Manoeuvring organo-electrocatalytic selective CO2 reduction to CO by terpyridine derivatives: DFT mechanistic exploration

Sk Samim Akhter a, Koushik Makhal b, Dev Raj a, Thillai Natarajan M. a, Palak Kumari Jaiswal a, Arun Biswas a, Bhabani S. Mallik b, Pankaj Kumar c and Sumanta Kumar Padhi *a
aArtificial Photosynthesis Laboratory, Department of Chemistry and Chemical Biology, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826004, Jharkhand, 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

Received 23rd August 2025 , Accepted 15th October 2025

First published on 21st October 2025


Abstract

The field of sustainable energy is rapidly growing, with electro-catalytic CO2 reduction becoming a key focus. This study examines the effectiveness of substituted terpyridine derivatives (L1L6) as electrocatalysts for CO2 reduction in a DMF and H2O mixture. Using polypyridyl ligands for CO2 electrochemical transformation into valuable products is relatively rare. Our investigation shows that terpyridine ligands are highly effective in converting CO2 into CO and H2via a 2H+/2e reduction process, with an overpotential of 650 to 850 mV vs. SCE. Notably, catalyst L6 demonstrated impressive faradaic efficiency (FE) for CO and H2, with 74% and 0.04% at −1.8 V vs. SCE, respectively, after 3 hours of electrolysis. All these catalysts, L1L6, exhibit over 90% selectivity for CO generation. The DFT studies reveal that the mechanism follows EECC pathways, where ‘E’ represents electrochemical steps and ‘C’ denotes chemical steps.


Introduction

The growing global demand for energy, primarily from non-renewable fossil fuels, contributes to air pollution and environmental toxins by continuing to emit carbon dioxide.1–5 Since CO2 is a thermodynamically inert molecule, its high overpotential makes reduction extremely difficult. On the other hand, electrochemical CO2 reduction has emerged as a widely used and economically feasible technique for converting greenhouse gases into useful commodities and fuels and using intermittent renewable energy.6–8 Reducing CO2 to other chemicals, such as CO, (HCOO/HCOOH), HCHO, CH3OH, and CH4, requires complex multi-electron processes.9–11 Molecular homogeneous catalysts are more selective than heterogeneous ones. They are essential catalysts due to this molecular property. Ruthenium, rhenium, rhodium and iridium-based catalysts have shown great promise in electrocatalytic CO2 reduction.12–15 However, from an economic point of view, the practical use of noble metal complexes is not viable for electrocatalytic CO2 reduction. Due to the growing demand for sustainable catalysis, we must focus on developing molecular catalysts that use readily available components to create a sustainable and economically feasible future.16 Metal-free electrocatalysts, composed solely of organic compounds devoid of metal centers, present distinct benefits in cost, abundance, and environmental footprint. These catalysts frequently demonstrate a wide range of functional groups that can enhance the electrochemical reduction of CO2 with exceptional selectivity. A promising path for sustainable chemical transformations has been previously described by certain groups that used organic catalysts, such as pyridinium salts and dihydropyridines (DHPs), to electrochemically reduce carbon dioxide.17 The ability of benzoate derivatives to reduce CO2 was examined by Gambino et al. with faradaic efficiencies of up to 99%; the study, which was carried out at a mercury (Hg) pool electrode in 0.1 M TBAP in DMF, found that monocarboxylic acid was the main result from benzoates along with dicarboxylic acid formation.18 Bocarsly and his team made a major breakthrough in the electrochemical reduction of CO2 in 1994 when they showed that the pyridinium ion could catalyze the creation of methanol on a hydrogenated platinum electrode. With a respectable faradaic efficiency of 30%, the investigation demonstrated the active involvement of the 1e reduced species originating from the pyridinium ion. Significantly, the absence of methanol formation occurred when pyridinium was substituted with N-methylpyridinium ions, highlighting the unique catalytic behaviour linked to various pyridinium derivatives in the electrochemical reduction of CO2.19 Aromatic nitriles and esters have the ability to preferentially convert CO2 to oxalate, which makes them promising homogeneous catalysts for CO2 reduction in aprotic environments, as later revealed by Gennaro et al. The catalytic activity is investigated by cyclic voltammetry, which shows nucleophilic addition to CO2 that creates a bond that regenerates the catalyst.20 With 0.5 M KCl at pH = 4.7, Portenkirchner et al. carried out a study showing how pyridine and pyridazine aid in the reduction of CO2. With a faradaic efficiency of 14% for the pyridine system and 3.6% for the pyridazine system, methanol was the final product that was separated from this process.21 Choudhury and his colleagues recently proposed a novel family of dicationic chemical compounds (PAC1–PAC4) as possible NADP+ mimics. By catalyzing electrochemical CO2 reduction (eCO2RR) to formic acid, these chemicals can be electrochemically reduced to produce NADPH mimics. Up to 35 ± 3% faradaic efficiency was attained with the PAC3 system. This marks a significant advancement in the use of metal-free organic catalysts to move from sub-stoichiometric to catalytic eCO2RR processes.22

A viable substitute that addresses issues with metal scarcity and environmental effects is the use of metal-free organocatalysts. Sustainable electrochemical CO2 reduction will be developed as a result of improving the performance of metal-free organocatalysts and comprehending their intricate mechanisms. The continuous research into metal-free and metal-complex catalysts highlights the multidisciplinary initiatives meant to reduce the negative effects of CO2 emissions on the environment while enhancing the potential of green and sustainable chemistry. Expanding on a similar design concept, we provide a novel class of organic CO2 reduction catalysts, [R-Tpy], where R is a phenyl derivative and Tpy is 4′-methyl-2,2′:6′,2′′-terpyridine. Since the terpyridine framework has a large π-conjugation, it is simple to add different functional groups to adjust its structural and electrical characteristics. Through multi-electron transfer and improved catalytic activity and selectivity, this versatility improves its capacity to stabilize radical intermediates during CO2 reduction. The electrochemical study involves featuring the general formula [R-Tpy], incorporating six unique terpyridine derivatives: [Cl-Tpy] (L1), [OMe-Tpy] (L2), [Ph-Tpy] (L3), [QCl-Tpy] (L4), [Me-Tpy] (L5), and [CN-Tpy] (L6), where QCl and Ph represent 3-chloroisoquinoline and phenyl groups, respectively. These compounds display remarkable selectivity towards the reduction of CO2 to CO, along with a negligible amount of H2 generation (Scheme 1).


image file: d5cy01027f-s1.tif
Scheme 1 Organo-electrocatalysis of CO2 to CO by R-Tpy.

Results and discussion

Synthesis and characterization of catalysts

Syntheses. The catalysts [R-Tpy], [Cl-Tpy] (L1), [OMe-Tpy] (L2), [Ph-Tpy] (L3), [QCl-Tpy] (L4), [Me-Tpy] (L5), and [CN-Tpy] (L6), were synthesized using a reported procedure (Scheme 2).23,24 Terpyridine is a tridentate redox-active chelating ligand, which was initially isolated by Morgan and Burstall.23–26 We have synthesized six terpyridine ligands (L1L6), with different substitutions at the 4′ position. The structure of the synthesized catalysts was confirmed through spectroscopic analyses, including 1H NMR, 13C NMR, and mass spectrometry (Fig. S2–S19).
image file: d5cy01027f-s2.tif
Scheme 2 Reaction scheme of Tpy derivatives (L1L6).

Electrochemical studies

The electrochemical properties of the catalysts were investigated by cyclic voltammetry using a three-electrode system with a glassy carbon electrode used as the working electrode, a calomel electrode as the reference, and a Pt wire as the counter electrode, and 0.1 M TBAP (tetra butyl ammonium perchlorate) was used as a supporting electrolyte under an inert atmosphere in a dry DMF solution. The voltammograms of all the ligands (L1L6) are shown in Fig. 1. In L1, there is one quasi-reversible peak at E1/2 = −1.78 (Ep,a = −1.71, Ep,c = −1.85) V vs. SCE, and other than that, there are three cathodic peaks at −1.44 V, −1.65 V, and −2.2 V vs. SCE and one anodic peak at −2.1 V vs. SCE. All these peaks were ascribed to the Tpy/Tpy˙ reduction process (Fig. 1).
image file: d5cy01027f-f1.tif
Fig. 1 (A)–(F) CVs of 1 mM L1L6 in dry DMF using 0.1 M TBAP as a supporting electrolyte under an N2 atmosphere.

The other three ligands L2, L3, and L5 show similar behavior, one sharp reduction peak appeared at −2.0 V, with two oxidation peaks at Ep,a = −1.93 V and −1.72 V, Ep,a = −1.86 V and −1.65 V, and Ep,a = −1.87 V and −1.69 V concerning L2, L3, and L5, respectively. These three ligands showed similar nature in CV due to the presence of electron donating groups OMe- and Me- in the ligand scaffold (Fig. 1). On the other hand, L4 shows three reduction peaks and two oxidation peaks at Ep,c = −1.68 V, −1.76 V, and −2.1 V, and Ep,a = −1.98 V and −1.63 V, respectively. However, L6 exhibits four reduction and oxidation peaks because of the CN moiety present in L6 at Ep,c = −1.51 V, −1.65 V, −1.85 V, and −2.07 V, and Ep,a = −1.96 V, −1.7 V, −1.5 V, and −1.31 V, respectively (Fig. 1). Further, we have checked the number of electron transfers in this electrochemical process. The coulometry of catalysts L1, L4, and L6 at −2.0 V and L2, L3, and L5 at −2.1 V vs. SCE reveals that two electrons are involved during electrochemical conversion (as shown in Fig. S21–S26). Moreover, the non-terpyridine-based ligand 2,2′-bipyridine (bpy) exhibits three cathodic-based peaks at Ep,c = −1.86 V, −2.2 V, and −2.54 V, and three anodic-based peaks at Ep,a = −2.0 V, −0.76 V, and −0.3 V. Similarly, 1,10-phenanthroline (phen) shows two cathodic-based peaks at Ep,c = −1.68 V and −2.02 V, and two anodic-based peaks at Ep,a = −0.6 V and −0.25 V (Fig. S27). A minimal increase in catalytic current has been noted when 2.2 M H2O has been added to the reaction mixture under an N2 atmosphere. Upon increasing the water concentration beyond 2.2 M, the catalytic current exhibits no further enhancement and remains essentially constant. This observation suggests that the electrode surface has reached a saturation limit with respect to proton availability, such that additional water does not contribute to an increase in proton-coupled electron transfer kinetics (Fig. S28). Moreover, it has been noted that catalytic current saturates rapidly under an N2 atmosphere upon addition of a small amount of phenol (PhOH) as the proton source indicating faster saturation of proton availability at the electrode surface (Fig. S29). It has been noted that peaks become shifted cathodically when potential has been measured with respect to ferrocene. The organo-catalyst L3 exhibits one sharp reduction peak appearing at −2.49 V and two oxidation peaks at Ep,a = −2.35 V and −2.1 V vs. Fc+/0 (Fig. S30C).

Electrochemical CO2 reduction

The evaluation of catalysts for electrochemical CO2 reduction centered on using terpyridine derivatives L1L6. The experimental conditions involved a solution of DMF saturated with CO2 gas at a concentration of approximately 0.20 M. A critical aspect of this process was the hydration of CO2, resulting in carbonic acid H2CO3. The generation of H2CO3 and its subsequent role as a proton donor were essential for facilitating the CO2RR.27,28 As shown in the cyclic voltammograms of catalysts L1L6 recorded in CO2-saturated DMF solution, the cathodic peak at −1.85 V shifts to −1.64 V vs. SCE for L1 (Fig. 2(A)), the peak at −2.0 V shifts to −1.76 V vs. SCE for L2 (Fig. 2(B)), the peak at −1.95 V shifts to −1.78 V vs. SCE for L3 (Fig. 2(C)), the peak at −1.7 V shifts to −1.61 V vs. SCE for L4 (Fig. 2(D)), the peak at −1.97 V shifts to −1.80 V vs. SCE for L5 (Fig. 2(E)), and the peak at −1.79 V shifts to −1.65 V vs. SCE for L6 (Fig. 2(F)). However, the peak shifts more cathodically (Ep,c = −2.35 V vs. Fc+/0) for L3 when measured with respect to ferrocene as a reference under a CO2 atmosphere. The addition of 2.2 M H2O as a proton source shifts the cathodic potential to Ep,c = −2.30 V vs. Fc+/0 when measured under a CO2 atmosphere (Fig. S30D).
image file: d5cy01027f-f2.tif
Fig. 2 (A)–(F) CVs of 1 mM catalysts L1L6 in DMF using 0.1 M TBAP as a supporting electrolyte under N2 and CO2 atmospheres and CO2 with 2.2 M H2O.

The 100–300 mV positive shift for L1L6 in the first cathodic process in the presence of CO2 exhibited cathodic current enhancement, as illustrated in Fig. 2. The mid-wave potential for this electrochemical process was determined to be E1/2 = −1.73 V, −1.71 V, −1.69 V, −1.78 V, −1.73 V, and −1.87 V vs. SCE for L1L6, respectively (Fig. S31). The L3 catalyst required less potential, and L6 had a higher potential than the other catalysts during the CO2RR. Cyclic voltammetry experiments were conducted under several conditions to elucidate further the electrocatalytic process, including N2, CO2, and CO2 with 2.2 M H2O in DMF. In each experiment, CO2 saturation was achieved after a 15 minute purge, revealing a moderate increase in catalytic current in the presence of 2.2 M H2O (Fig. 2).29 Moreover, upon CO2 purging a very minimal increase in catalytic current has been noted for both bpy and phen (Fig. S27B and D), implying that both catalysts are not effective for CO2 reduction as compared to terpyridine ligands. The absence of reactivity with bpy and phen can be attributed to their distinct electronic and structural features compared to terpyridine. Terpyridine, with three pyridyl units, provides stronger stabilization of radical/cationic intermediates and a more favorable geometry for transition-state organization. In contrast, bpy (bidentate) and phen (rigid, sterically congested) lack the same electronic and conformational flexibility, preventing efficient stabilization of the reactive intermediates. Addition of a small amount of a strong proton source such as PhOH leads to faster saturation of catalytic current (Fig. S32) in contrast to H2O (Fig. 2(A)–(F)).

Controlled potential electrolysis

Controlled potential electrolysis (CPE) was conducted at a glassy carbon plate used as a working electrode with a surface area of 1.0 cm2, utilizing 0.05 mM catalysts L1L6 in a CO2-saturated DMF/H2O mixture and employing a custom-made electrolyzer with 0.1 M TBAP as the supporting electrolyte. The CPE experiments, spanning three hours or until complete catalyst deactivation, revealed a noteworthy pattern in charge. The potential was systematically varied from −1.6 V to −1.9 V vs. SCE, revealing an increased charge build-up with higher applied potentials (Fig. S33). Upon conducting the rinse test after 3.0 h of electrolysis, followed by subsequent CPE without a catalyst at −1.9 V vs. SCE, no gaseous products were evolved during the bulk electrolysis of blank solution (electrolysis without using a catalyst) or in the rinse test (Fig. S34), which was further confirmed by gas chromatography. GC was employed to analyze the products formed during the electrolysis process. The analysis confirmed the presence of CO2-reducing products such as CO and H2. To validate the identification of CO and H2 produced by CO2 reduction, standard gases of CO and H2 were injected into the GC system to determine their retention times (Fig. S35). Gas chromatography plots for samples labelled L1 to L6 at a specific potential of −1.6 V vs. SCE are provided in Fig. S36–S41. Moreover, the GC plots for detection of CO and H2 using bpy and phen ligands are shown in Fig. S42. During the electrolysis process, gaseous samples were collected using a gas-tight syringe after 3.0 h of electrolysis. The collected samples were then injected into the gas inlet of the GC for analysis. CO and H2 production was quantified at each potential during electrolysis using response factors. The quantities of CO and H2 produced at a potential of −1.7 V vs. SCE are depicted in Fig. 3, providing a visual representation of the amount of each gas generated using organocatalysts L1L6.
image file: d5cy01027f-f3.tif
Fig. 3 Amount of CO (A) and H2 (B) produced after 3.0 h of CPE at −1.7 V vs. SCE of L1L6.

Further, the gaseous products were confirmed using an Omni Star™ Gas Analysis System GSD 320 (Pfeiffer) quadrupole mass spectrometer. Isotopic labelling experiments for catalysts L1L6 were carried out using an online mass analyzer, employing both 12CO2 and 13CO2 isotopes.30 During controlled potential electrolysis (CPE), the gas analyzer's sensor was connected to the headspace of the electrochemical cell. The detected products included hydrogen (H2, m/z = 2), 12CO (m/z = 28), and 13CO (m/z = 29) in the reaction under 12CO2 and 13CO2 atmospheres for the L1L6 catalysts (Fig. S43–S47). A typical online mass and gas measurement during the reduction of 13CO2 by catalysts L1 (Fig. 4(A)–(C)) and L4 (Fig. 4(D)–(F)) is depicted in Fig. 4. The highest faradaic efficiencies for CO and H2 were 59.0% and 0.02% for L6 at −1.6 V vs. SCE after 3 hours of electrolysis (Table 1).


image file: d5cy01027f-f4.tif
Fig. 4 The CPE under 13CO2 using L1 as a catalyst: (A) H2 detection, (B) 13CO detection, (C) and 13CO2 detection using an online gas and mass analyzer. The CPE under 13CO2 using L4 as a catalyst: (D) H2 detection, (E) 13CO detection, and (F) 13CO2 detection using an online gas and mass analyzer.
Table 1 Catalytic comparison for CO2 reduction among L1L6 at −1.6 V vs. SCE
Catalyst CO selectivity (%) FE for CO (%) H2 selectivity (%) FE for H2 (%)
L1 99.6 44.0 0.500 0.20
L2 99.9 22.0 0.046 0.01
L3 99.8 27.0 0.140 0.04
L4 99.9 29.0 0.070 0.02
L5 99.5 46.0 0.050 0.02
L6 99.9 59.0 0.0340 0.02


The faradaic efficiency, TON, and TOF for CO and H2 of all the catalysts (L1L6) after 3 h of CPE at varied potentials of −1.6 V to −1.9 V vs. SCE are summarized in Tables S2–S13 and those of bpy and phen are summarized in Tables S14 and S15 (detailed calculations are given in the SI).

Catalytic stability

The stability of catalysts was essential to ensure the efficiency and durability of the CO2RR process. A stable catalyst can maintain its activity over extended periods of operation. In this context, UV-vis spectroscopy was employed to analyze 0.5 mM L1L6 before and after bulk electrolysis (as shown in Fig. S10A), indicating no significant changes in the spectra (Fig. S48). This suggests that the L1L6 organocatalysts remained stable under the catalytic conditions of CO2 reduction. After 3 hours of electrolysis, a rinse test was conducted for all the catalysts (Fig. S34), confirming both the homogeneous nature of the catalytic conditions and the stability of the L1 to L6 organocatalysts throughout the process. Further, we have taken CVs before and after bulk electrolysis. No spectral changes were observed during the process (Fig. S49). All the catalysts exhibited stability throughout the catalytic process, confirming their reliability in the CO2RR. Furthermore, the homogeneity of the catalysts was validated using ESI mass spectrometry (Fig. S50–S55). The obtained spectra were consistent with the mass spectra recorded prior to electrolysis, confirming the stability of the catalysts.

Furthermore, cyclic voltammetry performed over 30 consecutive cycles under both N2 and CO2 atmospheres did not reveal the appearance of any new peaks for any of the ligands (Fig. S57, S60 and S62), further supporting their stability and homogeneity. Under an N2 atmosphere, the catalytic current remains constant after the addition of a proton source over successive cycles. This clearly indicates that the H2O concentration remains constant during the repetitive cycles. In contrast, when the solution is purged with CO2, a gradual decrease in the catalytic current is observed upon repetitive cycles (Fig. S60 and S62), which we attribute to the saturation of CO2 at the electrode surface.

Kinetic parameters and catalytic efficiency

We investigated the catalytic kinetics of catalysts L1L6, examining the impact of variables such as catalyst concentration and water content under different conditions. The study demonstrated a linear relationship between the inverse square root of the scan rate (ν−1/2) and the ratio of catalytic peak current under CO2 (ic) to peak current under N2 (ip), represented as (ic/ip) (Fig. S56–S62). At higher scan rates, a negative shift in the plateau potential was observed (Fig. S56, S58, S59 and S61). To further probe the catalytic mechanism, scan rate-dependent studies were performed (Fig. S56, S58, S59 and S61). In both N2 and CO2 atmospheres, no saturation of the catalytic current was observed, even at high scan rates, suggesting rapid and efficient electron transfer to CO2. Under N2, only a minimal increase in current was detected with an added proton source, in line with gas chromatography (vide infra), which confirmed negligible H2 evolution. Furthermore, the ratio of catalytic to peak current (ic/ip) decreased with increasing scan rate from 200–2000 mV s−1 (Fig. S63), indicating diffusion-limited catalysis, as expected for a system where substrate (CO2) mass transport governs the overall catalytic current.

To study the kinetics of electrocatalytic CO2 reduction, we analyzed the cyclic voltammograms of catalysts L1L6 with varying catalyst concentrations. We found that the rate of CO2 reduction increased linearly with higher catalyst concentrations (Fig. 5 and S64). The relationship between the catalytic current (icat) and its components is given by the equation icat = ncatFA[cat](DkCO2[CO2])1/2, where ncat = 2 for CO2 reduction to CO, F = 96[thin space (1/6-em)]500 C mol−1, A = 0.071 cm2, D is the diffusion coefficient, kCO2 is the rate constant, and [CO2] is the CO2 concentration in DMF. Moreover, there is a very negligible shift in the onset potential upon raising the catalyst concentration from 0.25 mM to 1.25 mM under a CO2 atmosphere (Fig. 5). Additionally, the current increased with the addition of H2O, which acts as a proton source in CO2 reduction. The choice of proton source significantly affects the efficiency, selectivity, and overall performance of the CO2 reduction reaction. Acidic environments help protonate intermediates like H2CO3, speeding up the reaction rates and optimizing CO2 conversion into desired products.


image file: d5cy01027f-f5.tif
Fig. 5 (A)–(F) CVs of catalysts L1L6 in DMF using 0.1 M TBAP as a supporting electrolyte under N2 (only 0.25 mM catalysts) and CO2 atmospheres with 2.2 M H2O [varying catalyst concentrations from 0.25 mM to 1.25 mM].

Foot of the wave analysis and catalytic Tafel plots

Investigating catalytic systems systematically through the utilization of turnover frequency and overpotential presents a significant challenge due to various factors, including intrinsic catalytic activities, dependence on transport parameters, and ohmic drops. Savéant et al. introduced a logical and non-destructive technique known as foot-of-the-wave analysis (FOWA), which is applied to routine cyclic voltammetry (CV) observations to estimate TOF quickly. This method focuses solely on the catalytic reaction rate in the electrode's reaction–diffusion layer, disregarding contributions from other phenomena like product inhibition, substrate consumption, and catalyst deactivation.31

 
image file: d5cy01027f-t1.tif (1)
The analytical approach from cyclic voltammetry (CV) measurements involves comparing data obtained in the absence and presence of CO2, relying on establishing a linear relationship between the current ratio (i/ip) and a potential-dependent term expressed as 1/{1 + exp[(nF/RT)(EEcat/2)]}. Here, i represents the catalytic current when CO2 is present, while ip denotes the peak current in the absence of CO2. Furthermore, F denotes the Faraday constant, R is the gas constant, T is the absolute temperature, n is the number of electrons required for CO2 to CO conversion, E is the applied potential, and Ecat/2 is the wave catalytic current. Plotting these terms against each other in the foot-of-the-wave region of the CV yields a linear relationship with a specific slope, which equals the expression 2.24(RT/nFν)1/2(kFOWA)1/2, where ν represents the scan rate in V s−1. From this slope, the catalytic rate constant, kFOWA, or TOFmax, for L1L6 was found to be 1105 s−1, 724 s−1, 1589 s−1, 2211 s−1, 1197 s−1 and 2511 s−1, respectively (Fig. S65).

The subsequent analysis revealed slope values and their respective catalytic rate constants for L1L6, indicating that L6 and L4 exhibit higher catalytic rates than the others. This difference can be attributed to the electron density, steric effects, and redox properties of L6 and L4 likely contributing to their superior catalytic activity towards electrochemical CO2 reduction compared to the other terpyridine derivatives. In-depth analysis involves the correlation of the turnover frequency, a kinetic parameter, with the thermodynamic overpotential (η) derived from the standard reduction potential for the CO2 reduction process image file: d5cy01027f-t2.tif and the applied potential (E). This relationship is expressed as image file: d5cy01027f-t3.tif. By plotting TOF against overpotential, a catalytic Tafel plot is generated, providing valuable insights into the electrochemical CO2 reduction process. In the investigation, the onset potential (Eonset) for L1L6 was determined to be −1.46 V, −1.41 V, −1.43 V, −1.46 V, −1.42 V, and −1.45 V vs. SCE, respectively. Additionally, the overpotential (η) and the catalytic half-wave potential (Ecat/2) were evaluated for each complex. These parameters provide crucial information for understanding each catalyst's catalytic activity and performance in electrochemical CO2 reduction (Table 2).

Table 2 List of onset potential, half-wave potential, TOFmax and overpotential of L1L6
Catalyst E onset (V vs. SCE) E 1/2 (V vs. SCE) Slope TOFmax (s−1) Overpotential (η/mV)
L1 −1.46 −1.73 27 1105 730
L2 −1.41 −1.72 22 724 715
L3 −1.43 −1.68 32 1589 687
L4 −1.46 −1.78 38 2211 780
L5 −1.42 −1.73 28 1197 730
L6 −1.45 −1.87 40 2511 870


For L1L6, the overpotentials were 730 mV, 715 mV, 687 mV, 780 mV, 730 mV, and 870 mV, respectively. Correspondingly, the catalytic half-wave potentials (E1/2) were determined to be −1.73 V, −1.72 V, −1.68 V, −1.78 V, −1.73 V, and −1.87 V vs. SCE for L1L6, respectively.

Utilizing catalytic Tafel plots as a benchmarking tool offers valuable insights into the intrinsic catalytic activities of various catalysts, aiding in identifying optimal operating conditions. The primary objective is to improve the catalytic rate (TOF) while minimizing overpotential, thus enhancing the efficiency of the electrochemical CO2 reduction process. An ideal catalyst is anticipated to demonstrate a catalytic Tafel plot that is shifted diagonally to the upper left, indicating superior performance and efficiency (Fig. 6). Upon comparing the catalytic Tafel plots for each catalyst within a single diagram, it becomes evident that L3 is suitable for CO2 reduction.


image file: d5cy01027f-f6.tif
Fig. 6 The catalytic Tafel plots of L1L6 for the comparison of catalytic activity.

The overpotential associated with L3 is measured to be 687 mV, suggesting that the multifaceted molecular structure of L3 endows it with a unique combination of properties that collectively enhance its catalytic activity and selectivity in electrochemical CO2 reduction. This compound represents a promising avenue for developing efficient CO2 conversion catalysts, contributing to advancing sustainable energy technologies.

Spectro-electrochemical studies

Investigating terpyridine derivatives for electrochemical CO2 reduction involves a comprehensive UV-vis spectroelectrochemistry study. This technique allows for real-time monitoring of the electronic changes occurring in the catalysts during the reduction process. In this study, UV-SEC was employed to analyze the behaviour of Tpy derivatives (L1L6) during CO2 reduction. The experiments were conducted in a DMF and H2O mixture solution, using 0.0156 mM (L1L6) with 0.1 M TBAP as a supporting electrolyte under N2 and CO2 atmospheres at −1.9 V vs. SCE (Fig. 7 and S66). Significant changes in the UV-vis spectra of the Tpy derivatives were observed after applying the potential during CPE. Initially, the Tpy derivatives exhibited characteristic absorption bands corresponding to their electronic transitions. During the reduction process, these bands underwent notable shifts and intensity changes, indicating the progression of the catalytic reaction. We noticed during UV-SEC that all the catalysts show a new band at around 310–380 nm, and the molar extinction value appears at around 3000 M−1 cm−1 for both conditions. This happened due to the 2e reduced R-Tpy transferring electrons to the CO2 molecule, and charge transfer occurred from the terpyridine nitrogen atom to CO2 due to n → π* transition. The appearance of new absorption bands and the shifting of existing bands provided insights into the mechanistic pathways of CO2 reduction by the terpyridine derivatives.
image file: d5cy01027f-f7.tif
Fig. 7 The spectral changes during spectro-electrochemical studies for 0.0156 mM under a CO2 atmosphere at −1.9 V vs. SCE using 0.1 M TBAP as a supporting electrolyte for (A) L1, (B) L2, (C) L3, (D) L4, (E) L5, (F) L6.

Infrared spectroelectrochemistry is a powerful tool for investigating the mechanistic pathways of electrochemical CO2 reduction. Organocatalysts have shown promise in CO2 reduction due to their ability to facilitate multi-electron transfer steps and stabilize reactive intermediates. This technique helps to track the formation and consumption of CO2-derived species, such as carbonates, carboxylates, or carbon monoxide. In situ FT-IR spectra were recorded for all organocatalysts (L1L6) using a high-sealing, optically transparent thin-layer electrochemical (OTTLE) cell equipped with CaF2 windows. A Pt mesh was employed as the working electrode, a Pt wire as the auxiliary electrode, and an Ag wire as the pseudo-reference electrode. The measurements were carried out with a 0.5 mM solution of the complex in 4.9 mL DMF containing 0.1 mL H2O and 0.1 M TBAP. Prior to data collection, the solution was purged with N2 followed by CO2 for 15 minutes. Controlled potential electrolysis was performed at −1.8 V vs. SCE. All the organocatalysts L1L6 exhibit bands that are indicative of C[double bond, length as m-dash]O stretching vibrations in protonated carboxylic acid species (N–CO2H) progressively appearing at υstr = 1700–1725 cm−1 corresponding to the formation of N–CO2H species (Fig. 8 and S67).32


image file: d5cy01027f-f8.tif
Fig. 8 ΔTransmittance vs. wavenumber plots, where the initial spectra before electrolysis have been subtracted from the observed spectra during electrolysis for 0.5 mM (A) L3 (red initial and blue final) and (B) L5 (blue initial and red final) under a CO2 atmosphere at −1.8 V vs. SCE using 0.1 M TBAP as a supporting electrolyte.

Density functional theory (DFT) analysis

DFT33 calculation based on Gaussian 16 (ref. 34) was performed to gain insight into the mechanism of the H2 formation and CO2 reduction reaction and the effect of different electron donating and withdrawing functional groups during the reaction. Geometry optimization and frequency calculations of all intermediates and transition states were performed using the MO6L functional.35 For instance, the formation of intermediate species could be inferred from the emergence of new bands in the UV-vis spectra. This approach underscores the potential of terpyridine derivatives as effective catalysts for CO2 reduction, contributing to the advancement of sustainable energy technologies.

Diffusion and polarization functions were utilized along with the Pople basis set36,37 6-31+G(d,p) to optimize all the geometries. The solvent effect of the reaction was considered by the self-consistent reaction field approach using the SMD solvation model38 in DMF solvent. The computational details are provided in the SI. In the electrochemical CO2 to CO formation, the reaction proceeds via several proton and electron transfer steps; hence, it is crucial to consider experimental potential and pH during the calculation of free energy and reduction potential.39 In this study, all free energy profile diagrams were constructed by implementing the experimental pH of 7.4 and applied potential of −1.60 V (vs. SCE). The H2 formation and CO2 reduction reactions were studied in the presence of six different reagents (L1 to L6). The possible reaction mechanism for the H2 formation reaction is shown in Scheme 3. The optimized structures for the intermediates are provided in Fig. S68.


image file: d5cy01027f-s3.tif
Scheme 3 Possible reaction mechanism of the H2 formation reaction on different catalysts at pH = 7.4 and an applied potential of −1.60 V (SCE) (all values are in kcal mol−1).

The H2 formation reaction generally proceeds through two electron and two proton addition reactions. Here, we consider two consecutive reduction reactions followed by two proton addition paths. The first reduction reaction of Int-1 is exogenic in nature and generates Int-2. The potential corrected free energy changes for the reaction vary from 1.9 kcal mol−1 to −8.4 kcal mol−1. The free energy of reduction of the para methoxybenzene derivative is the least exogenic with 1.9 kcal mol−1, and para cyano-benzene is the most exogenic. The result indicates that the electron-withdrawing substitution in the R group (like L6, L1, and L4) accelerates the reduction compared to the electron-donating substitution (L5 and L2). Int-2 is reduced to Int-3 in the subsequent reduction step, and the second reduction reaction is less exogenic than the first one. As expected, the second reduction reaction follows a similar substitution effect on reduction energy trends. In the protonation reaction, proton addition takes place at the N center to make a N–H bond in Int-4 with a 1.01 Å bond distance. All calculated pKa values of each protonation step are provided in Table S16.

Interestingly, the dehydrogenation reaction is more favourable for the electron-donating group than for the electron-withdrawing group. All transition states are confirmed by IRC calculation (Fig. S69). The reaction may proceed in another path through a nine-membered TS-2 transition state. In all cases, the activation energy viaTS-2 is more than the reaction viaTS-1. The reaction viaTS-1 is the most favourable reaction path (Fig. S70).

The DFT mechanistic study indicates that the most favourable path for the two proton and two electron addition reaction proceeds viaInt-5 protonation at two adjacent N-centers. The substrates are crucial in protonation, reduction, and activation energy in dehydrogenation reactions. Protonation at para methoxybenzene is the most favourable, with a reaction energy of −41.9 kcal mol−1. In the second protonation reaction, protonation occurs at adjacent N to make Int-5 or at the third nitrogen to make Int-6. In all the cases, the protonation at the adjacent N is more feasible than that at the third nitrogen atom (Table S17). After adding two protons and two electrons, we explore the dehydrogenation reaction. In the blue path, two N–H bonds dissociate via a six-membered TS-1 transition state to make an H2 molecule, followed by regenerating the active reagent Int-1. The activation energy for the dehydrogenation reaction varies for different reagents from 35.9 kcal mol−1 to 40.3 kcal mol−1 (Fig. 9).


image file: d5cy01027f-f9.tif
Fig. 9 Transition states and relative energy (a) transition state structures of the CO2 reduction reaction and (b) competition of the activation barrier for CO2 reduction and H2 formation reactions.

The high activation energy barrier for the H2 formation reaction indicates a significant competition between the HER and CO2 reduction reaction. The reagents L1L6 participate in the CO2 reduction reaction in the experimental study. Here, a detailed mechanism is proposed in Scheme 4 based on the experimental and theoretical studies. Like previous hydrogen formation reactions, Int-3 is generated after taking two electrons.


image file: d5cy01027f-s4.tif
Scheme 4 A possible CO2 reduction reaction path in the presence of different non-metal catalysts at pH = 7.4 and an applied potential of −1.60 V (SCE) (all values are in kcal mol−1).

When CO2 is introduced, Int-3 undergoes a reaction with CO2 to make Int-7. The CO2 addition reaction is exogenic for L3, L5, L2, and L1, with respective reaction energies of −5.0, −5.7, −6.3, and −4.8 kcal mol−1 (Fig. S71). The reaction is endothermic for L6 and L4 with respective reaction energies of 1.5 and 2.9 kcal mol−1. The electron-donating groups are more efficient for the CO2 capture reaction. Next, two consecutive proton additions occur to regenerate Int-1 with the formation of CO and H2O. In the first protonation reaction, captured CO2 in Int-7 takes one proton to generate Int-8, and the step is exothermic for all the catalysts. In the second protonation reaction, the OH group of Int-8 takes one proton and converts it into water and CO with a regenerating catalyst. The reaction proceeds through the TS-3 transition state with an activation barrier of 8.7, 7.1, 9.5, 8.3, 10.8, and 11.4 kcal mol−1 for catalysts L1L6, respectively. The free energy plot of the CO2 reduction reaction is illustrated in Fig. 10. The overall reaction is exothermic with a −96.4 kcal mol−1 reaction energy.


image file: d5cy01027f-f10.tif
Fig. 10 Energy profile diagram of the CO2 reduction reaction to CO on different non-metal catalysts at pH = 7.4 and an applied potential of −1.60 V (SCE) in the presence of the L1 catalyst.

After two reductions, the reduced intermediate can participate in the H2 evolution reaction, and in the presence of CO2, CO2 capture and CO formation take place. These non-metal catalysts are efficient for CO2 reduction reactions with a competitive H2 evolution process. Fig. 9(b) shows that the energy barrier for the CO2 to CO formation reaction is lower than that for the competitive H2 formation reaction for all six catalysts. The high activation energy barrier for the dehydrogenation step indicates that the CO2 reduction reaction is more feasible than the H2 formation reaction. Our experimental studies (vide supra, Fig. 5) are also nicely fitted with the DFT study in the presence of CO2 reduction reactions that are more feasible than H2 formation.

Catalytic comparison with reported catalysts

The catalytic performances of L1L6 for CO2 reduction were evaluated under identical electrochemical conditions (0.1 M Bu4NClO4 in DMF at −1.6 V vs. SCE), providing insights into their activity and selectivity. Among the terpyridine-based catalysts, L6 exhibited the highest faradaic efficiency for CO production at 59%, followed closely by L5 at 46%. L1 achieved an FE of 44%, while L4 and L3 produced CO with efficiencies of 29% and 27%, respectively. L2 displayed the lowest CO FE at 22% (Table 3). These results indicate that structural modifications, particularly the incorporation of electron-withdrawing or conjugation-extending substituents, significantly impact catalytic performance. When compared to reported catalysts, the terpyridine-based systems operate at a moderate potential of −1.6 V vs. SCE with competitive efficiencies. Benzoate derivatives and aromatic nitriles/esters, while achieving selectivity for oxalate (C2O42−) of 96–99%, require significantly more negative potentials (−1.98 to −2.2 V vs. Ag/AgI and SCE). Pyridinium and pyridine/pyridazine systems excel under milder aqueous conditions and produce MeOH with faradaic efficiencies of 14–30% at potentials as low as −0.55 to −0.75 V vs. SCE. The PAC3 system, an extensively π-conjugated heterohelicene, demonstrates a FE of 35% for HCOOH production at −2.26 V vs. Fc+/Fc.
Table 3 The comparison of the catalytic performances with reported organocatalysts
Catalyst Electrochemical conditions Reported potential (V) Products and selectivity Faradaic efficiency (%) CPE time (h) Ref.
Where PAC3 = bis-imidazolium embedded extensively π conjugated heterohelicene.
L1 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.6%) and H2 (0.5%) CO (44.0) and H2 (0.2) 3.0 This work
L2 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.9%) and H2 (0.046%) CO (22.0) and H2 (0.01) 3.0 This work
L3 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.8%) and H2 (0.14%) CO (27.0) and H2 (0.04) 3.0 This work
L4 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.9%) and H2 (0.07%) CO (29.0) and H2 (0.02) 3.0 This work
L5 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.5%) and H2 (0.05%) CO (46.0) and H2 (0.02) 3.0 This work
L6 0.1 M Bu4NClO4 in DMF −1.6 V vs. SCE CO (99.9%) and H2 (0.03%) CO (59.0) and H2 (0.02) 3.0 This work
bpy 0.1 M Bu4NClO4 in DMF −1.8 V vs. SCE CO (98.87%) and H2 (1.12%) CO (34.0%) and H2 (0.4%) 1.0 This work
phen 0.1 M Bu4NClO4 in DMF −1.8 V vs. SCE CO (99.8%) and H2 (0.19%) CO (36.0%) and H2 (0.7%) 1.0 This work
Benzoate derivatives 0.1 M Bu4NBr in DMF −1.98 V vs. Ag/AgI Dicarboxylic acid 75.0 18
Pyridinium ion over hydrogenated Pd electrodes 0.5 M NaClO4 at pH 5.4 −0.55 V vs. SCE MeOH 30 19.0 19
Aromatic nitriles and esters 0.1 M Bu4NClO4 in DMF −2.2 V vs. SCE C2O42− 96 20
Pyridine and pyridazine 0.5 M KCl at pH 4.7 −0.75 V vs. SCE MeOH 14 and 4 21
PAC3 0.1 M TBAPF6 in 10% H2O/MeCN −2.26 V vs. Fc+/Fc HCOOH 35 6.0 22


Overall, L5 and L6 stand out among the terpyridine-based catalysts due to their efficiency and moderate operating potential, offering a promising metal-free alternative for selective CO2 reduction to CO. Compared to other systems, their performance underscores the potential of structural tuning in organic frameworks to achieve competitive catalytic activity and selectivity.

Conclusion

The present study explains the synthesis and spectroscopic characterization of L1L6 terpyridine derivatives. These catalysts were harnessed for electrocatalytic CO2 reduction in a DMF/water mixture, utilizing TBAP as a supporting electrolyte. Controlled potential electrolysis (CPE) was conducted for all the catalysts, culminating in the generation of CO as the predominant end-product, exhibiting over 90% selectivity. A comparative analysis of the surface phenomenon unveiled a superior rate constant for L6 compared to others. Nevertheless, scrutinizing the catalytic Tafel plots revealed a heightened catalytic efficiency for L3 over all the other catalysts. The oblique displacement of the catalytic Tafel plot towards the upper left quadrant accentuates enhanced performance, indicating L3 as a preferred candidate for CO2 reduction. Furthermore, upon delving into the bulk phenomenon, L6 manifested augmented catalytic efficiency relative to the others, verified by accompanying density functional theory (DFT) calculations. In this context, the electrochemical and chemical (EECC) mechanistic paths are favoured.

Abbreviations

CO2RRCO2 reduction reaction
CPEControlled potential electrolysis
DFTDensity functional theory
FEFaradaic efficiency
FOWAFoot of the wave analysis
GCGas chromatography
HERHydrogen evolution reaction
SEM/EDXScanning electron microscopy with energy dispersive X-ray spectroscopy
SPESpectro electrochemistry
TBAPTetra butyl ammonium perchlorate
bpy2,2′-Bipyridine
phen1,10-Phenanthroline

Author contributions

SkSA: conducted the experiments and wrote the initial draft. KM: conducted the theoretical studies and wrote the manuscript. DR: curated the data and wrote the initial draft. TNM: conducted the labelling experiments. PKJ: conducted the experiments. AB: conducted the experiments. BSM: provided the resources for the DFT calculations and modified the manuscript. PK: provided the resources for online gas and mass analyses and labelled 13CO2 studies. SKP: conceived the idea, provided the resources, and wrote and modified the draft manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

Details of materials and instrumentation, characterisation of the compounds, electrochemistry experiments, bulk electrolysis, gas evolution and cartesian coordinates are available in the supplementary information (SI).

Supplementary information: the electronic spectra, HRMS, magnetic, electrochemical, GC, and geometry-optimized coordinates are provided in the SI. See DOI: https://doi.org/10.1039/d5cy01027f.

Acknowledgements

This work was conducted at the Artificial Photosynthesis Laboratory, Department of Chemistry and Chemical Biology, IIT(ISM) Dhanbad, Jharkhand, India. It was financially supported by CSIR grant 01(3062)/21/EMR-II awarded to SKP. The authors acknowledge the CRF, IIT(ISM) Dhanbad for the HRMS and SEM/EDX data. The theoretical studies were conducted at IIT Hyderabad. The online gas and mass analyses were conducted at IISER Tirupati.

References

  1. M. Aresta, A. Dibenedetto and A. Angelini, Chem. Rev., 2014, 114(3), 1709–1742 CrossRef.
  2. M. Robert, ACS Energy Lett., 2016, 1, 281–282 CrossRef.
  3. S. Ma and P. J. Kenis, Curr. Opin. Chem. Eng., 2013, 2(2), 191–199 CrossRef.
  4. S. Zhang, Q. Fan, R. Xia and T. J. Meyer, Acc. Chem. Res., 2020, 53(1), 255–264 CrossRef PubMed.
  5. S. J. Davis, K. Caldeira and H. D. Matthews, Sci., 2010, 329(5997), 1330–1333 CrossRef PubMed.
  6. K. Talukdar, S. Sinha Roy, E. Amatya, E. A. Sleeper, P. Le Magueres and J. W. Jurss, Inorg. Chem., 2020, 59(9), 6087–6099 CrossRef PubMed.
  7. M. Cheng, Y. Yu, X. Zhou, Y. Luo and M. Wang, ACS Catal., 2018, 9(1), 768–774 CrossRef.
  8. G. Centi and S. Perathoner, Catal. Today, 2009, 148(3–4), 191–205 CrossRef.
  9. W. H. Wang, Y. Himeda, J. T. Muckerman, G. F. Manbeck and E. Fujita, Chem. Rev., 2015, 115(23), 12936–12973 CrossRef PubMed.
  10. J. L. White, M. F. Baruch, J. E. Pander III, Y. Hu, I. C. Fortmeyer, J. E. Park, T. Zhang, K. Liao, J. Gu, Y. Yan and T. W. Shaw, Chem. Rev., 2015, 115(23), 12888–12935 CrossRef PubMed.
  11. C. Costentin, S. Drouet, M. Robert and J.-M. Savéant, Science, 2012, 338(6103), 90–94 CrossRef.
  12. (a) C. W. Machan, M. D. Sampson and C. P. Kubiak, J. Am. Chem. Soc., 2015, 137(26), 8564–8571 CrossRef; (b) H. Ishida, K. Fujiki, T. Ohba, K. Ohkubo, K. Tanaka, T. Terada and T. Tanaka, J. Chem. Soc., Dalton Trans., 1990,(7), 2155–2160 RSC.
  13. (a) M. R. Madsen, J. B. Jakobsen, M. H. Rønne, H. Liang, H. C. D. Hammershøj, P. Nørby, S. U. Pedersen, T. Skrydstrup and K. Daasbjerg, Organomettalics, 2020, 39(9), 1480–1490 CrossRef; (b) M. L. Clark, P. L. Cheung, M. Lessio, E. A. Carter and C. P. Kubiak, ACS Catal., 2018, 8(3), 2021–2029 CrossRef.
  14. (a) C. M. Bolinger, B. P. Sullivan, D. Conrad, J. A. Gilbert, N. Story and T. J. Meyer, J. Chem. Soc., Chem. Commun., 1985,(12), 796–797 RSC; (b) C. M. Bolinger, N. Story, B. P. Sullivan and T. J. Meyer, Inorg. Chem., 1988, 27(25), 4582–4587 CrossRef.
  15. S. T. Ahn, E. A. Bielinski, E. M. Lane, Y. Chen, W. H. Bernskoetter, N. Hazari and G. T. R. Palmore, Chem. Commun., 2015, 51(27), 5947–5950 RSC.
  16. R. Francke, B. Schille and M. Roemelt, Chem. Rev., 2018, 118(9), 4631–4701 CrossRef.
  17. (a) E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly, J. Am. Chem. Soc., 2010, 132, 11539–11551 CrossRef PubMed; (b) P. K. Giesbrecht and D. E. Herbert, ACS Energy Lett., 2017, 2(3), 549–555 CrossRef.
  18. S. Gambino, G. Filardo and G. Silvestri, J. Appl. Electrochem., 1982, 12(5), 549–555 CrossRef.
  19. G. Seshadri, C. Lin and A. B. Bocarsly, J. Electroanal. Chem., 1994, 372(1–2), 145–150 CrossRef.
  20. A. Gennaro, A. A. Isse, J. M. Savéant, M. G. Severin and E. Vianello, J. Am. Chem. Soc., 1996, 118(30), 7190–7196 CrossRef.
  21. E. Portenkirchner, C. Enengl, S. Enengl, G. Hinterberger, S. Schlager, D. Apaydin, H. Neugebauer, G. Knör and N. S. Sariciftci, ChemElectroChem, 2014, 1(9), 1543–1548 CrossRef.
  22. P. Karak, S. K. Mandal and J. Choudhury, J. Am. Chem. Soc., 2023, 145(13), 7230–7241 CrossRef PubMed.
  23. E. C. Constable, J. Lewis, M. C. Liptrot and P. R. Raithby, Inorg. Chem., 1990, 178(1), 47–54 Search PubMed.
  24. J. Patel, K. Majee, E. Ahmad, A. Vatsa, B. Das and S. K. Padhi, ChemistrySelect, 2017, 2(1), 123–129 CrossRef.
  25. G. T. Morgan and F. H. Burstall, J. Chem. Soc., 1932, 3, 20–30 RSC.
  26. G. Morgan and F. H. Burstall, J. Chem. Soc., 1937, 347, 1649–1655 RSC.
  27. C. Costentin, S. Drouet, M. Robert and J. M. Savéant, Science, 2012, 338(6103), 90–94 CrossRef.
  28. S. Roy, B. Sharma, J. Pécaut, P. Simon, M. Fontecave, P. D. Tran, E. Derat and V. Artero, J. Am. Chem. Soc., 2017, 139(10), 3685–3696 CrossRef PubMed.
  29. S. S. Akhter and S. K. Padhi, Eur. J. Inorg. Chem., 2022, 2022(20), e202200206 CrossRef.
  30. (a) M. Zabilskiy, V. L. Sushkevich, D. Palagin, M. A. Newton, F. Krumeich and J. A. van Bokhoven, Nat. Commun., 2020, 11, 2409 CrossRef; (b) R. Kumar, R. Babu, S. Chakrabortty, V. Madhu and E. Balaraman, J. Org. Chem., 2024, 89, 14720–14739 CrossRef PubMed; (c) P. Bhardwaj, T. Devi and P. Kumar, Inorg. Chem. Front., 2023, 10, 7285–7295 RSC; (d) Y. H. Hong, Y. M. Lee, S. Fukuzumi and W. Nam, Chem, 2024, 10, 1755–1765 CrossRef; (e) Kulbir, C. S. A. Keerthi, S. Beegam, S. Das, P. Bhardwaj, M. Ansari, K. Singh and P. Kumar, Inorg. Chem., 2023, 62, 7385–7392 CrossRef PubMed.
  31. C. Costentin, S. Drouet, M. Robert and J. M. Savéant, J. Am. Chem. Soc., 2012, 134(27), 11235–11242 CrossRef CAS.
  32. (a) S. C. Cheng, C. A. Blaine, M. G. Hill and K. R. Mann, Inorg. Chem., 1996, 35(26), 7704–7708 CrossRef; (b) V. Nikolaou, Z. M. Luo, M. Gil-Sepulcre, J. W. Wang, O. Rüdiger, I.-F. Ardoiz, M. Nicaso, J.-B. Buchholz and A. Llobet, Inorg. Chem., 2025, 64, 11796–11806 CrossRef.
  33. W. Kohn and L. J. Sham, Phys. Rev., 1965, 140(4A), A1133–A1138 CrossRef.
  34. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. S. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford CT Search PubMed.
  35. Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120(1), 215–241 Search PubMed.
  36. R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 2008, 72(1), 650–654 CrossRef.
  37. M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, M. S. Gordon, D. J. DeFrees and J. A. Pople, J. Chem. Phys., 1982, 77(7), 3654–3665 CrossRef.
  38. A. V. Marenich, C. J. Cramer and D. G. Truhlar, J. Phys. Chem. B, 2009, 113(18), 6378–6396 CrossRef PubMed.
  39. Y. Sakaguchi, A. Call, K. Yamauchi and K. Sakai, Dalton Trans., 2021, 50(44), 15983–15995 RSC.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.