Chanwoo Park,
Chaeyeon Kwon and
Young Hyun Hong*
Department of Chemistry and Center for Nano Materials, Sogang University, Seoul 04107, Republic of Korea. E-mail: yhhong@sogang.ac.kr
First published on 22nd April 2025
Artificial photosynthesis refers to a synthetic method of transforming solar radiation into storable fuels that are suitable for transport and practical applications, mimicking the natural photosynthesis found in plants and algae. Successfully replicating this natural energy conversion system could represent a major breakthrough in renewable energy technology, simultaneously providing clean fuel and reducing atmospheric carbon dioxide levels. Growing concerns over excessive CO2 emissions have led to considerable interest in developing technologies that convert CO2 into value-added fuels and raw materials by artificial photosynthesis. Essential processes such as photon absorption, charge transfer, water splitting, NAD(P)+ reduction, and carbon dioxide fixation have attracted significant research interest. Nevertheless, these individual processes have predominantly been studied independently. For the combination of NAD(P)H oxidation and CO2 reduction reactions, formate dehydrogenase (FDH) is a key enzyme in natural CO2 recycling systems, catalysing both the oxidation of formate to CO2 and the reduction of CO2 to formate via electron transfer involving NAD(P)H/NAD(P)+. Mimicking the metal active sites of such enzymes is crucial for designing efficient catalysts for CO2 conversion. However, no biomimetic in vivo catalysts have been reported for formate production from CO2 using NAD(P)H. This review focuses on catalytic studies involving the conversion of CO2 into formic acid using NAD(P)H with FDH as well as FDH-mimetic metal complexes. It covers enzymatic, photochemical, and electrochemical methods for CO2 reduction, highlighting the structure and mechanism of FDH and recent advances in the design of FDH mimetics. Additionally, this review explores strategies for enhancing the stability of the catalyst and catalytic performance through molecular tuning, offering insights into future research directions for developing efficient and sustainable CO2 reduction systems.
The development of efficient catalysts requires mimicking the metal active sites found in natural enzymes. Formate dehydrogenase (FDH) serves as a representative enzyme in CO2 recycling systems, catalysing the oxidation of formate to CO2 while transferring electrons to NAD(P)+ (nicotinamide adenine dinucleotide phosphate, Fig. 1) to produce NAD(P)H (Fig. 1) in the case of NADP-dependent FDH. Conversely, FDH can catalyse the reduction of CO2 to formate, transferring electrons from NAD(P)H to regenerate NAD(P)+ in the reverse reaction (see eqn (1)).21–25 Consequently, the most essential approach for biocatalytic conversion of CO2 is using in vivo biocatalysts and mimics of these catalysts. Despite the potential of these systems, there are no reported research studies on the development of mimetic in vivo biocatalysts capable of producing formate from CO2 using NAD(P)H.
![]() | (1) |
This review focuses on catalytic studies related to the conversion of CO2 into formic acid, as a value-added chemical, by NAD(P)H with the use of FDH and FDH mimetic metal complexes. Specifically, it describes studies on the production of formic acid in the reduction reaction of CO2 by NAD(P)H through photochemical, electrochemical, or enzymatic methods using enzymes or molecular catalysts. First, it highlights a representative biocatalyst, FDH, which enzymatically reduces CO2 to formic acid using NAD(P)H. The characteristics and structures of the active site of FDH, as well as the CO2 reduction reaction mechanism by NAD(P)H using FDH, are detailed. In addition, this review discusses the synthesis of FDH mimetics and the analysis of CO2 reduction reaction mechanisms and NAD(P)H oxidation/reduction reaction mechanisms via photochemical and electrochemical methods. Therefore, it not only provides an overview of FDH and FDH mimetics and recent progress in this field but also proposes research directions and prospects for developing catalysts. It also discusses the utilization of novel FDH mimetics through the tuning of molecular catalysts, to enhance the stability of molecular catalysts and their catalytic performance.
![]() | ||
Fig. 2 Metal-containing FDHs (formate dehydrogenases). The proposed structural configurations of metal active centres in their (a) oxidized and (b) reduced states. Additionally, the suggested catalytic mechanisms are presented for (c) the oxidation reaction converting formate into CO2, and (d) the reverse reaction, in which CO2 is reduced to formate by metal-dependent FDHs. Reprinted with permission from ref. 31 and 32. Copyright 2024, 2016, Multidisciplinary Digital Publishing Institute and American Chemical Society. |
The active site of the second type of FDH, which lacks a metal centre, is composed exclusively of amino acid components. A plausible configuration of the catalytic site for these non-metal FDHs is presented in Fig. 3.36,37 Typically, the active site comprises eight amino acids, such as Ile, Asn, Thr, Arg, Asp, Gln, His, and Ser.36,37 These residues collectively form binding sites for substrates such as NAD(P)+, formic acid, or CO2. Nearly all metal-independent FDHs identified in microbial systems utilize NAD(P)H as a necessary cofactor, enabling close docking with the active site of the enzyme and promoting the enzymatic reduction of CO2 to formate. In particular, NAD(P)H in the active site of CbFDH (Candida boidinii FDH) is stabilized by interactions with specific residues, predominantly Thr282, Asp308, Ser334, His332, and Gln313, positioning NAD(P)H proximal to CO2, thereby facilitating efficient hydride transfer for carbon reduction.36 Upon addition of CO2, NAD(P)H undergoes conformational adjustments, stabilizing the CO2 molecule through interactions with residues such as Asn146, Ile122, and Arg284. Subsequently, the hydride from the C4 position of the NAD(P)H nicotinamide ring transfers electrons directly to CO2, forming formate.38 The central catalytic step involves hydride transfer from the carbon atom of formate to the C4 atom on the pyridine ring of NAD(P)+, and this hydride transfer is the rate-determining step.38 Structural analyses demonstrate that the enzyme generates an optimal microenvironment that promotes hydride transfer and stabilizes interactions through targeted hydrogen bonding with the FDH peptide backbone.38,39 After hydride ion transfer, the resulting formate ions dissociate from the NAD(P)+-associated active centre of the FDH enzyme.38
![]() | ||
Fig. 3 Proposed structural configuration of the catalytic centre in metal-independent FDH. Reprinted with permission from ref. 36 and 37. Copyright 2008, 2015, American Chemical Society and the Royal Society of Chemistry. |
In the catalytic process of CO2 reduction, metal-free FDHs directly associate with NAD(P)H at their active sites, whereas metal-based FDHs indirectly utilize NAD(P)H via an electron-conducting subunit rather than through direct binding to the metal catalytic centre. The electrons originating from NAD(P)H traverse sequentially arranged [4Fe–4S] clusters before ultimately arriving at the metal-containing active site. Nevertheless, the detailed catalytic mechanism involving NAD(P)H in metal-dependent FDHs remains intricate and warrants further investigation.
Upon illumination, photosensitizers undergo excitation, resulting in the formation of electron–hole pairs, where the generated electrons subsequently transfer to NAD(P)+ through an electron mediator. Simultaneously, an electron-donating compound acting as a sacrificial reagent provides electrons to neutralize the holes generated in the photosensitizer (Fig. 4).57–61 To achieve compatibility with biological systems, molecular photosensitizers have been explored extensively as alternative semiconductor materials in photochemical NAD(P)H regeneration processes. Molecular photocatalysts frequently employed in enzymatic photocatalysis include xanthene-derived dyes such as fluorescein, eosin Y, and their modified analogues, along with porphyrin-based compounds, including porphyrin itself and its metal-containing derivatives.
![]() | ||
Fig. 4 Photocatalytic CO2 reduction to formate using NADH catalysed by a metal catalyst and FDH in the presence of a photosensitizer. |
The artificial photosynthetic production of 1-benzyl-1,4-dihydro nicotine amide (BNAH) or nicotinamide adenine dinucleotide (NADH) has been demonstrated under green- or red-light irradiation and an argon atmosphere, utilizing a photocatalytic system that integrates a tin(IV)-meso-tetrakis(N-methylpyridinium)-chlorin complex (SnC) (Scheme 1).62 This system enables efficient NADH analogue formation while recycling the selective hydride transfer mediator [Cp*Rh(bpy)H]+ in a neutral aqueous solution. The use of SnC as a multielectron transfer sensitizer selectively excited by red light allows for the formation of NADH analogues, thus enabling the photocatalytic reduction of NAD+ analogues. To facilitate the redox-mediated photocatalytic process, the rhodium complex [Cp*Rh(bpy)Cl]+ was introduced as a redox mediator.62 The process involves the formation of a rhodium-hydride intermediate [Cp*Rh(bpy)H]+, which generates the enzymatically active 1,4-NADH form of the reduced cofactor. This highlights the integration of transition metal catalysts with light-driven redox processes to efficiently and selectively generate 1,4-NADH. The photocatalytic process mimics the natural photosystem I (PSI), integrating a primary electron donor, SnC, and an additional redox mediator for NADH analogue formation.
![]() | ||
Scheme 1 Red-light-driven photocatalytic mechanism for BNA+ or NAD+ reduction using Sn(IV)-meso-tetrakis(N-methylpyridinium)-chlorin (SnC) as a multielectron transfer sensitizer, enabling selective hydride mediator [Cp*Rh(bpy)H]+ recycling in neutral aqueous solution. Reprinted with permission from ref. 62. Copyright 2008, 2015, American Chemical Society and the Royal Society of Chemistry. |
NADH is an essential cofactor involved in numerous biologically significant redox transformations. Consequently, the photochemical regeneration of its oxidized counterpart (NAD+) under biologically compatible conditions has become increasingly important in both photocatalytic and biocatalytic applications. Water-soluble, tri-anionic iridium complexes have shown remarkable efficiency as photocatalysts for reducing NAD+ analogues.63 The improved catalytic activity results from the robust electron-donating characteristics of the excited states of the tri-anionic IrIII complexes, their enhanced chemical stability, and beneficial electrostatic interactions with the positively charged rhodium co-catalyst (Scheme 2).63 The photocatalytic process utilizes triethanolamine (TEOA) as a sacrificial electron donor, ensuring continuous electron transfer. The rhodium co-catalyst, initially present as [CpRh(bpy)(H2O)]2+, undergoes a two-electron reduction coupled with protonation, forming the catalytically active RhIII hydride complex [CpRh(bpy)H]+. This complex is responsible for the selective reduction of the NAD+ mimic (BNA+) to its 1,4-reduced form (1,4-BNAH), which is enzymatically active. The IrIII photosensitizers showed significantly improved catalytic performance, with turnover frequencies (TOFs) ranging from 88 to 146 h−1, surpassing [RuII(bpy)3]2+ (TOF = 16 h−1). This enhancement directly relates to the substantially stronger electron-donating capacity of the excited-state IrIII species compared to [RuII(bpy)3]2+. Additionally, among the tested IrIII photosensitizers, a clear correlation emerged between the observed TOF and the facility of ground-state reduction from IrIV back to IrIII, indicating that efficient regeneration of the photosensitizer by TEOA could become a limiting factor. These outcomes underscore the necessity of optimizing both the excited-state electron-donating and ground-state electron-accepting properties of photosensitizers for effective catalytic activity.
![]() | ||
Scheme 2 Photochemical mechanism for 1,4-BNAH generation using TEOA as an electron and proton donor, a water-soluble iridium photosensitizer, and a rhodium co-catalyst for regioselective 1,4-reduction of BNA+. Reprinted with permission from ref. 63. Copyright 2023, American Chemical Society. |
The bimetallic Ru–Rh system enhances catalytic efficiency by enabling fast electron transfer, optimizing light absorption and catalysis separately, and ensuring compatibility with biological systems.64 In this system, a heterodinuclear Ru–Rh complex [(bpy)2Ru(tpphz)Rh(Cp)Cl]Cl3 is employed as a photocatalyst to drive the simultaneous production of NADH upon irradiation with visible light (Scheme 3).64 The Ru–Rh dyad functions through a photoinduced electron transfer mechanism, where the Ru-based chromophore absorbs light and transfers an electron through the bridging tetrapyridophenazine (tpphz) ligand to the Rh catalytic centre. This ultrafast intramolecular charge transfer facilitates the formation of the catalytically active [Cp*Rh(bpy)H]+ species, which selectively reduces NAD+ to 1,4-NADH.
![]() | ||
Scheme 3 Schematic depiction of the photo-biocatalytic system integrating photocatalytic NADH production from [(bpy)2Ru(tpphz)Rh(Cp)Cl]Cl3 with enzymatic reactions. Reprinted with permission from ref. 64. Copyright 2022, John Wiley & Sons. |
Efficient electron mediators can potentially replace NADH to improve the efficiency of formate production. Employing the electron mediator cobaloxime in combination with eosin Y as a photosensitizer successfully facilitated the conversion of NAD+ to NADH with a yield of approximately 36%, ultimately achieving a high conversion rate of 95% for CO2-to-formate synthesis (Scheme 4).65 Specifically, cobaloxime derivatives, including [Co(dmgH)2pyCl], [Co(dmgH)2(py-NMe2)Cl], and [Co(dmgH)2(py-COOMe)Cl] (where py = pyridine, py-NMe2 = 4-(dimethylamino)pyridine, and py-COOMe = methyl isonicotinate), acted as catalysts within a visible-light-driven photocatalytic system. This system utilized eosin Y as the photosensitizing agent and triethanolamine as the sacrificial electron donor to produce NADH from aqueous protons. The mechanistic pathway proposed involves sequential electron transfers from photoexcited eosin Y, reducing CoIII species first to CoI, followed by protonation to yield a CoIII-hydride intermediate. It is this hydride intermediate that is suggested to directly transfer a hydride to NAD+, thereby generating NADH.
![]() | ||
Scheme 4 Proposed catalytic cycles for the photo-driven generation of NADH using the cobalt catalyst/Eosin Y/TEOA system and subsequent enzymatic conversion of CO2 to formate using FDH. Reprinted with permission from ref. 65. Copyright 2012, American Chemical Society. |
Modified cobalt diamino-dioxime complexes, including [Co(DO)(DOH)pnX2] (where X = Cl and Br) and the BF2-bridged analog [Co((DO)2BF2)pnBr2], demonstrated notable photocatalytic efficiency in regenerating NADH (Scheme 5).66 These catalytic systems integrate diverse photosensitizers (PSs), such as [RuII(bpy)3]2+ (RuPS2+) and 4,8,12-tri-n-butyl-4,8,12-triazatriangulenium hexafluorophosphate (TATA+), and xanthene derivatives, including Rose Bengal (RB2−) and Eosin Y (EY2−), along with triethanolamine (TEOA) as an electron donor. Under visible-light exposure (λ = 420 nm), photoinduced electron transfer from the excited-state photosensitizer to the cobalt catalyst initiates a catalytic sequence, ultimately facilitating the selective reduction of BNA+ to its corresponding 1,4-dihydro derivative (1,4-BNAH).
![]() | ||
Scheme 5 Molecular structures of cobalt diimine–dioxime catalysts and photosensitizers. Reprinted with permission from ref. 66. Copyright 2020, the Royal Society of Chemistry. |
Recent studies have successfully reproduced the stoichiometric reaction of natural photosynthesis through photocatalytic systems that convert NAD(P)+ to NAD(P)H using water as the electron source, simultaneously producing oxygen (O2) (Scheme 6).67–72 Upon photoexcitation, Acr+-Mes forms a stable, long-lived charge-separated state (Acr˙-Mes˙+).67 Electron transfer from X-QH2, such as QH2 and Me4QH2, to the Mes˙+ moiety of Acr˙-Mes˙+, resulting in the formation of X-QH2˙+ and Acr˙-Mes, is energetically favorable, as indicated by the oxidation potentials of X-QH2 (Epa vs. SCE = 0.85–1.05 V) and the corresponding reduction potential of Mes˙+ within Acr˙-Mes˙+.67 Under these conditions, particularly at physiological pH, electron transfer from the Cl4QH− species to the Mes˙+ moiety of Acr˙-Mes˙+ generates Cl4QH˙ and neutral Acr˙-Mes, subsequently followed by electron transfer from Acr˙-Mes to CoIII(dmgH)2pyCl, forming [CoII(dmgH)2pyCl]− and regenerating Acr+-Mes.67 On the other hand, the generated X-QH2˙+ is rapidly deprotonated to produce X-QH˙. Then hydrogen atom transfer or proton coupled electron transfer from X-QH˙ to [CoII(dmgH)2pyCl]− produce [CoIII(H)(dmgH)2pyCl]− and X-Q. This [CoIII(H)(dmgH)2pyCl]− reacts with NAD+ to generate Co-NAD+ adduct to produce NADH.67
![]() | ||
Scheme 6 Illustrated mechanism for NAD+ photocatalytic reduction to NADH using X-QH2, with Acr+-Mes acting as a photoredox catalyst and CoIII(dmgH)2pyCl as an NAD+ reduction catalyst. Reprinted with permission from ref. 67. Copyright 2024, American Chemical Society. |
Although the overall photocatalytic mechanism closely resembles typical water-splitting reactions,66 achieving NAD(P)H synthesis directly from water marks a major advance. This is particularly significant since the resulting NAD(P)H can subsequently serve as a reducing agent for CO2 conversion into formate using formate dehydrogenase (FDH) (Scheme 7).73
![]() | ||
Scheme 7 Molecular photocatalytic cycle involving the photosensitized generation of NAD(P)H from water, in combination with enzymatic CO2 hydrogenation to formate. |
To enhance electron transfer efficiency from the excited-state photosensitizer to NAD+, the development and exploration of novel electron mediators remain essential. Effective photocatalytic systems and optimal electron mediators can then be applied to assess the enzymatic reduction of CO2 to formate.74 As previously described, molecular photocatalysts regenerate NAD(P)+ to NAD(P)H, which, in combination with FDH, facilitates the enzymatic reduction of CO2 to formate, regenerating NAD(P)+ in the process.75,76
Electrochemical CO2 reduction into formate can be carried out using two distinct strategies, which are the direct immobilization of the catalyst onto the electrode surface or indirect attachment via mediators. Formate dehydrogenase (FDH) serves as a representative example of direct immobilization, efficiently catalyzing the electrochemical reduction of CO2 to formate under mild conditions when immobilized on an electrode surface (Fig. 5).79 FDH-driven electrocatalysis is characterised by thermodynamic reversibility and minimal overpotential, with the equilibrium reduction potential determined by the zero net catalytic current point. Notably, formate is selectively generated as the exclusive product under mild reaction conditions. FDH thus demonstrates feasibility for reversible electrochemical interconversion between CO2 and formate and serves as a valuable model for the design and optimization of stable synthetic catalysts for real-world CO2 reduction technologies.79
![]() | ||
Fig. 5 Cartoon showing molybdenum/tungsten-containing FDH directly adsorbed onto an electrode surface catalysing the electrochemical reduction of CO2 to formate. Reprinted with permission from ref. 79. Copyright 2008, Proceedings of the National Academy of Sciences of the United States of America. |
The reduction of NAD+ was confirmed by attaching not only enzymes but also molecular metal catalysts to the electrode.80–82 In particular, electrode modifications employing noble metal-based catalysts are frequently studied to enhance NAD+ reduction efficiency. In such systems, mediators rapidly receive two electrons directly from the electrode surface along with a proton from the solution, generating hydride species capable of transferring electrons to NAD+ to yield NADH.83
Biomimetic metal complexes as additional mediator candidates have also been investigated for NAD+ regeneration in enzymatic CO2-to-formate conversion in an indirect way (Fig. 6).84–90 Indirect electrochemical methods utilizing electron-transfer mediators represent an effective strategy for cofactor regeneration.91 In such methods, electron mediators facilitate homogeneous reduction of the substrate, such as NAD+, and are subsequently regenerated at the electrode surface. This indirect pathway circumvents several limitations typically encountered in direct electrochemical approaches, including high overpotentials and the formation of inactive NAD dimers. Nonetheless, mediator selection is crucial for performance. Cp*RhIII(bpy)-based complexes are commonly employed. These mediators undergo a two-electron reduction at the cathode from RhIII to RhI, followed by protonation to yield a rhodium-hydride intermediate. The hydride intermediate subsequently transfers two electrons and a hydride to NAD(P)+, regenerating the initial RhIII catalyst (Scheme 8).90 Despite extensive investigation, direct electrochemical NAD+ reduction to enzymatically active NADH with complete yield, without electrode surface modification, has rarely been achieved. Following these foundational studies, subsequent research efforts have explored various electron mediators, evaluating their effectiveness in regenerating reduced nicotinamide cofactors and minimizing undesirable NAD dimerization.92–97 Furthermore, a rhodium-based complex incorporating bis(hydroxymethyl)-pentamethyl bipyridyl ligands exhibited a notably lower reduction peak at 0.549 V versus SHE, approximately 0.17 V more anodic than the NAD dimerization potential.97 Although NAD+ reduction by this mediator occurred at 0.592 V vs. SHE, the specific yield of the biologically active 1,4-NADH was not detailed. Generally, by applying a constant potential across the electrodes, electrons are transferred, allowing NAD+ conversion into NADH through direct, mediator-driven, or enzyme-coupled electrochemical processes.
![]() | ||
Scheme 8 Proposed catalytic mechanism for electrochemical NADH generation mediated by the rhodium-based complex [Cp*Rh(bpy)Cl]+. Reprinted with permission from ref. 90. Copyright 2011, Elsevier. |
To gain deeper mechanistic insights into the hydride-transfer process, an intermediate rhodium complex bearing a pentamethylcyclopentadiene ligand was successfully isolated from a short-lived rhodium hydride intermediate.101 Subsequently, the hydride-donating capability of this intermediate complex was evaluated, including its efficiency in reducing NAD+ to NADH. Remarkably, the direct involvement of the pentamethylcyclopentadienyl ligand in hydride transfer represented a previously unrecognized mechanistic feature of these rhodium-catalysed NAD+ reductions. The mechanism proposed was also supported by density functional theory (DFT) computational studies conducted on related rhodium and iridium organometallic complexes (Scheme 9).101
![]() | ||
Scheme 9 Proposed catalytic mechanism for NAD+ reduction involving a [(Cp*H)Rh(bpy)]+ intermediate species. Reprinted with permission from ref. 101. Copyright 2016, the Royal Society of Chemistry. |
Analogous to enzymatic methods employing formate as a hydride donor, ruthenium102 complexes of the general type [(η6-arene)Ru(en)Cl]PF6 (arene = hexamethylbenzene, p-cymene, or indan; en = ethylenediamine) have also been investigated for catalysing the reduction of NAD+ to 1,4-NADH using formate as the hydride source.103 The catalytic efficiency of these [(η6-arene)Ru(en)Cl]+ complexes is influenced by two critical factors.103 First, the identity of the coordinated arene significantly affects the turnover frequency (TOF), with performance declining in the order: hexamethylbenzene > indan > p-cymene.103 This trend highlights the importance of the arene ligand in facilitating hydride transfer from formate to the ruthenium centre, a process likely serving as the rate-limiting step. Second, substituting the chelating ligand also strongly impacts catalyst performance. Specifically, the replacement of the neutral N,N-chelating ligand ethylenediamine with the negatively charged O,O-chelating ligand acetylacetonate (acac) in p-cymene complexes resulted in a threefold reduction in catalytic efficiency under the same experimental conditions.103 This decreased activity was initially explained by stronger binding interactions between the adenine portion of NAD+ and the ruthenium catalyst, potentially obstructing access for formate. While the [(η6-arene)Ru(en)Cl]+ complexes exhibit relatively weak adenine affinity, derivatives such as [(η6-p-cymene)Ru(acac)Cl] display notably enhanced binding to adenine.
In particular, in studies mimicking photosynthesis, two catalytic pools namely the oxygen-evolving complex (OEC)108,109 and ferredoxin/ferredoxin-NADP+ reductase110–114 play key roles by providing viable kinetic pathways for both water oxidation and NADP+ reduction.115 The reaction of water with NADP+, which represents the net electron transfer reaction in the Z-scheme of oxygenic photosynthetic organisms, results in the formation of NADPH, which is subsequently used to reduce CO2 in light-independent reactions, along with the production of O2 (Fig. 7).116 Many organometallic complexes have been successfully utilized to independently catalyse water oxidation,108–120 the hydrogenation of NAD+,100,121–124 and the dehydrogenation of NADH.125–128
In the final stage of photosynthesis, the Calvin cycle, CO2 reduction takes place.129 Developing a catalytic system that integrates the aforementioned ferredoxin/ferredoxin-NADP+ reductase with CO2 reduction would represent a significant advancement toward artificial photosynthesis. No biomimetic metal catalysts combining NAD(P)H oxidation and CO2 reduction reactions have been reported, highlighting the need for research in this area. Establishing such a system requires the utilization of previously reported CO2 reduction catalysts, which are FDH mimetic metal complexes.130–148 The development of an artificial photosynthesis system could enable practical fuel production by employing an efficient and eco-friendly catalytic system, contributing to both the understanding of biological reaction processes and industrial applications. In this context, we discuss biomimetic metal complexes that catalyse the oxidation of NAD(P)H to NAD(P)+ and those that catalyse the reduction of CO2 to formate, and we propose the potential for developing a catalyst capable of facilitating both reactions simultaneously.
![]() | ||
Fig. 8 Suggested mechanism for hydride transfer from 1,4-NADH to (a) RuII complexes and (b) IrIII complex. Reprinted with permission from ref. 149. Copyright 2003, John Wiley & Sons. |
IrIII cyclopentadienyl complexes can mediate the oxidation of NADH to NAD+ via a hydride transfer mechanism (Fig. 9a).150 In experiments with the chloro complex, [(η5-Cpxbiph)Ir(phpy)(Cl)], the complex was found to accept a hydride from NADH, thereby converting it into its oxidized form, NAD+.150 Similar investigations with the pyridine-bound complex, [(η5-Cpxbiph)Ir(phpy)(py)]+ (Fig. 9b and 9c), demonstrate that the nature of the ancillary ligand significantly influences redox behavior and may also lead to the generation of reactive oxygen species. These findings underscore the importance of hydride transfer in the redox cycling of NADH/NAD+ and offer insights into the oxidant mode-of-action (MoA) of these IrIII complexes.
![]() | ||
Fig. 9 (a) Cartoon showing the potent oxidant and anticancer activity of organoiridium catalysts. (b) Synthetic pathway of 1-py·PF6. (c) X-ray crystal structure of [(η5-Cpxbiph)Ir(phpy)(py)]PF6·(CH3OH)0.5. Reprinted with permission from ref. 150. Copyright 2014, John Wiley & Sons. |
As shown in Fig. 10a, under mildly acidic conditions, the complex [Cp*Ir(pica)Cl] (1; with pica representing the picolinamidate, or κ2-pyridine-2-carboxamide ion) catalyses the dehydrogenation of β-NADH.151 In this process, β-NADH is oxidized to NAD+, which occurs via the removal of a hydride ion along with two electrons. This redox transformation, central to cellular metabolism, underscores the ability of complex 1 to modulate the NADH/NAD+ cycle effectively (Fig. 10b). Based on the mechanism, previously reported kinetic experiments and the supporting literature, a catalytic cycle has been proposed for the reduction of NAD+ via transfer hydrogenation using potassium formate (HCOO-K+) (Fig. 10c). Initially, substitution of the chloride ligand by water forms an aquo complex (A). This species can then either react with formate to yield an intermediate (B) or reversibly bind NAD+ to form an off-cycle adduct (E). The formation of adduct E appears to slow the reaction when NAD+ concentrations are high, suggesting that NAD+ can approach the catalyst in two distinct orientations. Additionally, the generation of intermediate B is likely favored by a hydrogen bond between the carboxylate oxygen and a coordinated –NH group. Subsequent β-elimination from B, accompanied by the release of CO2, produces a hydridic species (C) that rapidly transfers a hydride to NAD+, thereby forming NADH and regenerating the aquo complex A.
![]() | ||
Fig. 10 (a) Suggested mechanism for water oxidation to molecular oxygen and NAD+/NADH transformations using complex 3. (b) Structure of Ir complexes 1–3. (c) Illustrated mechanism for NAD+ hydrogenation using HCOOK as a hydrogen source, catalysed by an iridium complex. Reprinted with permission from ref. 151. Copyright 2017, American Chemical Society. |
A series of RuII monocarbonyl complexes were reported, all incorporating the bidentate phosphine 1,4-bis(diphenylphosphino)butane (dppb) ligand alongside various bidentate nitrogen ligands (Fig. 11).152 The general formula of these complexes is [Ru(OAc)CO(dppb)(N^N)]n (with n = + 1 or 0, and OAc representing acetate). Three complexes were synthesised in which the dppb framework was maintained, while the N^N ligands were varied among 2,2′-bipyridine (bipy), 1,10-phenanthroline (phen), and pyrazino[2,3-f][1,10]phenanthroline (pzphen). Mechanistically, the catalytic cycle for these RuII complexes is initiated either directly from the parent complex or from an aquo-species formed upon ligand substitution.152 This aquo-species reacts with formate to yield a hydride intermediate via β-hydride elimination with concurrent CO2 release. The resulting hydride complex is then capable of transferring hydrogen to NAD+, generating 1,4-NADH and regenerating the aquo-species. In a complementary set of experiments, the catalytic oxidation of NADH was also examined. Under appropriate conditions, such as in a buffered aqueous medium at pH 7.4, formation of the RuII hydride species facilitates the oxidation of NADH back to NAD+.
![]() | ||
Fig. 11 Proposed mechanism for NADH oxidation catalysed by ruthenium complexes and bidentate N^N donor ligands. Reprinted with permission from ref. 152. Copyright 2023, American Chemical Society. |
Recent investigations have demonstrated that [Cp*Ir(pyza)Cl] complexes (with pyza denoting pyrazine amidate) function as efficient reversible electrocatalysts for NAD+/NADH interconversion (Fig. 12).153 Designed to lower the overpotential required for NAD+ reduction relative to the earlier [Cp*Ir(pica)Cl] (pica = picolinamidate) system, the unique electronic properties of pyza, which is a weaker σ-donor yet a stronger π-acceptor compared to pica, enable an electronic reduction process at a notably lower potential (approximately −0.29 V) that is very close to the equilibrium potential of the NAD+/NADH redox couple (−0.32 V vs. NHE, 298 K, pH 7).153 Moreover, the catalyst is capable of driving both NAD+ reduction and NADH oxidation with a bias towards reduction, as evidenced by a catalytic current ratio (|ipred/ipox|) of 6.2 at 333 K. The reversible nature of this redox process is underscored by the rapid attainment of equilibrium between the hydride species and NAD+ with an equilibrium constant of about 3 and a ΔGrxn of −0.6 kcal mol−1 at 298 K, along with similar hydridicity values for NADH (28.9 kcal mol−1) and the corresponding iridium hydride (28.3 kcal mol−1).
![]() | ||
Fig. 12 Proposed thermochemical mechanism for calculating ![]() |
Ru(II) trimethylphosphine complexes have been identified with the general formula RuX(Y)(PMe3)4 (where X, Y = H, Cl, or O2CMe) as exceptionally active catalyst precursors for this reaction (Fig. 13a).130 In particular, the complex RuCl(O2CMe)(PMe3)4 has been recognized for its robust stability and high catalytic efficiency. Moreover, employing the RuCl(OAc)(PMe3)4 precursor in conjunction with the salt formed from DBU (1,8-diazabicyclo[5.4.0]undec-7-ene) and methylcarbonic acid has yielded remarkable improvements in reaction rates.130 This system outperforms other bases by delivering an 8-fold enhancement over tetramethylethylenediamine (TMEDA) and a 10-fold increase compared to triethylamine (Et3N), which was used in earlier studies. Under these optimized conditions, turnover frequencies have reached up to 95000 h−1, and there is potential for even greater activity through further optimization of cocatalyst and amine combinations.141
![]() | ||
Fig. 13 Structure and TOF of (a) RuCl(OAc)(PMe3)4, (b) [Cp*Ir(H2L1)Cl]+ and (c) PNP-IrIII trihydride complex. |
Deprotonation of the phenolic hydroxyl group occurs in the 4,4′-dihydroxy-2,2′-bipyridine iridium complex [Cp*Ir(H2L1)Cl]Cl, resulting in a stabilized catalyst with a characteristic pKa value of −2.30 (Fig. 13b).131 These iridium-based catalysts have demonstrated turnover frequencies (TOF) reaching 42000 h−1, highlighting their impressive catalytic efficiency in hydrogenation processes.142 Investigations have revealed that a PNP-IrIII trihydride complex bearing isopropyl groups on its phosphorus atoms exhibits superior catalytic performance, with turnover numbers (TONs) and turnover frequencies (TOFs) exceeding those of any other catalysts reported to date (Fig. 13c).132 Deprotonation of the PNP ligand is proposed to play an essential role in the catalytic cycle. Under conditions of 200 °C, this complex achieves a TON of 150
000 h−1.132
Scheme 10 illustrates the proposed catalytic cycle for the reduction of carbon dioxide by an iron catalyst.133 In this cycle, the active iron hydride species is generated initially from the precatalyst.133 This iron hydride then engages in the insertion of CO2 to form a metal-bound formate intermediate. Depending on the reaction conditions and the presence of appropriate substrates, such as alcohols or amines, the formate intermediate can be transformed further. Protonation releases free formate, and reaction with alcohols leads to the formation of alkyl formats. The cycle is closed by regeneration of the active iron hydride.144
![]() | ||
Scheme 10 Proposed mechanism for CO2 hydrogenation catalysed using Fe(BF4)2·6 H2O/PP3. Reprinted with permission from ref. 133. Copyright 2010, John Wiley & Sons. |
The FeII pincer complex trans-[(tBu-PNP)Fe(H)2(CO)] has been identified as a highly active catalyst for the hydrogenation of both carbon dioxide and sodium bicarbonate to formate salts (Scheme 11).134 Under aqueous conditions at 80 °C and at low hydrogen pressures (6–10 bar), complex 4 achieves a turnover number of up to 788 and a turnover frequency as high as 156 h−1.134 Initial investigations with complex 4 for the hydrogenation of sodium bicarbonate revealed that significant activity is observed only when water is used as the solvent with small amounts of THF as a co-solvent. Experiments conducted at an H2 pressure of 8.3 bar indicated that the highest activity is achieved at 60 °C with TON = 200.134 A proposed mechanism for the hydrogenation of carbon dioxide (Scheme 11) begins with the direct attack of CO2 on the hydride ligand of complex 4 to form an oxygen-bound formate intermediate (complex 5).134 The formate ligand is then readily displaced by a water molecule to yield a cationic species (complex 6). Formation of a dihydrogen-coordinated intermediate (A) is envisaged, which subsequently regenerates the active catalyst. This regeneration may proceed via heterolytic cleavage of coordinated H2 by hydroxide (pathway B) or through a mechanism involving dearomatization followed by proton migration (pathway C), thereby completing the catalytic cycle.134
![]() | ||
Scheme 11 Proposed reaction mechanism for CO2 reduction to produce formate using an Fe pincer complex. Reprinted with permission from ref. 134. Copyright 2011, John Wiley & Sons. |
Using well-established thermodynamic parameters for hydricity (ΔGH−) and acidity (pKa), researchers have designed a cobalt-based catalyst system for converting CO2 and H2 into formate (Scheme 12).135 In this system, the complex Co(dmpe)2H [with dmpe = 1,2-bis(dimethylphosphino)ethane] serves as the active catalyst for CO2 hydrogenation. The proposed catalytic cycle involves three key reactions: hydride transfer from the metal hydride to CO2, forming a metal-bound formate intermediate, subsequent addition of H2 to the resulting metal complex and regeneration of the metal hydride via deprotonation.
![]() | ||
Scheme 12 Proposed mechanism for CO2 hydrogenation using Co(dmpe)2H. Reprinted with permission from ref. 135. Copyright 2013, American Chemical Society. |
The reaction mechanisms for the hydrogenation of carbon dioxide catalysed by PNP-ligated metal pincer complexes such as (PNP)IrH3 (PNP = 2,6-bis(di-iso-propylphosphinomethyl)pyridine) are shown in Scheme 13.136 As shown in Scheme 13, two reaction pathways were considered for formic acid formation: one involving direct H2 cleavage by hydroxide (OH−) without any participation from the PNP ligand, and the other involving H2 cleavage through a mechanism that involves the aromatization and dearomatization of the pyridine ring in the PNP ligand. This finding underscores the vital role played by OH− as a base in the catalytic CO2 reduction cycle.147 The computed overall enthalpy barriers for the formation of formic acid from H2 and CO2 catalysed by (PNP)IrH3 were calculated to be 18.6 kcal mol−1.147
![]() | ||
Scheme 13 Proposed mechanism for CO2 hydrogenation using Co(dmpe)2H. Reprinted with permission from ref. 136. Copyright 2011, American Chemical Society. |
The FeII formate carbonyl hydride complexes, (RPNP)Fe(H)CO(HCO2) (RPNP = HN{CH2CH2(PR2)}2; R = iPr), bearing a bifunctional amine ligand, have been shown to achieve up to 106 turnovers for formate production with a turnover frequency near 200000 h−1 (Scheme 14).137 A plausible catalytic pathway for CO2 hydrogenation starting from these complexes involves an initial 1,2-addition of H2 across the Fe–N bond, followed by the insertion of CO2 into an Fe–H bond, and finally N–H deprotonation accompanied by formate extrusion to regenerate the active species.148
![]() | ||
Scheme 14 Proposed mechanism for (RPNP)Fe(H)CO/Li+ catalysed CO2 hydrogenation. Reprinted with permission from ref. 137. Copyright 2015, the Royal Society of Chemistry. |
In a comprehensive study on the catalytic reduction of CO2 to formate, it was observed that the IrIII trihydride complex with a bifunctional PNP ligand shows exceptional selectivity for formate formation, even in the presence of water as a solvent (Fig. 14).138–140 The reduction proceeds with a faradaic yield of 93%, with formate being the only reduced carbon product, and no formation of carbon monoxide (CO) (Fig. 14a).138 In a related investigation, an InIII trihydride complex ((PNHP)IrH3) supported by a bifunctional PNP ligand was reported as an exceptionally mild electrocatalyst for this transformation, delivering near-quantitative faradaic efficiencies (up to 99%) for CO2 reduction to formate (Fig. 14a and 14c).139 The catalytic cycle for the reduction of CO2 to formate involves the extrusion of the formate product and subsequent proton–electron transfer steps to regenerate the tri-hydride iridium complex and complete the cycle. These findings are consistent with a similar reaction mechanism proposed for a (POCOP)IrH2 catalyst (Fig. 14a and 14b), where a cationic acetonitrile adduct of iridium is implicated as the initial formate-lost product in the cycle.141 Mechanistic insights indicate that, following CO2 insertion into an Ir–H bond, the extrusion of the formate product and subsequent proton–electron transfer to the iridium centre regenerate the active Ir(III) trihydride species. A similar sequence has been proposed for (POCOP)IrH2, where a cationic acetonitrile adduct is suggested to form initially upon formate loss; however, hydrogen bonding between the bound formate and an ancillary N–H moiety can impede product release.139 The product release from this iridium complex is believed to be influenced by hydrogen bonding between the bound formate and the N–H group in the ancillary ligand, which could potentially hinder the release process. A parallel electrocatalytic mechanism has been seen in the (POCOP)IrH2(MeCN) complex (Fig. 14a), which exhibits an impressive 85% faradaic efficiency. The proposed mechanism involves the formation of an acetonitrile adduct of the dihydride species, which exists in rapid equilibrium with the cationic species. Upon electron reduction, cationic species undergoes a sequential two-electron, one-proton reduction to generate the dihydride species again, resulting in the formation of bicarbonate (HCO3−) as CO2 reacts with hydroxide.140 Both the dihydride species and its cationic form in the form of BArF4− salts have been independently synthesised and characterised, further validating this catalytic cycle. Overall, these studies not only unravel the operational mechanisms of highly efficient iridium catalysts for CO2 reduction but also provide critical insights into the roles of auxiliary ligands, proton sources, and the equilibrium dynamics that govern their catalytic activity. Additionally, another study on an IrIII trihydride complex with a bifunctional PNP ligand confirmed its remarkable selectivity for formate production, achieving a faradaic efficiency as high as 99%.139
![]() | ||
Fig. 14 (a) Molecular structures of four Ir complexes. (b) Proposed mechanism for electrocatalytic CO2 reduction catalysed by (b) the (POCOP)IrH2(MeCN) complex or (c) the (PNHP)IrH3 complex. Reprinted with permission from ref. 137. Copyright 2020, American Chemical Society. Reprinted with permission from ref. 138–140. Copyright 2013, 2015 and 2012, the Royal Society of Chemistry and American Chemical Society. |
cis-H,T-Rh2(L)2(phen)2 (phen = 1,10-phenanthroline) complexes with L being trifluoroacetamidate (complex 1), N-tolylacetamidate (complex 2), or N,N′-bis(tolyl)ethanimidamidate (complex 3) were prepared as shown in Fig. 15.141 The Rh2(II,I)-hydride acts as a key intermediate. Electrochemical data indicate that, at around 1.6 V, complex 2 undergoes a three-electron reduction followed by protonation in the presence of water to form an axial Rh2(II,I)-hydride.141 This intermediate is proposed to transfer a hydride to CO2, thereby releasing formic acid upon subsequent protonation. In contrast, complex 1 is inactivated via oligomer formation during reduction, and complex 3 exhibits the lowest reactivity and selectivity toward CO2 reduction.141 The reduced performance of complex 3 is attributed to the steric hindrance imposed by two tolyl groups at the axial sites, which likely interferes with the approach of CO2 to the active centre.141
![]() | ||
Fig. 15 (a) Schematic representation of the molecular structures of 1–3. (b) The proposed electrocatalytic cycle for CO2-to-HCOOH conversion using complexes 1 and 2. Reprinted with permission from ref. 141. Copyright 2021, the Royal Society of Chemistry. |
A series of manganese carbonyl complexes bearing elaborated bipyridine or phenanthroline ligands have been developed that selectively reduce CO2 to either formic acid or CO, depending on the ligand design (Figs. 16a and 16b).142 When the ligand framework incorporates strategically positioned tertiary amines, the complexes preferentially produce formic acid. Conversely, if the amine groups are absent or located far from the metal centre, CO is the predominant reduction product.142 Notably, the amine-modified complexes rank among the most active catalysts for CO2 reduction to formic acid, achieving turnover frequencies as high as 5500 s−1 at an overpotential of 630 mV.142 Furthermore, it was observed that decorating the bipyridine ligand with benzylic diethylamine groups at the ortho positions shifts the product selectivity from CO to formic acid. Mechanistic investigations indicate that these amine groups play a critical role by delivering a proton to the manganese centre to generate a Mn-hydride intermediate. This intermediate then undergoes CO2 insertion and, following an additional protonation step, yields formic acid. As shown in Fig. 16c, two pathways were proposed for forming [Mn(OCOH)]− from CO2 reduction. In one route (referred to as the LO pathway), a Mn-H intermediate undergoes CO2 insertion and then is reduced, thereby yielding [Mn(OCOH)]− species. Alternatively, in the HO pathway the same intermediate is first reduced and then undergoes CO2 insertion, with protonation to produce formic acid. Another suggested role for the amine group is that it could capture CO2, forming a carbamate intermediate. However, carbamates were not detected in the other amine-containing complexes exposed to CO2, which suggests that this pathway is unlikely to occur. A more plausible role of the amine groups is to function as Brønsted bases that sequester protons and deliver them to the metal centre, thereby generating a manganese-hydride intermediate. This proton relay mechanism has been invoked to explain the formation of a minor amount of formic acid under electrocatalytic conditions and has been observed in photochemical CO2 reduction studies. Importantly, the acidity of the reaction medium is critical; for instance, using a strong acid such as HCl may favor H2 evolution via Mn-hydride intermediates over the desired CO2 reduction.
![]() | ||
Fig. 16 (a) Proposed mechanism illustrating ligand-dependent product selectivity in Mn-bipyridine-catalysed electrochemical CO2 reduction. (b) Chemical structures of modified bipyridine–Mn complexes. (c) Proposed catalytic cycle for CO2 reduction to HCOOH. Reprinted with permission from ref. 142. Copyright 2020, American Chemical Society. |
The iron carbonyl cluster, [Fe4N(CO)12]−, has been identified as an effective electrocatalyst for the selective reduction of CO2 to formate across a wide pH range (5–13) (Scheme 15).143 Under conditions with low applied overpotentials (230–440 mV), formate is produced with a high current density of 4 mA cm−2 and boasts a faradaic efficiency of 96%. These remarkable catalytic performances are further complemented by the long-term stability of the catalyst, with a lifetime exceeding 24 hours. As shown in Scheme 15, the CO2 reduction mechanism involves the reduction of the electrocatalyst from 1− to an oxidation state, followed by protonation to form the hydride species (H-1)−. Analysis of the solution and gas phase reveals that (H-1)− selectively interacts with CO2, facilitating the formation of a C–H bond and subsequently producing formate as the final product. Operating at low overpotentials (230–440 mV), this catalyst produces formate with impressive metrics, achieving a current density of 4 mA cm−2 and a faradaic efficiency of 96%, while maintaining a long operational lifetime (>24 h).
![]() | ||
Scheme 15 Mechanistic proposal for CO2 reduction to formate by 1− in a proton-rich environment. Reprinted with permission from ref. 154. Copyright 2015, American Chemical Society. |
Moreover, the proposed mechanism for the electrocatalytic reduction of CO2 is shown in Scheme 16,144 integrating both experimental electrochemical findings and DFT computational insights. In this mechanism, the catalytically active species is a CoII-hydride intermediate that plays a central role in both H2 evolution and CO2 reduction. The cycle is initiated by a CoIII precursor that undergoes two successive electron transfers to generate a CoI species.144 This low-valent species is then protonated and further reduced, leading to the formation of a CoII–hydride intermediate. DFT calculations, performed on models of the acetonitrile derivative of the catalyst, support a concerted proton-coupled electron transfer (PCET) process in this conversion.144 In particular, computed redox potentials that account for the involvement of a pendant amine group align closely with the experimentally observed potential, reinforcing the role of PCET in a purely sequential electron–proton transfer. Once formed, the CoII-hydride intermediate reacts with CO2 to produce a CO2 adduct. This intermediate subsequently undergoes an internal hydride transfer, yielding a formate-bound species. The final step involves the extrusion of formic acid, which regenerates the active cobalt species and completes the catalytic cycle.
![]() | ||
Scheme 16 DFT-optimized mechanism for CO2 reduction using the CpCo(PCy2NBn2)I2MeCN-real model. Reprinted with permission from ref. 144. Copyright 2017, American Chemical Society. |
One of the CO2 reduction catalysts to produce formate, an iron complex, [Fe(PP3)(MeCN)2](BF4)2, was reported.156 The proposed mechanism for electrocatalytic CO2 reduction is shown in Scheme 17 via the formate pathway.145 In this pathway, free CO2 is directly reduced to formate. Under reductive conditions, [Fe(PP3)(MeCN)2](BF4)2 is reduced by 2e− and Fe(PP3)+ to generate Fe(PP3)+.145 The resulting Fe(PP3)+ then reacts with CO2 to form the formate adduct [(PP3)Fe(HCO2)]+. Finally, the formate is dissociated, regenerating [Fe(PP3)(MeCN)2](BF4)2 and thereby completing the formate pathway.
![]() | ||
Scheme 17 Proposed mechanism for CO2 reduction to formate catalysed by [Fe(PP3)(MeCN)2](BF4)2. Reprinted with permission from ref. 145. Copyright 2019, the Royal Society of Chemistry. |
A series of cobalt(III) pyridine–thiolate complexes have been investigated as electrocatalysts for CO2 reduction to produce formate.146 The detection of CO as a minor product suggests that an alternative pathway may be operative, in which CO2 initially binds to a reduced CoI centre and subsequently undergoes proton-mediated CO bond cleavage (Scheme 18).146 These experimental observations, together with the DFT calculations performed, led to the proposal of an overall reaction mechanism for the reduction of CO2 to formic acid (HCOOH), H2, and CO, as illustrated.
![]() | ||
Scheme 18 Proposed reaction mechanism for the generation of HCOOH and H2, with relative Gibbs free energies (ΔG, kcal mol−1) and transition state barriers (ΔG‡, kcal mol−1) referenced to the preceding intermediate. Reprinted with permission from ref. 146. Copyright 2020, John Wiley & Sons. |
The proposed reaction mechanisms for CO2 reduction to formate and hydrogen by complex 3 derived from the pre-catalyst [(bdt)MoVI(O)S2CuICN]2 offer an exciting advancement in the field of electrocatalysis using Earth-abundant metals (Scheme 19).147 In this system, the active catalyst species is generated in situ by the loss of an oxo group, which, in other contexts, would be involved in carbonate formation.147 This unprecedented strategy creates a vacant coordination site that facilitates the formation of a highly reactive metal hydride intermediate. Notably, the CuI centre is critical for the activity of the catalyst, although further studies are needed to fully understand the interplay between the molybdenum and copper sites. This work not only enriches the family of electrocatalysts that catalyze the conversion of CO2 to formate but also paves the way for designing new systems based on sustainable metal resources.
![]() | ||
Scheme 19 Proposed reaction mechanisms for CO2 reduction to formate and hydrogen catalysed by complex 3, derived from pre-catalyst 1. Relative Gibbs free energies (ΔG, kcal mol−1) and transition state barriers (ΔG‡, kcal mol−1) are referenced to the preceding intermediate, with standard one-electron reduction potential (E°, V) given vs. Fc/Fc+. The added electron in complex 3red is localized on the S3p orbitals of the ligands. Reprinted with permission from ref. 147. Copyright 2020, the Royal Society of Chemistry. |
Stable Ni-dithiolene complexes, structurally similar to molybdopterin, were prepared and characterised for the electroreduction of CO2 (Fig. 17a).148,160 These complexes efficiently catalyse CO2 reduction, with formic acid as the major product. Among them, [NiIII(qpdt)2]− achieves 13 turnovers in 4 hours, with a faradaic yield of formic acid of 60%.148 Notably, [NiIII(2H-qpdt)2]− exhibited the highest selectivity for CO2 reduction to produce formic acid, with a faradaic yield of 70%, significantly outperforming hydrogen reduction, which showed a faradaic yield of only 4%. In parallel, the corresponding Mo(O)(dithiolene)2 complexes have also been synthesised.160 These represent the first functional and stable catalysts inspired by the Mo centre of formate dehydrogenases. Under photoreduction conditions, the Mo-based catalysts predominantly generate formic acid, with carbon monoxide as a minor product, achieving more than 100 turnover numbers in approximately 8 hours (Fig. 17b).
![]() | ||
Fig. 17 (a) Structures of [NiIII(qpdt)2]− and [NiIII(2H-qpdt)2]−. (b) Scheme for electrocatalytic CO2 reduction to formate using Mo complexes. Reprinted with permission from refs. 148 and 160. Copyright 2019 and 2018, American Chemical Society and John Wiley & Sons. |
A mechanistic pathway for the electrochemical reduction of CO2 to formate using the [Co(triphos)(bdt)]+ complex has been established through a combination of experimental electrochemical data and DFT calculations.161 Cyclic voltammetry (CV) measurements show an anodic shift at the [Co(triphos)(bdt)]0/− couple under a CO2 atmosphere and in the presence of a proton source, indicating favorable interactions between the catalyst and reaction substrates. DFT calculations suggest that the catalytic cycle proceeds via an ECEC mechanism, wherein an initial one-electron reduction of [Co(triphos)(bdt)]0 (II) generates [Co(triphos)(bdt)]− (III), a necessary precursor for catalytic activity (Fig. 18). A subsequent protonation step leads to the formation of the cobalt-hydride intermediate [Co(triphos)(bdt)(H)]0 (IV-H), with a computed hydricity (ΔGH−) of 58.7 kcal mol−1. However, this hydricity value is insufficient to drive direct CO2 reduction thermodynamically. Instead, an additional electron transfer is required, reducing IV-H to [Co(triphos)(bdt)(H)]− (V), which exhibits a significantly lower hydricity of 37.8 kcal mol−1, making it competent for CO2 reduction. The reduction potential of the IV-H/V couple was calculated to be −1.61 V vs. Fc0/+, well within the electrochemical window where catalytic turnover occurs. This suggests that electron transfer to the metal hydride species is an essential step in the catalytic cycle, facilitating the hydride transfer to CO2. The formation of the formato complex [Co(triphos)(bdt)(HCOO)] (VI) is favored, followed by dissociation of formate and regeneration of the active catalyst.
![]() | ||
Fig. 18 Proposed mechanism for electrocatalytic CO2 reduction to formate using [Co(triphos)(bdt)]+. Reprinted with permission from ref. 161. Copyright 2024, the Royal Society of Chemistry. |
To the best of our knowledge, there has been no report on the use of molecular biomimetic compounds as catalysts in catalytic CO2 reduction to formate using NAD(P)H. In order to develop a catalyst and clarify the mechanism by capturing the intermediates formed in CO2 reduction to formate using NAD(P)H, experiments should be conducted using the biomimetic metal complexes mentioned above.162–164 This will be a process of developing the most effective catalyst using all chemical, photochemical, and electrochemical catalytic reactions. In particular, the optimal ligand system and experimental reaction conditions related to the stability of the molecular catalyst will be a big key to the success of the catalytic reaction. By anchoring effective homogeneous catalysts on heterogeneous supports, highly efficient hybrid catalytic systems can be developed that combine the advantages of both the tunability of homogeneous catalysts and the practicality of heterogeneous supports (Fig. 19). Such systems are expected to evolve significantly in the future, as exemplified by recent studies reporting the highest selectivity for CO2-to-formate conversion using hybrid catalysts.165 Therefore, this strategy could serve as a foundational platform for the next generation of catalyst design and development.
![]() | ||
Fig. 19 Cartoon showing hybrid catalytic systems by anchoring effective homogeneous catalysts on heterogeneous catalysts. |
Future research should focus on designing and optimizing biomimetic metal complexes with an emphasis on stabilizing molecular catalysts through optimal ligand design and reaction conditions. Furthermore, enzymatic catalysis integrated with photochemical or electrochemical technologies is anticipated to serve as a leading approach for efficient, selective, and economically viable conversion of CO2 to formate. Such systems leverage renewable energy sources, enabling sustainable and affordable CO2 transformation. Specifically, formate generation catalysed by engineered enzymes combined with renewable photo/electrochemical systems could facilitate the subsequent synthesis of valuable industrial platform chemicals. Consequently, substantial efforts are being dedicated to developing catalytic systems characterised by affordability, selectivity, durability, and sustainability, thereby advancing practical whole-cell CO2 fixation strategies. This review may hopefully provide more productive insights into the successful development of such catalysts and could pave the way for efficient artificial photosynthesis systems and sustainable fuel production, contributing significantly to global carbon management and industrial applications.
This journal is © the Partner Organisations 2025 |