Sk Samim
Akhter
a,
Koushik
Makhal
b,
Dev
Raj
a,
Manaswini
Raj
a,
Thillai
Natarajan M
a,
Bhabani S.
Mallik
b,
Prabhakar
Bhardwaj
c,
Pankaj
Kumar
c,
Ebbe
Nordlander
d and
Sumanta Kumar
Padhi
*a
aArtificial Photosynthesis Laboratory, Department of Chemistry and Chemical Biology, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826004, India. E-mail: sumanta@iitism.ac.in
bDepartment of Chemistry, Indian Institute of Technology Hyderabad, Sangareddy, 502284, Telangana, India
cDepartment of Chemistry, Indian Institute of Science Education and Research (IISER), Tirupati 517507, India
dChemical Physics, Department of Chemistry, Lund University, Box 124, SE-221 00 Lund, Sweden
First published on 24th October 2025
An approach to reducing greenhouse gas emissions that shows promise is the electrochemical conversion of CO2 to products with added value. Here, we present [Co(8HQ-Tpy)(H2O)]2(PF6)2 ([Co1]), a cobalt-based molecular electrocatalyst that can convert CO2 to CO in a DMF/H2O mixture (4.8
:
0.2 v/v) in a selective manner (8HQ-Tpy = 2-([2,2′:6′,2′′-terpyridin]-4′-yl)quinolin-8-ol). At an overpotential of 760 mV, the catalyst shows a TOFmax of 2575 s−1 and a high Faradaic efficiency of 94 ± 2%. The CO2 reduction follows both ECEC and EECC-type routes, involving stepwise proton and electron transfer, according to a mechanistic investigation that combines DFT calculations, infrared spectroelectrochemistry (IR-SEC), and kinetic isotope effect (KIE) observations. Sequential protonation and CO2 activation are made possible by the reduction of a hexa- to penta-coordinate Co centre. According to DFT studies, protonation at the ligand O− site, which takes place before CO2 coordination and favours an EECC pathway, becomes thermodynamically favourable following reduction. Both deprotonated and protonated CO2-derived intermediates are captured by IR-SEC measurements, and proton transfer is not rate-limiting as the KIE is low (kH/kD = 1.17). When taken as a whole, these results offer a comprehensive mechanistic understanding of CO2-to-CO conversion as well as design guidelines for creating advanced molecular electrocatalysts for carbon capture and utilization.
Transition-metal complexes are uniquely suited as catalysts for CO2 reduction because they can store and transfer various electrons, thus circumventing the high-energy CO2 radical intermediate. Numerous molecular catalysts based on transition metal complexes have been proposed for CO2 reduction, generally exhibiting good product selectivity.20–22 However, a significant challenge is that the reduction of CO2 is kinetically restrained by multiple electron-transfer processes accompanied by the highly competitive hydrogen evolution reaction. Molecular catalysts based on Ru, Ir, Pd, and Re have been investigated for converting CO2 to CO and formate. Catalysts based on earth-abundant transition metals,23–25 such as Fe, Co, and Ni, could provide inexpensive materials for large-scale use.26,27 However, such base metal catalysts are relatively labile and prone to generating H2, resulting in low selectivity for CO2 reduction. Redox-active ligands with extensive π–π conjugation can facilitate electron transfer and storage, thus accelerating electrocatalytic kinetics. Cobalt complexes with redox-active ligands prefer to reduce CO2 to CO.28–31 Several molecular cobalt-based catalysts with well-defined coordination environments have been reported for CO2 reduction. Also, polypyridine ligands, represented by, inter alia, bipyridine (bpy) and terpyridine (tpy), have gained recognition for their inherent stability and ability to facilitate these electrochemical reduction of CO2 in organic solvents. Elgrishi and collaborators focused on synthesizing [Co(tpy)2]2+ complexes and thoroughly examined their electrochemical properties. Their electrochemical investigations were conducted in an electrolyte consisting of 0.1 M [NBu4](ClO4) dissolved in DMF. The findings from the electrochemical studies revealed the generation of CO and H2 with Faradaic efficiencies ranging from 16% (CO) to 21% (H2).32 Later, the same group explored the potential of [Co(tpy)2]2+ to enhance CO generation despite its initially low Faradaic efficiency. The study involved comprehensive electrochemical analyses of the [Co(tpy)2]2+ and [Co(tpyY2X)2]2+ systems, where substituents X, such as C6H5Cl, C6H5CH3, H, OCH3, and C4H9, were examined alongside Y, which could be H or C4H9. The primary reaction products observed were CO and H2, with the substituents dictating the pathway of the reaction.33
In the endeavour to achieve efficient CO2 reduction catalysis, cooperative reactivity between two metal centers has emerged as a feasible strategy for enhancing the catalytic effectiveness of molecular catalysts. When compared to single-metal counterparts, complexes comprising bimetallic Ni,34 Fe,35 Co,36–38 Pd,39 and Re40,41 centers have shown modest to considerable improvements in catalytic rates and turnover numbers for CO2 reduction. Cobalt dimer complexes have shown promise in CO2 reduction due to their ability to facilitate multi-electron transfer steps and stabilize reactive intermediates.36–38
As demonstrated by Ni–Fe carbon monoxide dehydrogenases (CODHs), which accomplish reversible CO2-to-CO conversion through spatial and functional cooperation between two metals, nature commonly uses bimetallic active sites to mediate challenging multielectron reactions.7,34 This inspired us to employ a unique ligand 2-([2,2′:6′,2′′-terpyridin]-4′-yl)quinolin-8-ol (8HQ-Tpy, Scheme 1) to create a fully bridged bimetallic cobalt complex, motivated by this biological precedent. In the present study, we have synthesized a novel and robust cobalt dimer complex [Co(8HQ-Tpy)(H2O)]2(PF6)2[Co1] designed explicitly for electrochemical CO2 reduction. In this architecture, the ligand acts as a structural and electronic scaffold, coordinating both cobalt centers through the quinoline and terpyridine moieties to establish octahedral coordination environments at each metal. The rigid 8HQ backbone, serves not only as a chelating unit, but also introduces asymmetry by bridging the second Co center. This system shows weak antiferromagnetic coupling (vide infra), indicating little electronic communication, in contrast to enzymatic systems where high metal–metal cooperativity directly catalyzes the reaction. Rather, metal–ligand cooperativity, in which the ligand promotes substrate binding, electron delivery, and geometric control, appears to be the source of the catalytic performance, especially the excellent selectivity for CO over H2. By using this method, we can investigate how carefully crafted ligand frameworks can be used to tune activity and selectivity in molecular CO2 reduction in place of direct M–M synergy.
![]() | ||
| Scheme 1 A schematic depiction of the synthetic route to the [CoII(8HQ-Tpy)(H2O)]2(PF6)2[Co1] complex. | ||
:
4, v/v) containing one equivalent of the ligand 8HQ-Tpy (0.1 g, 0.21 mmol). The reaction mixture was stirred at room temperature for 4–5 hours, after which the solvents were removed under vacuum to reduce the volume. An aqueous solution of KPF6 was then added, resulting in the formation of an orange-red precipitate (Scheme 1). The precipitate was isolated from the reaction solution by reducing the volume under vacuum, then washing with ice-cold methanol and diethyl ether. Brown-coloured single crystals were obtained through slow evaporation of a DMF solution of the complex, yielding 55% with respect to the Co starting material. The isolated crystals of [Co1] were further analysed by single crystal X-ray diffraction. The UV-Visible spectrum and high-resolution mass spectrum of the obtained complex are depicted in Fig. S2 and S3. The calculated mass for [Co2C48H30N8O2]2+ was determined to be 434.058, while the found mass was 434.070.
The Co1–N4 bond length (2.214 (2) Å) by the 8-HQ fragment is longer than the Co1–NTpy bond distances. The Co1–O2 (2.136(2) Å) for the coordinated aqua ligand is significantly longer than the Co1–O1 bond (1.965(2) Å) involving the deprotonated oxygen of the 8-HQ moiety, which is the shortest bond in the cobalt coordination sphere. Among the three five-membered bite-angles, the ∠N2Co1N angles (∠N2Co1N3 (74.99(7)°), ∠N2Co1N1 (74.66(7)°)) generated by the Tpy moiety are more constrained than ∠N4Co1O1 (79.26(7)°) generated by the 8-HQ moiety.
![]() | (1) |
and
(N = Avogadro's number; F1, F2 and F3 are the parameters used in ref. 42e–g).
In addition, the magnetic susceptibility data were fitted using a model consisting of homo-dinuclear high-spin cobalt(II) octahedral complexes with little distortion. The magnetic susceptibility was estimated using the Sakiyama susceptibility equation based on the model outlined in ref. 42j–k. A weak antiferromagnetic interaction, J = −1.0 cm−1 was also found in this case, with the κ and λ values equivalent to a free Co(II) ion (Fig. S5).
![]() | ||
| Fig. 3 Cyclic voltammograms in dry DMF using 0.1 M [NBu4](ClO4) as a supporting electrolyte under N2 atmosphere of 1.0 mM of (A) ligand 8HQ-Tpy (B) 1.0 mM of [Co1] complex. | ||
The cyclic voltammogram revealed three metal-centered and three ligand-centered redox couples of the [Co1] complex. Fig. 3(B) depicts a quasi-reversible wave at E1/2 = −0.14 V vs. Fc+/0; Fig. S6(D), assigned to the CoIIICoIII/CoIICoII redox couple. The broad nature of this peak was attributed to the large reorganization energy involved in the transition between CoIICoII and CoIIICoIII redox states, corresponding to high spin d7 to low spin d6 configurations, respectively.42d–f It seems that both the CoII-centres in [Co1] oxidise simultaneously to CoIII, and this was verified by chronoamperometry at Eapp = +0.33 V vs. Fc+/0 (Fig. S7(A) and S7(B)). Furthermore, the [Co1] complex exhibits other metal-based quasi-reversible peaks at E1/2 = −1.26 V vs. Fc+/0 and E1/2 = −1.41 V vs. Fc+/0 which were attributed to CoIICoII/CoIICoI and CoICoII/CoICoI redox couples, respectively. The two one-electron steps involved in the reduction of the metal centers were verified via chronoamperometry at Eapp = −1.5 V vs. Fc+/0 (Fig. S7(C) and S7(D)). Another two peaks at E1/2 = −1.28 V vs. Fc+/0 and E1/2 = −1.52 V vs. Fc+/0 in Fig. S6(D) correspond to reductions of the 8HQ and Tpy moieties of the 8HQ-Tpy ligand, respectively. Moreover, upon complexation the metal lowers the energy of the ligand's π* orbitals via metal–ligand orbital interactions. This stabilization makes it thermodynamically difficult to add electrons, causing the ligand-based reduction peaks to shift to more negative (cathodic) potentials which is seen in Fig. 3(B).
vs. Fc+/0 (pKa value of 7.37 for H2CO3 in DMF).43a Accordingly, [Co1] produces CO from CO2 at an overpotential of 760 mV vs. Fc+/0, with catalytic mid-wave potentials (Ecat/2) at −2.23 V vs. Fc+/0, at a scan rate of 0.1 V s−1.44
Catalysis only starts once both cobalt centres are reduced to the +1 oxidation state, which is followed by the reduction of the 8HQ ligand, according to cyclic voltammetry conducted in a CO2 environment. The fully reduced [Co(I)Co(I)(8HQ˙−-Tpy)] complex is the catalytically active species, according to this sequence. The requirement for this ligand-centered reduction suggests that the 8HQ unit plays a significant part in promoting CO2 activation, either by enhancing nucleophilicity or via electronic delocalization through metal–ligand cooperativity. These results are consistent with the observed low kinetic isotope effect (kH/kD = 1.17, vide infra) and little hydrogen evolution activity, suggesting an EECC-type process in which several electron accumulation stages precede CO2 binding and protonation.
:
0.2 v/v). The experiments were performed in a custom-made electrolyzer with 0.1 M [NBu4](ClO4) as the supporting electrolyte, applying potentials ranging from −2.0 V to −2.3 V vs. Fc+/0 (Fig. S8A). As the applied potential increased, the charge accumulation also increased (Fig. S8A). After 2.0 h of electrolysis, a rinse test was conducted on [Co1], followed by a CPE experiment without the complex at −2.3 V vs. Fc+/0. No gaseous products were detected in the bulk electrolysis without a catalyst or in the rinse test (Fig. S8B). Additionally, NMR spectra of the solution of [Co1] after 2.0 h of electrolysis did not exhibit any peaks attributable to the HCOO− ion.
Gaseous products, including CO and H2, were detected by gas chromatography using a flame ionization detector (FID) and a thermal conductivity detector (TCD). The retention times and the calibration plots of standard CO and H2 gases are provided in Fig. S9 and S10, respectively. Gas chromatography plots for CO2 reduction by [Co1] at −2.0 V to −2.3 V vs. Fc+/0 are displayed in Fig. S11 and S12. After 2.0 h of electrolysis, gaseous samples were collected using a gas-tight syringe and injected into the GC. The amounts of CO and H2 produced were quantified using the calibration plot. The quantities of CO and H2 generated at each potential during electrolysis are presented in Fig. 5.
![]() | ||
| Fig. 5 Amount of CO and H2 produced after 2.0 h of CPE at different potentials ranging from −2.0 V to −2.3 V vs. Fc+/0 for [Co1] complex. | ||
Furthermore, the gaseous products were confirmed using an online Mass Analyzer equipped with an Omni StarTM Mass Analysis System GSD 320 (Pfeiffer) quadrupole mass spectrometer apparatus. In addition, a labeling experiment was conducted using an online gas analyzer to identify 12CO2 and 13CO2 molecules.45,46 The gas analyzer sensor was attached to the headspace of the electrochemical cell during CPE.
During electrolysis, the evolution of H2 was initially detected ahead of CO, indicating the consecutive reduction of protons to H2 followed by the reduction of CO2 to CO. The detection of H2, 12CO, and 13CO (H2, m/z = 2; 12CO, m/z = 28; and 13CO, m/z = 29) during controlled potential electrolysis of [Co1] under 12CO2 and 13CO2 atmosphere is depicted in Fig. S13 and Fig. 6.
![]() | ||
| Fig. 6 The controlled potential electrolysis of [Co(8HQ-Tpy)(H2O)]2(PF6)2; [Co1] as catalyst under 13CO2 (A) 13CO detection, (B) H2 detection by online mass analyzer. | ||
The Faradaic Efficiency (FE) for CO evolution for the [Co1] complex was determined to be 94 ± 2% at −2.2 V vs. Fc+/0 (Fig. 7). In comparison, the Faradaic efficiencies (FE) for H2 evolution for the same complex were calculated to be 1.00% at −2.0 V vs. Fc+/0 after 2.0 h of electrolysis (Fig. 7). The TONs and TOFs for CO2 reduction by [Co1] after 2 h of CPE at different potentials are summarized in Tables S4 and S5. A comparison with the reported mononuclear and dinuclear Co-based catalysts are provided in Table 1.
![]() | ||
| Fig. 7 Faradaic efficiency of CO and H2 production after 2.0 h of CPE at potentials varying from −2.0 V to −2.3 V vs. Fc+/0 for [Co1]. | ||
| Catalyst system | FE(CO) | FE(H2) | Main product(s) | Design principle | Ref. |
|---|---|---|---|---|---|
| DBPy-PyA = (1-([2,2′-bipyridin]-6-yl)-N-([2,2′-bipyridin]-6-ylmethyl)-N-(pyridin-2-ylmethyl) methanamine; TPA = tris(2-pyridylmethyl)amine; N5 ligand (2,13-dimethyl-3,6,9,12,18-pentaazabicyclo[12.3.1]octadeca-1(18),2,12,14,16-pentaene); qpy = 2,2′:6′,2′′:6′′,2′′′-quaterpyridine); biqpy = 4,4′′′′-(2,7-di-tert-butyl-9,9-dimethyl-9H-xanthene-4,5-diyl)di-2,2′:6′,2′′:6′′,2′′′-quaterpyridine. | |||||
| [Co(tpy)2]2+ | 31% | 23% | CO + H2 | Ligand-tuned redox behavior | 32b |
| [Co(DBPy-PyA)]2+ | ∼6%(CO) | 78% | H2 | Coordination geometry control | 33a |
| +7% HCOOH | |||||
| Capsule-like Co complex | ∼90% | <10% | CO | Proton management via cavity control | 33b |
| [CoII(TPA)Cl][Cl] | TON > 900 | — | CO | Rigid ligand architecture | 33c |
| [Co(N5)]2+ | ∼82% | Low | CO | Metal–ligand electronic interplay | 33d |
| [Co(qpy)(OH2)2]2+ | 94% | 1% | CO | Extended quaterpyridine system, enhanced redox stability/selectivity | 33e |
| [Co2(biqpy)Cl]3+ | ∼90%(CO) | 0.5% | CO | Bimetallic synergy for activation | 36 |
| +7%(HCOOH) | |||||
| [Co(8HQ-Tpy)(H2O)]2(PF6)2 | 94% | 1% | CO | Asymmetric metal–ligand cooperation | This work |
The cyclic voltammetry experiments further confirmed the homogeneous nature of the reaction. For a better understanding of the homogeneity and stability of the complex during the catalytic process, Scanning Electron Microscopy (SEM) (Fig. S16 and S17), and Energy-Dispersive X-Ray Spectroscopy (EDX) (Fig. S18 and S19) analysis were carried out on the working electrode surface, specifically the glassy carbon plate. Controlled potential electrolysis was performed for 2.0 hours using 0.1 M [NBu4](ClO4) as the electrolyte for the blank experiment (without catalyst) and, in another experiment, with the [Co1] complex at an applied potential of −2.2 V vs. Fc+/0. SEM and EDX analyses revealed only trace amounts of cobalt metal deposited on the working electrode surface after electrolysis, more precisely 0.08 atom% for the [Co1] complex under the specified reaction conditions, indicating a homogeneous nature of the catalysis effected by [Co1]. Moreover, from XPS studies, it has also been confirmed that deposition of metal catalyst does not occur on the electrode surface upon performing CPE for 2.0 h in a CO2 atmosphere using 0.5 M [Co1] complex (Fig. S20).
A Cottrell plot, which shows the peak current versus the square root of the scan rate (ν1/2) for the CoIICoII/CoIICoI and CoIICoI/CoICoI redox processes, was found to be linear for [Co1]. This observation was made during cyclic voltammetry experiments conducted with 1.0 mM solutions of complex in DMF under N2 or CO2 atmospheres, with scan rates ranging from 0.025 to 0.25 V s−1 (Fig. S21 and S22). The diffusion coefficients for the CoIICoII/CoIICoI and CoIICoI/CoICoI redox events under N2 were determined for the cathodic peaks D0,c = 0.6 × 10−6 and 1 × 10−6 cm2 s−1, respectively. For the anodic peak, the diffusion coefficients were D0,a = 0.3 × 10−6 and 0.25 × 10−6 cm2 s−1, respectively. These values indicate that the redox events were diffusion-controlled and homogeneous in DMF solution under a nitrogen atmosphere. Similarly, diffusion-controlled behaviour was observed for the CoIICoII/CoIICoI and CoIICoI/CoICoI redox events under a CO2 atmosphere (Fig. S21 and S22). The diffusion coefficients for the cathodic peak under CO2 were D0,c = 0.8 × 10−6 and 1.4 × 10−6 cm2 s−1, respectively. The diffusion coefficients for the anodic peak under CO2 were D0,a = 0.3 × 10−6 and 0.5 × 10−6 cm2 s−1, respectively. These values indicate that the redox events were diffusion-controlled and occur uniformly in DMF solution under a CO2 atmosphere.
Subsequently, cyclic voltammograms were recorded by varying scan rate from 0.1 V s−1 to 1.5 V s−1 in DMF medium under N2 and CO2 atmosphere (Fig. S23). The plateauing of the catalytic current at scan rates beyond 1.0 V s−1 can be attributed to kinetic limitations of the CO2 reduction reaction at the electrode surface. At lower scan rates, the current increases proportionally because the reaction is under diffusion-controlled conditions. However, at higher scan rates (>1.0 V s−1), the system becomes electron-transfer limited, and the mass transport of CO2 to the electrode surface cannot keep up with the faster potential sweep. This results in a saturation of the catalytic current (Fig. S23B and S24). Moreover, TOF has been extracted for each scan rate using the equation;47
{where, R = universal gas constant; T = temperature; n = number of electrons involved; F = Faraday's constant; υ = scan rate; kobs = rate constant}. Using this equation actual rate constant (TOFmax) of the catalyst [Co1] has been calculated in CO2 saturated atmosphere and found to be 2575.0 s−1 (Fig. 8).
![]() | ||
| Fig. 8 Plot of TOF vs. scan rate for the calculation of TOFmax for catalyst [Co1] under CO2 saturated atmosphere in DMF medium. | ||
Investigations into the kinetics of electrocatalytic CO2 reduction involved analyzing the catalytic cyclic voltammograms of the [Co1] complex. The rate of CO2 reduction was enhanced upon rising catalyst concentrations, following a linear trend. According to the equation icat = ncatFA[cat](DkCO2[CO2])1/2, there is a first-order dependence of catalytic activity on catalyst concentration (Fig. S25(A) and S25(B)).
Moreover, the catalytic current exhibits a linear dependence on the concentration of H2O. The choice of proton source in electrochemical CO2 reduction was crucial for determining the efficiency, selectivity, and overall effectiveness of the reaction. Acidic environments, for example, enhance the protonation of intermediates like H2CO3, formed during CO2 reduction, thereby accelerating the reaction rates. Such conditions were essential for optimizing the efficiency of CO2 conversion to the desired products. In this study, we varied the H2O concentration up to 2.2 M to serve as a proton source that participates in the rate-determining steps (Fig. S26(A) and S26(B)).
Analyzing the kinetic isotope effect (KIE) by monitoring the catalytic current (icat/ip) at different H2O and D2O concentrations revealed a linear relationship with proton donor concentration (Fig. S27). The [Co1] complex exhibited a modest KIE value of 1.17 ± 0.05, suggesting that proton transfer is not the rate-limiting step. An EECC-type mechanism, in which electron accumulation comes before CO2 activation and subsequent protonation, is supported by this minor isotope effect as well as electrochemical evidence that catalysis only starts after complete reduction of both cobalt centres and the 8HQ ligand. In accordance with metal–ligand cooperativity, the 8HQ moiety of the ligand most likely promotes CO2 binding through increased nucleophilicity or electronic delocalization. Although the proton source is necessary to complete the catalytic cycle, its limited kinetic contribution indicates that the chemical step that limits turnover is CO2 binding or C–O bond cleavage.
![]() | ||
| Fig. 9 Spectral changes during spectroelectrochemical studies for [Co1] at −2.3 V vs. Fc+/0 using 0.1 M [NBu4](ClO4) as supporting electrolyte under N2 and CO2 atmosphere. | ||
Infrared spectroelectrochemical (IR-SEC) measurements (Fig. 10(A)) were performed in a sealed OTTLE cell with CaF2 windows, using a Pt mesh working electrode, Pt wire auxiliary, and Ag wire pseudo-reference. A 0.5 mM complex solution in 5 mL of DMF/H2O (4.8
:
0.2 v/v) with 0.1 M TBAP was purged with N2, then CO2 for 15 min, and electrolysis was conducted at −2.3 V vs. Fc+/0. The emergence of a deprotonated CO2-derived intermediate, such as M–CO2− or formate, is indicated by the IR bands seen at 1670 and 1388 cm−1 during electrolysis under CO2 (Fig. 10(B)).37,47b–dFig. 10(C) presents the ΔTransmittance (normalized) vs. wavenumber plot, showing the difference between spectra recorded during electrolysis and the initial pre-electrolysis spectrum, highlighting spectral changes associated with electrochemical transformations. These bands correspond to the asymmetric and symmetric stretching modes of the carboxylate group (vasym and νsym COO−). These characteristics imply that CO2 is first activated in its anionic form at the metal centre and then stabilised by coordination. As the reaction continues, additional bands that are indicative of C
O stretching vibrations in protonated carboxylic acid species (M–CO2H) progressively appear at 1704, 1715, and 1724 cm−1 (Fig. 10(C)). A shift from a loosely linked or deprotonated intermediate to a more distinct, protonated species is reflected in this spectral development. Collectively, these findings demonstrate the dynamic coordination environment at the metal centre under catalytic circumstances and offer compelling evidence for a sequential method of CO2 activation, including both anionic and protonated intermediates.
The sextet state of Int-1 is the most stable spin state of the molecule and is 9.4 kcal mol−1 and 10.6 kcal mol−1 more stable than the corresponding doublet and quartet states, respectively. In an aqueous medium, a water exchange reaction may occur with reduced intermediates and catalysts (Fig. S29). The DFT study indicates that the hexacoordinate Co aqua complexes are less stable than the corresponding pentacoordinate complexes. In addition, for the reduction from Int-1 to Int-2, the H2O–Co bond distance increases, and the formation of the corresponding pentacoordinate complex by reduction of the hexacoordinate Co aqua complex becomes more exothermic than the corresponding reduction of a pentacoordinate Co complex.
In the first step in the electrochemical reduction of CO2 to CO, Int-1′ is reduced, followed by water desorption to produce Int-1 (E° = −0.44 V vs. SHE). Subsequently, Int-1 captures one electron to form Int-2 (E° = −0.60 V vs. SHE) via an exergonic step with a reaction energy of −22.1 kcal mol−1 (Scheme 2).
In the next step, Int-2 is further reduced by one electron to form Int-3 with a relatively higher reduction potential of −1.59 V (vs. SHE) and reaction energy values (0.7 kcal mol−1) than the first and second reduction reactions. In the CO2 reduction reaction, a competitive hydrogen evolution reaction is also possible. All feasible stepwise proton and electron addition reaction paths in a hydrogenation reaction and dehydrogenation via dissociation of O–H and Co–H were computationally modelled (Scheme S1). In the hydrogen evolution reaction, a proton addition reaction either takes place at the 8-hydroxyquinoline oxygen (to form Int-7) or at the cobalt ion (to form Int-9) in a competitive hydrogen evolution reaction. The formation of Int-7 is thermodynamically favored relative to Int-9 with respective reaction energies of −9.3 and −4.5 kcal mol−1 (Fig. 11). The HOMO–LUMO analysis indicates that the HOMO is primarily distributed on the quinoline, while the LUMO is located on the terpyridine fragment in Int-1 (Fig. S30). Compared to Int-1, the reduced intermediate Int-3 has a smaller HOMO–LUMO energy gap, which may facilitate protonation at the 8-hydroxyquinoline oxygen in Int-3.
A spin density distribution analysis was performed to study the protonation reaction steps. In Int-3, the spin densities are mainly located at Co, and the adjacent oxygen and nitrogen are symmetrically distributed between the two fragments (Fig. 12). In the proton addition reaction, the spin density of the participating oxygen decreases from 0.055 to 0.001 (Table S8) due to its direct participation in the protonation reaction.
In Scheme S1, the formation of Int-7 from Int-2 can proceed via two possible stepwise pathways: proton transfer followed by electron transfer (PT → ET) i.e.,
or electron transfer followed by proton transfer (ET → PT), i.e.,
. In the PT → ET pathway, Int-2 undergoes a proton transfer to form Int-10 (ΔG = −2.4 kcal mol−1), which is then reduced to form Int-7via an electron transfer (ΔG = −6.2 kcal mol−1). Alternatively, in the ET → PT pathway, Int-2 is first reduced to form Int-3 (ΔG = +0.7 kcal mol−1), which then undergoes proton transfer to generate Int-7 (ΔG = −9.3 kcal mol−1). The energy difference between Int-10 and Int-3 is relatively small (3.1 kcal mol−1), suggesting that the CO2 reduction reaction (CO2RR) may initially follow an ECEC pathway starting from Int-2. However, as the reaction progresses, particularly after the reduction of Int-2 to Int-3, it is likely to shift toward an EECC pathway. This mechanistic switch is supported by the fact that proton addition to Int-3 is significantly more favorable (ΔG = −9.3 kcal mol−1) compared to proton addition to Int-2 (ΔG = −2.4 kcal mol−1), as shown in Scheme S1. The enhanced protonation at Int-3 can be attributed to its increased electron density relative to Int-2, which facilitates proton addition at the O−-center following reduction.
The reduced Int-3 reacts with CO2 and generates C-coordinated metal CO2 complex Int-4. Fig. S31 suggests that with reduction, the Co metal center becomes more electron-rich, which in turn facilitates the CO2 addition reaction with a 6.9 kcal mol−1 reaction energy. After CO2 addition, the spin density at the participating Co metal center decreases from 2.551 to 1.072, indicating electron transfer from the metal to the CO2 moiety. Eventually, one of the oxygen atoms of the CO2 molecule Int-4 can acquire one proton and convert to Int-5. The protonation at CO2 is thermodynamically feasible with a reaction energy of −10.5 kcal mol−1 (Scheme 2). In the second protonation step, the oxygen atom of the 8-hydroxyquinoline entity receives a proton to make an O–H bond in Int-6. The proton addition step is energetically downhill by −0.4 kcal mol−1.
In an alternative reaction pathway, the reduced intermediate Int-3 can undergo protonation prior to CO2 addition, leading to the formation of Int-7. Computational results suggest that protonation at the O−-center in Int-3 is more favorable than CO2 addition, with reaction energies of −9.3 kcal mol−1 and +6.9 kcal mol−1, respectively. This indicates a stronger thermodynamic preference for protonation over CO2 coordination at this stage. Subsequently, Int-7 reacts with CO2 to form Int-8, which, upon an additional protonation step, yields Int-6.
Further, the hydroxyquinoline O–H and formate C(
O)–OH bonds in Int-6 dissociate through a five-membered transition state, TS-1, with an activation energy barrier of 10.3 kcal mol−1. In TS-1, the C–O, O–H, and ligand O–H bond distances are elongated by 1.65 Å, 1.23 Å, and 1.18 Å, respectively, relative to their corresponding equilibrium bond lengths (Fig. S32). The transition state calculation was performed using the QST2 method without guessing the transition state and was further confirmed by the IRC calculation (Fig. 13). In the last step of the reaction, CO and H2O dissociate to regenerate the active catalyst (Int-1) with a reaction energy of −32.0 kcal mol−1.
![]() | ||
| Fig. 13 Structure of transition state TS-1 and corresponding intrinsic reaction coordinate path (all bond distances are in Å). | ||
In summary, the mechanistic DFT study suggests that upon reduction of the Co-center, coordinated water molecules dissociate from the metal complex, converting the hexa-coordinate species into a penta-coordinate one. The catalytic CO2 reduction reaction (CO2RR) proceeds via initial reduction of the Co-based catalyst, followed by protonation at the ligand O−-center. In the stepwise proton and electron addition process, the reaction initially follows an ECEC pathway. However, due to the high proton affinity of the oxygen center in Int-3, formed after reduction of Int-2, the mechanism may shift to an EECC pathway, as supported by both experimental observations and DFT calculations. After the formation of Int-3, the most favorable pathway involves protonation of the Co-bound O−-center prior to CO2 coordination. Subsequent CO2 addition and a final protonation step led to the formation of CO and H2O, with a relatively low activation energy barrier of 10.3 kcal mol−1.
Protonation of the reduced deprotonated 8HQ oxygen is more advantageous than CO2 coordination at Int-3, according to thermodynamic data, suggesting that protonation takes place before CO2 binding. This sequence of stepwise reduction, protonation, and CO2 activation is compatible with a mechanism of the EECC type. Further evidence that proton transfer is not the rate-limiting step and that CO2 binding or subsequent bond cleavage most likely controls the total catalytic rate is provided by the experimentally found low kinetic isotope effect (kH/kD = 1.17).
Collectively, electrochemical, spectroscopic, and computational data support the idea that [Co1] predominantly operates through an EECC-type mechanism, providing important information for creating effective CO2-to-CO molecular catalysts in moderate environments.
| CO2RR | CO2 reduction reaction |
| CPE | Controlled potential electrolysis |
| DFT | Density functional theory |
| FE | Faradaic efficiency |
| GC | Gas chromatography |
| HER | Hydrogen evolution reaction |
| SEM/EDX | Scanning electron microscopy with energy dispersive X-ray spectroscopy |
| SEC | Spectroelectrochemistry |
| SQUID | Superconducting quantum interference device |
| TBAP | Tetra butyl ammonium perchlorate |
| ILCT | Intra ligand charge transfer |
| MLCT | Metal-to-ligand charge transfer |
| PCET | Proton-coupled electron transfer |
| ECEC | Electron transfer–chemical reaction–electron transfer–chemical reaction |
| EECC | Electron transfer–electron transfer–chemical reaction–chemical reaction |
| This journal is © The Royal Society of Chemistry 2025 |