Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Feb 26;63(10):4474–4481. doi: 10.1021/acs.inorgchem.3c04246

CO2 Reduction by an Iron(I) Porphyrinate System: Effect of Hydrogen Bonding on the Second Coordination Sphere

Chengxu Zhu §,, Carmine D’Agostino ⊥,, Sam P de Visser §,⊥,*
PMCID: PMC10934816  PMID: 38408891

Abstract

graphic file with name ic3c04246_0008.jpg

Transforming CO2 into valuable materials is an important reaction in catalysis, especially because CO2 concentrations in the atmosphere have been growing steadily due to extensive fossil fuel usage. From an environmental perspective, reduction of CO2 to valuable materials should be catalyzed by an environmentally benign catalyst and avoid the use of heavy transition-metal ions. In this work, we present a computational study into a novel iron(I) porphyrin catalyst for CO2 reduction, namely, with a tetraphenylporphyrin ligand and analogues. In particular, we investigated iron(I) tetraphenylporphyrin with one of the meso-phenyl groups substituted with o-urea, p-urea, or o-2-amide groups. These substituents can provide hydrogen-bonding interactions in the second coordination sphere with bound ligands and assist with proton relay. Furthermore, our studies investigated bicarbonate and phenol as stabilizers and proton donors in the reaction mechanism. Potential energy landscapes for double protonation of iron(I) porphyrinate with bound CO2 are reported. The work shows that the bicarbonate bridges the urea/amide groups to the CO2 and iron center and provides a tight bonding pattern with strong hydrogen-bonding interactions that facilitates easy proton delivery and reduction of CO2. Specifically, bicarbonate provides a low-energy proton shuttle mechanism to form CO and water efficiently. Furthermore, the o-urea group locks bicarbonate and CO2 in a tight orientation and helps with ideal proton transfer, while there is more mobility and lesser stability with an o-amide group in that position instead. Our calculations show that the o-urea group leads to reduction in proton-transfer barriers, in line with experimental observation. We then applied electric-field-effect calculations to estimate the environmental effects on the two proton-transfer steps in the reaction. These calculations describe the perturbations that enhance the driving forces for the proton-transfer steps and have been used to make predictions about how the catalysts can be further engineered for more enhanced CO2 reduction processes.

Short abstract

Density functional theory calculations show the importance of second-coordination-sphere effects for positioning of the substrate and proton donor in CO2 reduction processes.

Introduction

The cumulative emission of CO2 is raising environmental concerns for its role in global warming, and consequently CO2 reduction is an urgent global issue. In the face of these serious environmental problems, the capture of CO2 has become a hot issue in recent years.19 However, apart from CO2 capture and storage, research has started into the utilization of CO2 for the synthesis of valuable chemicals and materials.19 Because CO2 is a cheap, abundant, and nontoxic C1 feedstock, it can be converted into various carbon resources such as alcohols, acids, esters, and hydrocarbons.1018 This approach has 2-fold advantages, namely, CO2 reduction in the atmosphere as well as reduction in the use of unrenewable fossil fuels for synthesis of these compounds. Therefore, transforming CO2 into valuable chemicals may provide a solution to not only environmental problems caused by the emission of CO2 but also the anticipated depletion of fossil resources.

Because CO2 reduction is a challenging reaction, often transition-metal catalysts are used for this process.1022 In recent years, however, a number of catalysts with first-row transition-metal elements such as iron or manganese have been identified.2331 Particularly promising CO2 reduction catalysts contain an iron(I) porphyrinate group.32,33 Thus, Chang et al. took an iron(I) tetraphenylporphyrin (TPP) complex and inserted substituents on the ortho and para positions of one of the phenyl groups (Scheme 1).34 These porphyrin derivatives with amide groups in the second coordination sphere were shown to give an enhanced reduction in CO2 to CO and water. They proposed a better positioning of CO2 on the iron center that enabled more efficient proton transfer from solution. Subsequent, density functional theory (DFT) calculations of some of us on CO2 reduction by [Fe(TPP)] and [Fe(o-2-amide-TPP)] showed the same electronic configurations and mechanisms for the catalytic cycle with small proton-transfer barriers from phenol.35 Recent work of Chang and co-workers used a system with a urea group linked to one of the meso-phenyl groups.36 In combination with bicarbonate as the proton donor, the authors measured a 1500-fold rate enhancement for CO2 reduction compared to that of a system without bicarbonate. They also determined a crystal structure of the bicarbonate-bound o-urea-TPP with Zn2+ as the central ion. To understand how and why bicarbonate can enhance the reaction rates for CO2 reduction processes with iron porphyrins, we decided to do a DFT study into the mechanism of the CO2 reduction process using an iron(I) porphyrin system with either p-urea, o-2-amide, or o-urea as substituents to one of the meso-phenyl groups of the porphyrin system with bicarbonate as the proton donor (models IIII in Scheme 1) and the models with o-2-amide and o-urea but with phenol as the proton donor (models II,phenol and III,phenol). The work shows that a hydrogen-bonding interaction in the second coordination sphere enhances proton-transfer rates. Moreover, the urea group can lock bicarbonate into a double hydrogen-bonding interaction mimicking a salt bridge in enzymatic structures that stabilizes the system and provides the best network for proton relay.

Scheme 1. Models Investigated in This Work.

Scheme 1

Methods

The model was based on the CO2-bound optimized geometry of our previous studies on CO2 reduction by an [FeI(TPP)] complex, where TPP = meso-tetraphenylporphyrin,35 whereby we manually inserted the p-urea, o-2-amide, or o-urea groups into one of the meso-phenyl substituents to create models IIII (Scheme 1) and replaced the other three meso-phenyl groups by a hydrogen atom. In addition, a molecule of bicarbonate was added to the distal site to create a cluster model with overall charge of 3– and odd multiplicity. To test whether the large negative charge in the model influences the structure and kinetics, we also used phenol as a proton donor because it has a pKa value similar to that of bicarbonate, which gave the model an overall charge of 2–. Calculations were performed in Gaussian 09 using the unrestricted B3LYP density functional method supplemented with the GD3 dispersion correction of Grimme, with Becke–Johnson damping included.3740 An all-electron def2-SVP41 basis set was selected for all atoms during the geometry optimizations, analytical frequency calculations, and constraint geometry scans designated as BS1. Single-point energy calculations were performed with the def2-TZVP basis set on all atoms designated as BS2. Although the experimental work was performed in a dimethylformamide (DMF) solution, we tested several implicit solvent models with a range of dielectric constants of 37.2 (DMF) and 78.4 (water).42 In general, a change in the dielectric constant for the solvent model gives minor differences to the energies and optimized geometries and therefore does not change the results dramatically; see the Supporting Information. Calculations were performed on all low-lying triplet and quintet-spin-state surfaces. These methods and approaches were used before in our group and reproduced experimental free energies of activation within a few kilocalories per mole and predicted product selectivities in line with experimental observation.43,44

Results and Discussion

The catalytic cycle for the reduction of CO2 to CO and water on an iron(I) porphyrinate complex as calculated here is shown in Scheme 2. The cycle starts from an iron(I) porphyrinate system with nearby bicarbonate, structure A or [FeI(HCO3)]2–. This complex is reduced from iron(I) to iron(0) to form B and triggers the binding of a molecule of CO2 and its reaction to give C or [FeI(CO2)(HCO3)]3–. An internal proton transfer from bicarbonate to CO2 is expected to lead to protonated CO2 and carbonate in complex D: [FeII(OCOH)(CO32–)]3–. The carbonate picks up a proton from the solvent/buffer and assists with the second protonation step of CO2 that leads to the formation of CO and water in complex E. The cycle is closed with the release of CO and the reduction of iron(II) to iron(I). In our previous studies,35 we showed that the first two steps in the CO2 reduction cycle, i.e., AB and BC, are little affected by second-coordination-sphere effects and substitution of the meso-phenyl groups of the porphyrin complex. Therefore, our work presented here is mostly focused on the protonation steps in the cycle and on how bicarbonate is involved in the process. In particular, we created CO2-bound iron(II) porphyrinate complexes with p-urea, o-2-amide, and o-urea as substituents to one of the phenyl groups on the meso position: structures CI, CII, and CIII. Optimized geometries of structures C of models IIII in the triplet and quintet spin states are shown in Figure 1. The spin multiplicity is given as a superscript in front of the label and the model type as a subscript after the label.

Scheme 2. Catalytic Cycle of CO2 Reduction on a Bicarbonate-Bound Iron Porphyrinate System Studied in This Work.

Scheme 2

Formal oxidation states of iron are given in square brackets.

Figure 1.

Figure 1

UB3LYP-GD3/BS1-optimized geometries of structure C for models IIII. Relative energies include energies at the BS2 level of theory and zero-point corrections in kilocalories per mole. Bond lengths are in angstroms.

In all structures, the triplet and quintet spin states are within 1.5 kcal mol–1, which implies that both may be accessible. Structures II and III have the same spin-state ordering and spin ground state; hence, these are not affected by the modification in the second coordination sphere. In the p-urea system (I), we find the quintet spin below the triplet spin state by 1.3 kcal mol–1 for structure C. Nevertheless, for all structures C, the triplet–quintet energy gap is within 1.3 kcal mol–1; consequently, the spin states are degenerate, and both will be accessible at room temperature conditions. We also calculated the CO2 binding energies for all complexes. In previous work, we reported CO2 binding energies of ΔG = 11.3 kcal mol–1 for the difference in energy between [FeI(CO2)(o-2-amide-TPP)]2– and an isolated CO2 and [Fe0(o-2-amide-TPP)]2– complex. With additional bicarbonate in the model in hydrogen-bonding interactions with CO2 in the CII and CIII structures, CO2 release requires a free energy of ΔG = 9.6 and 9.2 kcal mol–1 in the triplet spin state, whereas we find ΔG = 4.6 and 3.5 kcal mol–1 in the quintet spin state. As such, the double hydrogen bond from urea to bicarbonate does not appear to dramatically influence the CO2 binding energy compared to the o-2-amide system. Despite the fact that the bicarbonate anion forms hydrogen-bonding interactions with the CO2 group bound with o-urea and o-2-amide, there does not appear to be a major stabilization energy as a result of this. As a matter of fact, our previous study also had a hydrogen bond to the bound CO2 group in [FeII(CO2)(o-2-amide-TPP)]2– but directly with the amide proton. However, the hydrogen-bonding network will facilitate the CO2 binding and lock it in position.

The structures shown in Figure 1 give tight binding of CO2 and bicarbonate between the iron(II) and hydrogen-bonding o-amide and o-urea groups. In the p-urea structure, the hydrogen-bonding groups are relatively far from the iron center by more than 10 Å, and, consequently, a bridge will require multiple water molecules, which were not included in the model. Nevertheless, the hydrogen-bonding network to the CO2 group only has a minor effect on the Fe–C bond lengths, which is 2.085 Å in 5CI and 2.018 Å in 3CI, whereas the quintet-spin-state structures for 5CII and 5CIII have values within 0.02 Å. These Fe–C distances are in line with those reported previously for the interaction of iron with a first-row element such as oxygen, nitrogen, or carbon.45 The optimized geometries compare well to those calculated before using the B97-D approach with strong hydrogen-bonding interactions between bicarbonate and the two NH groups of the urea substituent and the oxygen atom of CO2.36 The porphyrin scaffold is close to planarity, similar to previous calculations on porphyrin models in the gas phase.4650

In model III, the bicarbonate ion is locked in position through two hydrogen-bonding interactions with the urea N–H groups, while its OH group forms a hydrogen bond to the CO2 group. This orientation is reminiscent of salt bridges in protein structures with a negatively charged carboxylate group opposite of a positively charged arginine side chain.5153 However, different from a salt bridge in a protein structure, the urea system investigated here is charge-neutral; hence, the binding strength of the interaction with bicarbonate will be much weaker than that in a typical salt bridge. The two urea hydrogen-bonding interactions are different in length; namely, the one closest to the porphyrin is shortest at 1.690 Å in 5CIII, while the other one has a distance of 1.737 Å. At the same time, the OH group of bicarbonate forms a hydrogen bond with one of the oxygen atoms of CO2 at 1.718 Å. These distances are very similar for the o-2-amide complex, but because it can form only two hydrogen bonds, the bicarbonate is lesser restraint.

Next, we studied the proton-transfer step from 5,3C (models IIII) to form an iron(II) with a protonated CO2 group (structures IM1) via transition state TS1. Thereafter, we took structures 5,3IM1 and added a proton to the carbonate group to form the 5,3RC2 complexes and searched for a proton-transfer transition state (TS2) to form water and CO products (structures IM2). The optimized transition-state structures, i.e., TS1 and TS2, for the model III pathway are shown in Figure 2, together with the calculated energy landscapes. Both proton-transfer free energies of activation are small, namely, ΔG = 5.4 kcal mol–1 for the first proton transfer with respect to 5CIII and ΔG = 5.7 kcal mol–1 with respect to 5RC2III (or ΔG = 11.0 kcal mol–1 with respect to 5CIII). On the triplet spin state, the barriers are considerably higher in energy, namely, ΔG > 14 kcal mol–1 for the first proton transfer to reach 3IM1III in an endergonic step of 13.1 kcal mol–1. On the triplet spin state, we were not able to fully characterize it transition state, but a constraint geometry scan shows it to be only slightly higher in energy than the value for 3IM1III; hence, a value of <14 kcal mol–1 is reported. The second proton transfer, similar to the quintet-spin-state surface, is 5.7 kcal mol–1. This is not surprising as no electron transfer happens in these steps and all group spin densities remain the same. Interestingly, for the IM2III complex, the quintet-spin-state structure is higher in energy than the corresponding triplet spin complex. Thermal collisions of 5IM2III in solution, however, are likely to revert the spin to the triplet spin ground state for the CO-bound structures IM2.

Figure 2.

Figure 2

UB3LYP-GD3/BS2//UB3LYP-GD3/BS1 calculated the free energy profile for the two proton-transfer steps from the CO2-bound iron(II) porphyrin complex [FeII(CO2)(o-urea-TPP)(HCO3)]3–. Free energies (ΔG) were calculated at 298 K and include energies at the BS2 level of theory and zero-point, solvent, thermal, and entropic corrections in kilocalories per mole relative to 5CIII, whereby calculations in the water solvent are the data outside parentheses and those in the DMF solvent are the data inside parentheses. Optimized transition-state geometries for the first and second proton transfer for model III are shown with bond lengths in angstroms, angles in degrees, and the imaginary frequency in reciprocal centimeters. In RC2, a proton is added to the model of IM1 and the energy is set to the value of IM1 for comparison.

Structurally, the proton-transfer transition state 5TS1III has the transferring proton midway between the donor and acceptor groups at distances of 1.184 and 1.246 Å. The O–H–O angle is close to linearity (177°), and the imaginary frequency of the transition state is i871 cm–1 for the proton-transfer mode. This is a relatively low imaginary frequency for proton transfer because usually values well over i1000 cm–1 are found.5456 For 5TS2III, the imaginary frequency is similar (i730 cm–1) and represents simultaneous proton transfer and C–O cleavage, i.e., water release. The distance of the proton from the carbonate in 5TS2III is 1.242 Å, while the distance to the OH group is 1.178 Å. Note that the Fe–C distances in 5TS1III and 5TS2III are similar (2.138 vs 2.129 Å, respectively), and consequently no electron transfer has happened in this process. Both energy landscapes were calculated with either water or DMF as an implicit solvent model. Both sets of data are shown in Figure 2, and as can be seen, the energy and structural differences are small when the solvent model is changed from water to DMF. These results match previous conclusions drawn from changing the value of the dielectric constant in implicit solvent calculations.57

To find out how bicarbonate as a proton donor reacts compared to alternative proton donors, we removed bicarbonate from structures 3,5CIII and replaced it with a phenol molecule: 3,5CIII,phenol. Subsequently, the proton-transfer pathways of CO2 were investigated, leading to CO and water as products. Figure 3 shows the optimized transition-state geometry for the first and second proton transfer on the quintet-spin-state surface with phenol as a proton donor. The obtained landscape for the phenol-bound structures compared to the bicarbonate reaction is shown in Figure 3, and the data for all local minima and transition states for all calculated models are reported in Table 1. The free energy of activation for the first proton transfer from 5CIII,phenol is ΔG = 8.0 kcal mol–1, while the second proton transfer was calculated at ΔG = 6.6 kcal mol–1 with respect to 5CIII,phenol. Previously, for proton transfer from phenol to CO2 bound to FeII(o-2-amide-TPP), we found a proton-transfer free energy of activation of 7.0 kcal mol–1 for the second proton-transfer step in the catalytic cycle, in agreement with what we find here with different methods and models.35 Taken together, the two consecutive proton-transfer steps for 5CIII have a maximum barrier of 11.0 kcal mol–1 using bicarbonate as a proton donor, while with phenol as the proton donor, the barrier is reduced to 8.0 kcal mol–1. Although the energy differences between the bicarbonate and phenol models are small, they do indicate a slowing down of the rate by a factor of 600 for the replacement of phenol by bicarbonate. This is not a surprising result because the pKa value of the free bicarbonate/carbonate couple is 10.3, while the deprotonation of phenol has a pKa value of 9.95.58 As such, free phenol should react with lower proton-transfer barriers than free bicarbonate, as indeed is observed here. These pKa values, of course, will be affected upon binding to a hydrogen-bonding scaffold, such as o-urea, which can hold bicarbonate stronger than phenol. Despite the strong binding of bicarbonate to the urea scaffold, we see little effect on the rates based on the pKa differences with phenol. The pKa values of the o-urea and o-2-amide groups were calculated and found to be much higher than those for phenol. Therefore, the o-urea and o-2-amide groups cannot serve as proton donors in the reaction mechanisms because their N–H bonds are too strong to break. In none of the geometry optimizations reported here was a proton transfer from the o-urea or o-2-amide groups to either bicarbonate, CO2, or phenol observed. Therefore, the calculations show that a proton donor is needed and that phenol and bicarbonate are strong enough bases to provide protons for the CO2 reduction reaction.

Figure 3.

Figure 3

UB3LYP-GD3/BS2//UB3LYP-GD3/BS1 calculated the free energy profile for the two proton-transfer steps from the CO2-bound iron(II) porphyrin complex [FeII(CO2)(o-urea-TPP)(phenol)]2–. The data for bicarbonate are in blue, whereas the data for phenol are in green. Free energies (ΔG) were calculated at 298 K and include energies at the BS2 level of theory and zero-point, solvent, thermal, and entropic corrections in kilocalories per mole relative to 5CIII,phenol. Optimized transition-state geometries for the first and second proton transfer for model III are shown with bond lengths in angstroms, angles in degrees, and the imaginary frequency in reciprocal centimeters. In RC2, a proton is added to the model of IM1, and the energy is set to the value of IM1 for comparison.

Table 1. Free Energies (BS2//BS1) for Calculated Reaction Pathways for Double Protonation of CO2 Bound to Various Porphyrin Modelsa.

  III III,phenol II II,phenolb
5RC 0.0 0.0 0.0 0.0
5TS1 5.4 8.0 <5 <1
5IM1 5.3 –0.4 4.6 –11.4
5RC2 5.3 –0.4 4.6 –11.4
5TS2 11.0 6.6 30.7 7.0
5IM2 7.5 1.0 7.3 –14.4
a

ΔG(BS2) values in kilocalories per mole.

b

Data from ref (35).

The optimized transition-state structures 5TS1III,phenol and 5TS2III,phenol are also shown in Figure 3. The first proton-transfer transition state (5TS1III,phenol) has a small imaginary frequency of i183 cm–1 for proton transfer from phenol to CO2. The structure is relatively central, with the transferring proton midway between the donor (by 1.190 Å) and acceptor (by 1.140 Å) oxygen atoms, respectively. The angle O–H–O is close to linearity (173°), as is expected of a proton-transfer transition state. In the second proton-transfer transition state, the structure is more upright and the C–O bond has elongated significantly to 2.097 Å. The imaginary frequency of i660 cm–1 indicates dominant proton transfer from phenol to oxygen. The transferring proton is at a distance of 1.258 Å from phenolate and 1.162 Å from the accepting oxygen atom. Nevertheless, the calculations presented in Figures 2 and 3 show that both bicarbonate and phenol are strong enough acids to donate a proton to an FeII-CO2 complex efficiently and initiate the CO2 reduction process. We observe a lowering of the first proton-transfer barrier for o-urea versus o-2-amide by 0.4 kcal mol–1, which would correspond to a rate enhancement of a factor of 2.4. Our work, therefore, is in good quantitative agreement with the experimental work of Chang et al., which showed enhanced reactivity by 2-fold between the two complexes with either o-urea or o-2-amide in the second coordination sphere.36

Overall, the calculations presented in this work indicate that an o-urea substituent to a TPP scaffold enables strong hydrogen-bonding interactions that hold and position distal ligands to the iron center better than o-2-amide. To fully understand the positional differences, we created an overlay of the 3CII and 3CIII optimized geometries and present the results in Figure 4. As can be seen, the two structures are seemingly similar and have most groups in a similar position and orientation. However, there are differences, as highlighted in the figure. Thus, the O–H–O angle is close to linearity in 3CIII at 172°, while the angle is 168° in 3CII. The closer to linearity the O–H–O angle is, the easier it is for proton transfer to take place. Indeed the smallest proton-transfer barrier is found for 3CII. Further differences between 3CII and 3CIII relate to the positioning of the bicarbonate ion in the complex, as seen from the dihedral angles Fe–C–O–H and O–C–O–H in Figure 4. Both of these differ by more than 12° and show that bicarbonate is better positioned for proton transfer in 3CIII than in 3CII.

Figure 4.

Figure 4

Overlay of the UB3LYP-GD3/BS1-optimized geometries. Angles and dihedral angles are in degrees.

To find out how these iron porphyrin systems can be improved for more efficient proton transfer, we decided to apply electric-field-effect perturbations to the optimized geometries of 5CIII, 5IM1III, 5RC2III, and 5IM2III. In particular, single-point calculations at the UB3LYP-GD3/BS2 level of theory were run in Gaussian with an electric field located along the molecular x, y, or z axis, with magnitudes ranging from −200 to +200 au, and the results are shown in Figure 5. These electric-field perturbations were used previously in our group and shown to influence charge distributions in complexes and bifurcation patterns in chemical catalysis.59,60 In particular, recent work showed that charged groups in proteins can influence the strength of the C–H bonds in substrates and direct a reaction selectivity to a specific bond in a substrate.61,62 Thus, an electric-field perturbation has a major effect on the thermodynamics for proton transfer and the reaction energy. To be specific, an electric-field effect along the negative z axis is along the proton-transfer axis and makes the first proton-transfer step more exergonic, while a field in the opposite direction makes the reaction more endergonic, i.e., less likely. Interestingly, electric-field effects in the x and y directions do not appear to have a major effect on the first proton-transfer step unless very large electric fields are used. For the second proton-transfer step, a field along the positive y axis is favorable, while in the negative y direction, the reaction becomes more endothermic. The vector in the imaginary frequency of TS2 for this reaction step is along the y axis and shows motions for the C–O stretch vibration, i.e., cleavage of the C–O bond as well as proton transfer. Also, for the second proton-transfer step, fields orthogonal to the proton transfer show few effects on the reaction thermochemistry. Overall, these calculations show that the first proton-transfer step is improved with an electric-field effect along the negative z axis, while the second proton-transfer step is probably faster with a field along the positive y axis. Therefore, engineering of the iron porphyrin complex by, for instance, latching the porphyrin to a metal surface through its axial ligand, may help with the first proton-transfer reaction in the CO2 activation reaction, while perturbations with charged groups along the y axis will enhance the second proton-transfer step.

Figure 5.

Figure 5

Electric-field effects on the stabilization free energies of 5IM1 and 5IM2 for model II. Electric fields are as defined in Gaussian.

Conclusions

In this work, a computational study is presented on CO2 reduction to CO and water on several iron porphyrinate complexes. The calculations show that an iron porphyrinate system is an efficient catalyst for CO2 reduction reactions and particularly systems with a hydrogen-bonding donor in the second coordination sphere because that helps to position and tighten the substrate. The tight binding of the CO2 substrate enables two low-energy proton-transfer steps to form CO and water efficiently. Furthermore, bicarbonate can be locked in a position with an o-urea group attached to the ortho meso position of the ligand to provide a second-coordination-sphere environment for locking the proton donor (bicarbonate) and substrate (CO2) in a tight orientation for efficient proton shuttle and ultimately CO2 reduction purposes. Finally, electric-field calculations were performed on the complexes for the successive proton-transfer steps in the catalytic cycle. These calculations predict that the proton-transfer steps will be sensitive to local perturbations, and an electric-field-effect perturbation can influence driving forces for these steps and make the CO2 reduction reaction more exothermic and efficient.

Acknowledgments

The Computational Shared Facilities at the University of Manchester are acknowledged for computational support.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c04246.

  • Raw data including absolute and relative energies, group spin densities and charges, and Cartesian coordinates of optimized geometries discussed in this work (PDF)

Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ic3c04246_si_001.pdf (864.9KB, pdf)

References

  1. Liu Z.; Qian J.; Zhang G.; Zhang B.; He Y. Electrochemical CO2-to-CO conversion: comprehensive review of recent developments and emerging trends. Separ. Purif. Technol. 2024, 330, 125177 10.1016/j.seppur.2023.125177. [DOI] [Google Scholar]
  2. Wang C.; Lv Z.; Feng X.; Yang W.; Wang B. Recent advances in electrochemical CO2-to-multicarbon conversion: from fundamentals to industrialization. Adv. Energy Mater. 2023, 13, 2302382 10.1002/aenm.202302382. [DOI] [Google Scholar]
  3. Darji H. R.; Kale H. B.; Shaikh F. F.; Gawande M. B. Advancement and state-of-art o heterogeneous catalysis of selective CO2 hydrogenation to methanol. Coord. Chem. Rev. 2023, 497, 215409 10.1016/j.ccr.2023.215409. [DOI] [Google Scholar]
  4. Younus H. A.; Ahmad N.; Ni W.; Wang X.; Al-Abri M.; Zhang Y.; Verpoort F.; Zhang S. Molecular catalysts for CO2 Electroreduction: Progress and prospects with pincer type complexes. Coord. Chem. Rev. 2023, 493, 215318 10.1016/j.ccr.2023.215318. [DOI] [Google Scholar]
  5. Chatterjee S.; Sengupta K.; Mondal B.; Dey S.; Dey A. Factors Determining the Rate and Selectivity of 4-/4H+ Electrocatalytic Reduction of Dioxygen by Iron Porphyrin Complexes. Acc. Chem. Res. 2017, 50, 1744–1753. 10.1021/acs.accounts.7b00192. [DOI] [PubMed] [Google Scholar]
  6. Rao H.; Bonin J.; Robert M. Toward Visible-Light Photochemical CO2-to-CH4 Conversion in Aqueous Solutions Using Sensitized Molecular Catalysis. J. Phys. Chem. C 2018, 122, 13834–13839. 10.1021/acs.jpcc.8b00950. [DOI] [Google Scholar]
  7. Franco F.; Pinto M. F.; Royo B.; Lloret-Fillol J. A Highly Active N-Heterocyclic Carbene Manganese(I) Complex for Selective Electrocatalytic CO2 Reduction to CO. Angew. Chem., Int. Ed. 2018, 57, 4603–4606. 10.1002/anie.201800705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Fernández S.; Franco F.; Martínez Belmonte M.; Friães S.; Royo B.; Luis J. M.; Lloret-Fillol J. Decoding the CO2 Reduction Mechanism of a Highly Active Organometallic Manganese Electrocatalyst: Direct Observation of a Hydride Intermediate and Its Implications. ACS Catal. 2023, 13, 10375–10385. 10.1021/acscatal.3c01430. [DOI] [Google Scholar]
  9. Zhu C.; D’Agostino C.; de Visser S. P. Mechanism of CO2 Reduction to Methanol with H2 on an Iron(II)-scorpionate Catalyst. Chem. Eur. J. 2023, 29, e202302832 10.1002/chem.202302832. [DOI] [PubMed] [Google Scholar]
  10. He M.; Sun Y.; Han B. Green Carbon Science: Scientific Basis for Integrating Carbon Resource Processing, Utilization, and Recycling. Angew. Chem., Int. Ed. 2013, 52, 9620–9633. 10.1002/anie.201209384. [DOI] [PubMed] [Google Scholar]
  11. Aresta M.; Dibenedetto A.; Quaranta E. State of the art and perspectives in catalytic processes for CO2 conversion into chemicals and fuels: The distinctive contribution of chemical catalysis and biotechnology. J. Catal. 2016, 343, 2–45. 10.1016/j.jcat.2016.04.003. [DOI] [Google Scholar]
  12. Franco F.; Fernández S.; Lloret-Fillol J. Advances in the Electrochemical Catalytic Reduction of CO2 with Metal Complexes. Curr. Opin. Electrochem. 2019, 15, 109–117. 10.1016/j.coelec.2019.04.002. [DOI] [Google Scholar]
  13. Modak A.; Bhanja P.; Dutta S.; Chowdhury B.; Bhaumik A. Catalytic Reduction of CO2 into Fuels and Fine Chemicals. Green Chem. 2020, 22, 4002–4033. 10.1039/D0GC01092H. [DOI] [Google Scholar]
  14. Ding P.; Zhao H.; Li T.; Luo Y.; Fan G.; Chen G.; Gao S.; Shi X.; Lu S.; Sun X. Metal-Based Electrocatalytic Conversion of CO2 to Formic Acid/Formate. J. Mater. Chem. A 2020, 8, 21947–21960. 10.1039/D0TA08393C. [DOI] [Google Scholar]
  15. Zhang R.; Warren J. J. Recent Developments in Metalloporphyrin Electrocatalysts for Reduction of Small Molecules: Strategies for Managing Electron and Proton Transfer Reactions. ChemSusChem. 2021, 14, 293–302. 10.1002/cssc.202001914. [DOI] [PubMed] [Google Scholar]
  16. Ghosh A. C.; Duboc C.; Gennari M. Synergy Between Metals for Small Molecule Activation: Enzymes and Bio-inspired Complexes. Coord. Chem. Rev. 2021, 428, 213606 10.1016/j.ccr.2020.213606. [DOI] [Google Scholar]
  17. Kinzel N. W.; Werlé C.; Leitner W. Transition Metal Complexes as Catalysts for the Electroconversion of CO2: An Organometallic Perspective. Angew. Chem., Int. Ed. 2021, 60, 11628–11686. 10.1002/anie.202006988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Lei K.; Yu Xia B. Electrocatalytic CO2 Reduction: from Discrete Molecular Catalysts to Their Integrated Catalytic Materials. Chem. Eur. J. 2022, 28, e202200141 10.1002/chem.202200141. [DOI] [PubMed] [Google Scholar]
  19. Francke R.; Schille B.; Roemelt M. Homogeneously Catalyzed Electroreduction of Carbon Dioxide-Methods, Mechanisms, and Catalysts. Chem. Rev. 2018, 118, 4631–4701. 10.1021/acs.chemrev.7b00459. [DOI] [PubMed] [Google Scholar]
  20. Abdinejad M.; Hossain M. N.; Kraatz H.-B. Homogeneous and heterogeneous molecule arcatalysts for electrochemical reduction of carbon dioxide. RSC Adv. 2020, 10, 38013–38023. 10.1039/D0RA07973A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Fors S. A.; Malapit C. A. Homogeneous Catalysis for the Conversion of CO2, CO, CH3OH, and CH4 to C2+ Chemicals via C–C Bond Formation. ACS Catal. 2023, 13, 4231–4249. 10.1021/acscatal.2c05517. [DOI] [Google Scholar]
  22. Sadique A. R.; Brennessel W. W.; Holland P. L. Reduction of CO2 to CO using low-coordinate iron: Formation of a four-coordinate iron dicarbonyl complex and a bridging carbonate complex. Inorg. Chem. 2008, 47, 784–786. 10.1021/ic701914m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Wei D.; Sang R.; Moazezbarabadi A.; Junge H.; Beller M. Homogeneous Carbon Capture and Catalytic Hydrogenation: Toward a Chemical Hydrogen Battery System. JACS Au 2022, 2, 1020–1031. 10.1021/jacsau.1c00489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Zhao H.-Z.; Chang Y.-Y.; Liu C. Electrodes modified with iron porphyrin and carbon nanotubes: application to CO2 reduction and mechanism of synergistic electrocatalysis. J. Solid State Electrochem. 2013, 17, 1657–1664. 10.1007/s10008-013-2027-1. [DOI] [Google Scholar]
  25. Mohamed E. A.; Zahran Z. N.; Naruta Y. Efficient electrocatalytic CO2 reduction with a molecular cofacial iron porphyrin dimer. Chem. Commun. 2015, 51, 16900–16903. 10.1039/C5CC04273A. [DOI] [PubMed] [Google Scholar]
  26. Okabe Y.; Lee S. K.; Kondo M.; Masaoka S. Syntheses and CO2 reduction activities of π-expanded/extended iron porphyrin complexes. J. Biol. Inorg. Chem. 2017, 22, 713–725. 10.1007/s00775-017-1438-3. [DOI] [PubMed] [Google Scholar]
  27. Sinha S.; Warren J. J. Unexpected solvent effect in electrocatalytic CO2 to CO conversion revealed using asymmetric metalloporphyrins. Inorg. Chem. 2018, 57, 12650–12656. 10.1021/acs.inorgchem.8b01814. [DOI] [PubMed] [Google Scholar]
  28. Fukuzumi S.; Lee Y.-M.; Ahn H. S.; Nam W. Mechanisms of catalytic reduction of CO2 with heme and nonheme metal complexes. Chem. Sci. 2018, 9, 6017–6034. 10.1039/C8SC02220H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Ogawa A.; Oohora K.; Gu W.; Hayashi T. Electrochemical CO2 reduction by a cobalt bipyricorrole complex: decrease of an overpotential value derived from monoanionic ligand character of the porphyrinoid species. Chem. Commun. 2019, 55, 493–496. 10.1039/C8CC08876D. [DOI] [PubMed] [Google Scholar]
  30. Zhang R.; Warren J. J. Recent Developments in Metalloporphyrin Electrocatalysts for Reduction of Small Molecules: Strategies for Managing Electron and Proton Transfer Reactions. ChemSusChem. 2021, 14, 293–302. 10.1002/cssc.202001914. [DOI] [PubMed] [Google Scholar]
  31. Domingo-Tafalla B.; Chatterjee T.; Palomares E. Recent advances in the rational designing of metalloporphyrinoid-based CO2 reduction catalysts: From molecular structural tuning to the application in catalysis. J. Porphyrins Phthalocyanines 2023, 27, 23–46. 10.1142/S1088424623300033. [DOI] [Google Scholar]
  32. Bhugun I.; Lexa D.; Savéant J.-M. Catalysis of the Electrochemical Reduction of Carbon Dioxide by Iron(0) Porphyrins: Synergystic Effect of Weak Brönsted Acids. J. Am. Chem. Soc. 1996, 118, 1769–1776. 10.1021/ja9534462. [DOI] [Google Scholar]
  33. Costentin C.; Passard G.; Robert M.; Savéant J.-M. Ultraefficient homogeneous catalyst for the CO2-to-CO electrochemical conversion. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 14990–14994. 10.1073/pnas.1416697111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Nichols E. M.; Derrick J. S.; Nistanaki S. K.; Smith P. T.; Chang C. J. Positional effects of second-sphere amide pendants on electrochemical CO2 reduction catalyzed by iron porphyrins. Chem. Sci. 2018, 9, 2952–2960. 10.1039/C7SC04682K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Davethu P. A.; de Visser S. P. CO2 reduction on an iron-porphyrin center: A computational study. J. Phys. Chem. A 2019, 123, 6527–6535. 10.1021/acs.jpca.9b05102. [DOI] [PubMed] [Google Scholar]
  36. Derrick J. S.; Loipersberger M.; Nistanaki S. K.; Rothweiler A. V.; Head-Gordon M.; Nichols E. M.; Chang C. J. Templating Bicarbonate in the Second Coordination Sphere Enhances Electrochemical CO2 Reduction Catalyzed by Iron Porphyrins. J. Am. Chem. Soc. 2022, 144, 11656–11663. 10.1021/jacs.2c02972. [DOI] [PubMed] [Google Scholar]
  37. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand R.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas Ö.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2009.
  38. Becke A. D. Density-Functional thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
  39. Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti Correlation-energy Formula Into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
  40. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  41. Weigend F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. 10.1039/b515623h. [DOI] [PubMed] [Google Scholar]
  42. Tomasi J.; Mennucci B.; Cammi R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
  43. Mokkawes T.; de Visser S. P. Caffeine Biodegradation by Cytochrome P450 1A2. What Determines the Product Distributions?. Chem. Eur. J. 2023, 29, e202203875 10.1002/chem.202203875. [DOI] [PubMed] [Google Scholar]
  44. Bagha U. K.; Yadav R.; Mokkawes T.; Satpathy J. K.; Kumar D.; Sastri C. V.; de Visser S. P. Defluorination of Fluorophenols by a Nonheme Iron(IV)-Oxo Species: Observation of a New Intermediate Along the Reaction. Chem. Eur. J. 2023, 29, e202300478 10.1002/chem.202300478. [DOI] [PubMed] [Google Scholar]
  45. Faponle A. S.; Seebeck F. P.; de Visser S. P. Sulfoxide Synthase Versus Cysteine Dioxygenase Reactivity in a Nonheme Iron Enzyme. J. Am. Chem. Soc. 2017, 139, 9259–9270. 10.1021/jacs.7b04251. [DOI] [PubMed] [Google Scholar]
  46. Colomban C.; Tobing A. H.; Mukherjee G.; Sastri C. V.; Sorokin A. B.; de Visser S. P. Mechanism of oxidative activation of fluorinated aromatic compounds by N-bridged diiron-phthalocyanine. What determines the reactivity?. Chem. Eur. J. 2019, 25, 14320–14331. 10.1002/chem.201902934. [DOI] [PubMed] [Google Scholar]
  47. Ali H. S.; Henchman R. H.; de Visser S. P. Lignin biodegradation by a cytochrome P450 enzyme: A computational study into syringol activation by GcoA. Chem. Eur. J. 2020, 26, 13093–13102. 10.1002/chem.202002203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Louka S.; Barry S. M.; Heyes D. J.; Mubarak M. Q. E.; Ali H. S.; Alkhalaf L. M.; Munro A. W.; Scrutton N. S.; Challis G. L.; de Visser S. P. The catalytic mechanism of aromatic nitration by cytochrome P450 TxtE: Involvement of a ferric-peroxynitrite intermediate. J. Am. Chem. Soc. 2020, 142, 15764–15779. 10.1021/jacs.0c05070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Hermano Sampaio Dias A.; Yadav R.; Mokkawes T.; Kumar A.; Skaf M. S.; Sastri C. V.; Kumar D.; de Visser S. P. Biotransformation of Bisphenol by Human Cytochrome P450 2C9 Enzymes: a Density Functional Theory Study. Inorg. Chem. 2023, 62, 2244–2256. 10.1021/acs.inorgchem.2c03984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Zhang Y.; Mokkawes T.; de Visser S. P. Insights into Cytochrome P450 Enzyme Catalyzed Defluorination of Aromatic Fluorides. Angew. Chem., Int. Ed. 2023, 62, e202310785 10.1002/anie.202310785. [DOI] [PubMed] [Google Scholar]
  51. Berman H. M.; Westbrook J.; Feng Z.; Gilliland G.; Bhat T. N.; Weissig H.; Shindyalov I. N.; Bourne P. E. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Cantú Reinhard F. G.; Lin Y.-T.; Stańczak A.; de Visser S. P. Bioengineering of cytochrome P450 OleTJE: How does substrate positioning affect the product distributions?. Molecules 2020, 25, 2675–2697. 10.3390/molecules25112675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Pickl M.; Kurakin S.; Cantú Reinhard F. G.; Schmid P.; Pöcheim A.; Winkler C. K.; Kroutil W.; de Visser S. P.; Faber K. Mechanistic studies of fatty acid activation by CYP152 peroxygenases reveal unexpected desaturase activity. ACS Catal. 2019, 9, 565–577. 10.1021/acscatal.8b03733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Li X.-X.; Postils V.; Sun W.; Faponle A. S.; Solà M.; Wang Y.; Nam W.; de Visser S. P. Reactivity patterns of (protonated) Compound II and Compound I of Cytochrome P450: Which is the better oxidant?. Chem. Eur. J. 2017, 23, 6406–6418. 10.1002/chem.201700363. [DOI] [PubMed] [Google Scholar]
  55. Ghafoor S.; Mansha A.; de Visser S. P. Selective Hydrogen Atom Abstraction from Dihydroflavonol by a Non-Heme Iron Center Is the Key Step in the Enzymatic Flavonol Synthesis and Avoids Byproducts. J. Am. Chem. Soc. 2019, 141, 20278–20292. 10.1021/jacs.9b10526. [DOI] [PubMed] [Google Scholar]
  56. Ali H. S.; Henchman R. H.; de Visser S. P. What Determines the Selectivity of Arginine Dihydroxylation by the Nonheme Iron Enzyme OrfP?. Chem. Eur. J. 2021, 27, 1795–1809. 10.1002/chem.202004019. [DOI] [PubMed] [Google Scholar]
  57. Cantú Reinhard F. G.; Faponle A. S.; de Visser S. P. Substrate sulfoxidation by an iron(IV)-oxo complex: benchmarking computationally calculated barrier heights to experiment. J. Phys. Chem. A 2016, 120, 9805–9814. 10.1021/acs.jpca.6b09765. [DOI] [PubMed] [Google Scholar]
  58. Lide D. R., Ed. Handbook of Chemistry and Physics, 76th ed.; CRC Press: Boca Raton, FL, 1996. [Google Scholar]
  59. de Visser S. P.; Mukherjee G.; Ali H. S.; Sastri C. V. Local charge distributions, electric dipole moments and local electric fields influence reactivity patterns and guide regioselectivities in α-ketoglutarate-dependent nonheme iron dioxygenases. Acc. Chem. Res. 2022, 55, 65–74. 10.1021/acs.accounts.1c00538. [DOI] [PubMed] [Google Scholar]
  60. Gérard E. F.; Mokkawes T.; Johannissen L. O.; Warwicker J.; Spiess R. R.; Blanford C. F.; Hay S.; Heyes D. J.; de Visser S. P. How Is Substrate Halogenation Triggered by the Vanadium Haloperoxidase from Curvularia inaequalis?. ACS Catal. 2023, 13, 8247–8261. 10.1021/acscatal.3c00761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Lin Y.-T.; Ali H. S.; de Visser S. P. Electrostatic perturbations from the protein affect C–H bond strengths of the substrate and enable negative catalysis in the TmpA biosynthesis enzyme. Chem. Eur. J. 2021, 27, 8851–8864. 10.1002/chem.202100791. [DOI] [PubMed] [Google Scholar]
  62. Ali H. S.; Warwicker J.; de Visser S. P. How Does the Nonheme Iron Enzyme NapI React Through L-Arginine Desaturation Rather Than Hydroxylation? A QM/MM Study. ACS Catal. 2023, 13, 10705–10721. 10.1021/acscatal.3c02262. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic3c04246_si_001.pdf (864.9KB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES