable intermediate, [Ru(CO)]0. The [Ru(CO)]0 then undergoes electrophilic attack by CO2, followed by protonation and dehydration, to regenerate the original Ru(II) complex. In the HCOO cycle, product selectivity is controlled by the equilibrium among [Ru(bpy)2(CO)(COO)]+, [Ru(bpy)2(CO)(C(O)OH)]+, [Ru(bpy)2(CO)2]2+. The formation of formate is favored by the reduction of [Ru(bpy)2(CO)(C(O)OH)]+ to [Ru(bpy)2(CO)]0 []. The alternative formate cycle (Figure 5b) begins with the one-electron reduction of [Ru(CO)H]+ to [Ru(CO)H]0, which reacts with CO2 to form [Ru2(CO)(OCHO)]0. Subsequent electron transfer and coordination with an MeCN molecule leads to the release of HCOO and the formation of [Ru(CO)(NCCH3)]0. Final protonation and solvent dissociation regenerate the initial hydride species [,].
The simplest way to characterize these cycles is as competing processes as opposed to simultaneous ones []. While the HCOO cycle (1e) progresses under less reducing potentials or in the presence of H+ donors that stabilize the HCOO intermediate, the CO cycle (2e) typically predominates under higher reducing potentials and in the presence of electron-rich ligands where CO2 binding to the Ru center is favored [].
While the present work does not include detailed density functional theory (DFT) data, previous computational studies suggest that the CO-forming pathway generally has a slightly higher activation barrier but leads to a more stable product under strongly reducing conditions, whereas the HCOO pathway is kinetically more accessible under moderate potentials. For example, a comprehensive DFT analysis of the HCOO, CO and H2, formation pathways by Zhang and colleagues [] suggests that the catalytic process begins at the Ru complex 11-S, with two sequential one-electron reductions and MeCN dissociation producing active species 13 (Figure 6).
Figure 6. CO2 reduction mechanism catalyzed by the Ru CNC pincer complex [].
Protonation by TEOA–H+ gives hydride 13pt′, which reacts with CO2 to form Ru(II)-carboxylate (11-OCHO). Reduction of 11-OCHO by BI(OH)H produces 22-OCHO, releasing -HHCOO- and regenerating 22. Alternatively, 13pt′ can react with TEOA–H+ releasing H2. For CO formation, nucleophilic attack of 13 on CO2 forms 11-CO2, which undergoes two protonations and C4–O2 cleavage to form 11-CO. CO dissociation from 3-CO (+22.2 kcal mol−1) is the rate-determining step (RDS). On the other hand, kinetic analysis shows that the HCOO pathway is favored, with a lower RDS barrier (10.7 kcal mol−1) than CO (27.2 kcal mol−1) and H2 (11.1 kcal mol−1) generation. The minor barrier difference between HCOO and H2 (ΔΔG⧧ = 0.4 kcal mol−1) suggests selectivity can be controlled through catalyst modification.
For CO2 reduction via 1-CO, protonation at bipyridyl N5 forms 1OC-3pte, followed by orthogonal Proton-Coupled Electron Transfer (PCET) to generate HCOO. This pathway exhibits a lower activation barrier (ΔG⧧ = 17.9 kcal mol−1) than CO production. Protonation at bipyridyl N (ΔG = −2.6 kcal mol−1) acts as a thermodynamic sink and kinetic bottleneck, while PCET-mediated HCOO formation (ΔG = −27.7 kcal mol−1) drives product release and catalyst regeneration [].
While similar mechanisms have been proposed in recent studies [,,,], no universally accepted mechanism yet, fully explains the photocatalytic activity and product selectivity of Ru(II) complexes. Additionally, a common limitation of Ru(CO) is its tendency to polymerize into [Ru(CO)]ₙ precipitates, which poorly accept electrons from the PS []. In such cases, the efficiency of CO2 photoreduction can be improved by adding excess PS and SD [].

4. Systematic Approaches to Ru-Driven CO2 Photoreduction

Ru-based photocatalytic systems for CO2 conversion can be divided into three categories: (a) PS/Cat mixed systems, (b) dyad systems, and (c) hybrid or supramolecular catalyst systems, as illustrated in Figure 7.
Figure 7. Photocatalytic CO2 reduction systems: (a) PS/Cat mixture, (b) dyad, and (c) hybrid/supramolecular system [].
In mixed PS/Cat systems, electron transfer (ET) from the PS* to the Cat is often rate-limiting, as it depends on physical contact. Performance can be hindered by PS instability in both PS* and PS•−, particularly in homogenous systems where interactions are limited by diffusion []. Similar ET constraints can potentially appear in heterogeneous systems, depending on redox stability and interfacial charge transfer dynamics. Dyad systems, by contrast, constitute a covalently linked PS and Cat within a single molecule, for example, a binuclear Ru complex, [Ru(bpy)2(bpm)Ru(CO)2Cl2] (RuRu), in which both Ru centers are bridged via a bpm ligand []. Upon photoexcitation, an SD donates an electron to the PS, forming PS•−, which then transfers the electron intramolecularly to the Cat. According to Ishizuka et al. [], the overall efficiency of dyads depends on two successive ET steps: from the donor to the PS* and from PS•− to the catalytic site. On the other hand, supramolecular photocatalysts integrate both PS and Cat units into multinuclear complexes (i.e., MOF-253-Ru(dcbpy)2Ru-(dcbpy)2Cl2) (MOF: metal–organic framework) [], promoting short intramolecular ET distances. This integration improves CO2 photoreduction performance by reducing mass transfer limitations and enhancing ET efficiency, in contrast to systems with physically separated components.

4.1. Photosensitizer/Catalyst Mixed Systems

Table 2 summarizes recent research focused on mixed PS/Cat systems with Ru complexes as PSs for CO2 photoreduction.
Table 2. Metal molecular systems for CO2 photoreduction with Ru complexes as photosensitizers.
From Table 2, CO2 reduction can be performed using either homogeneous or heterogeneous systems and the product distribution appears to be independent of the process type. For example, CO and/or HCOO were produced with high TON and selectivity in homogeneous (entries 1–8, Table 2). While CO was produced in heterogeneous (entries 9–13, Table 2) systems, the corresponding data for TON and selectivity is not available. However, the overall results suggest that product selectivity is probably influenced by the reaction conditions (Cat, PS, SD, solvent). It is important to mention that homogeneous molecular complexes are less attractive for industrial or large-scale applications because of stability, recovery, and scalability challenges. In contrast, heterogeneous systems are more suited for large-scale applications due to their increased photostability, simpler separation, and recyclability [,].
In a typical breakaway from the conventional Ru(II)-based PS, a heteroleptic osmium (II) complex (Os) with two different tridentate ligands was utilized as a PS in the presence of BI(OH)H to reduce CO2 to CO and HCOO at λ > 700 nm (entry 8, Table 2). Since the excited Os complex (Os) can be reductively quenched by BI(OH)H, Os shows strong potential as a panchromatic photosensitizer, exhibiting an extended excited-state lifetime that enhances HCOO production []. However, potential light absorption by other components (e.g., catalysts or substrates) induces side reactions and inner-filter effects, reducing light availability to the PS and lowering efficiency []. Therefore, developing panchromatic redox PS capable of absorbing longer-wavelength light is crucial.
While CO and HCOO were the main CO2 reduction products in most photocatalytic systems, syngas (CO/H2) was produced in some cases with high TON (entries 3, 4, Table 2). Using Ru(bpy)3Cl2 as the PS, the Co(bpy)2Cl2 catalyst demonstrated exceptional stability and high efficiency for photocatalytic CO2 and proton reduction to CO and H2 []. The proposed mechanism of formation of syngas gas is illustrated in Figure 8.
Figure 8. Proposed photocatalytic conversion of CO2 using Co(bpy)2Cl2 in the presence of TEOA [].
A single-state electron transfer quenching pathway was proposed whereby after photoexcitation PS forms PS*, which undergoes oxidative quenching by Co(bpy)2Cl2, generating a Co0 substrate. The Co0 then binds CO2 forming a Co0–CO2 adduct, which is protonated to produce CO and H2O, regenerating Co(bpy)22+. The oxidized PS (PS+) is then reduced by TEOA, completing the catalytic cycle (Figure 7). The mechanism correlates with the reported Cyclic Voltammetry (CV) data of Co(bpy)2Cl2, which indicates that the singlet-state potential (Es(PS*/PS+) = −1.53 V vs. Fc+/Fc) is sufficient to drive electron transfer from the PS to the catalyst, whereas the triplet state (ET(PS*/PS+) = −0.82 V vs. Fc+/Fc) is inadequate [,]. Additionally, catalytic current appeared only after the second reduction wave, suggesting that two successive electron transfers are required to initiate CO2 reduction.
To overcome the limitations of PS/Cat-mixed systems, a more effective approach involves dyads or hybrid/supramolecular complexes that integrate both PS and Cat functions within a single framework. These self-photosensitizing catalysts have gained significant attention, particularly for eliminating the need for external electron transfer [].

4.2. Dyad CO2 Reduction Systems

In self-photosensitizing catalytic systems, performance of molecular complexes for CO2 reduction has been shown to depend on the nature of the bridging ligand that connects the PS to the catalytic unit []. Hence, it is not surprising that recent research focus on catalytic performance has shifted towards developing new and improving or modifying the ligand structure of photocatalysts. To that effect, several dyad systems have been investigated for photocatalytic CO2 reduction under visible light, as shown in Table 3.
Table 3. CO2 photoreduction using Ru(II)-based dyad systems.
While dimers have proven highly effective in CO2 reduction (entries 5, 6, 7), a new class of dyad systems incorporating Ru-pincer complexes as reductive units for CO2 conversion has emerged in recent years. These pincer-based photocatalysts have demonstrated high selectivity and turnover numbers (TONs) under mild conditions (entries 2, 3, 4, Table 3). Notably, incorporating phenyl groups and a bpy coligand into the NHC ligand of the CNC pincer framework (entry 3, Table 3) enhances CO2 photoreduction efficiency []. The increased TON observed at low concentrations is likely due to greater photon availability, while the decrease at higher concentrations may result from reduced light transmittance and photoactivity.
In a related study, a shift from traditional α-diimine ligands such as cis-[Ru(bpy)2(CO)2]2+ to non-diimine complexes was demonstrated in entry 2 (Table 3). Here, the Ru(II) catalyst was supported by N,N′-bis(diphenylphosphino)-2,6-diaminopyridine (PNP) ligands instead of the [Ru(bpy)2(CO)]2+ complex, forming a PNP pincer. Paired with [Ru(bpy)3]2+ as the photosensitizer (PS), the pincer complex achieved 100% selectivity for HCOO using TEOA as the donor (entry 2, Table 3). Pincer-supported Ru complexes possess an electron-rich π-conjugated system (26e) around the metal center, which promotes excellent activity and selectivity for CO2 reduction []. However, photodecomposition of [Ru(bpy)3]2+ was observed in aqueous BNAH systems, where the resulting byproducts facilitated CO2-to-HCOO conversion, as previously reported []. This issue was mitigated by switching to anhydrous solvents (entries 2, 3, 4, Table 3). Figure 9 depicts a plausible mechanism for CO2 photoreduction using a Ru-pincer complex and illustrates CO2 photoreduction in Ru(II)-based dyad systems.
Figure 9. Plausible mechanism for CO2 photoreduction using a Ru–pincer complex [].
Figure 9 depicts the proposed reaction pathway for the Ru–pincer complex, [Ru(k3-{2,6-(Ph2PNMe)2NC5H3})(CO)2Cl]+ (1+). The photocatalytic cycle starts with the reduction of the cationic complex A, which spontaneously loses a Cl to form a square-pyramidal Ru0 species (Figure 9). The subsequent protonation of Ru0 generates the hydride complex Ru–H, which undergoes CO2 insertion to produce a Ru–COOH intermediate. Upon reduction, the intermediate yields formate and forms Ru+, which is then reduced by the PS, regenerating Ru0 and completing the cycle. The liberation of H2 may result from further protonation of Ru–H, dimerization of hydrides, or dehydration of HCOOH. Noteworthy, protonation of Ru–H is difficult to suppress and provides a competing pathway to H2 evolution, acting as a shortcut in the catalytic cycle [].

4.3. Hybrid/Supramolecular Catalyst Systems

Table 4 summarizes CO2 reduction systems employing Ru(II)-based hybrid/supramolecular photocatalysts.
Table 4. CO2 reduction systems employing Ru(II)-based hybrid/supramolecular photocatalysts.
Table 4 shows that Ru supramolecular systems are very effective in CO2 photoreduction, with high TON and product selectivity. For example, MOF-253-Ru(dcbpy)2 (Table 4, entry 1) acts as a bifunctional photocatalyst, facilitating simultaneous CO2 reduction to CO and HCOO, and semidehydrogenation of THIQ to 3,4-dihydroisoquinoline (DHIQ) []. The supramolecular catalyst exhibits superior performance than the Ru-doped MOF-253 (Ru-MOF-253), illustrating the benefits of using open coordination sites for fabricating surface-supported frameworks. The results highlight the potential of MOFs as building blocks for multifunctional Ru supramolecular frameworks, enabling a green and cost-effective pathway for parallel CO2 photoreduction and selective organic transformations.
In a similar study, the supramolecular coordination polymer gel (CPG) with Ru(II), and porphyrin-based tetrapodal gelator (TPY-POR) units (Ru-TPY-POR-CPG) achieved >99% selectivity for CO using TEA as a donor. More than 95% selectivity for methane was obtained when TEA was replaced with BNAH (entry 4, Table 4). In this system, the POR acts as the PS, while covalently attached [Ru(TPY)2]2+ serves as the Cat. The functionalization of POR core with four TPY moieties through the alkyl amide chain allows for additional metal-binding, which acts as a catalytic site for CO2 reduction. Additionally, BNAH’s lower oxidation potential (Eoₒₓ = 0.57 V vs. SCE) compared to TEA (0.69 V), may have supported the donor to be more readily oxidized, facilitating faster charge transfer during photoreduction []. The reduction pathway for supramolecular systems is demonstrated in Figure 10 using the Ru-TPY-POR CPG.
Figure 10. Proposed photocatalytic CO2 reduction pathways to (a) CO and (b) CH4 over Ru-TPY-POR CPG [].
Noteworthy, only the CH4 cycle will be discussed, since CO production follows a similar pathway, as already discussed for other systems. Thus, in the presence of BNAH the conversion of CO into CH4 follows a series of proton-coupled electron transfer steps (VII to X) (Figure 9). The steps result in the formation of key intermediates, including Ru–CH2O, Ru–OCH3, and Ru–CH3, with progressively favorable free energy changes (ΔG = −1.21 to −3.54 eV). The final intermediate, Ru–CH3, generates CH4 via a highly exothermic step (ΔG = −3.29 eV), regenerating the active Ru catalyst and completing the cycle. The outlined pathway demonstrates the crucial role of BNAH in facilitating efficient methane formation under visible light [].
While typical CO2 photoreduction conditions were employed in most studies, Zhang and colleagues [] explored a different approach to synthesize CH4. Using the hybrid catalyst Ru–H bipyridine complex-grafted TiO2 nanohybrids, CH4 was produced from CO2 via CO2 methanation, with H2 acting as a proton and electron donor in the absence of a photosensitizer (entry 6, Table 4). Ligand exchange of surface Cp–RuH complexes with 4,4′-dimethyl-2,2′-bipyridine (4,4′-bpy) significantly enhanced CO2 methanation, increasing selectivity and reaction rate to 93.4% and 241 μL·g−1·h−1, respectively. The CO2•− radicals produced at TiO2 oxygen vacancies reacted with Ru–H to form Ru–OOCH intermediates, which subsequently produced CH4 and H2O in the presence of H2. Selective CH4 production relies on directional proton-coupled electron transfer (PCET) to CO2, stabilizing intermediates such as H2COO*, HCOO*, and H2CO* [].
Another CO2 conversion approach employed cycloaddition with an epoxide cocatalyst (entry 10, Table 4). The reaction produced cyclic carbonates instead of the conventional organic fuels, which is an intriguing deviation from typical CO2 photoreduction methods. The RuPyL2/Mn2Ni4 photocatalytic complex was used, achieving very high catalytic activity under mild conditions, with a CO2 cycloaddition conversion rate of ~99% and a TON of 1723 compared to single Mn2Ni4 (TON = 656) []. The mechanism of CO2 reduction using supramolecular systems is represented in Figure 11 using the RuPyL2/Mn2Ni4 system.
Figure 11. Plausible pathway of CO2 cycloaddition with epoxides to form cyclic carbonates [].
Figure 11 demonstrates a plausible reaction pathway with coordination synergism for CO2 cycloaddition using the RuPyL2/Mn2Ni4. Upon light exposure, RuPyL2/Mn2Ni4 promotes electron transfer to the epoxide, activating the molecule through coordination interactions. This is followed by a nucleophilic attack and opening of the epoxy ring. Subsequently, the oxygen anion from the ring-opening intermediate attacks CO2, producing a bromocarbonate intermediate. The closure of the ring forms cyclic carbonates and the catalyst is regenerated. The synergistic effects of photo- electrons, coordination bonds that promote electron transfer, and bromide ions (Br) acting as nucleophiles to rupture the epoxide ring collectively facilitate the effective progress of cycloaddition reactions [].

4.4. Artificial Photosynthesis Systems

Mimicking natural photosynthesis using metal molecular complexes presents a promising strategy to overcome the high energy barrier of CO2 reduction and advance next-generation renewable energy solutions. Artificial photosynthetic systems typically integrate a chromophore or PS for light absorption and charge separation with a catalytic center capable of facilitating multielectron CO2 reduction []. Analogous to natural photosynthesis in plants, CO2 is reduced by electrons and protons, often derived from water, resulting in the generation of oxygen (O2) and energy rich-chemical products, all under ambient temperature and pressure [,]. Some examples of artificial photosynthesis systems used for the photocatalytic conversion of CO2 are given in Table 5.
Table 5. CO2 photoreduction using Ru-based artificial photosynthetic systems.
While highly efficient and selective molecular systems have been reported in organic media [], achieving photocatalytic CO2 reduction in aqueous media is more appealing for practical applications due to easier handling and coordination with artificial photosynthesis objectives. This perspective is reflected in Table 5, where most of the studies focused on artificial photosynthetic systems in aqueous media. Entry 3, (Table 5) demonstrates artificial photosynthetic conversion of CO2 into methanol using water as the sacrificial electron donor. The proposed mechanism for this reaction was selected to represent CO2 photoreduction using Ru-based artificial photosynthetic systems (Figure 12).
Figure 12. Proposed mechanism of CO2 photoreduction on the RuSA-mC3N4 catalyst surface [].
In the hybrid catalyst, RuSA–mC3N4, Ru atoms are anchored via Ru–C/N and possibly Ru–O bonds. During aqueous CO2-to-HCOO reduction with water as the electron donor, these coordination sites may bridge electron transfer and increase charge density on Ru, reducing the photocarrier transfer barrier and enhancing photocatalytic efficiency. Although the role of Ru–O moieties remain unclear, the atomic dispersion of Ru strengthens Ru–C/N interactions, promoting charge separation and suppressing electron–hole recombination. Consequently, RuSA–mC3N4 shows improved photocatalytic activity compared to pristine C3N4 [].
Noteworthy, in aqueous systems, water can act as the electron donor, enabling net energy production. However, the proton (H+)-rich environment poses a major setback, as it facilitates competitive H+ reduction, often resulting in H2 generation instead of CO []. To address this challenge, semiconductor-based molecular catalysts have been employed to promote CO2 photoreduction, with their band gaps and energy levels adjustable through crystal structure, elemental composition, and ligand ratios. In particular, Ru complexes with bpy ligands have demonstrated promising CO2 reduction efficiency when integrated into various semiconductors, including doped tantalum oxide (Ta2O5), paving the way for more efficient and selective aqueous-phase systems.
A recent innovative advancement in semiconductor-based molecular CO2 photoreduction was reported by Morikawa et al. [], who developed a monolithic “artificial leaf” by integrating N-Ta2O5 with [Ru(dcbpy)2(CO)2]2+ (entry 4, Table 5). Leveraging the low overpotential of CO2 reduction at metal complexes, this single-device system efficiently produces formate from CO2 and H2O in a membrane-free, bias-free reactor, achieving a solar-to-chemical conversion efficiency of 4.6%, surpassing that of natural photosynthesis. The CO2 reduction rate increased with the number of dicarboxylic acid anchors, attributed to enhanced electron transfer from the conduction band of N-Ta2O5 to the Ru complex. Due to stronger nonadiabatic coupling, the COOH anchoring groups facilitate faster electron transfer (7.5 ps) than PO3H2 groups (56.7 ps), highlighting the importance of anchoring group selection in optimizing the performance of CO2 photoreduction systems []. Figure 13 illustrates a typical energy level diagram of hybrid photocatalysis with a semiconductor and a metal-complex photocatalyst.
Figure 13. Energy diagram of hybrid photocatalysis using a semiconductor-metal complex [].
Upon irradiation, the semiconductor generates charge carriers, electrons/holes (e/h+). The photoexcited electrons migrate from the semiconductor CB to the metal complex for selective CO2 reduction at the metal-complex catalyst (Figure 13). The CB electrons drive the CO2 reduction reaction, while the photogenerated holes are neutralized by an electron donor (ED) []. Efficient electron transfer requires that the CB minimum of the semiconductor is more negative than the LUMO of the metal complex or its CO2 reduction potential. Additionally, the development of intermediate bands (IB) between the VB and CB under photoirradiation facilitates the redshift frequently seen in Ru-doped semiconductors. These new energy levels have two functions: they facilitate photoactivation during excitation and may also be charge carrier recombination sites [].
The problem of competing H2 evolution in aqueous media has also been reported in conjugated polymer photocatalysts, where residual palladium from synthesis can further promote proton reduction. Thus, redirecting highly active polymer photocatalysts toward CO2 reduction instead of proton reduction remains a key yet unresolved challenge. In their attempt to provide a solution, Sakakibara et al. [] developed a supramolecular photocatalyst (RuRu′/Ag/P10), comprising a Ag-loaded conjugated polymer and a binuclear Ru(II) complex. This system displayed high catalytic activity and durability for formate production (entry 7, Table 5). While residual Pd in the polymer supported H2 evolution in aqueous media, Ag nanoparticle loading enhanced CO2 reduction selectivity by effectively suppressing H2 generation. Functioning as a semiconductor, the RuRu′/Ag/P10 framework facilitates sequential visible-light photoexcitation, boosting oxidation power and enabling the use of mild reductants such as water. Mimicking a Z-scheme in natural photosynthesis, the system drives selective CO2 reduction via an efficient electron cascade [].

6. Effect of Parameters

Ru(II) complexes can convert CO2 into CO, HCOO, or CH4, with product distribution and selectivity strongly influenced by the ligand environment and reaction conditions.

6.1. Effect of Donors

Recent reports demonstrated that selectivity and efficiency of the CO2 photoreduction could be modulated with electron donors []. Several authors have investigated the effect of sacrificial donors on CO2 photoreduction with varying results [,,,]. Most recently, Sakakibara et al. [] investigated the effect of MeCN and different electron donors on CO2 reduction using the RuRu′/Ag/P10 system under 5 h of visible light irradiation (entry 7, Table 5). Among the tested SDs, TEA in H2O/MeCN exhibited the highest formate yield, TON, and selectivity, implying optimal electron transfer and catalyst activity. In contrast, TEOA and ASc showed moderate to low formate yield, while no CO2 reduction products were observed with EDTA·2Na in the absence of MeCN. TEA in MeCN/H2O emerged as the most effective condition, which contradicts earlier findings by Hammed et al. [] using a PNP pincer system. In the latter, substituting DMF with MeCN reduced activity, while other donors such as ASc, NaASc, and TEA showed poor performance. Overall, the results suggest that the SD type and solvent environment have a significant effect on the efficiency and selectivity of CO2 photoreduction, highlighting the importance of controlling these factors to suppress competitive H2 evolution and enhance formate selectivity.
In another study, the homobimetallic Cu and Co bisquaterpyridine catalysts paired with Ru(phen)32+ showed that full catalytic performance was only achieved with a specific combination of BIH, TEOA, and H2O. In contrast, monometallic systems were less effective, highlighting the importance of cooperative metal centers for multielectron transfer processes []. Similarly, the Ru-TPY-POR-CPG system achieved variable selectivity, favoring CO in the presence of TEA and shifting to CH4 with BNAH []. The contrast in product distribution can be explained by the oxidation potentials of the donors, with BNAH enabling faster electron transfer and deeper reduction pathways. While the Co(II) complexes and the CPG frameworks benefit from well-designed ligand environments and sacrificial donors, the CPG system additionally leverages structural integration and supramolecular assembly for improved charge separation and metal site availability. These results align with the afore mentioned findings of [], demonstrating the pivotal role of catalyst structure, PS, and ED in determining the efficiency and product selectivity of CO2 photoreduction.

6.2. Effect of Solvent

The reaction medium critically influences the efficiency and product distribution in CO2 photoreduction. Solvent properties can lower the activation energy for CO2 reduction and affect selectivity by modulating the adsorption of the CO2 intermediate on catalyst surfaces []. In low-polarity solvents, CO2 binds more strongly via the carbon atom, favoring CO and formate production []. High-dielectric solvents, however, better stabilize CO2 through solvation, weakening its catalyst interaction and promoting formate production. In metal-based systems, solvent characteristics also govern the formation and stability of M-CO2 intermediates, further emphasizing their role in shaping catalytic pathways. Table 6 summarizes findings by [] on solvent effects during 2 h of visible-light-driven CO2 reduction.
Table 6. Effect of solvents on photocatalytic CO2 reduction (adopted and modified from [].
As shown in Table 6, photocatalytic CO2 reduction is highly influenced by the reaction solvent. Aprotic solvents such as MeCN and DMF promote higher CO production, with yields correlating closely to CO2 solubility rather than solvent viscosity. However, the MeCN results contradict the later findings of Sakakibara et al. [], where the removal of MeCN from the MeCN/TEOA/H2O solution was found to enhance CO production, presumably due to improved dispersibility of the photosensitizer and increased stability of the Ru complex. On the other hand, DMSO and H2O show poor performance due to low CO2 solubility or unfavorable coordination environments. Notably, THF achieves the highest CO selectivity (90.5%) despite moderate CO yield, possibly due to its strong coordination ability with the Co-bpy complex. These results emphasize the important role of solvent viscosity and CO2 solubility in boosting product yield and suppressing competing H2 evolution. However, in a recent study, Das et al. [] cautioned that organic solvents such as ethyl acetate (EtOAc), CH3CN, TEOA and TEA can undergo photolysis upon UV-Visible light exposure, producing CO, CH4, C2H4, and H2 even without a catalyst. This may result in overstated catalyst performance and product selectivity. Therefore, careful selection of solvents and SDs is crucial to avoid misinterpretation in photocatalytic CO2 reduction studies.

6.3. Effect of Substituents Groups/Ligands

Catalytic activity in CO2 photoreduction strongly depends on the catalyst’s structure, particularly the coordination environment of the metal center []. Improving performance requires strategic selection and modification of metal centers and ligands, as structural changes to primary ligands affect the properties and efficiency of metal complexes. In their related studies, Kuramochi et al. [,] reported that TOFs were significantly affected by the type and position of amide groups and other substituents on the bipyridyl ligand in trans-(Cl)-[Ru(bpy)(CO)2Cl2]-type complexes. Despite differences in hydrophilicity and molecular size, reaction rates correlated consistently with the first reduction potential, offering a useful design strategy for Ru-based catalysts. Furthermore, the molecular ratio of Ru(PS) to Ru(Cat) influenced product selectivity between CO and HCOO. Similar findings were reported in other studies [,].
Additionally, in metal-based CO2 reduction systems, the adsorption strength of CO2 on the catalyst surface has a profound effect on product selectivity. For example, HCOO is the major product in the RuC/mpg-C3N4 (mpg-C3N4: mesoporous graphitic carbon nitride; C: COOH) or RuP/mpg-C3N4 systems (P: PO3H2), regardless of Ru(II) complex loading or solvent. On the contrary, CO selectivity increased markedly (40–70%) when mpg-C3N4 loaded with higher amounts of RuCP (CP: CH2CH2PO3H2) was used in solvents with high donor numbers (i.e., DMA, DMF, DMSO) and TEOA as the electron donor []. The latter indicates that substrate concentration affects process reactivity, which agrees with the earlier findings of Kuramochi et al. []. The effect of ligand substituents was further demonstrated using dinuclear nickel(II)-bipyridine [], and cobalt (II) tripodal complexes [].
On the other hand, Hameed et al. [] showed that the choice of metal center has a significant influence on product distribution in CO2 reduction. Bidentate N-(diphenylphosphino)-2-aminopyridine ligands ({Ph2PNR}NC5H4; R=H, CH3) were studied with Re(I) and Mn(I) complexes in DMF using [Ru(bpy)3](PF6)2 PS and TEOA donor under visible light. The Mn complex [Mn{κ2-(Ph2P)NH-(NC5H4)}(CO)3Br], (1) produced CO with a TON of 55, >99% selectivity, and an AQY of 0.75. Replacing Mn with Re (complex (2)) shifted the product to HCOOH with a TON of 343, >96% selectivity, and an AQY of 4.7. This shift in reactivity demonstrates the critical role of CO2 coordination: in complex (1), CO2 probably binds via carbon, resulting in O-protonation and CO formation, whereas in complex (2), O-bound CO2 undergoes C-protonation, favoring HCOOH production. Br- addition further increased TON, supporting the hypothesis that Br− remains coordinated after the second electron transfer, facilitating CO or HCOOH generation [].
The effect of Lewis acids on visible-light-driven CO2 reduction have also been investigated in a previous study using Ni(II) complexes with pyridine pendants []. The enhanced selective CO generation was probably due to the formation of a Mg2+-bound Ni complex, which strengthened cooperative interactions between the Ni and Mg centers, with the Lewis acidic Mg2+ stabilizing the Ni-CO2 intermediate [].

7. Limitations and Perspectives

Ru(II)-based molecular photocatalysts for CO2 reduction have made great strides in recent years, but a number of drawbacks still prevent widespread use. The instability of essential components, such as the Cat and PS, is a major problem, particularly in mixed systems where diffusion limitations restrict electron transfer because of the physical separation of PS and Cat. Catalytic turnover and overall efficiency are also compromised by the PS’s proneness to deterioration in both its excited and reduced states. Long-term stability and recyclability are further restricted by catalyst deactivation via photolysis, precipitation (such as the polymerization of Ru(CO) into inactive species), or ligand dissociation. For example, it has been demonstrated that [Ru(CO)]n aggregates prevent effective electron transfer from the PS, requiring the use of SDs and extra PS.
The choice of solvent and SDs is another crucial factor. Even without a catalyst, organic solvents like MeCN and EtOAc and EDs like TEOA and TEA can undergo photo decomposition under UV-visible light to produce CO, CH4, or H2. This could lead to overstated product selectivity and catalyst efficiency. Furthermore, the production and stability of M-CO2 intermediates are directly affected by the solvent’s dielectric characteristics and CO2 solubility, impacting both the reaction rate and product selectivity. The limited product selectivity of many systems often requires subtle modifications to the ligand structure, SD type and metal coordination. While CO and HCOO are major products, further reduction to other hydrocarbons such as CH4 remains challenging and often necessitates careful reaction parameter control.
Although dyad and supramolecular catalysts provide enhanced electron transfer through intramolecular pathways, their development can be complex and expensive from a design standpoint. Additionally, high selectivity toward CO2 reduction products is impeded by competitive side reactions such as H2 evolution, which are more prevalent in aqueous conditions.
Developing more resilient and scalable systems in the future will necessitate: (i) designing Cat and PS units with improved photochemical and electrochemical stability; (ii) designing ligand assemblies that facilitate effective and selective multi-electron transfer; (iii) implementing SDs that are photostable and water-tolerant; and (iv) investigating earth-abundant metal alternatives to minimize costs and increase sustainability. New approaches that could speed up these initiatives and assist in overcoming present obstacles in photocatalytic CO2 conversion include single-site hybrid photocatalysts, nanostructured supports, and machine learning-driven molecular designs.

Author Contributions

Conceptualization, writing—original draft preparation, writing—review and editing, P.N.; review and editing, M.J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors are grateful to the University of South Africa (UNISA). In particular, the Department of Chemistry.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kumar, A.; Kumar, P.; Aathira, M.S.; Singh, D.P.; Behera, B.; Jain, S.L. A Bridged Ruthenium Dimer (Ru–Ru) for Photoreduction of CO2 under Visible Light Irradiation. J. Ind. Eng. Chem. 2018, 61, 381–387. [Google Scholar] [CrossRef]
  2. Barrocas, B.; Oliveira, M.C.; Nogueira, H.I.S.; Fateixa, S.; Monteiro, O.C. A comparative study on emergent pollutants photo-assisted degradation using ruthenium modified titanate nanotubes and nanowires as catalysts. J. Environ. Sci. 2020, 92, 38–51. [Google Scholar] [CrossRef]
  3. Fang, S.; Rahaman, M.; Bharti, J.; Reisner, E.; Robert, M.; Ozin, G.A.; Hu, Y.H. Photocatalytic CO2 reduction. Nat. Rev. Methods Primers 2023, 3, 61. [Google Scholar] [CrossRef]
  4. Barrocas, B.T.; Přech, J.; Filip Edelmannová, M.; Szaniawska, E.; Kočí, K.; Čejka, J. Titanosilicates enhance carbon dioxide photocatalytic reduction. Appl. Mater. Today 2022, 26, 101392. [Google Scholar] [CrossRef]
  5. Gu, J.; Chen, W.; Shan, G.G.; Li, G.; Sun, C.; Wang, X.L.; Su, Z. The Roles of Polyoxometalates in Photocatalytic Reduction of Carbon Dioxide. Mater. Today Energy 2021, 21, 100760. [Google Scholar] [CrossRef]
  6. Kuriki, R.; Maeda, K. Development of Hybrid Photocatalysts Constructed with a Metal Complex and Graphitic Carbon Nitride for Visible-Light-Driven CO2 Reduction. Phys. Chem. Chem. Phys. 2017, 19, 4938–4950. [Google Scholar] [CrossRef]
  7. Wang, Q.; Fang, Z.; Zhang, W.; Zhang, D. High-Efficiency g-C3N4 Based Photocatalysts for CO2 Reduction: Modification Methods. Adv. Fiber Mater. 2022, 4, 342–360. [Google Scholar] [CrossRef]
  8. Yoshino, S.; Takayama, T.; Yamaguchi, Y.; Iwase, A.; Kudo, A. CO2 Reduction Using Water as an Electron Donor over Heterogeneous Photocatalysts Aiming at Artificial Photosynthesis. Acc. Chem. Res. 2022, 55, 966–977. [Google Scholar] [CrossRef] [PubMed]
  9. Samanta, S.; Srivastava, R. Catalytic Conversion of CO2 to Chemicals and Fuels: The Collective Thermocatalytic/Photocatalytic/Electrocatalytic Approach with Graphitic Carbon Nitride. Mater. Adv. 2020, 1, 1506–1545. [Google Scholar] [CrossRef]
  10. Maeda, K. Metal-Complex/Semiconductor Hybrid Photocatalysts and Photoelectrodes for CO2 Reduction Driven by Visible Light. Adv. Mater. 2019, 31, 1808205. [Google Scholar] [CrossRef]
  11. Pirzada, B.M.; Dar, A.H.; Shaikh, M.N.; Qurashi, A. Reticular-Chemistry-Inspired Supramolecule Design as a Tool to Achieve Efficient Photocatalysts for CO2 Reduction. ACS Omega 2021, 6, 29291–29324. [Google Scholar] [CrossRef]
  12. Chen, H.; Chen, L.; Chen, G.; Robert, M.; Lau, T.C. Electrocatalytic and Photocatalytic Reduction of Carbon Dioxide by Earth-Abundant Bimetallic Molecular Catalysts. ChemPhysChem 2021, 22, 1835–1843. [Google Scholar] [CrossRef] [PubMed]
  13. Kuramochi, Y.; Itabashi, J.; Toyama, M.; Ishida, H. Photochemical CO2 Reduction Catalyzed by Trans(Cl)-[Ru(2,2′-Bipyridine)(CO)2Cl2] Bearing Two Methyl Groups at 4,4′-, 5,5′- or 6,6′-Positions in the Ligand. ChemPhotoChem 2018, 2, 314–322. [Google Scholar] [CrossRef]
  14. 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. [Google Scholar] [CrossRef]
  15. 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. [Google Scholar] [CrossRef] [PubMed]
  16. Fujita, E.; Grills, D.C.; Manbeck, G.F.; Polyansky, D.E. Understanding the Role of Inter- and Intramolecular Promoters in Electro- and Photochemical CO2 Reduction Using Mn, Re, and Ru Catalysts. Acc. Chem. Res. 2022, 55, 616–628. [Google Scholar] [CrossRef]
  17. Sampaio, R.N.; Grills, D.C.; Polyansky, D.E.; Szalda, D.J.; Fujita, E. Unexpected Roles of Triethanolamine in the Photochemical Reduction of CO2 to Formate by Ruthenium Complexes. J. Am. Chem. Soc. 2020, 142, 2413–2428. [Google Scholar] [CrossRef]
  18. Kumagai, H.; Tamaki, Y.; Ishitani, O. Photocatalytic Systems for CO2 Reduction: Metal-Complex Photocatalysts and Their Hybrids with Photofunctional Solid Materials. Acc. Chem. Res. 2022, 55, 978–990. [Google Scholar] [CrossRef]
  19. Irikura, M.; Tamaki, Y.; Ishitani, O. Development of a Panchromatic Photosensitizer and Its Application to Photocatalytic CO2 reduction. Chem. Sci. 2021, 12, 13888–13896. [Google Scholar] [CrossRef]
  20. Deng, X.; Qin, Y.; Hao, M.; Li, Z. MOF-253-Supported Ru Complex for Photocatalytic CO2 Reduction by Coupling with Semidehydrogenation of 1,2,3,4-Tetrahydroisoquinoline (THIQ). Inorg. Chem. 2019, 58, 16574–16580. [Google Scholar] [CrossRef] [PubMed]
  21. 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. [Google Scholar] [CrossRef]
  22. Livingstone, S.E. The Chemistry of Ruthenium, Rhodium, Palladium, Osmium, Iridium and Platinum, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 1–122. [Google Scholar]
  23. Hameed, Y.; Gabidullin, B.; Richeson, D. Photocatalytic CO2 Reduction with Manganese Complexes Bearing a Κ2-PN Lig and: Breaking the α-Diimine Hold on Group 7 Catalysts and Switching Selectivity. Inorg. Chem. 2018, 57, 13092–13096. [Google Scholar] [CrossRef]
  24. 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. [Google Scholar] [CrossRef]
  25. Liao, W.M.; Zhang, J.H.; Hou, Y.J.; Wang, H.P.; Pan, M. Visible-Light-Driven CO2 Photo-Catalytic Reduction of Ru(II) and Ir(III) Coordination Complexes. Inorg. Chem. Commun. 2016, 73, 80–89. [Google Scholar] [CrossRef]
  26. Huang, A.; Zhou, T.; Zhang, J.; Zhang, Y.; Wu, Y.; Wang, Y.; Luo, W. Competing CO and HCOOH Pathways in CO2 Electroreduction. ChemCatChem 2024, 16, e202400504. [Google Scholar] [CrossRef]
  27. Ishida, H. Electrochemical/Photochemical CO2 Reduction Catalyzed by Transition Metal Complexes. In Carbon Dioxide Chemistry, Capture and Oil Recovery; Karamé, I., Shaya, J., Srour, H., Eds.; IntechOpen: Rijeka, Croatia, 2018. [Google Scholar]
  28. Zhang, Y.-Q.; Wu, T.; Zhang, Y.; Liao, R.-Z. Theoretical Study on Photocatalytic CO2 Reduction to Formate by a Ruthenium CNC Pincer Complex. J. Phys. Chem. A 2025, 129, 8357–8369. [Google Scholar] [CrossRef]
  29. Bharti, J.; Chen, L.; Guo, Z.; Cheng, L.; Wellauer, J.; Wenger, O.S.; von Wolff, N.; Lau, K.C.; Lau, T.C.; Chen, G.; et al. Visible-Light-Driven CO2 Reduction with Homobimetallic Complexes. Cooperativity between Metals and Activation of Different Pathways. J. Am. Chem. Soc. 2023, 145, 25195–25202. [Google Scholar] [CrossRef]
  30. Ma, K.; Li, J.; Liu, J.; Li, C.; Chen, X.B.; Li, Z.; Wang, L.; Shi, Z.; Feng, S. Covalent Triazine Framework Featuring Single Electron Co2+ Centered in Intact Porphyrin Units for Efficient CO2 Photoreduction. Appl. Surf. Sci. 2023, 629, 157453. [Google Scholar] [CrossRef]
  31. Xiong, Z.; Huang, L.; Peng, J.; Hou, Y.; Ding, Z.; Wang, S. Spinel-Type Mixed Metal Sulfide NiCo2S4 for Efficient Photocatalytic Reduction of CO2 with Visible Light. ChemCatChem 2019, 11, 5513–5518. [Google Scholar] [CrossRef]
  32. Ishizuka, T.; Hosokawa, A.; Kawanishi, T.; Kotani, H.; Zhi, Y.; Kojima, T. Self-Photosensitizing Dinuclear Ruthenium Catalyst for CO2 Reduction to CO. J. Am. Chem. Soc. 2023, 145, 23196–23204. [Google Scholar] [CrossRef]
  33. Li, N.X.; Chen, Y.M.; Xu, Q.Q.; Mu, W.H. Photocatalytic Reduction of CO2 to CO Using Nickel(II)-Bipyridine Complexes with Different Substituent Groups as Catalysts. J. CO2 Util. 2023, 68, 102385. [Google Scholar] [CrossRef]
  34. Hong, D.; Kawanishi, T.; Tsukakoshi, Y.; Kotani, H.; Ishizuka, T.; Kojima, T. Efficient Photocatalytic CO2 Reduction by a Ni(II) Complex Having Pyridine Pendants through Capturing a Mg2+ Ion as a Lewis-Acid Cocatalyst. J. Am. Chem. Soc. 2019, 141, 20309–20317. [Google Scholar] [CrossRef]
  35. Yao, Y.; Gao, Y.; Ye, L.; Chen, H.; Sun, L. Highly Efficient Photocatalytic Reduction of CO2 and H2O to CO and H2 with a Cobalt Bipyridyl Complex. J. Energy Chem. 2018, 27, 502–506. [Google Scholar] [CrossRef]
  36. Sun, M.; Wang, C.; Sun, C.Y.; Zhang, M.; Wang, X.L.; Su, Z.M. Ultra Stable Multinuclear Metal Complexes as Homogeneous Catalysts for Visible-Light Driven Syngas Production from Pure and Diluted CO2. J. Catal. 2020, 385, 70–75. [Google Scholar] [CrossRef]
  37. Wang, J.W.; Huang, H.H.; Sun, J.K.; Ouyang, T.; Zhong, D.C.; Lu, T.B. Electrocatalytic and Photocatalytic Reduction of CO2 to CO by Cobalt(II) Tripodal Complexes: Low Overpotentials, High Efficiency and Selectivity. ChemSusChem 2018, 11, 1025–1031. [Google Scholar] [CrossRef] [PubMed]
  38. Qin, Y.; Chen, L.; Chen, G.; Guo, Z.; Wang, L.; Fan, H.; Robert, M.; Lau, T.C. A Highly Active and Robust Iron Quinquepyridine Complex for Photocatalytic CO2 reduction in Aqueous Acetonitrile Solution. Chem. Commun. 2020, 56, 6249–6252. [Google Scholar] [CrossRef]
  39. Dong, Y.L.; Liu, H.R.; Wang, S.M.; Guan, G.W.; Yang, Q.Y. Immobilizing Isatin-Schiff Base Complexes in NH2-UiO-66 for Highly Photocatalytic CO2 Reduction. ACS Catal. 2023, 13, 2547–2554. [Google Scholar] [CrossRef]
  40. Zhong, W.; Sa, R.; Li, L.; He, Y.; Li, L.; Bi, J.; Zhuang, Z.; Yu, Y.; Zou, Z. A Covalent Organic Framework Bearing Single Ni Sites as a Synergistic Photocatalyst for Selective Photoreduction of CO2 to CO. J. Am. Chem. Soc. 2019, 141, 7615–7621. [Google Scholar] [CrossRef]
  41. Li, M.; Fu, Y.; You, S.; Yang, Y.; Qin, C.; Zhao, L.; Su, Z. Hexanuclear Nickel-Based [P4Mo11O50] with Photocatalytic Reduction of CO2 Activity. Inorg. Chem. Commun. 2021, 134, 109009. [Google Scholar] [CrossRef]
  42. Wang, Y.; Liu, T.; Chen, L.; Chao, D. Water-Assisted Highly Efficient Photocatalytic Reduction of CO2 to CO with Noble Metal-Free Bis(Terpyridine)Iron(II) Complexes and an Organic Photosensitizer. Inorg. Chem. 2021, 60, 5590–5597. [Google Scholar] [CrossRef]
  43. 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. [Google Scholar] [CrossRef]
  44. Kuramochi, Y.; Fukaya, K.; Yoshida, M.; Ishida, H. Trans-(Cl)-[Ru(5,5′-Diamide-2,2′-Bipyridine)(CO)2Cl2]: Synthesis, Structure, and Photocatalytic CO2 Reduction Activity. Chem. Eur. J. 2015, 21, 10049–10060. [Google Scholar] [CrossRef]
  45. Hameed, Y.; Rao, G.K.; Ovens, J.S.; Gabidullin, B.; Richeson, D. Visible-Light Photocatalytic Reduction of CO2 to Formic Acid with a Ru Catalyst Supported by N,N″-Bis(Diphenylphosphino)-2,6-Diaminopyridine Ligands. ChemSusChem 2019, 12, 3453–3457. [Google Scholar] [CrossRef]
  46. Das, S.; Rodrigues, R.R.; Lamb, R.W.; Qu, F.; Reinheimer, E.; Boudreaux, C.M.; Webster, C.E.; Delcamp, J.H.; Papish, E.T. Highly Active Ruthenium CNC Pincer Photocatalysts for Visible-Light-Driven Carbon Dioxide Reduction. Inorg. Chem. 2019, 58, 8012–8020. [Google Scholar] [CrossRef]
  47. Chen, S.; Kong, P.; Niu, H.; Liu, H.; Wang, X.; Zhang, J.; Li, R.; Guo, Y.; Peng, T. Co-Porphyrin/Ru-Pincer Complex Coupled Polymer with Z-Scheme Molecular Junctions and Dual Single-Atom Sites for Visible Light-Responsive CO2 Reduction. Chem. Eng. J. 2022, 431, 133357. [Google Scholar] [CrossRef]
  48. Kuramochi, Y.; Sekine, M.; Kitamura, K.; Maegawa, Y.; Goto, Y.; Shirai, S.; Inagaki, S.; Ishida, H. Photocatalytic CO2 Reduction by Periodic Mesoporous Organosilica (PMO) Containing Two Different Ruthenium Complexes as Photosensitizing and Catalytic Sites. Chem. Eur. J. 2017, 23, 10301–10309. [Google Scholar] [CrossRef] [PubMed]
  49. Benseghir, Y.; Solé-Daura, A.; Mialane, P.; Marrot, J.; Dalecky, L.; Béchu, S.; Frégnaux, M.; Gomez-Mingot, M.; Fontecave, M.; Mellot-Draznieks, C.; et al. Understanding the Photocatalytic Reduction of CO2 with Heterometallic Molybdenum(V) Phosphate Polyoxometalates in Aqueous Media. ACS Catal. 2022, 12, 453–464. [Google Scholar] [CrossRef]
  50. Elcheikh Mahmoud, M.; Audi, H.; Assoud, A.; Ghaddar, T.H.; Hmadeh, M. Metal-Organic Framework Photocatalyst Incorporating Bis(4′-(4-Carboxyphenyl)-Terpyridine)Ruthenium(II) for Visible-Light-Driven Carbon Dioxide Reduction. J. Am. Chem. Soc. 2019, 141, 7115–7121. [Google Scholar] [CrossRef]
  51. Wang, J.; Li, X.; Chang, C.H.; Zhang, T.; Guan, X.; Liu, Q.; Zhang, L.; Wen, P.; Tang, I.; Zhang, Y.; et al. Engineering Single Ni Sites on 3D Cage-like Cucurbit[n]Uril Ligands for Efficient and Selective CO2 Photocatalytic Reduction. Angew. Chem. Int. Ed. 2025, 64, 202417384. [Google Scholar] [CrossRef]
  52. Verma, P.; Rahimi, F.A.; Samanta, D.; Kundu, A.; Dasgupta, J.; Maji, T.K. Visible-Light-Driven Photocatalytic CO2 Reduction to CO/CH4 Using a Metal–Organic “Soft” Coordination Polymer Gel. Angew. Chem. Int. Ed. 2022, 61, e202116094. [Google Scholar] [CrossRef]
  53. 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. [Google Scholar] [CrossRef]
  54. Zhang, P.; Sui, X.; Wang, Y.; Wang, Z.; Zhao, J.; Wen, N.; Chen, H.; Huang, H.; Zhang, Z.; Yuan, R.; et al. Surface Ru-H Bipyridine Complexes-Grafted TiO2 Nanohybrids for Efficient Photocatalytic CO2 Methanation. J. Am. Chem. Soc. 2023, 145, 5769–5777. [Google Scholar] [CrossRef] [PubMed]
  55. Guo, K.; Zhu, X.; Peng, L.; Fu, Y.; Ma, R.; Lu, X.; Zhang, F.; Zhu, W.; Fan, M. Boosting Photocatalytic CO2 Reduction over a Covalent Organic Framework Decorated with Ruthenium Nanoparticles. Chem. Eng. J. 2021, 405, 127011. [Google Scholar] [CrossRef]
  56. Kumar, A.; Ananthakrishnan, R. Visible Light-Assisted Reduction of CO2 into Formaldehyde by Heteroleptic Ruthenium Metal Complex-TiO2 Hybrids in an Aqueous Medium. Green Chem. 2020, 22, 1650–1661. [Google Scholar] [CrossRef]
  57. Maeda, K.; An, D.; Kuriki, R.; Lu, D.; Ishitani, O. Graphitic Carbon Nitride Prepared from Urea as a Photocatalyst for Visible-Light Carbon Dioxide Reduction with the Aid of a Mononuclear Ruthenium(II) Complex. Beilstein J. Org. Chem. 2018, 14, 1806–1812. [Google Scholar] [CrossRef] [PubMed]
  58. Wei, Y.; Du, L.; Gao, H. Coordination Synergy between Metal Nanoclusters and Ruthenium Photosensitizers for Photocatalytic Cycloaddition of CO2 under Mild Conditions. Mol. Catal. 2025, 577, 114977. [Google Scholar] [CrossRef]
  59. Saito, D.; Tamaki, Y.; Ishitani, O. Photocatalysis of CO2 Reduction by a Ru(II)-Ru(II) Supramolecular Catalyst Adsorbed on Al2O3. ACS Catal. 2023, 13, 4376–4383. [Google Scholar] [CrossRef]
  60. Muraoka, K.; Uchiyama, T.; Lu, D.; Uchimoto, Y.; Ishitani, O.; Maeda, K. A Visible-Light-Driven Z-Scheme CO2 Reduction System Using Ta3N5 and a Ru(II) Binuclear Complex. Bull. Chem. Soc. Jpn. 2019, 92, 124–126. [Google Scholar] [CrossRef]
  61. Morikawa, T.; Sato, S.; Sekizawa, K.; Suzuki, T.M.; Arai, T. Solar-Driven CO2 Reduction Using a Semiconductor/Molecule Hybrid Photosystem: From Photocatalysts to a Monolithic Artificial Leaf. Acc. Chem. Res. 2022, 55, 933–943. [Google Scholar] [CrossRef]
  62. Meng, X.; Li, R.; Yang, J.; Xu, S.; Zhang, C.; You, K.; Ma, B.; Guan, H.; Ding, Y. Hexanuclear Ring Cobalt Complex for Photochemical CO2 to CO Conversion. Chin. J. Catal. 2022, 43, 2414–2424. [Google Scholar] [CrossRef]
  63. Qin, J.; Lin, L.; Wang, X. A Perovskite Oxide LaCoO3 Cocatalyst for Efficient Photocatalytic Reduction of CO2 with Visible Light. Chem. Commun. 2018, 54, 2272–2275. [Google Scholar] [CrossRef] [PubMed]
  64. Sharma, P.; Kumar, S.; Tomanec, O.; Petr, M.; Zhu Chen, J.; Miller, J.T.; Varma, R.S.; Gawande, M.B.; Zbořil, R. Carbon Nitride-Based Ruthenium Single Atom Photocatalyst for CO2 Reduction to Methanol. Small 2021, 17, 2006478. [Google Scholar] [CrossRef] [PubMed]
  65. Zhao, Y.; Kim, S.; Eom, Y.K.; Valandro, S.R.; Schanze, K.S. Polymer Chromophore-Catalyst Assembly for Photocatalytic CO2 Reduction. ACS Appl. Energy Mater. 2021, 4, 7030–7039. [Google Scholar] [CrossRef]
  66. Sakakibara, N.; McQueen, E.; Sprick, R.S.; Ishitani, O. Photocatalytic CO2 Reduction in Aqueous Media Using a Silver-Loaded Conjugated Polymer and a Ru(II)-Ru(II) Supramolecular Photocatalyst. Bull. Chem. Soc. Jpn. 2025, 98, uoaf017. [Google Scholar] [CrossRef]
  67. Chen, Z.; Hu, Y.; Wang, J.; Shen, Q.; Zhang, Y.; Ding, C.; Bai, Y.; Jiang, G.; Li, Z.; Gaponik, N. Boosting Photocatalytic CO2 Reduction on CsPbBr3 Perovskite Nanocrystals by Immobilizing Metal Complexes. Chem. Mater. 2020, 32, 1517–1525. [Google Scholar] [CrossRef]
  68. Wang, B.; Zhao, J.; Chen, H.; Weng, Y.X.; Tang, H.; Chen, Z.; Zhu, W.; She, Y.; Xia, J.; Li, H. Unique Z-Scheme Carbonized Polymer Dots/Bi4O5Br2 Hybrids for Efficiently Boosting Photocatalytic CO2 Reduction. Appl. Catal. B 2021, 293, 120182. [Google Scholar] [CrossRef]
  69. Lin, J.; Qin, B.; Zhao, G. Effect of Solvents on Photocatalytic Reduction of CO2 Mediated by Cobalt Complex. J. Photochem. Photobiol. A Chem. 2018, 354, 181–186. [Google Scholar] [CrossRef]
  70. Kuriki, R.; Ishitani, O.; Maeda, K. Unique Solvent Effects on Visible-Light CO2 Reduction over Ruthenium(II)-Complex/Carbon Nitride Hybrid Photocatalysts. ACS Appl. Mater. Interfaces 2016, 8, 6011–6018. [Google Scholar] [CrossRef]
  71. Thompson, W.A.; Sanchez Fernandez, E.; Maroto-Valer, M.M. Review and Analysis of CO2 Photoreduction Kinetics. ACS Sustain. Chem. Eng. 2020, 8, 4677–4692. [Google Scholar] [CrossRef]
  72. Das, R.; Chakraborty, S.; Peter, S.C. Systematic Assessment of Solvent Selection in Photocatalytic CO2 Reduction. ACS Energy Lett. 2021, 6, 3270–3274. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.