Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Photocatalytic conversion of CO2 to CO by Ru(II) and Os(II) octahedral complexes: a DFT/TDDFT study

Athanassios C. Tsipis * and Antonia A. Sarantou
Laboratory of Inorganic Chemistry, Department of Chemistry, University of Ioannina, 45110, Greece. E-mail: attsipis@uoi.gr

Received 15th January 2024 , Accepted 21st March 2024

First published on 22nd March 2024


Abstract

The reaction mechanisms of the photocatalytic reduction of CO2 to CO catalyzed by [(en)M(CO)3Cl] complexes (M = Ru, Os, en = ethylenediamine) in the presence of triethanolamine (TEOA), R3N (R = –CH2CH2OH), in DCM and DMF solvents, were studied by means of DFT/TDDFT electronic structure calculations. The geometric and free energy reaction profiles for two possible reaction pathways were calculated. Both reaction pathways studied, start with the 17e, catalytically active intermediate, [(en)M(CO)3+ generated from the first triplet excited state, T1 upon reductive quenching by TEOA which acts as a sacrificial electron donor. In the first possible pathway, TEOA anion binds to the metal center of the catalytically active intermediate, [(en)M(CO)3+ followed by CO2 insertion into the M-OCH2CH2NR2 bond. The latter upon successive protonations releases a metal ‘free’ [R2NCH2CH2OC(O)(OH)] intermediate which starts a new and final catalytic cycle, leading to the formation of CO and H2O while regenarating TEOA. In the second possible pathway, the 17e, catalytically active intermediate, [(en)M(CO)3+ captures CO2 molecule, forming an η1-CO2 complex. Upon 2H+/2e successive protonations and reductions, CO product is obtained along with regenarating the catalytically active intermediate [(en)M(CO)3+. The nature of the proton donor affects the reaction profiles of both mechanisms. The nature of the solvent does not affect significantly the reaction mechanisms under study. Finally, since photoexcitation and T1 reductive quenching are common to both pathways, we have srutinized the photophysical properties of the [(en)M(CO)3Cl] complexes along with their T1 excited states reduction potentials, image file: d4dt00125g-t1.tif. The [(en)M(CO)3Cl] complexes absorb mainly in the UV region while the absolute image file: d4dt00125g-t2.tif are in the range 6.4–0.9 eV.


1. Introduction

The burning of fossil fuels is the main source of carbon dioxide, CO2 released into the atmosphere by anthropogenic activity. Eighty per cent, 80% of the concentration of greenhouse gasses in the atmosphere is due to the human factor and more specifically to combustion for energy production. The energy sector, is to a large extent, based on the direct combustion of fuels, a process that in turn leads to large CO2 emissions.1 Carbon dioxide is one of the most important greenhouse gasses that contributes to the strengthening of this phenomenon on our planet. The main greenhouse gases are the following: carbon dioxide (CO2), water vapor (H2O), methane (CH4), ozone (O3), nitrogen dioxide (N2O) and chlorofluorocarbons (CFC). Of these greenhouse gases, especially water vapor and carbon dioxide play a decisive role in regulating the temperature of the earth's surface and its atmosphere.2 Many studies have been devoted to the way in which the problem of the continuous increase in atmospheric CO2 can be resolved while at the same time there are continuous efforts and studies not only to bind CO2 but at the same time to convert it into “fine chemicals”.3

Among various strategies to achieve this goal, the biggest and most attractive challenge is the efficient conversion of CO2 into useful compounds using sunlight as an energy source.4 Many complexes with a transition metal element as metal center (such as Ni, Fe, Re, Cr, Ir, Mo etc.) have been extensively studied for the homogeneous electrocatalytic and photocatalytic reduction of carbon dioxide.5–16

The majority of research has focused on the photocatalytic reduction of CO2, which has been based on systems that include a light unit consisting of a photosensitizer, (PS) and two catalytic sites (a reduction site and an oxidation site) (Scheme 1).7–10,13–16


image file: d4dt00125g-s1.tif
Scheme 1 Photocatalytic CO2 reduction by a system comprising a PS and two catalytic sites.

At the oxidation site the donor provides an e to the PS after its excitation to the triplet excited state (3MLCT) which is then reductively quenched at the reduction site and finally an e is transferred to CO2. In many cases, however, PS acts not only as a PS but also as a reducing agent.17

So far, the majority of research efforts on the electro/photo-catalytic reduction of CO2, has been put on 4d and 5d transition metal polypyridyl complexes as catalysts.13 Among them, Re and Ru polypyridyl complexes have attracted the major concern, being the subject of numerous studies. In pioneering works,18–20 Lehn et al., studied the electro/photo-catalytic reduction of CO2 to CO by Re(I) and Ru(II) bipyridyl complexes. Although, numerous mechanisms have been proposed for the photochemical reduction of CO2 catalyzed by Re(I) and Ru(II) complexes, none of them could be regarded as universal.16 Among the various mechanisms proposed, the most widespread is based on the formation of either [Re(bpy)(CO)3] or [Re(bpy)(CO)3 very reactive, five-coordinated key intermediates, able to capture CO2.21,22 Another proposed mechanism, based upon experimental23 as well as theoretical24 studies, involves the formation of a Re(I) dimer species i.e. [Re(dmb)(CO)3]2(OCO), (dmb = 4,4′-dimethyl-2,2′-bipyridine) intermediate. In this dimeric Re(I) complex, from which CO is released, the CO2 molecule bridges the two metal centres. Notice however, that the theoretical study24 points out that formation of the Re(I) dimer is entropically unfavourable, while experiment23 showed that this process is relatively slow. Nevertheless, the mechanism of CO2 to CO reductions by Re(I) still remains elusive, especially the CO2 addition to the metal centre.16 Ishitani et al.,23 proposed that even the non-reduced [ReI(N^N)(CO)3]+ 17e species is able to capture CO2 with the assistance of TEOA. Accordingly, they found that during the photocatalytic reduction of CO2 by the fac-[ReI(bpy)(Br)] complex, TEOA is coordinated to the metal centre of the non-reduced [ReI(bpy)(CO)3]+ 17e intermediate forming another intermediate namely the fac-[ReI(bpy)(CO)3(OCH2CH2NR2)] (R = CH2CH2OH). In a subsequent step, CO2 is inserted into the Re–O coordination bond of the fac-[ReI(bpy)(CO)3(OCH2CH2NR2)] intermediate yielding the fac-[ReI(bpy)(CO)3(R2NCH2CH2O-COO)] insertion product, which is proposed to be the key intermediate for the CO2 photocatalytic reduction.23

The other most studied catalysts for electro/photo-catalytic CO2 conversion are the Ru(II) polypyridyl complexes. The light-induced reduction of CO2 to CO in an acetonitrile/water/triethanolamine solution in the presence of the [Ru(2,2′-bipyridine)3]2+/Co2+ system was first reported by Lehn and Ziessel.18 Since then, numerous studies have been devoted to the CO2 reduction by Ru(II) catalysts. The most representative complexes studied are the cis-[Ru(bpy)2(CO)X]n+, [Ru(tpy)(bpy)X]n+ and trans (Cl)-Ru(bpy)(CO)2Cl2 complexes as well as their derivatives.16

The earliest study of electrochemical CO2 reduction by a Ru(II) complex was reported by Tanaka et al.,25 which used a homogeneous [Ru(bpy)2(CO)2]2+ catalytic system to obtain a mixture of CO/formate from CO2 upon electrolysis. The proposed mechanism, involves formation of an unsaturated five coordinate species which after reduction is attacked by CO2 to form the [Ru(bpy)2(CO)(η1-CO2)] intermediate (η1-CO2 adduct). Formation of the η1-CO2 adduct is thought to occur also in the CO2 electrocatalytic reduction by the cis-[Ru(bpy)2(CO)Cl]+ catalysts, which after reduction, loses the chloride ligand forming the unsaturated five coordinate species which in turn is attacked by CO2.26 An alternate combined η1-CO2 complex/CO2 insertion (hydride) mechanism for CO2 reduction has been proposed for Ru(II)-hydride catalysts.16,27–29 Finally, Os(II) complexes are far less studied either as photosensitizers30,31 or catalysts32,33 for CO2 conversion.

Taking into account the scarcity of theoretical studies of CO2 to CO photocatalytic conversion by Ru or Os complexes we instigated to contact a theoretical study, by means of DFT/TDDFT electronic structure calculations, in order to delineate the mechanistic details of the CO2 to CO photocatalytic conversion by simple Ru(II) and Os(II) octahedral model complexes of the general formula [(en)M(CO)3Cl] (en = ethylenediamine, M = Ru, Os) in the presence of TEOA. The aim of our work is twofold: (a) to expand our previous molecular modeling, mechanistic investigation on the photocatalytic CO2 to CO reduction by Re(I) complexes17 to cover also Ru(II) and Os(II) analogue complexes and (b) to theoretically delineate the possibility for the mechanism proposed by Ishitani's seminal work23 on similar Re(I) bipy octahedral complexes to be valid also for Ru(II)/Os(II) complexes. Solvent effects on the reaction mechanisms of the CO2 to CO conversion by [(en)M(CO)3Cl] complexes, were also studied, upon conducting our calculations in the non polar DCM solvent as well as in the polar DMF solvent. Finally, we set out to explore the photophysical properties of these complexes as well as their excited state electrochemistry.

2. Computational details

The geometries of all species have been fully optimized, without symmetry constraints, using the 1997 hybrid functional of Perdew, Burke and Ernzerhof34–39 as implemented in the program Gaussian16.40 This functional uses 25% exchange and 75% weighting correlation and is denoted as PBE0. The Def2-TZVP basis set for all atoms was used for the geometry optimizations. The method used in the DFT calculations will be abbreviated PBE0/Def2-TZVP. All stable structures have been identified as energy minima with number of imaginary frequencies NImag = 0. In detail, the frequencies were calculated at the same level as in theory, and the nature of the fixed points was determined in each case according to the number of negative eigenvalues of the Hessian. The Gibbs free energy was calculated to be 298.15 K and 1 atm pressure. Solvent effects were calculated via the Polarizable Continuum Model (PCM) using the integral equation formalism variant (IEF-PCM), which is the default method (self-consistent reaction field (SCRF)).41 DCM and the more polar solvent DMF were used as solvents. Natural Bond Orbital (NBO) population analysis was performed using the Weinhold methodology.42,43 Time-dependent density functional theory (TD-DFT) calculations44–47 were performed on the ground-state, S0 equilibrium geometries in DCM and DMF solvents using the PBE0/Def2-TZVP/PCM computational protocol, taking account the first 30 excited states.

3. Results and discussion

Based upon previous experimental as well as theoretical evidences16 we employed DFT electronic structure calculations in order to explore two possible reaction mechanisms for the photocatalytic CO2 conversion by [(en)M(CO)3Cl] octahedral complexes. The two possible mechanisms under investigation are: (a) The CO2 to CO conversion, based on experimental studies by Ishitani et al.,23 on Re(I) analogue complexes, where TEOA is thought to act, not only as a ‘sacrificial donor’, but also as a ligand as well, coordinated to the metal center of the 17e very reactive intermediate and (b) a more conventional mechanism where the 17e intermediate, after reduction, captures directly CO2, to form the [(en)M(CO)31-CO2)] intermediate (η1-CO2 adduct).16 Let us start however, with the computational study of the initial step of the reaction, being common to both possible reaction mechanisms under consideration, namely the photo-excitation and subsequent reductive quenching processes of the starting photocatalyst (or pre-catalyst).

3.1 Initial step

The initial step of the photocatalytic CO2 to CO conversion by the [(en)M(CO)3Cl] complexes, is depicted schematically in Fig. 1.
image file: d4dt00125g-f1.tif
Fig. 1 Initial step of the photocatalytic CO2 to CO conversion by the [(en)M(CO)3Cl] complexes.

In the initial step, the [(en)M(CO)3Cl] complexes, upon irradiation, are excited in their first singlet excited state, S1 which via Intersystem Crossing (ISC) yields their first triple excited states, T1. Finally, the [(en)M(CO)3Cl] complexes in their T1 state, receive an e from TEOA forming the so called One Electron Reduced (OER), 19e complex, OER_M.

The optimized geometries of all species in DCM, involved in the initial step of the CO2 → CO conversion by [(en)Ru(CO)3Cl], 1 and [(en)Os(CO)3Cl], 2 complexes are depicted in Fig. 2 while those in DMF are given in Fig. S1 (see ESI). Perusal of Fig. 2 and ES1 reveals that the structural parameters of 1 and 2 in their S0 ground states are not affected and practically remain the same in either the non polar DCM solvent or in the polar DMF solvent. The same holds true also for 1 and 2 in their T1 states. Upon S0 → T1 excitation, the most striking structural change is observed for the axial CO ligand i.e. the Ru–CO bond is significantly elongated by 0.2–0.4 Å while the ∠Ru–C–O bond angle departs from linearity by up to 47°. Also, there is a small change of the Ru–CO bonds of the equatorial CO ligands in the range 0.01–0.07 Å while the ∠COax–Ru–COax and ∠COax–Ru–COeq bond angles change is in the range 1.9–5.4°. The rest of the structural parameters around the coordination sphere exhibit only minor changes.


image file: d4dt00125g-f2.tif
Fig. 2 Optimized geometries of 1 and 2 in their S0 and T1 states and respective OER species in DCM solvent at the PBE0/Def2-TZVP level.

On the other hand, the optimized geometry of the OER_M species in the two solvents, DCM and DMF remains practically unaltered. However, the structural parameters in the coordination sphere of OER_M species are significantly changed as compared to those found for the respective S0 and T1 states. Accordingly, the Ru–Cl bond in OER_M is elongated as compared to the respective bond in the S0 and T1 states. The same holds also for the two Ru–N bonds formed between the metal centers and the en ligand. In contrast, one Ru–COeq equatorial bond in OER_M is lengthen as compared to both the S0 and T1 states. This equatorial COeq ligand is also found to be significantly bended compared to S0 and T1 states. The other Ru–COeq equatorial bond in OER_M is shortened while retaining linearity. Finally, the Ru–COax axial bond in OER_M is similar to that of S0 state while both differ significantly compared to that in T1 state. The bond angles around the coordination sphere in OER_M are similar to those found for the S0 rather the T1 state.

3.1.1 Photophysical properties. Since excitation of the initial photocatalyst/precatalyst, upon light irradiation, is a prerequisite in order to obtain the 17e catalytically active intermediate, we set out to calculate the respective absorption spectra. In Fig. 3 are depicted schematically the simulated absorption spectraof 1 and 2 in both DCM and DMF solvents.
image file: d4dt00125g-f3.tif
Fig. 3 Simulated absorption spectra of 1 and 2 in DCM and DMF solvents at the PBE0/Def2-TZVP level.

In Table 1 are given the respective electronic transitions related to the simulated absorption spectra of 1 and 2. Inspection of Fig. 3 and Table 1 reveals that the two complexes in either DCM or DMF solvents, absorb in the UV region. Upon replacing the non-polar DCM solvent with the polar DMF solvent, has practically no effect on the absorption spectra which remain unaltered. The absorption spectra of 1, in either DCM or DMF solvents, exhibit one band with peak around 180 nm and a shoulder at around 250 nm. The band arises mainly from an electronic excitation at 179 nm which based upon the shapes of the MOs involved in the respective electronic excitation (Fig. S2), could be assigned as MLCT/MC/IL. The same holds also true for the shoulder peaking around 250 nm.

Table 1 Principal singlet–singlet electronic transitions in the simulated absorption spectra, for 1 and 2 calculated in DCM and DMF solvents at the PBE0-GD3BJ/Def2-TZVP level of theory
Excitation (%composition) E (eV) λ (nm) f
1_DCM
H−5 → L (32%), H−3 → L (23%), H−3 → L+2 (11%) 6.92 179 0.354
H−1 → L+1 (70%) 4.83 257 0.015
1_DMF
H−5 → L (43%), H−3 → L (15%) 6.96 178 0.373
H−1 → L+1 (73%) 4.86 255 0.014
2_DCM
H−4 → L+2 (12%), H−3 → L+2 (30%), H−2 → L+7 (13%) 7.32 169 0.060
H−5 → L+1 (15%), H−2 → L+6 (48%) 7.25 171 0.081
H−2 → L+4 (25%), H−2 → L+5 (10%), H−2 → L+6 (25%) 7.16 173 0.074
H−1 → L+7 (19%), H → L+7 (20%) 6.94 179 0.110
H−1 → L+2 (57%), H → L+2 (18%) 5.83 213 0.014
H−2 → L (36%), H−2 → L+1 (36%) 5.79 214 0.020
H−1 → L+1 (53%), H → L (30%) 5.13 242 0.043
2_DMF
H−4 → L+2 (15%), H−3 → L+2 (36%) 7.33 169 0.057
H−5 → L (12%), H−2 → L+6 (43%) 7.26 171 0.067
H−2 → L+4 (−23%), H−2 → L+6 (33%) 7.17 173 0.064
H−1 → L+7 (17%), H → L+7 (20%) 6.97 178 0.113
H−2 → L (14%), H−2 → L+1 (57%) 5.79 214 0.020
H−1 → L+1 (54%), H → L (22%) 5.16 240 0.042


The simulated absorption spectra of 2 in DCM and DMF solvents exhibit one high energy strong band, peaking around 170 nm and a lower energy, weak band at around 240 nm. In addition, there is a shoulder around 215 nm. Based upon the shapes of the MOs involved in the electronic excitations (Fig. S2) the bands and the shoulder are assigned as MLCT/MC/IL.

Since the T1 excited states of the complexes under study is of central importance in Step A, we set out to explore their properties in more detail. In Fig. 4 are depicted the spin density along with the SOMOs for the T1 excited state of 1 and 2.


image file: d4dt00125g-f4.tif
Fig. 4 (a) Spin density isosurfaces (0.005 au) along with atomic spin densities of T1 state of complexes 1 and 2 in DCM solvent (numbers in parenthesis refer to DMF solvent) and (b) 3D contour plots of the relevant SOMOs at the PBE0-GD3BJ/Def2-TZVP level of theory.

Inspection of Fig. 4 reveals that the spin density in the T1 state of 1 and 2 is mainly accumulated on the metal center and to a lesser degree on the axial chloride and carbonyl ligands. The spin density distribution resembles the 3D shapes of the SOMOs of 1 and 2 in the T1 state.

3.1.2 Excited state redox potentials. Taken into account that formation of the 17e catalytically active intermediate is preceded by the reductive quenching of the first triplet excited state, T1 of the pre-catalyst, we set out to calculate the redox potential for this excited state. The T1 excited-state redox potentials were estimated upon employing the procedure followed by Hansen et al.48 First, we calculated the ground-state reduction potentials, image file: d4dt00125g-t3.tif based upon the Born–Haber cycle shown in Scheme 2. Then, we calculated the excited state reduction potential, image file: d4dt00125g-t4.tif according to the Latimer diagram depicted also in Scheme 2.
image file: d4dt00125g-s2.tif
Scheme 2 Born–Haber cycle employed in the calculation of the ground state reduction potentials, image file: d4dt00125g-t11.tif (above) and Latimer diagram employed for the calculation of T1 excited state reduction potentials, image file: d4dt00125g-t12.tif (below).

The standard absolute ground-state reduction potential, image file: d4dt00125g-t5.tif is calculated by the following equation:

 
image file: d4dt00125g-t6.tif(1)
where F is the Faraday constant (23.061 kcal per volt gram equivalent) and Z is unity for one-electron redox processes. ΔG°(soln., redox) is obtained from the following equation:
 
image file: d4dt00125g-t7.tif(2)

The calculated Gibbs free energies appearing in Scheme 2 are given in Tables S1 and S2 of the ESI along with 0–0 transition energies for the T1 states, E0–0 in DCM and DMF solvents. The E0–0 are obtained as the differences of the zero point energy corrected total electronic energies of 1 and 2 in their S0 and T1 states in DCM and DMF solvents. The calculated T1 excited state absolute reduction potentials, image file: d4dt00125g-t8.tif are 6.4 and 6.3 V for 1 in DCM and DMF respectively, while for 2 are 6.4 and 0.9 V in DCM and DMF respectively. Taking into account that SHE potential is −4.281 V,49 the image file: d4dt00125g-t9.tifvs. SHE for 1 in DCM and DMF are 2.1 and 2.0 Volt respectively while for 2 in DCM and DMF are 2.0 and −3.4 V.

3.2 Catalytic cycle

After the photoexcitation and reductive quenching processes in the Initial step of the photocatalytic CO2 to CO conversion, it follows the ‘actual’ catalytic cycle, starting with the five coordinated, 17e intermediate. We examined two possible mechanisms for the catalytic cycle. The first possible mechanism is based on the concept of the formation of a complex between the 17e catalytically active intermediate and TEOA, acting as a ligand. Henceforth, we will call this the TEOA complex mechanism. The second possible mechanism under investigation, is based on the direct CO2 capture by the 17e catalytically active intermediate which henceforth we will call it η1-CO2 complex mechanism.
3.2.1 TEOA complex mechanism. This proposed mechanism is based on the formation of a complex between TEOA and the 17e five coordinated intermediate obtained from OER_M. This concept is thought23,50 to be important for the CO2 reduction in low concentrations e.g. in the air, and its utilization from dilute CO2 sources. However, obtaining CO via this mechanism still remains elusive. Accordingly, we employed DFT electronic structure calculations to study in more depth the concept of the TEOA complex mechanism applying it for the Ru(II) and Os(II) catalysts under consideration. The proposed catalytic cycle is depicted in Fig. 5 and comprising two steps: (a) the TEOA complex formation step and (b) the CO release step.
image file: d4dt00125g-f5.tif
Fig. 5 Proposed TEOA complex mechanism for CO2 to CO conversion by 1 and 2.

3.2.1.1 TEOA complex formation step. The TEOA complex formation step starts with the 17e five coordinated, very reactive intermediates, ImA_M, obtained from OER_M, and considered to be the ‘true’ catalysts. Next, TEOA upon deprotonation, is coordinated, as an anionic ligand, TEOA, to the metal center of ImA_M, forming the [(en)(CO)3M(OCH2CH2NR2)] intermediates, ImB_M. The CO2 molecule could then be captured via an insertion reaction, where it is inserted into the M–O bond of ImB_M yielding the [(en)(CO)3MO(O)COCH2CH2NR2] key intermediates, ImC_M which are of significant importance.23 The latter, upon protonation, dissociate to form the [R2NCH2CH2OC(O)OH]+ species, ImD_M and giving back the ‘true’ catalysts, ImA_M. The free energy, ΔG profiles calculated for the TEOA complex formation cycle are depicted schematically in Fig. 6 and 7 for the Ru(II) complex 1 and Os(II) complex 2 respectively, in both DCM and DMF solvents.
image file: d4dt00125g-f6.tif
Fig. 6 Free energy, ΔG (in kcal mol−1), reaction profiles of TEOA complex cycle for 1 in DCM and DMF (numbers in parenthesis) solvents calculated at the PBE0/Def2-TZVP level.

image file: d4dt00125g-f7.tif
Fig. 7 Free energy, ΔG (in kcal mol−1), reaction profiles of TEOA complex cycle for 1 in DCM and DMF (numbers in parenthesis) solvents calculated at the PBE0/Def2-TZVP level.

Notice that, in the mechanistic profiles depicted in Fig. 6 and 7, the ‘magic’ TEOA sacrificial electron donor is thought of acting as the proton donor as well. Perusal of Fig. 6 and 7 reveals that the TEOA complex formation step is expected to be endergonic in the case where TEOA is thought to be the proton donor source. Initially, coordination of TEOA ligand to the metal center of either ImA_Ru or ImA_Os is exergonic in either DCM or DMF solvents. This process is slightly more favorable for Os(II) complex, 2 as compared to the Ru(II) complex, 1. In addition, the TEOA coordination to the metal centers is clearly more favorable in DCM rather in DMF solvent. Next, the CO2 insertion into the M–O bond of ImB_M species proceeds via a transition state, TS1_M, leading to intermediate ImC_M (Fig. 6 and 7). The calculated energy barrier for the formation of ImC_M from ImB_M is estimated to be in the range 36–42 kcal mol−1 at the PBE0/Def2-TZVP level of theory. It should be noticed that, this process is slightly more favorable for Ru(II) complex, 1 rather for Os(II) complex, 2, since the energy barrier for the former is lower than for the latter. Finally, upon protonation of ImC_M, with TEOA as the proton source, we obtain the ImD intermediate and ImA_Mvia a somewhat endergonic process.

Taking into account that for similar Re(I) complexes the proton obtained from the deprotonated form of TEOA should be trapped by another TEOA molecule23 we set out to explore this possibility also for the Ru(II) and Os(II) complexes under study. Inspection of Fig. 6 and 7 reveals that protonation with TEOA(H)+, as the proton source, is exergonic in contrast to the process where TEOA is used as a proton source.

The relatively weak Ru–O coordination bond, formed upon coordination of the R2NCH2CH2O (R = CH2CH2OH) ligand to the Ru metal center of the [(en)Ru(CO)3] intermediate, ImA_Ru, is reflected to the relatively small Wiberg Bond Indices (WBI) obtained from the NBO analysis of ImB_Ru, being equal to 0.582 and 0.581 in DCM and DMF solvents respectively.

In addition, NBO population analysis revealed that in DCM and DMF solvents both the Ru metal center and the O donor atom of the R2NCH2CH2O ligand acquire negative natural atomic charges equal to −0.850 and −0.707|e| respectively, resulting in a repulsive electrostatic interaction. On the other hand, in both DCM and DMF solvents, the covalent component of the Ru–O coordination bond arises from a σ(Ru–O) bonding NBO, formed upon interaction of the sp4.90d4.24 hybrid orbitals (48.3% p and 41.7% d character) of Ru with the sp2.60 hybrid orbital (72.2% p-character) of the oxygen donor atom of R2NCH2CH2O ligand and is described as σ(Ru–O) = 0.429hRu + 0.903hO. The occupation number of σ(Ru–O) NBO is 0.958|e|. The σ(Ru–O) NBO is depicted schematically in Scheme 3.


image file: d4dt00125g-s3.tif
Scheme 3 3D surface of the σ(Ru–O) bonding NBO of ImB_Ru.

It follows the insertion of a CO2 molecule into the Ru–O bond of ImB_Ru yielding ImC_Ruvia an exergonic process with ΔG equal to −31 to −21 kcal mol−1 in DCM and DMF solvents respectively. Formation of ImC_Ru procceds via transition state TS1_Ru, exhibiting a relatively low energy activation barrier calculated to be 8 and 19 kcal mol−1 in DCM and DMF solvents respectively (Fig. 6). Subsequently, ImC_Ru could undergo protonation at three possible O atom sites of the R2NCH2CH2O(O)CO ligand denoted as Oa, Ob and Oc (Fig. 5). The natural atomic charges on the Oa and Ob atoms are −0.719 and −0.649 respectively in DCM solvent and −0.724 and −0.648 respectively in DMF solvent. Protonation of either Oa or Ob of the R2NCH2CH2OC(O)O ligand leads to cleavage of the Ru–O bond, yielding ImD and regenerating the initial 17e catalytic species ImA_Ru, thus closing the catalytic cycle (Fig. 5).

The third possible protonation site i.e. Oc acquires a natural atomic charge of only −0.519, thus being less prone to protonation as compared to the other two possible protonation sites, Oa and Ob of ImC_Ru. However, the protonation of the third possible site, Oc yields back ImA_Ru, CO2 and TEOA.

Protonation of ImC_Ru, with TEOA proton donor, at Oa and Ob sites, leading to Ru–O bond cleavage to yield ImA_Ru and ImD, is estimated to be a slightly endergonic process with ΔG around to 12 kcal mol−1 in both DCM and DMF solvents respectively (Fig. 6). In contrast, if TEOA(H)+ is the proton donor, this process is exergonic with ΔG equal to 77 and 62 kcal mol−1 in DCM and DMF solvents respectively.

The calculated free energy reaction profile for 2 is quite similar with that calculated for 1 (Fig. 7). The anionic TEOA ligand interacts with the Os metal center of the highly active 17e intermediate ImA_Os, giving the neutral radical [(en)Os(CO)3(TEOA)]˙ intermediate, ImB_Os. The estimated interaction energy (IE) of TEOA ligand with the Os metal center of ImB_Os is equal to −44.6 kcal mol−1 in DCM solvent and −34.3 kcal mol−1 in DMF solvent.

Obviously, TEOA interacts more strongly with Os rather than the Ru metal center of ImA_M. The conversion of ImA_Os to the ImB_Os intermediate is slightly more exergonic (ΔG = −33 and −22 kcal mol−1 in DCM and DMF solvents respectively) as compared to the respective process calculated for the Ru counterparts.

NBO analysis indicates that the TEOA ligand coordinates to the Os metal center of ImA_Os, forming a relatively weak Os–O bond in the ImB_Os intermediate. The estimated WBI for Os–O bond is 0.528 in DCM solvent and 0.529 in DMF solvent. In DCM solvent both the Os metal center and the O donor atom of the coordinated TEOA ligand acquire negative natural atomic charges of −0.397 and −0.761|e| respectively. In DMF solvent, the respective charges are −0.400 and −0.757|e| respectively. The electrostatic interaction between TEOA ligand with Os metal center of ImB_Os is unfavorable similar to that found for the respective Ru intermediate, ImB_Ru.

Formation of ImC_Os for 2i.e. CO2 insertion into the Os–O(TEOA) bond is calculated to be slightly more exergonic as compared to its Ru counterpart (Fig. 6 and 7). It follows protonation of ImC_Os at either Oa or Ob sites leading ultimately to the formation of ImA_Os and ImE intermediates (Fig. 7). The protonation sites Oa and Ob acquire natural atomic charges of −0.712 and −0.728|e| respectively in DCM solvent and −0.728 and −0.697|e| respectively in DMF solvent. Protonation of ImC_Os, with TEOA, accompanied by the cleavage of the Ru–O bond, is slightly endergonic, similar to that found for the protonation of the respective Ru intermediate, ImC_Ru. In contrast, protonation with TEOA(H)+ is exergonic just as for its Ru counterpart.

Notice that, as in the case of 1, there is a third protonation site at Oc (Fig. 5) being less probable however, due to the lower negative natural charge calculated for Oc, being equal to −0.516 and −0.508|e| in DCM and DMF solvents respectively. The protonation of the third possible site, Oc yields back ImA_Os, CO2 and TEOA similar with the respective process being observed for 1 (vide supra). Finally, the optimized geometries of all species involved in the catalytic cycle of TEOA complex formation step (Fig. 5) along with selected structural parameters are given in Fig. S3 and S4 of the ESI.


3.2.1.2 CO release step. The second step of the TEOA complex mechanism, starts with the [R2NCH2CH2OC(O)OH]+ species, ImD, being the product of protonation of ImC_M during the preceding TEOA complex formation step (Fig. 5). The energetic profile calculated for the CO release step is depicted in Fig. 8. As for the TEOA complex formation step (vide supra), we examined two possible reaction mechanisms i.e. one where TEOA acts as the proton source and another where protonated TEOA, TEOA(H)+ is the proton source. There are three possible protonation sites in ImD which are denoted as Oa, Ob and Oc (Fig. 8). The natural atomic charges of the Oa, Ob and Oc atoms of ImD in DCM are −0.644, −0.680 and −0.476|e| respectively while in DMF are −0.648, −0.679 and −0.477|e| respectively. Thus, based upon the natural charges, the most probable protonation site is Ob followed by Oa and Oc. Protonation at Ob yields directly a protonated derivative of ImD namely intermediate ImF. The latter could also be formed, via a transition state TS, upon protonation at the second most probable site of ImDi.e. Oa (Fig. 8). Finally, protonation of ImF at site Oc (protonation at site Oa does not lead to products formation) yields CO and H2O while regenerating TEOA.
image file: d4dt00125g-f8.tif
Fig. 8 Free energy, ΔG (in kcal mol−1), reaction profiles for CO release step in DCM and DMF (numbers in parenthesis) solvents, with TEOA and TEOA(H)+ as proton donors (blue and green lines respectively), calculated at the PBE0/Def2-TZVP level.

It should be stressed here, that the most facile route to products formation in the TEOA complex formation step is expected to proceed via protonation of ImD at site Ob, with TEOA(H)+ as the proton source since in this case direct formation of ImF exhibits a relatively low energy barrier around 20 kcal mol−1. All the other paths examined i.e. protonation of ImD at either site Ob, with TEOA proton donor or at site Oa, with TEOA/TEOA(H)+ proton donors exhibit very high energetic barriers rendering them unfavorable.

The TEOA complex formation step yielding CO, H2O and regenerating TEOA is calculated to be strongly exergonic for all routes examined with ΔG ranging from −98 to −268 kcal mol−1.

The optimized geometries of all species involved in the catalytic cycle of CO release step (Fig. 8) are given in Fig. S5 of the ESI.

3.2.2 η1-CO2 complex mechanism. The second possible mechanism, examined for the CO2 electro/photocatalytic conversion by Ru(II)/Os(II) complexes under study, is based on the direct CO2 capture by the 17e catalytically active intermediate. This mechanism is the most popular and widespread to people working in this area. The proposed catalytic cycle, involving the η1-CO2 complex formation, is depicted schematically in Fig. 9.
image file: d4dt00125g-f9.tif
Fig. 9 Proposed η1-CO2 complex mechanism for CO2 to CO conversion by 1 or 2.

It can be seen that, in the η1-CO2 complex mechanism, the five coordinated, 17e catalytically active intermediate, ImA_M is one electron reduced with TEOA to yield the five coordinated intermediate ImB′_M. This step has been proposed also to occur during the photocatalytic CO2 to CO reduction by the trans-Ru(bpy)(CO)2Cl2 complex.51 Next, CO2 is captured by the reactive ImB′_M species forming the η1-CO2 complex, ImC′_M which is considered to be a key intermediate. The CO2 capture/activation step proceeds via transition state TS1′_M. The formation of ImC′_M is followed by two successive protonations of ImC′_M with TEOA acting as the proton source, yielding subsequently ImD′_M and ImE′_M intermediates (Fig. 9). The ImE′_M intermediate loses a water molecule, producing the tetracarbonyl intermediate, ImF′_M. This process proceeds via transition state TS2′_M. Finally, ImF′_M receives an e from TEOA forming the reduced ImG′_M intermediate which is then yields CO and the initial catalytic species ImA_M.

The optimized geometries of all the species participating in the catalytic cycles depicted in Fig. 9, along with selected structural parameters are given in Fig. S6 and S7 for Ru and Os species respectively in both DCM and DMF solvents. The calculated free energy, ΔG profiles for 1 and 2 in DCM and DMF solvents are given in Fig. 10 and 11 respectively. Inspection of Fig. 10 and 11 reveals that the one electron reduction of ImA_M with TEOA to yield ImB′_M is endergonic. Next, we have the formation of ImC′_Mvia transition state TS1′_M with an energy barrier estimated to be in the range 48–50 kcal mol−1. The protonation of ImC′_M to yield ImD′_M could proceed either with TEOA or with protonated TEOA(H)+. However, protonation with the former is predicted to be unfavourable while for the latter it is favourable. Subsequent protonation of ImD′_M, with TEOA (blue lines, Fig. 10 and 11) to yield ImE′_M as well as formation of the tetracarbonyl intermediates, ImF′_M and ImG′_Mvia transition state TS2′_M are expected to be highly unfavourable exhibiting quite high energy barriers. Finally, formation of CO giving back also the initial intermediate ImA_M is an endergonic process. In contrast to TEOA, the path with TEOA(H)+ (green lines, Fig. 10 and 11) is expected to be more favourable.


image file: d4dt00125g-f10.tif
Fig. 10 Free energy, ΔG (in kcal mol−1), reaction profile for the CO2 to CO conversion by 1, following the η1-CO2 complex mechanism, in DCM and DMF (numbers in parenthesis) solvents, with TEOA as proton donor (blue lines) and with TEOA(H)+ (green lines), calculated at the PBE0/Def2-TZVP level.

image file: d4dt00125g-f11.tif
Fig. 11 Free energy, ΔG (in kcal mol−1), reaction profile for the CO2 to CO conversion by 2, following the η1-CO2 complex mechanism, in DCM and DMF (numbers in parenthesis) solvents, with TEOA as proton donor (blue lines) and with TEOA(H)+ (green lines), calculated at the PBE0/Def2-TZVP level.

Accordingly, protonation of ImD′_M, with TEOA(H)+ is almost barrierless, while formation of the tetracarbonyl intermediates proceeds viaTS2′_M exhibiting a relatively low energy activation barrier. The path followed upon employing TEOA(H)+ is predicted to be slightly endergonic.

Since ImC′_M is of key importance for CO2 to CO conversion we set out to analyse in more depth the nature of the M–CO2 bond by means of NBO analysis method. Thus, the natural charges on Ru and C atoms of the Ru–CO2 bond in ImC′_Ru are equal to −1.307 and 0.794|e| in DCM and −1.302 and 0.790|e| in DMF solvent. On the other hand, the natural charges on Os and C atoms of the Os–CO2 bond in ImC′_Os are equal to −0.817 and 0.685|e| in DCM and −0.813 and 0.682|e| in DMF solvent. Obviously, there is favourable electrostatic attraction between the constituent atoms of the M–CO2 bond. In addition, the M–CO2 bond is expected to have a small covalent component as well, since the WBIs are 0.660 and 0.671 for the Ru–CO2 bond in DCM and DMF solvents respectively while the respective WBIs for the Os–CO2 bond are 0.672 and 0.679. The covalent nature of the M–CO2 is demonstrated also by the existence of a bonding NBO (Scheme 4).


image file: d4dt00125g-s4.tif
Scheme 4 3D surface of the σ(Ru–C) bonding NBO of ImC_Ru.

The above bonding NBO is formed upon interaction of the sp3.28d2.70 hybrid orbitals (47.0% p and 38.6% d character) of Ru with the sp2.35 hybrid orbital (70.1% p-character) of the carbon atom of CO2 and is described as σ(Ru–O) = 0.702hRu + 0.713hC. The occupation number of σ(Ru–C) NBO is 1.752|e|.

4. Conclusions

The photocatalytic CO2 to CO reduction by [(en)M(CO)3Cl], Ru(II) and Os(II) octahedral complexes was studied by means of DFT calculations. In an initial step, the [(en)M(CO)3Cl] complexes, upon irradiation, are excited in their T1 state which upon one electron reduction with TEOA, are converted to the OER species. It follows the main catalytic cycle which starts with a 17e five coordinated species, obtained from OER upon losing its chloride ligand. Two possible mechanisms were examined for the main catalytic cycle: (a) the TEOA complex mechanism and (b) the η1-CO2 complex mechanism. At the beginning, we scrutinized the initial step of the photocatalytic process under study. The TDDFT simulated absorption spectra of the initial photocatalysts, [(en)M(CO)3Cl] in DCM and DMF solvents, exhibit absorption bands mainly in the UV region. The polarity of the solvent does not affect the absorption spectra. In contrast, the calculated T1 excited state reduction potentials, image file: d4dt00125g-t10.tif depend upon the solvent and reveal that 1 and 2 in DCM are the easiest to be reductively quenched.

The TEOA complex mechanism proposed for the main catalytic cycle, succeeding the Initial Step, is based on the formation of the complex [(en)M(CO)3(OCH2CH2NR2)] between the 17e five coordinated catalytically active intermediate with TEOA anion acting as O-donor ligand. A similar Re(I) complex as well as its CO2 insertion product have already been experimentally observed.23 The TEOA complex mechanism is thought to occur in two steps. In the first step, the catalytically active intermediate Im_A is regenerated and a metal ‘free’ intermediate, Im_D is also produced. The latter, starts the second step were via two possible pathways, produces CO, TEOA and H2O. It is anticipated that the proton donor source should play a crucial role in the TEOA complex mechanism. Thus, if TEOA is the proton donor, the TEOA complex formation step is predicted to be slightly endergonic while if TEOA(H)+ is the proton donor, this step is strongly exergonic. The CO release step is exergonic using either TEOA or TEOA(H)+ proton donors. However, use of TEOA proton donor in the CO release step, results in very high activation energy barriers in contrast to TEOA(H)+ proton donor where the activation energy barriers are much lower. Therefore, this step is expected to be kinetically more favorable if the proton source is TEOA(H)+.

The η1-CO2 complex mechanism is based on earlier studies proposing a two electron – two proton process with TEOA acting as a sacrificial anode. This mechanism starts also with the 17e catalytically active intermediate which upon one electron reduction, captures CO2 and forming a loosely bound η1-CO2 complex. Successive protonations and one electron reduction yields CO while regenerating the 17e catalytically active intermediate. The whole process is predicted to be endergonic though it has to surmount a significant activation barrier. The latter should be alleviated in the case of electrocatalytic CO2 to CO reduction by the complexes under study. As is for the TEOA complex mechanism, the η1-CO2 complex mechanism is favored with TEOA(H)+ proton donor.

The geometric and the energetic profiles of both mechanisms are not significantly affected by the nature of the solvent. We believe that, the TEOA complex mechanism should be of importance in low CO2 concentrations23 while the η1-CO2 complex mechanism is a more general approach for CO2 to CO electrocatalytic/photocatalytic conversion.

Overall, the present theoretical study, being a follow up of our previous work on Re(I) complexes,17 is an extension to cover also the photocatalytic CO2 to CO conversion by similar Ru(II)/Os(II) simple model complexes, highlighting the role of the TEOA ‘magic’ sacrificial donor. Our molecular modeling works on Re(I), Ru(II) and Os(II) complexes proposes for the first time an alternative ‘unconventional’ mechanism which could trigger further experimental works aiming at fully delineating the photocatalytic CO2 to CO reduction using TEOA.

Conflicts of interest

There are no conflicts to declare.

References

  1. R. Quadrelli and S. Peterson, Energy Policy, 2007, 35, 5938–5952 CrossRef.
  2. B. J. Mason, Contemp. Phys., 1989, 30, 417–432 CrossRef.
  3. M. Cokoja, C. Bruckmeier, B. Rieger, W. A. Herrmann and F. E. Kühn, Angew. Chem., Int. Ed., 2011, 50, 8510–8537 CrossRef CAS PubMed.
  4. A. J. Morris, G. J. Meyer and E. Fujita, Acc. Chem. Res., 2009, 42, 1983–1994 CrossRef CAS PubMed.
  5. J. A. Keith, K. A. Grice, C. P. Kubiak and E. A. Carter, J. Am. Chem. Soc., 2013, 135, 15823–15829 CrossRef CAS PubMed.
  6. C. Riplinger, M. D. Sampson, A. M. Ritzmann, C. P. Kubiak and E. A. Carter, J. Am. Chem. Soc., 2014, 136, 16285–16298 CrossRef CAS PubMed.
  7. H. Koizumi, H. Chiba, A. Sugihara, M. Iwamura, K. Nozaki and O. Ishitani, Chem. Sci., 2019, 10, 3080–3088 RSC.
  8. Y. Asai, H. Katsuragi, K. Kita, T. Tsubomura and Y. Yamazaki, Dalton Trans., 2020, 49, 4277–4292 RSC.
  9. I. Choudhuri and D. G. Truhlar, J. Phys. Chem. C, 2020, 124, 8504–8513 CrossRef CAS.
  10. Y. S. Gong, J. Fan, V. Cecenc and C. Huang, Chem. Eng. J., 2021, 405, 12613–12627 Search PubMed.
  11. K. S. Rawat, A. Mahata, I. Choudhuri and B. Pathak, J. Phys. Chem. C, 2016, 120, 8821–8831 CrossRef CAS.
  12. D. H. Apaydin, S. Schlager, E. Portenkirchner and N. S. Sariciftci, ChemPhysChem, 2017, 18, 3094–3116 CrossRef CAS PubMed.
  13. N. Elgrishi, M. B. Chambers, X. Wang and M. Fontecave, Chem. Soc. Rev., 2017, 46, 761–796 RSC.
  14. K. E. Dalle, J. Warnan, J. J. Leung, B. Reuillard, I. S. Karmel and E. Reisner, Chem. Rev., 2019, 119, 2752–2875 CrossRef CAS PubMed.
  15. Y. Yamazaki, H. Takeda and O. Ishitani, J. Photochem. Photobiol., C, 2015, 25, 106–137 CrossRef CAS.
  16. Y. Kuramochi, O. Ishitani and H. Ishida, Coord. Chem. Rev., 2018, 373, 333–356 CrossRef CAS.
  17. A. C. Tsipis and A. A. Sarantou, Dalton Trans., 2021, 50, 14797–14809 RSC.
  18. J.-M. Lehn and R. Ziessel, Proc. Natl. Acad. Sci. U. S. A., 1982, 79, 701–704 CrossRef CAS PubMed.
  19. J. Hawecker, J.-M. Lehn and R. Ziessel, J. Chem. Soc., Chem. Commun., 1983, 536–538 RSC.
  20. J. Hawecker, J.-M. Lehn and R. Ziessel, Helv. Chim. Acta, 1986, 69, 1990–2012 CrossRef CAS.
  21. J. M. Smieja and C. P. Kubiak, Inorg. Chem., 2010, 49, 9283–9289 CrossRef CAS PubMed.
  22. H. Takeda, K. Koike, H. Inoue and O. Ishitani, J. Am. Chem. Soc., 2008, 130, 2023–2031 CrossRef CAS PubMed.
  23. T. Morimoto, T. Nakajima, S. Sawa, R. Nakanishi, D. Imori and O. Ishitani, J. Am. Chem. Soc., 2013, 135, 16825–16828 CrossRef CAS PubMed.
  24. J. Agarwal, E. Fujita, H. F. Schaefer and J. T. Muckerman, J. Am. Chem. Soc., 2012, 134, 5180–5186 CrossRef CAS PubMed.
  25. H. Ishida, K. Tanaka and T. Tanaka, Chem. Lett., 1985, 14, 405–406 CrossRef.
  26. H. Ishida, K. Tanaka and T. Tanaka, Organometallics, 1987, 6, 181–186 CrossRef CAS.
  27. P. Kang, C. Cheng, Z. Chen, C. K. Schauer, T. J. Meyer and M. Brookhart, J. Am. Chem. Soc., 2012, 134, 5500–5503 CrossRef CAS PubMed.
  28. T. Matsubara and K. Hirao, Organometallics, 2001, 20, 5759–5768 CrossRef CAS.
  29. P. Voyame, K. E. Toghill, M. A. Méndez and H. H. Girault, Inorg. Chem., 2013, 52, 10949–10957 CrossRef CAS PubMed.
  30. M. Irikura, Y. Tamaki and O. Ishitani, Chem. Sci., 2021, 12, 13888–13896 RSC.
  31. Y. Tamaki, K. Koike, T. Morimoto, Y. Yamazaki and O. Ishitani, Inorg. Chem., 2013, 52, 11902–11909 CrossRef CAS PubMed.
  32. C. E. Castillo, J. Armstrong, E. Laurila, L. Oresmaa, M. Haukka, J. Chauvin, S. Chardon-Noblat and A. Deronzier, ChemCatChem, 2016, 8, 2667–2677 CrossRef CAS.
  33. N. W. Kinzel, C. Werlé and W. Leitner, Angew. Chem., Int. Ed., 2021, 60, 11628–11686 CrossRef CAS PubMed.
  34. V. Vetere, C. Adamo and P. Maldivi, Chem. Phys. Lett., 2000, 325, 99–105 CrossRef CAS.
  35. C. Adamo and V. Barone, Theor. Chem. Acc., 2000, 105, 169–172 Search PubMed.
  36. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
  37. M. Ernzerhof and G. E. Scuseria, J. Chem. Phys., 1999, 110, 5029–5036 CrossRef CAS.
  38. C. Adamo, G. E. Scuseria and V. Barone, J. Chem. Phys., 1999, 111, 2889–2899 CrossRef CAS.
  39. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  40. 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. 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, 2016 Search PubMed.
  41. J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999–3093 CrossRef CAS PubMed.
  42. A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899–926 CrossRef CAS.
  43. F. Weinhold, Natural Bond Orbital Methods, in Encyclopedia of Computational Chemistry, ed. P. V. R. Schleyer, N. L. Allinger, T. Clark, J. Gasteiger, P. A. Kollman, H. F. Schaefer III and P. R. Schreiner, John Wiley and Sons, Chichester, UK, 1998, vol. 3, pp. 1792–1811 Search PubMed.
  44. S. J. A. van Gisbergen, F. Kootstra, P. R. T. Schipper, O. V. Gritsenko, J. G. Snijders and E. J. Baerends, Phys. Rev. A: At., Mol., Opt. Phys., 1998, 57, 2556–2571 CrossRef CAS.
  45. C. Jamorski, M. E. Casida and D. R. Salahub, J. Chem. Phys., 1996, 104, 5134–5147 CrossRef CAS.
  46. R. Bauernschmitt and R. Ahlrichs, Chem. Phys. Lett., 1996, 256, 454–464 CrossRef CAS.
  47. C. Adamo and D. Jacquemin, Chem. Soc. Rev., 2013, 42, 845–856 RSC.
  48. T. B. Demissie, K. Ruud and J. H. Hansen, Organometallics, 2015, 34, 4218–4228 CrossRef CAS.
  49. A. A. Isse and A. Gennaro, Absolute Potential of the Standard Hydrogen Electrode and the Problem of Interconversion of Potentials in Different Solvents, J. Phys. Chem. B, 2010, 114, 7894–7899 CrossRef CAS PubMed.
  50. T. Nakajima, Y. Tamaki, K. Ueno, E. Kato, T. Nishikawa, K. Ohkubo, Y. Yamazaki, T. Morimoto and O. Ishitani, J. Am. Chem. Soc., 2016, 138, 13818–13821 CrossRef CAS PubMed.
  51. Y. Kuramochi, J. Itabashi, K. Fukaya, A. Enomoto, M. Yoshida and H. Ishida, Chem. Sci., 2015, 6, 3063–3074 RSC.

Footnote

Electronic supplementary information (ESI) available: 3D surfaces of the MOs involved in the most relevant electronic excitations in the simulated absorption spectra, Selected structural parameters of the various intermediates involved in step B of the catalytic cycle for the CO2 to CO conversion, Gibbs free energies, G for all species involved in the Born–Haber cycle for redox potentials, parameters relevant to redox potentials, Cartesian coordinates and energetic data. See DOI: https://doi.org/10.1039/d4dt00125g

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.