You are currently viewing a new version of our website. To view the old version click .
Chemistry
  • Article
  • Open Access

22 September 2025

Theoretical Study on the Electronic Structure of Fe(0)–, Pd(0)–, and Pt(0)–Phosphine–Carbon Dioxide Complexes

and
1
National Laboratory of Renewable Energy, University of Pécs, H-7624 Pécs, Hungary
2
Department of General and Inorganic Chemistry, University of Pécs, H-7624 Pécs, Hungary
3
HUN-REN Research Group for Selective Chemical Syntheses, H-7624 Pécs, Hungary
4
Szentágothai Research Centre, H-7624 Pécs, Hungary
This article belongs to the Section Theoretical and Computational Chemistry

Abstract

The activation of carbon dioxide by transition metal complexes is a fundamental process in catalysis and carbon capture. In this study, density functional theory (DFT) calculations, combined with Quantum Theory of Atoms in Molecules (QTAIM) and Natural Orbitals for Chemical Valency (NOCV) analyses, were employed to investigate the bonding characteristics and electronic structure of Fe(0)–, Pd(0)–, and Pt(0)–phosphine complexes with CO2. The Fe(0) complexes exhibited the strongest CO2 activation, characterized by substantial C=O bond elongation, significant charge transfer, and strong π -backdonation. In contrast, Pd(0) complexes showed minimal CO2 activation, while Pt(0) complexes displayed intermediate behavior. The electronic effects of phosphine ligands were also analyzed, revealing that electron-donating phosphines enhance CO2 activation, whereas electron-withdrawing phosphines weaken metal–CO2 interactions. These findings provide key insights into the design of transition-metal-based catalysts for CO2 conversion and utilization.

1. Introduction

In recent years, the urgent need to address environmental concerns, particularly global climate change and greenhouse gas emissions, has led to significant advancements in the field of green chemistry and sustainable technology. One of the exciting areas of research within this domain is the study of transition metal–carbon dioxide complexes [,,,]. These complexes offer promising opportunities for both catalysis and carbon capture, potentially paving the way towards a more sustainable future.
Transition metals (TMs) exhibit unique electronic properties due to their partially filled d-orbitals. These properties make them excellent candidates for forming coordination complexes with various ligands, including CO2. Carbon dioxide, a greenhouse gas notorious for its role in global warming, is challenging to capture and convert [,,,] due to its inert nature. Transition metal-phosphine complexes have demonstrated the ability to interact with CO2 and these interactions are mainly governed—in a simplified manner—by the Lewis acid–base chemistry, where the transition metal acts as a Lewis acid, accepting electron pairs from the Lewis base CO2 [,,,,,,,,]. For instance, the trigonal bipyramidal Fe(η2−CO2)(depe)2 complex was synthesized from the dinitrogen complex Fe(N2)(depe)2 []. Complex Fe(N2)(depe)2 was fully characterized by 31P NMR spectroscopy, revealing four chemically different phosphorus atoms in the coordination sphere of iron, with chemical shifts in the range of 68.13 to 92.44 ppm, and ddd spin patterns with a large (J = 127 Hz) coupling constant for the P atoms accommodating trans positions. The NMR study of this Fe–CO2 complex is interesting to compare with that of the oldest reported transition metal–carbon dioxide complex (Aresta-complex, Ni(PCy3)2(CO2)) []. In solution, a single sharp resonance at 36.1 ppm was found in the 31P NMR spectrum; however, two different isotropic peaks (at 22.0 and 45.2 ppm) were reported for the solid-state CPMAS 31P NMR measurement suggesting fluxional behavior only in solution.
Transition metal complexes have shown immense potential as catalysts for the conversion of CO2 into useful chemicals. One of the most significant achievements in this area is the development of catalytic systems that can selectively reduce CO2 to carbon monoxide or even further to methanol [,,,,]. Both CO and methanol are essential feedstocks in the chemical industry, and their production from CO2 could significantly reduce greenhouse gas emissions while simultaneously providing valuable resources. The activation of CO2 using transition metal complexes to facilitate the synthesis of cyclic carbonates has already been also explored []. These compounds have applications as solvents, electrolytes, and even in the production of biodegradable plastics, presenting an exciting avenue for sustainable material development. Transition metal complexes have been proved to be effective systems for CO2 coordination and reduction, offering pathways for converting CO2 into valuable feedstocks not only as carbon monoxide (CO) and methanol, but cyclic carbonates as well [,,,].
The interaction between CO2 and a transition metal center is governed by a balance of electrostatic interactions, π backdonation, and charge transfer effects. To gain deeper insight into these interactions, Quantum Theory of Atoms in Molecules (QTAIM) [,] and Natural Orbitals for Chemical Valency (NOCV) [,,] methods provide an advanced means to analyze metal–ligand interactions at the electronic structure level. This study focuses on Fe(0)–, Pd(0)–, and Pt(0)–phosphine complexes, aiming to understand how metal identity and phosphine ligands influence CO2 activation.
Our focus on Fe(0), Pd(0), and Pt(0) phosphine systems reflects a design choice to probe CO2 activation across complementary late-transition-metal manifolds while keeping the ligand class fixed (phosphines). Together with our previous studies [,,,,], they span 3d/4d/5d chemistry and distinct d-electron counts (Fe0: d 8 ; Pd0/Pt0: d 10 ), allowing us to disentangle metal-identity and ligand-basicity effects under phosphine-saturated conditions. The set is anchored experimentally by structurally and spectroscopically well-characterized Fe(0)–phosphine–CO2 adducts and through the extensive Pd(0)/Pt(0)–phosphine literature, providing robust external reference points for our computed trends [].

2. Computational Details

All the structures were optimized without symmetry constraints with tight convergence criteria using the Gaussian suite [] of programs with the exchange and correlation functional denoted as B97-D3 []. For all atoms, the def2-TZVP [] basis set was employed. According to our previous benchmark studies, this model chemistry proved robust among several DFT methods in terms of predicting vibrational frequencies of TM–carbonyl complexes [], and for producing reliable geometries and reaction energetics []. All IR data were obtained from harmonic vibrational frequency calculations on fully optimized geometries at the same level of theory used for optimization. Analytical second derivatives were evaluated to yield normal-mode frequencies and IR intensities (from dipole-moment derivatives). Each stationary point was verified as a true minimum by the absence of imaginary frequencies.
All singlet structures were optimized with spin-restricted Kohn–Sham (RKS) DFT. For every reported singlet minimum, we ran internal stability analyses in Gaussian (Stable = Opt) and then seeded UKS (singlet) from the converged RKS (Guess = Mix, Stable = Opt). In our four representative cases (Fe, Pt; ligands dmpe, dtfmpe), UKS (singlet) collapsed back to RKS in all instances (identical energies within SCF precision. Triplet-state test calculations were treated with spin-unrestricted Kohn–Sham (UKS) DFT.
For the complexes containing conformationally flexible ligands, conformational analyses have been performed in a similar manner reported earlier and the lowest energy species were considered [,]. QTAIM (Quantum Theory of Atoms in Molecules) [,] analyses of the wave function were carried out with AIMAll [] software. For the EDA-NOCV [,,] calculations, the ADF software [] from the AMS 2025 suite was used, employing the triple- ζ STO basis set for all atoms with one set of polarization functions (denoted as TZP) and no frozen core []. Total and partial density-of-states (TDOS and PDOS) calculations were performed with the Multiwfn program [,] using the Hirshfeld method, with the Gaussian broadening function width 1.36057 eV as the full width at half maximum.

3. Results and Discussion

The QTAIM and NOCV analyses provide significant insights into the electronic structure and bonding characteristics of Fe(0)–, Pd(0)–, and Pt(0)–phosphine complexes coordinated with CO2 (Figure 1). Through a careful examination of electron density distributions, orbital interactions, and charge transfer effects, all governed by the ligand basicity and bite angle, we can discern the relative abilities of these transition metal complexes to activate CO2 and engage in catalytic transformations.
Figure 1. Generic 2D structure of the carbon dioxide complexes.

3.1. QTAIM Analysis: Bonding Trends and Metal—CO2 Interactions

The QTAIM analysis allowed us to determine how electron density is distributed within the metal–CO2 complexes, highlighting the bonding strength and activation of CO2 upon coordination. The electron density ( ρ B C P ) at the metal-carbon bond critical point (M–C) was highest in Fe(0) complexes, indicating a strong covalent interaction between Fe and CO2. In contrast, Pd(0) exhibited the weakest interaction, as evidenced by the significantly lower ρ B C P values. This suggests that Fe(0) is the most efficient at activating CO2, whereas Pd(0) plays a relatively passive role in CO2 coordination. Pt(0), as expected, exhibited intermediate bonding characteristics, forming a moderate interaction with CO2.
All structures considered in this study were singlet-state complexes. Triplet-state structures were only obtained for the species with strong electron-donating phosphines. With less basic ligands, such as dtfmpe, or dppe, the carbon dioxide ligand left the metal center during geometry optimization. Interestingly, the adducts Fe(dmpe)2(CO2) and Pt(dmpe)(CO2) possess a CO2 ligand coordinated in a η 1 manner (see Figure S1 in the Supporting Information). These structures are, however, not feasible energetically, as their relative free energies are higher by 6.7 and 30.8 kcal/mol, respectively, compared to their singlet-state counterparts.
Further evidence for CO2 activation was obtained from the analysis of C=O bond weakening upon coordination. The ρ B C P values for the C=O bonds decreased significantly in Fe(0) complexes, demonstrating the strongest activation and partial reduction of CO2. This trend correlates with the ability of Fe(0) to donate electron density into the antibonding orbitals of CO2. On the other hand, Pd(0) retained high ρ B C P values for C=O bonds, suggesting that CO2 remains largely unactivated in these complexes upon coordination.
According to the delocalization indices, the change in bite angle (i.e., going from bis-dimethylphosphino methane (dmpm) to bis-dimethylphosphino propane (dmpp)) results in a decrease in bond orders inside the CO2 ligands. On the other hand, the metal–carbon bond order shows a minimum for the ethandiphos-type ligands, that is, when the diphosphine forms a five-membered ring with the metal. A notable exception is the electron-withdrawing dtfmpe ligand, when attached to platinum, as it reveals the strongest metal–CO2 bond order compared to the other ligands containing trifluoromethyl substituents.
The V/G value, that is, the ratio of the potential and kinetic energy densities provided additional insights into the nature of the bonding. Fe(0) complexes exhibited the highest V/G values (1.55–1.56), confirming a stronger covalent component in the Fe–CO2 interaction. Conversely, Pd(0) had the lowest V/G values (1.45–1.50), indicating a somewhat more ionic character. Pt(0) complexes displayed intermediate values, consistent with their moderate CO2 activation capability (Table 1).
Table 1. QTAIM parameters of the CO2 complexes (in atomic units).

3.2. Geometric and IR Trends: CO2 Activation Efficiency

The extent of CO2 activation can also be understood from the geometric parameters of the complexes. The CO2 bond angle is a key indicator of activation, as a larger deviation from linearity suggests a stronger interaction with the metal center. Fe(0) complexes exhibited the smallest bond angles (132–142°), indicating substantial activation, while Pd(0) and Pt(0) complexes retained larger bond angles (145–152°), implying weaker activation (Table 2 and Figure 2).
Table 2. Geometric parameters of the different carbon dioxide complexes. Bond lengths are in Å, bond angles are in degrees.
Figure 2. Computed geometries of complexes Fe(dmpe)2(CO2) (a) Fe(dtfmpe)2(CO2) (b) Pt(dmpe)(CO2) (c), and Pt(dtfmpe)(CO2) (d). Bond distances are in Å, bond angles are in degrees.
The C=O bond lengths further reinforced this trend, as they were most elongated in Fe(0) complexes, signifying strong π -backdonation from Fe to CO2. Pd(0) complexes exhibited minimal elongation, indicating a notably weaker electronic interaction.
As expected, the asymmetric CO2 stretching frequencies show strong correlation with the Lewis-basicity of the diphosphine ligands. For instance, in the presence of the basic depe ligand, a stretching frequency of 1695 cm−1 was calculated. Comparing the ligands with a similar bite angle, but different basicity, the ν (CO2) value increased to 1718, 1731, and 1899 cm−1 for dmpe, dppe, and dtfmpe, respectively.

3.3. EDA-NOCV Analysis: Charge Transfer and Orbital Interactions

The EDA-NOCV analysis (energy decomposition analysis in combination with the natural orbitals of chemical valence methodology) provides critical insight into the charge transfer mechanisms that govern metal-ligand interactions, in our case the CO2 coordination.
As it was reported previously [,,,,] the dominating contribution of the orbital interaction energy is the back-donation part, initiated by a d-orbital of the metal center directing towards the π * orbitals of the CO2 ligand. Depending on the metal, and on the spectator ligands, the back-donation interaction energy can exceed −100 kcal/mol, whereas the π -donation component originating from carbon dioxide towards a low-population orbital of the metal is usually weaker than −20 kcal/mol.
The orbital interaction energy ( Δ E o r b ) was highest in Fe(0) complexes (−156.9 kcal/mol for Fe–dmpe), indicating a significant charge transfer between Fe and CO2. Pt(0) complexes showed a lower interaction energy (−131.5 kcal/mol), and Pd(0) exhibited the weakest charge transfer (−109 kcal/mol). As the Pd(0)-CO2 complexes were found to be the least stable according to our QTAIM studies [], we restrict here our NOCV studies to the Pt and Fe complexes, taking one example each for strong electron-donating spectator ligands (dmpe) and one as an electron withdrawing ligand (dtfmpe).
The deformation densities from the ETS-NOCV (Figure 3) analysis revealed that Fe(0) exhibited the largest π -backdonation component (−77.1 and −121.3 kcal/mol, for ligands dtfmpe and dmpe, respectively), confirming that bisligated Fe(diphosphine)2 types of complexes donate electron density efficiently into the antibonding π * orbitals of CO2. With a single coordinated diphosphine, Pt(0) displayed moderate charge flow for the back-donation interaction (−65.1 and −97.8 kcal/mol, respectively) (Table 3).
Figure 3. NOCV deformation density maps of complexes Fe(dmpe)2(CO2) (a,b) and Fe(dtfmpe)2(CO2) (c,d) describing the leading π –donor (a,c) and π –acceptor (b,d) interactions. The color code also indicates the direction of the charge transfer between fragments; the charge flows from the gold to the green regions.
Table 3. Ziegler-Rauk energy decomposition analyses for selected complexes. Energies are given in kcal/mol.
The influence of phosphine ligands was also apparent for the other energy components, such as Pauli-repulsion and electrostatic contribution (Table 4). Electron-donating phosphines (e.g., dmpe, depe) enhanced CO2 activation by increasing the overall interaction energy. In contrast, electron-withdrawing phosphines (e.g., dtfmpe) reduced Δ E i n t , thus results in a weaker meta–CO2 interaction.
Table 4. Orbital energy decomposition values for the leading deformation density contributions (NOCV), and Hirshfeld charges of the CO2 fragment. Energies are given in kcal/mol.
Overall, the combination of QTAIM and NOCV analyses reveals Fe(0) as the most effective metal center for CO2 activation, with Pt(0) as a moderate activator and Pd(0) as the least effective, providing valuable insights for designing CO2-reducing catalysts.
The Hirshfeld charge of a fragment is an indicator for the charge transfer capability of a metal-containing fragment strongly influenced by its spectator ligands. In Table 3, the Hirshfeld charge (QH) is also displayed for the dmpe and dtfmpe ligands in the case of the iron and platinum centers. The difference in the Lewis-basicity of the diphosphines is clearly reflected by the almost 0.2 difference in QH for both metals. As expected, CO2 is more negative in the Fe(0) complexes where there are two diphosphines to increase the electron density on the metal.

3.4. DOS/PDOS Analysis: Electronic Structure Depending on Phosphine Substituents

Figure 4 compares the total and projected densities of states (DOS/PDOS) for two Fe–CO2 complexes containing diphosphines with either electron-donating (methyl, EDG, top) or electron-withdrawing (trifluoromethyl, EWG, bottom) substituents. The dashed line marks the HOMO edge, and representative HOMO–1/HOMO/LUMO isosurfaces are shown.
Figure 4. DOS/PDOS plots with selected molecular orbitals of Fe(dmpe)2(CO2) (top) and Fe(dtfmpe)2(CO2) (bottom).
Replacing EDGs by EWGs shifts the Fe-centered PDOS envelope to lower energies relative to the Fermi-level ( E F ). This stabilization of the d-manifold is the expected response to reduced σ donation and enhanced ligand electronegativity. Conversely, EDGs raise the metal d levels, rendering the Fe center more electron-rich. Near the frontier region, the overlap of Fe (red) and CO2 (blue) PDOS is markedly larger for EDGs. In the EDG complex, the carbon dioxide intensity that appears in the LUMO region tracks with a non-negligible red (Fe) contribution, and the LUMO isosurface shows clear mixing of Fe d with the CO2 π * ) framework. In contrast, for EWGs, the blue peak nearest to the LUMO is weaker and less coincident with Fe intensity; the LUMO is more ligand-localized with the diminished Fe admixture.
Stronger back-donation (EDG case) implies a more activated CO2: greater O–C–O bending, longer C–O bonds, and a lower ν (C=O) stretching frequency. The EWG complex should display the opposite: less bending/elongation and higher ν (C=O). These trends are consistent with the qualitative frontier-orbital pictures shown (HOMO/HOMO-1 bearing more Fe–phosphine character, LUMO with larger CO2 π * ) weight for EDGs).

4. Conclusions

This study highlights the critical role of transition metal identity and phosphine ligand effects in CO2 activation. The Fe(0) complexes demonstrated the strongest CO2 activation, as evidenced by the significant elongation of C=O bonds, high charge transfer, and substantial π -backdonation. In contrast, Pd(0) complexes exhibited the weakest CO2 activation, indicating limited catalytic potential for CO2 reduction. Pt(0) complexes displayed an intermediate degree of activation, suggesting their potential for selective CO2 functionalization. Furthermore, the choice of a phosphine ligand was shown to be crucial, with electron-donating phosphines enhancing CO2 activation, whereas electron-withdrawing phosphines weakened the metal–CO2 interaction. These findings provide valuable insights for designing transition metal catalysts optimized for CO2 reduction and conversion, contributing to the advancement of sustainable chemistry.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry7050152/s1, The internal energies and Cartesian coordinates of all species occuring in this study, as well as Figure S1 (UB97-D3/def2tzvp triplet-state structures).

Author Contributions

Conceptualization and methodology, T.K.; calculations, T.R.K. and T.K.; writing, review, and editing, T.R.K. and T.K. All authors have read and agreed to the published version of the manuscript.

Funding

Project no. RRF-2.3.1-21-2022-00009, titled National Laboratory for Renewable Energy, has been implemented with the support provided by the Recovery and Resilience Facility of the European Union within the framework of Programme Széchenyi Plan Plus; Project number 014_2023_PTE_RK/14.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yin, X.; Moss, J.R. Recent developments in the activation of carbon dioxide by metal complexes. Coord. Chem. Rev. 1999, 181, 27–59. [Google Scholar] [CrossRef]
  2. Braunstein, P.; Matt, D.; Nobel, D. Reactions of carbon dioxide with carbon-carbon bond formation catalyzed by transition-metal complexes. Chem. Rev. 1988, 88, 747–764. [Google Scholar] [CrossRef]
  3. Behr, A. Carbon dioxide as an alternative C1 synthetic unit: Activation by transition-metal complexes. Angew. Chem. Int. Ed. 1988, 27, 661–678. [Google Scholar] [CrossRef]
  4. Paparo, A.; Okuda, J. Carbon dioxide complexes: Bonding modes and synthetic methods. Coord. Chem. Rev. 2017, 334, 136–149. [Google Scholar] [CrossRef]
  5. Yang, Z.Z.; He, L.N.; Gao, J.; Liu, A.H.; Yu, B. Carbon dioxide utilization with C–N bond formation: Carbon dioxide capture and subsequent conversion. Energy Environ. Sci. 2012, 5, 6602–6639. [Google Scholar] [CrossRef]
  6. Yuan, Z.; Eden, M.R.; Gani, R. Toward the development and deployment of large-scale carbon dioxide capture and conversion processes. Ind. Eng. Chem. Res. 2016, 55, 3383–3419. [Google Scholar] [CrossRef]
  7. Lu, B.; Fan, Y.; Zhi, X.; Han, Z.; Wu, F.; Li, X.; Luo, C.; Zhang, L. Material design and prospect of dual-functional materials for integrated carbon dioxide capture and conversion. Carbon Capture Sci. Technol. 2024, 12, 100207. [Google Scholar] [CrossRef]
  8. Cokoja, M.; Bruckmeier, C.; Rieger, B.; Herrmann, W.A.; Kühn, F.E. Transformation of carbon dioxide with homogeneous transition-metal catalysts: A molecular solution to a global challenge? Angew. Chem. Int. Ed. 2011, 50, 8510–8537. [Google Scholar] [CrossRef]
  9. Aresta, M.; Nobile, C.F.; Albano, V.G.; Forni, E.; Manassero, M. New nickel–carbon dioxide complex: Synthesis, properties, and crystallographic characterization of (carbon dioxide)-bis (tricyclohexylphosphine) nickel. J. Chem. Soc. Chem. Commun. 1975, 636–637. [Google Scholar] [CrossRef]
  10. Durfy, C.S.; Zurakowski, J.A.; Jobin, G.; Drover, M.W. An Investigation of Allyl-Substituted Bis (Diphosphine) Iron Complexes: Towards Precursors for Cooperative CO2 Activation. Chem. J. 2024, 30, e202302721. [Google Scholar] [CrossRef]
  11. Metters, O.J.; Forrest, S.J.; Sparkes, H.A.; Manners, I.; Wass, D.F. Small molecule activation by intermolecular Zr (IV)-phosphine frustrated Lewis pairs. J. Am. Chem. Soc. 2016, 138, 1994–2003. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, X.; Gong, J.K.; Collins, A.W.; Grove, L.J.; Seyler, J.W. Theoretical study of carbon dioxide coordination in palladium complexes. Appl. Organomet. Chem. 2001, 15, 95–98. [Google Scholar] [CrossRef]
  13. Kégl, T.; Ponec, R.; Kollár, L. Theoretical Insights into the Nature of Nickel Carbon Dioxide Interactions in Pt(PH3)22-CO2). J. Phys. Chem. A 2011, 115, 12463–12473. [Google Scholar] [CrossRef] [PubMed]
  14. Kégl, T.R.; Carrilho, R.M.; Kégl, T. Theoretical insights into the electronic structure of nickel (0)-diphosphine-carbon dioxide complexes. J. Organomet. Chem. 2020, 924, 121462. [Google Scholar] [CrossRef]
  15. Kégl, T.R.; Kégl, T. Comparative Analysis of Nickel–Phosphine Complexes with Cumulated Double Bond Ligands: Structural Insights and Electronic Interactions via ETS-NOCV and QTAIM Approaches. Molecules 2024, 29, 324. [Google Scholar] [CrossRef]
  16. Kégl, T.R.; Kollár, L.; Kégl, T. DFT Study on the Mechanism of Iron-Catalyzed Diazocarbonylation. Molecules 2020, 25, 5860. [Google Scholar] [CrossRef]
  17. Pálinkás, N.; Kollár, L.; Kégl, T. Nature of the Metal-Ligand Interactions in Complexes M(PH3)2(η2-L)(M = Ni, Pd, Pt; L = CO2, COS, CS2): A Theoretical Study. ChemistrySelect 2017, 2, 5740–5750. [Google Scholar] [CrossRef]
  18. Komiya, S.; Akita, M.; Kasuga, N.; Hirano, M.; Fukuoka, A. Synthesis, structure and reactions of a carbon dioxide complex of iron (0) containing 1, 2-bis (diethylphosphino) ethane ligands. J. Chem. Soc. Chem. Comun. 1994, 1115–1116. [Google Scholar] [CrossRef]
  19. Jegat, C.; Fouassier, M.; Tranquille, M.; Mascetti, J.; Tommasi, I.; Aresta, M.; Ingold, F.; Dedieu, A. Carbon dioxide coordination chemistry. 3. Vibrational, NMR, theoretical studies of (carbon dioxide)bis(tricyclohexylphosphine)nickel. Inorg. Chem. 1993, 32, 1279–1289. [Google Scholar] [CrossRef]
  20. Navarro-Jaén, S.; Virginie, M.; Bonin, J.; Robert, M.; Wojcieszak, R.; Khodakov, A.Y. Highlights and challenges in the selective reduction of carbon dioxide to methanol. Nat. Rev. Chem. 2021, 5, 564–579. [Google Scholar] [CrossRef]
  21. Boutin, E.; Wang, M.; Lin, J.C.; Mesnage, M.; Mendoza, D.; Lassalle-Kaiser, B.; Hahn, C.; Jaramillo, T.F.; Robert, M. Aqueous electrochemical reduction of carbon dioxide and carbon monoxide into methanol with cobalt phthalocyanine. Angew. Chem. Int. Ed. 2019, 58, 16172–16176. [Google Scholar] [CrossRef] [PubMed]
  22. Shen, J.; Kortlever, R.; Kas, R.; Birdja, Y.Y.; Diaz-Morales, O.; Kwon, Y.; Ledezma-Yanez, I.; Schouten, K.J.P.; Mul, G.; Koper, M.T. Electrocatalytic reduction of carbon dioxide to carbon monoxide and methane at an immobilized cobalt protoporphyrin. Nat. Commun. 2015, 6, 8177. [Google Scholar] [CrossRef]
  23. Agarwala, H.; Chen, X.; Lyonnet, J.R.; Johnson, B.A.; Ahlquist, M.; Ott, S. Alternating Metal-Ligand Coordination Improves Electrocatalytic CO2 Reduction by a Mononuclear Ru Catalyst. Angew. Chem. Int. Ed. 2023, 135, e202218728. [Google Scholar] [CrossRef]
  24. Yao, Y.; Wu, J.H.; Liu, G.; Zhang, R.; Yang, Z.S.; Gao, S.; Lau, T.C.; Zhang, J.L. A Bio-Inspired Bimetallic Fe-M Catalyst for Electro-and Photochemical CO2 Reduction. ChemCatChem 2024, 16, e202301705. [Google Scholar] [CrossRef]
  25. Guo, L.; Lamb, K.J.; North, M. Recent developments in organocatalysed transformations of epoxides and carbon dioxide into cyclic carbonates. Green Chem. 2021, 23, 77–118. [Google Scholar] [CrossRef]
  26. Cokoja, M.; Wilhelm, M.E.; Anthofer, M.H.; Herrmann, W.A.; Kühn, F.E. Synthesis of cyclic carbonates from epoxides and carbon dioxide by using organocatalysts. ChemSusChem 2015, 8, 2436–2454. [Google Scholar] [CrossRef]
  27. Lopes, E.J.; Ribeiro, A.P.; Martins, L.M. New trends in the conversion of CO2 to cyclic carbonates. Catalysts 2020, 10, 479. [Google Scholar] [CrossRef]
  28. Alves, M.; Grignard, B.; Méreau, R.; Jerome, C.; Tassaing, T.; Detrembleur, C. Organocatalyzed coupling of carbon dioxide with epoxides for the synthesis of cyclic carbonates: Catalyst design and mechanistic studies. Catal. Sci. Technol. 2017, 7, 2651–2684. [Google Scholar] [CrossRef]
  29. Richard, F.; Bader, R. Atoms in Molecules: A Quantum Theory; Oxford Academic: Oxford, UK, 1990. [Google Scholar]
  30. Bader, R.F. Bond paths are not chemical bonds. J. Phys. Chem. A 2009, 113, 10391–10396. [Google Scholar] [CrossRef]
  31. Mitoraj, M.; Michalak, A. Natural orbitals for chemical valence as descriptors of chemical bonding in transition metal complexes. J. Mol. Model. 2007, 13, 347–355. [Google Scholar] [CrossRef]
  32. Mitoraj, M.; Michalak, A. Applications of natural orbitals for chemical valence in a description of bonding in conjugated molecules. J. Mol. Model. 2008, 14, 681–687. [Google Scholar] [CrossRef]
  33. Mitoraj, M.P.; Michalak, A.; Ziegler, T. A combined charge and energy decomposition scheme for bond analysis. J. Chem. Theory Comput. 2009, 5, 962–975. [Google Scholar] [CrossRef]
  34. 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.; et al. Gaussian 09, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  35. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  36. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  37. Kégl, T.; Pálinkás, N.; Kollár, L.; Kégl, T. Computational Characterization of Bidentate P-Donor Ligands: Direct Comparison to Tolman’s Electronic Parameters. Molecules 2018, 23, 3176. [Google Scholar] [CrossRef]
  38. Papp, T.; Nagy, P.R.; Kegl, T. Advanced computation of enthalpies for a range of hydroformylation reactions with a predictive power to match experiments. Chem. Phys. Lett. 2025, 861, 141833. [Google Scholar] [CrossRef]
  39. Felgueiras, A.P.; Rodrigues, F.M.; Carrilho, R.M.; Cruz, P.F.; Rodrigues, V.H.; Kégl, T.; Kollár, L.; Pereira, M.M. Stereoisomeric Tris-BINOL-Menthol Bulky Monophosphites: Synthesis, Characterisation and Application in Rhodium-Catalysed Hydroformylation. Molecules 2022, 27, 1989. [Google Scholar] [CrossRef]
  40. Keith, T.A. AIMAll, Version 15.05.18; TK Gristmill Software: Overland Park, KS, USA, 2015; Available online: https://aim.tkgristmill.com (accessed on 3 May 2024).
  41. AMS2025; SCM, Theoretical Chemistry; Vrije Universiteit: Amsterdam, The Netherlands, 2025; Available online: http://www.scm.com (accessed on 8 August 2025).
  42. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  43. Lu, T. A comprehensive electron wavefunction analysis toolbox for chemists, Multiwfn. J. Chem. Phys. 2024, 161, 082503. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.