Next Article in Journal
Structural Diversity of Heteroleptic Cobalt(II) Dicyanamide Coordination Polymers with Substituted Pyrazines and Pyrimidines as Auxiliary Ligands
Previous Article in Journal
Quantitative Profiling of Phenolic Constituents in Hypericum perforatum L. via HPLC–PDA and HPLC–ECD: A Chemometric Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Thorough Understanding of Methylrhodium(III)–Porphyrin Photophysics: A DFT/TDDFT Study

Institute of Chemistry, University of Silesia in Katowice, Szkolna 9, 40-006 Katowice, Poland
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(19), 3855; https://doi.org/10.3390/molecules30193855
Submission received: 12 August 2025 / Revised: 15 September 2025 / Accepted: 18 September 2025 / Published: 23 September 2025

Abstract

Rhodium–porphyrin complexes are characterised by their ability to activate C-H and C-C bonds and, therefore, find applications in synthesis and catalysis. Axial rhodoporphyrin ligands are susceptible to photodissociation under the influence of light. DFT and TDDFT calculations were performed to investigate the mechanism of photodissociation of the methyl ligand from the methylrhodium(III)–porphyrin complex (MeRhPor). Various photolysis pathways of the rhodium–methyl bond were investigated, including photolysis from states in the Q and Soret bands. Photolysis from triplet states was also considered. Based on the calculations, the most probable mechanism for photodissociation of the methyl ligand was proposed. The methyl-rhodium binding energy in the methylrhodium(III)–porphyrin complex and the energy of formation of the rhodium–porphyrin radical dimer formed by methyl dissociation were also calculated.

Graphical Abstract

1. Introduction

Metalloporphyrins are complexes of porphyrins with transition metal ions and primarily perform important biological functions in nature, including oxygen transport using haem proteins, photosynthesis involving chlorophylls, and oxidation reactions with cytochromes P450, and are involved in numerous redox processes [1,2,3,4,5,6,7,8,9,10,11]. Many synthetic metalloporphyrins with various metal ions have been synthesised to develop artificial molecular receptors, enzyme mimics, DNA cleavers, photodynamic therapeutic agents, chiral catalysts, new materials, etc. [12,13,14,15,16,17,18,19,20,21,22]. Rhodium–porphyrins have been extensively studied over the last few decades, mainly in aspects of selective catalytic activation of C–H and C–C bonds [23,24,25,26,27].
Considering the wide range of applications of metalloporphyrins, the photophysics and photochemistry of rhodium–porphyrins play a key role in the research and application of these complexes [28,29,30,31,32,33]. Rhodium–porphyrins display strong luminescent properties due to the large aromatic network of the porphyrin ring. Alkyl substituents on the porphyrin, e.g., in OEP (octaethylporphyrin), have been shown to play a small role in the photophysical properties of the resulting complexes [34,35], whereas aryl substituents such as in TPP (meso-tetraphenylporphyrin) significantly increase the phosphorescence and fluorescence properties of these complexes. Rhodium–porphyrins exhibit moderate-to-strong phosphorescence and only weak fluorescence properties [36,37]. For five- or six-coordinate complexes, photodissociation of axial ligands is also possible. Interestingly, ligand photodissociation can occur via either a singlet or triplet photochemical pathway [38,39]. Specifically, the experimentally observed photodissociation processes involving the singlet or triplet state depend on the nature of the ligand. In the case of MeRhOEP (methylrhodium(III)–octaethylporphyrin), photodissociation of the methyl ligand is observed exclusively for the singlet photophysical channel [39]. Similarly, photodissociation of CO from (X)(CO)RhTPP (carbonylrhodium(III)–meso-tetraphenylporphyrin, X = I, Cl) occurs solely from the singlet state [38]. Since the carbonyl ligand is a strong π acceptor, it is postulated that the singlet channel applies to the photodissociation of the acceptor-type ligands. Conversely, for (Cl)(Py)RhTPP (pyridinerodium(III)–meso-tetraphenylporphyrin), photodissociation of pyridine occurs exclusively from the triplet state. Given that pyridine is a donor ligand, it is believed that the triplet state is responsible for the dissociation of the σ-type donor ligand [38]. In a situation where a ligand has both donor σ and acceptor π features, it is assumed that the ligand dissociation can occur through both singlet and triplet photophysical channels. This property was observed for the photodissociation of the isonitrile ligand (R-NC, 2,6-Dimethylphenyl isocyanide) in the (I)(R-NC)RhTPP complex [38].
The methylrhodium(III)–octaethylporphyrin (MeRhOEP) complex, due to the simple electronic structure of the methyl axial ligand, can be considered as a model system for photophysical processes. Experimental studies of MeRhOEP photochemistry reveal the occurrence of different photophysical channels, leading to processes such as photodissociation and phosphorescence [39]. The basic photophysical paths for this complex are presented in Figure 1.
The MeRhOEP molecule absorbs light at two wavelengths, λ = 543 nm and 395 nm, which correspond to the SQ and SSoret singlet states, respectively. In a degassed benzene solution, there are two possible ways of deactivating excited singlet states of MeRhOEP: radiationless intersystem crossing (ISC) to the lowest triplet state and photochemical cleavage of the Rh-CMe bond with involvement of only singlet states. Since, according to laser photolysis studies, there is no transition from the SSoret to the SQ states, the photodissociation of the Rh-CMe bond occurs in two independent channels depending on the excitation wavelength region.
MeRhOEP + (λ > 410 nm) → SQ → X → RhOEP + Me
MeRhOEP + (λ < 410 nm) → SSoret → RhOEP + Me
For irradiation with light of λ > 410 nm, the temperature dependence of the photoreaction indicates that the dissociation of the Rh-CMe bond in the excited singlet state requires activation energy, i.e., the intermediate X is formed thermally from the SQ states, and the activation energy is estimated to be 7.4 kcal/mol [39]. Low-temperature photolysis studies revealed that photocleavage of the Rh-CMe bond with irradiation light of λ < 410 nm takes place even at low temperatures, pointing to the lack of an energy barrier on the photoreaction pathway. The long-lived species X is regarded as an intermediate on the singlet path of photodissociation from the SQ band; however, the structure or the electronic state of X is not known [39]. Based on the analysis of the experimental results of the quantum yield of the dissociative process and the triplet state occupation, it was concluded that the triplet state is not responsible for the photochemical reaction induced by irradiation with light of both wavelength regions λ < 410 nm and λ > 410 nm. The triplet state is deactivated through phosphorescence, which is observed at a wavelength of λ = 650 nm, and it is not involved in the process of breaking the rhodium–carbon bond.
Among the possible photophysical pathways, the channel leading to photolytic axial bond cleavage appears to be one of the most beneficial processes from the point of view of practical applications. This is related, firstly, to the relatively high energy efficiency in the case of a sufficiently high quantum yield of photohomolysis and, secondly, to the ease of control of the bond cleavage on an “on–off” basis. As long as the first aspect is important in designing photocatalytic systems, the second will find application in situations requiring the controlled release of radicals generated during photolysis. It should be emphasised here that, in general, the efficiency of a given photophysical process, including photolysis, is the result of the electronic structure of the complex but is also determined by the environment of the complex and is dependent on factors such as the nature of the solvent, the viscosity of the medium, the presence of oxygen, etc. Explaining the influence of these and other factors on photophysics is difficult without understanding the mechanistic, submolecular interaction of electronic states.
The present work describes the results of DFT and TDDFT calculations for the model structure of methylrhodium(III)–porphyrin (MeRhPor) on issues such as the equilibrium geometry of the ground state and the first excited state, the rhodium–carbon bond energy, the electronic structure and energetics of vertical excited states, and the energetics of the excited states as a function of the Rh-CMe bond length, as well as the interaction of singlet and triplet states of the methylrhodium(III)–porphyrin complex.

2. Results

2.1. Geometry of MeRhPor Complex and Rh-CMe Bond Dissociation Energy

The most relevant geometrical parameters of the coordination sphere for the methylrhodium(III)–porphyrin complex (MeRhPor) and the product of the homolytic photodissociation of the methyl radical, rhodium(II)–porphyrin (RhPor), are shown in Table 1. As expected from the crystallographic data, the structure of the optimised equatorial sphere of the MeRhPor complex in the ground state S0, comprising the central ion and the porphyrin ligand, is almost flat. The calculated geometrical parameters of the equatorial sphere fully agree with the experimental values. The optimised axial bond length of Rh-CMe is only slightly larger, by about 0.03 Å, as compared to the crystallographic data. The calculated length of this bond practically does not change for the optimised geometries of the lowest excited singlet (S1) and triplet (T1) states. Rh-N bond lengths slightly increase by ~0.01 Å and ~0.02 Å for the S1 and T1 states, respectively. At the same time, the valence and torsion angles of the coordination sphere in the excited states S1 and T1 are not significantly different compared to the geometry of the ground state (see also Table S1 in the Supplementary Materials). A superimposition of the optimised geometries in the S0 and S1 electronic states of the MeRhPor complex, presented in Figure 2, clearly shows that electronic excitation to the S1 state does not involve a significant change in the geometry of the complex. Therefore, it can be noted here that as a result of vertical electronic excitation S0 → S1, the geometry relaxation process in the excited S1 state should occur to a minimal extent. For the photolysis product, RhPor, the optimised interatomic distances of Rh-N increase minimally, on average by about 0.006 Å in relation to the optimised geometry of MeRhPor in the ground state; at the same time, the values of the calculated angles indicate a completely flat geometry of the coordination sphere. In the case of the lowest singlet excited states of RhPor, the geometry of the complex changes only very slightly in relation to the ground state S0. In the coordination sphere, the changes are essentially in the lengths of the Rh-N coordination bonds and are most probably caused by a change in the distribution of electronic density on the central ion as a result of electronic excitation.
The determined value of the Rh-CMe binding energy using the PBE0 functional is 49.9 kcal/mol, and it is 8.1 kcal/mol and 4.4 kcal/mol lower than the reported experimental BDE values. Taking into account the ZPE correction and thermal corrections to the energy increases this difference by another 4 kcal/mol, giving a computationally estimated dissociation energy value of 45.8 kcal/mol. Dispersion correction considerably reduces this difference, giving values of 54.7 kcal/mol and 50.5 kcal/mol for ΔEDE and ΔEBDE, respectively, which brings the estimated binding energy closer to the experimental values. The calculated value of 39.9 kcal/mol for the bond dissociation free energy (ΔGBDFE) also coincides with the experiment.
Test calculations for a widely used B3LYP hybrid functional produce a slightly larger discrepancy between the experimental data and the calculated BDE value. Including the dispersion contribution in DFT/B3LYP calculations gives practically the same ΔEDE, ΔEBDE, and ΔGBDFE values as the PBE0 functional. The tested functionals of the GGA types, BP86 and PBE, give higher ΔEDE bond energy values than hybrid functionals, whereas taking into account the ZPE value and thermal corrections, a bond dissociation energy (ΔEBDE) close to the experimental data of 54 kcal/mol is obtained. Taking dispersion corrections into account significantly increases the calculated ΔEDE bond energy, and the obtained ΔEBDE values are in this case close to the experimental value of 58 kcal/mol.

2.2. Simulated UV/VIS Spectrum of MeRhPor and Character of Excited Electronic States

Based on the results of calculations at the TDDFT level of theory, the absorption spectrum of the MeRhPor complex was simulated. The simulated absorption spectrum, along with the calculated excitation wavelengths and oscillator strengths, is shown in Figure 3. The simulated spectrum contains two main bands at 463 nm and 330 nm and one additional band at 364 nm. The bands at 463 nm and 330 nm can be interpreted as characteristic Q and Soret bands in the experimental spectrum. From the experiment, the position of the bands for methylrhodium(III)–octaethylporphyrin, MeRhOEP, corresponds to the following wavelengths: 543 nm, 510 nm, and 395 nm for Q(0,0), Q(1,0), and Soret(0,0), respectively [39]. Comparing the above values with each other, it is clear that the primary absorption bands in the simulated spectrum obtained from the TDDFT/PBE0 results are consistently shifted toward shorter wavelengths, including the Q band by 80 nm and the Soret band by 65 nm. Converted into energy units, this corresponds to a shift of 0.4 eV and 0.6 eV for the Q band and Soret band, respectively. Relative to the experiment, usually, such a blue shift of 50–100 nm for the main bands of the spectrum is very characteristic when hybrid functionals are used in calculations for transition metal complexes with tetrapyrrole ligands.
According to the results in Table 2, the simulated Q band is the result of electronic excitations to the first two degenerate singlet states S1 and S2, while the Soret band is a consequence of electronic transition to the degenerate states S9 and S10. The additional band in the simulated spectrum at 364 nm appears due to the presence of two states, S7 and S8, for which the calculated oscillator strength takes on a relatively higher value. It is likely that in the experimental spectrum, transitions to the S7–S10 states may form only one visible absorption band, or the actual moment of transition to the S7 and S8 states is much smaller than TDDFT calculations would suggest.
It is generally recognised that the characteristic Q and Soret absorption bands in the spectra of metalloporphyrins are the result of π → π* electronic excitations between occupied and unoccupied π orbitals localised on the porphyrin ring [44,45,46]. Also, it is assumed that the position of these bands is to some extent the result of the central ion’s interaction with the ligand. This interpretation regarding the electronic structure of singlet excited states, whose occupation as a result of photon absorption is observed in the form of the mentioned Q and Soret bands, results from the very characteristic energetic distribution of occupied and unoccupied π molecular orbitals. For the MeRhPor complex under consideration, the distribution of orbital energies within important frontier orbitals is shown in Figure 4. Due to the fact that the porphyrin ligand has a fully conjugated π-electronic system, the two highest occupied, degenerate orbitals H and H-1 and the two lowest unoccupied, also degenerate orbitals L and L+1 are π-type orbitals localised on the porphyrin ligand and have practically no admixtures of the orbitals of the central ion. Such a system of four π orbitals is commonly called Gouterman’s four-orbital model and is used to explain the absorption spectra of porphyrins. Among the frontier orbitals presented in Figure 4, the contribution of rhodium d orbitals is evident for H-4, H-3, H-2, and L+2. Orbital H-4 is a doubly occupied dx2-y2 orbital, and H-3 and H-2 are a combination of the dxz or dyz orbital with the π orbital of the porphyrin ligand, while L+2 is an unoccupied σ* orbital, which contains the dz2 orbital.
According to the TDDFT results given in Table 2, the first two degenerate singlet states, S1 and S2, correspond to an excitation energy of 2.68 eV and are characterised by a low value of the calculated oscillator strength. Both states are combinations of one-electron excitations between occupied H and H-1 orbitals and unoccupied L and L+1 orbitals; they are states of π → π* type. States S9 and S10 have excitations of an analogous nature; however in this case, the oscillator strengths are two orders of magnitude larger. The calculated excitation energy for these states is 3.75 eV. The remaining singlet states, presented in Table 2, are characterised by zero or medium oscillator strength and correspond to electronic excitations of varying character, but their common feature is that the Kohn–Sham orbitals involved in these transitions contain dxz and dyz rhodium orbitals. The S3–S6 states have a mixed π/d → π* character, whereas degenerate states S7 and S8 with an excitation energy of 3.40 eV and medium oscillator strength are of π/d → σ* type. The characterisation of singlet excited states, performed on the basis of natural transition orbital (NTO) analysis, is presented in Table S2 in the Supplementary Materials. For excited states S1–S12, the charge-transfer process in the particle–hole picture remains fully consistent with the excitation components expressed by the transitions between the frontier Kohn–Sham orbitals. Only in the case of the S9 and S10 states can a small contribution of electronic donation of the d → σ* type be observed, co-occurring with the dominant charge transfer of the π → π* character.
In Table S3 (Supplementary Materials), the triplet excited states are presented. The vertical triplet states form groups of a different character, π → π*, π/d → π*, and π/d → σ*. Below the lowest singlet states S1 and S2, four triplet states T1–T4 can be identified, appearing as two pairs of degenerate states with calculated excitation energies of 2.14 eV and 2.28 eV. The distribution of one-electron transitions characterising the electronic structure of these four states is analogous to that in the case of the singlet states S1 and S2 and S9 and S10 and corresponds to a combination of π → π* excitations. The subsequent triplet states have higher excitation energies relative to the lowest singlet states S1 and S2. Their electronic structure is more complex due to the fact that the d orbitals of rhodium are involved in electronic excitations. Triplet states T5–T8 located in the excitation energy range from 2.58 eV to 2.70 eV are of π/d → π* character. The next two states with excitation energies of 3.03 eV, T9 and T10, are of π/d → σ* type. The T11 and T12 states with an excitation energy of 3.58 eV are generally of π → π* type, and their electronic structure is similar to that of the T1–T4 states. From the point of view of MeRhPor photophysics, it is worth noting T34 σ→σ*, which is the non-bonding triplet state. In the paper [33], a state of this type was indicated as the state potentially responsible for the photoinduced homolysis of the Rh-C bond. However, in light of the experimental results in the work of [39], the above conclusion seems rather problematic. Based on the analysis of the decay rate of the triplet state and the quantum yield of the photodecomposition of the Rh-C bond for the aerated and degassed solution, respectively, it was concluded that the triplet state is not responsible for the photochemical reaction induced by irradiation with light of a wavelength in the regions λ < 410 nm and λ > 410 nm [39]. Thus, the mechanism of photodissociation of the axial bond emerging from the experiment indicates the singlet nature of photohomolysis. It should also be noted that the degassed benzene solution of MeRhOEP showed no ESR signal before and after photolysis. At the same time, the irradiated solution, when exposed to air, exhibits an intense ESR signal, and the ESR spectrum is in good agreement with that of the oxygen adduct of RhOEP, O2-RhOEP, produced from the [RhOEP]2 dimer [39].

2.3. Excited States of RhPor Complex and Dimerisation of Radical Complex

The RhPor complex is a direct product of photolytic Rh-CMe bond rupture. In the ground state D0, the complex is a radical in which the unpaired electron occupies the dz2 orbital, Figure S1 in the Supplementary Materials. This location of the electronic density of the unpaired electron is the most favourable energetically and allows the formation of a two-centre Rh-Rh bond. The formation of the (RhPor)2 dimer can be easily confirmed computationally, as shown in Figure S2 in the Supplementary Materials. Of course, in the axial bond photolysis process, the D0 state is not directly accessible because the methyl group photolysis in the dissociation limit leads to an electronic state with a doubly occupied dz2 orbital. This fact is obvious from the point of view of the electronic excitations obtained from TDDFT calculations, the results of which are shown in Table S4 in the Supplementary Materials. First of all, the five low-lying excited states D1–D5 are of interest. The first two states are degenerate and have a π/d → dz2 character, which can be referred to as LMCT/LF states. The next state D3 is an LF-type state d → d. The next two are π → dz2 LMCT-type transitions. The character of these doublet states can be considered as rhodium electronic states with different electronic configurations on d orbitals. The form of the Kohn–Sham frontier orbitals involved in electronic excitations is presented in the Supplementary Materials in Figures S3 and S4. For the RhPor complex, the π Guterman’s orbital system is analogous to the π H-1, H, L, and L+1 orbital system for the MeRhPor complex (Figure S4 vs. Figure 4). The main difference between the frontier orbitals of the RhPor and MeRhPor complex is the occurrence of a single occupied dz2 orbital in the RhPor radical (H-7 alpha and L beta orbitals in Figure S3). In the case of the MeRhPor complex, the elongation of the Rh-CMe axial bond practically does not change the character and energy position of the frontier orbitals; only in the case of σ and σ* orbitals localised on the axial bond, a significant change in orbital energy is observed, as shown in Figure S5 in the Supplementary Materials. The energy of the σ orbital increases towards the HOMO orbital, while the energy of the σ* orbital decreases, and with a significant elongation of the rhodium–carbon bond, the σ* orbital becomes the LUMO orbital. After breaking the Rh-CMe bond, the dz2 orbital of the rhodium in the RhPor complex becomes the equivalent of the σ and σ* orbitals.
The calculated excitation energies for the D1–D2, D3, D4, and D5 states are, respectively, 0.42 eV, 1.17 eV, 1.76 eV, and 1.85 eV. For relaxed geometries of the D1, D3, and D4 states, the corresponding excitation energies are 0.40 eV, 1.17 eV, and 1.65 eV. Geometry optimisation of the D2 and D5 states at the level of the TDDFT method is impossible due to the close degeneracy of the D1 and D2 and D4 and D5 states, but it can be assumed that both the energy and geometry of these two states are the same or very close to the energy and equilibrium geometry of the D1 and D4 states, respectively. Energetically, due to the low excitation energy of about 10 kcal/mol, the D1 and D2 states can be characterised as “hot” states, whose deactivation to the ground state is a thermal process of energy dissipation due to transitions between vibronic states. From the perspective of the above calculational results, the most probable states in the photodissociation limit of the Rh-CMe bond are D1 and D2 states of π/d → dz2 character. The relatively small difference in the energy of the states D1 and D2 and the ground state D0 gives rise to the conclusion that the lowest doublet excited states of the RhPor radical are quenched by internal conversion, IC, D1, and D2/D0, with the formation of the ground state with the configuration of the central ion (dx2−y2)2 (dxz,dyz)4 (dz2)1.

2.4. PECs as a Function of the Rh-C Bond Length

In order to determine the photophysical mechanisms of the Rh-CMe bond photodissociation process and the quenching of excited states, potential energy curves (PECs) were determined at the TDDFT level of theory as a function of the rhodium–methyl distance, shown in Figure 5 and Figure S6 in the Supplementary Materials.
A predominant number of PECs for excited singlet and triplet states have minima in the vicinity of the Rh-CMe equilibrium distance for the S0 ground state, i.e., ~2.00 Å. Only in the case of a few electronic states it can be observed that their PECs have broad, shallow minima at much larger distances, usually around 2.30 Å. Among these states, the most noteworthy are the S7 and S8 states, with the π/d → σ* excitation electronic structure, which at distances greater than 2.2 Å become the lowest, excited singlet states. A similar shape of PECs is found for the S13, S14, and S17 states. The S13 and S14 states are of π → σ* character, and the S17 state is of d → σ* type. Around the 2.35 Å distance, two degenerate triplet states, T9 and T10, with the character of π/d → σ* excitation, have extensive shallow minima. In the range of distances from 2.35 Å to 2.60 Å, these are the lowest triplet states. Near the distance of 2.60 Å, the PECs of these states intersect with the curve of the T34 state, which is a non-bonding σ → σ* state. This curve satisfactorily converges to the calculated Rh-CMe bond dissociation limit of 2.37 eV (54.7 kcal/mol). In Figure 5, a PEC fragment of the lowest triplet state, obtained at the UKS level of theory, is visible, slightly below the value of the dissociation limit. All the above-mentioned excited electronic states with a minimum on the PEC above 2.20 Å are states with electron donation to the σ* orbital of the Rh-CMe bond. Thus, they are potentially dissociative states for this bond and can lead to its photolytic cleavage in the MeRhPor complex. However, the occurrence of minima on the considered PECs, especially in the case of singlet states, can be a problematic issue. First of all, it should be noted that the reference function in TDDFT calculations is a closed-shell, single-determinant wave function. Such a wave function does not correctly describe the process of homolytic dissociation of the Rh-CMe bond, and in the limit of dissociation, it describes the ionic states of the products. As a result, the PESs of the excited states determined by TDDFT will not correctly describe the homolytic dissociation of the Rh-CMe bond at larger rhodium–carbon distances. Calculations using a broken-symmetry wave function (BS WF) show that the PEC of the ground state S0 separates at a distance of 2.6 Å, and at a distance of 3.5 Å, the energy of the S0,BS state is only about 0.25 eV (5.8 kcal/mol) lower than the calculated dissociation energy. For distances smaller than 2.55 Å, the BS wave function adopts the form of a closed-shell function and is equivalent to the solution of the Kohn–Sham equations in the restricted variant, giving the same total energy values as in the RKS calculations. Based on the BS function, for a distance of 3.5 Å using the TDDFT method, the excitation energy was calculated, and the character of the five lowest excited states S1–S5 was determined. The energies of the relevant states are presented in Figure 6 (R(Rh-CMe) = 3.50 Å, BS WF). The energy distribution of these states correlates well with the energies of the corresponding vertical states in the dissociation limit, R(Rh-CMe) = . That is, the energy order of the states considered here coincides with the energy order of the five lowest excited states for the isolated RhPor homolysis product. On the PECs, at a rhodium–carbon distance of 3.50 Å, for five further singlet states S1–S5, the energy order of the states is different than that resulting from calculations using the BS wave function; however, above a distance of 2.20 Å, the two states of π/d → σ* character remain consistently the lowest-energy singlet states. Thus, it can be concluded that the PECs for these states as a function of rhodium–carbon bond length may not have a minimum, or the minimum is very shallow, and singlet S1 and S2 π/d → σ* states (S7 and S8 at equilibrium geometry Req(Rh-CMe) = 2.00 Å) will be responsible for homolytic photodissociation. Preliminary calculations at the CASSCF/NEVPT2 level of theory [47,48,49,50], the results of which are presented in Figure S7 in the Supplementary Materials, reveal that the lowest singlet excited state in the Rh-CMe distance range from approximately 2.40 Å to the end of the considered distance range at 4.00 Å is a state of d → σ* character. The PEC of the excited singlet state in the indicated rhodium–carbon distance range is characterised by a broad and very shallow minimum with a depth of ~0.5 kcal/mol. This confirms the above-formulated predictions that the PECs of states S7 and S8 π/d → σ* should have an almost completely dissociative character. Given the character of the next excited singlet state, S3 d → σ*, it can be assumed that this electronic state could also participate in photohomolysis, and in the dissociation limit, it would lead to the population of the D3 d → d state on the RhPor radical. Since the energy of the D3 state of 1.17 eV (27 kcal/mol) is relatively high, the homolysis of the Rh-CMe bond with the occupation of this excited state of the RhPor radical seems to be less likely than the direct occupation of the D1 and D2 states, but ultimately, this possibility cannot be ruled out.
The potential energy curve for the optimised geometry of the first excited singlet state, S1,opt in Figure 5, essentially coincides with the curve obtained for the vertical state S1. Understandably, the energy of this state for the optimised geometry is consistently slightly lower relative to that of the vertical state. The shape of this PEC is the result of the intersection of two singlet state curves, i.e., the lowest singlet state of π → π* character with a minimum at 2.00 Å and the higher, excited singlet state of π/d → σ* character. Although, at the TDDFT level of theory, in this case, only the one, lowest singlet state can be optimised, due to degeneracy, this curve describes the PECs of four states, S1 and S2 in the range of 1.80 Å to 2.25 Å and S7 and S8 above 2.25 Å.

2.5. Singlet–Triplet Interaction

For the MeRhPor complex, the occurrence of phosphorescence at wavelengths of 650 nm unambiguously indicates the presence of singlet–triplet interactions in the intersystem crossing (ISC) process. According to the Landau–Zener theory [51,52], the transition probability between the singlet and triplet states is proportional to the square of the modulus of the spin–orbital coupling integral and inversely proportional to the difference in the gradients of the intersecting PESs of the electronic states and the gradient of the change in the active coordinate over time (Appendix A). Thus, it can be stated that a large value of spin–orbital coupling, expressed by a significant value of the H S T S O integral, transfers into an increase in the probability of the ISC S/T transition, and the probability of a singlet-to-triplet population can be estimated via the evaluation of the SOC constant (SOCC).
Radiation absorption corresponding to the Q band of the MeRhPor spectrum is associated with the occupancy of the lowest two singlet excited states S1 and S2. In the vicinity of the minimum energy of these states, four closely located T5–T8 triplet states with a similar PEC shape can be identified, as seen in Figure 7.
Calculated values of spin–orbit coupling constants (SOCCs) for selected states of the MeRhPor complex at certain Rh-CMe distances are shown in Table S5 in the Supplementary Materials. The calculated SOCC value for the coupling of S1 and S2 and T5–T8 states for a distance of R(Rh-CMe) = 2.00 Å is in the range of 31 cm−1 to 35 cm−1. These values are not significantly large, but they indicate the possibility of weak interactions of singlet states excited in the Q band with closely located triplet states. For the two higher-energy triplet states T9 and T10, the calculated SOCC values are very small and amount to 2 cm−1–4 cm−1. In the literature [39], the possibility of a direct transition between the S1 and S2 states and lower T1–T4 states has been postulated; however, given the significant energy difference between the above-mentioned singlet and triplet states, 0.40–0.54 eV (9–12 kcal/mol), the direct interaction of these states seems unlikely. Elongation of the Rh-CMe bond to 2.10 Å has practically no effect on the calculated SOCC values compared to those obtained for the equilibrium distance of 2.00 Å. Further elongation of the rhodium–carbon axial bond slightly increases the SOCC values for some triplet states of π/d → π* excitation character. However, in the case of π/d → σ* states, the SOC coefficient remains very small, ranging from 7 cm−1 to 5 cm−1 (SOCC values for S1/T6,T7 and S2/T6,T7 states at R(Rh-CMe) = 2.15 Å, Table S5 in the Supplementary Materials). There are two states, T9 and T10, at an equilibrium distance of 2.00 Å, whose energy rapidly decreases with an increasing rhodium–carbon bond length. Finally, at a distance of approximately 2.15 Å, the PEC of states S1 and S2 and the PEC of T9 and T10 intersect. The estimated energy barrier to the intersection point is only about 0.13 eV (~3 kcal/mol), but the direct interaction of states S1,S2  T9,T10 via ISC is rather negligible. Firstly, this is indicated by the low SOCC value, and secondly, it must be taken into account that the PECs intersect at a significant angle. The difference in PEC gradients at the intersection point is therefore large, which, according to the Landau–Zener theory, reduces the probability of triplet states being occupied on this path. For a distance of 2.30 Å, the π → π* states S1 and S2 are close to the triplet states T6–T8 of the π/d → π* type and T34 with the σ → σ* character. At this distance, the calculated SOCC for the coupling between states S1 and S2 and T6–T7 is slightly higher than for the equilibrium geometry at R(Rh-CMe) = 2.00 Å and ranges from 45 cm−1 to 51 cm−1. In the case of the T34 state, the SOC coefficient of the S1/T34 and S2/T34 coupling at this Rh-CMe bond length reaches a value of 69 cm−1. Whereas from the point of view of the estimated S/T coupling values, the probability of interaction between states S1 and S2 and triplet states may be relatively high for such an elongated axial bond, from the perspective of the photophysical mechanism, it is of rather minor importance. Elongation of the axial bond to a distance of 2.30 Å is associated with an increase in the energy of states S1 and S2 by approximately 0.46 eV (11 kcal/mol). Furthermore, at a distance of 2.25 Å, the curves of these states intersect with states S7 and S8. Since the barrier to the intersection point S1,S2/S7,S8 is ~0.32 eV (7.4 kcal/mol), it is rather unlikely that the axial bond would stretch beyond 2.25 Å along the potential energy surface of the π → π* states S1 and S2. It is more likely that IC between states S1 and S2 and S7 and S8 is the main photophysical channel for the photolysis of the rhodium–carbon bond. For the Rh-CMe bond elongated to a length of 2.30 Å, the π/d → π* states S7 and S8 become the lowest excited singlet states. The interaction of these states with the π/d → π* triplet states T5–T8 can be relatively effective, considering that the calculated SOCC values range from 84 to 118 cm−1 depending on the specific triplet state. However, it should be considered that very flat singlet state curves intersect with triplet state curves at a relatively large angle; hence, the gradient factor at the intersection points of the curves can significantly reduce the efficiency of the ISC process for the non-radiative transition S7 and S8  T5–T8. For a distance of 2.40 Å, the PEC of the lowest-energy singlet states, S7 and S8, intersects the curves of the π → π* states of T3 and T4, as well as the T34 non-bonding state. For the T3 and T4 states, the calculated SOCC value for the interaction with the S7 and S8 states is not large, ranging from 21 cm−1 to 34 cm−1, while for the T34 state, it is very large, equal to 533 cm−1. This result may suggest that the probability of occupying the triplet state is high and that a non-radiative transition to the triplet homolysis pathway would be possible during singlet photodissociation of the methyl group. However, as in the case of the other triplet states whose PECs intersect the curves of the S7 and S8 states, the S7,S8/T34 intersection also occurs at a large angle, which certainly increases the gradient factor and thus reduces the probability of a direct transition of the singlet states to the triplet σ → σ* state. With further elongation of the axial bond to a distance of 2.45 Å, the intersection of the singlet state curves with the PEC of T1 and T2 triplets is apparent. The calculated value of the SOCC for the coupling of these states is very small and not greater than 14 cm−1. Thus, a direct interaction of the singlet states with the lowest triplet states is unlikely to be possible in the photophysical path corresponding to the elongation of the rhodium–carbon bond.
Excitation into the Soret band corresponds to the occupancy of mainly two singlet electronic states, S9 and S10 of the π → π* type. Near the energy minimum of these states, nine closely spaced T13-T21 triplet states appear, of which the PECs of the T13–T18 states have a curvature similar to the PECs of the singlet states, with a minimum for R(Rh-CMe) = ~2.00 Å, as depicted in Figure 8. In general, for states T13 to T21, the calculated SOC coefficient value varies greatly, from low to medium to relatively high. In addition, these values vary somewhat depending on the singlet state interacting with a given triplet. The calculated SOC coefficients of medium value mainly refer to the interaction with the triplet states T16, T18, and T19 with π → σ*, π/d → d, and d → σ* characters, respectively, and their range is between 21 and 52 cm−1. The higher value is reached by the SOC coefficient for the coupling of the S9 and S10 states with the T16 and T21 states with the π → σ* character, and it is in the range of 21 cm−1 to 75 cm−1, while the highest value of the coefficient corresponds to the coupling of the S9 and S10 singlet states with the T17 π/d → d state, where the SOCC reaches about 200 cm−1. Thus, analysing the singlet–triplet interactions within the equilibrium geometry of the S9 and S10 states, it can be concluded that the probability of occupying the triplet states after excitation in the Soret band is relatively high.
Elongation of the axial bond slightly increases the SOCC for the interaction of singlet π → π* states with triplet states of type π → σ*, and at a distance of 2.15 Å, the value of the coupling coefficient is 84 cm−1–92 cm−1. Also, at the same distance, the SOCC for the interaction with triplet T19 of type d → σ* increases and is 77 cm−1 and 107 cm−1 for the coupling of states S9/T19 and S10/T19, respectively. As the bond length of Rh-CMe increases, the energy difference between the singlet and triplet states considered above grows, and the PECs of these states move away from each other, so a virtually barrier-free, direct crossing of the potential surface of the triplet states T16, T19, and T21 with the surface of the states S9 and S10 occurs in the vicinity of the equilibrium geometry of the singlets. Due to the strong compaction of triplet states in the vicinity of the minimum and the occurrence of the PEC intersections in close proximity to it, qualitative estimation of the influence of the gradient factor on the probability of S T transitions is basically impossible. At a distance of 2.15 Å, a pronounced coupling of singlet states with the non-bonding triplet state σ → σ* also appears, and the SOCC for this coupling is about 108 cm−1. The estimated energy barrier to the PEC intersection of S9 and S10 and T34 in this case is ~0.1 eV (~2.3 kcal/mol). Although the energy barrier is not significant, the PEC curves intersect at a rather large angle, which increases the gradient factor and reduces the probability of ISC between the S9, S10, and T34 states.

3. Discussion

Based on the obtained results, a mechanism can be proposed for singlet photophysical pathways leading to internal conversion (IC) and photodissociation of the rhodium–methyl bond, which is schematically shown in Figure 9.
Absorption in the Q band (exp. 543 nm, calc. 463 nm) excites the MeRhPor molecule to two degenerate π → π* states, S1 and S2 (generally SQ). After excitation, the molecular geometry practically does not relax, as the equilibrium geometries of the ground state S0 and the excited states S1 and S2 are not significantly different from each other. In these two lowest electronically excited singlet states, the molecule is stabilised in a relatively deep minimum on the potential energy surface. The S1 and S2 states could possibly be quenched by the S1 and S2 S0 fluorescence process, but experimental studies do not indicate the occurrence of this process. Alternatively, deactivation may occur by elongating the axial bond with the methyl ligand. Elongation of the Rh-CMe bond leads to internal conversion between the S1 and S2 and π/d → π* S7 and S8 states above 2.2 Å. The computationally estimated energy barrier of this IC process is 7.4 kcal/mol and correlates well with the experimental value of 7.4 ± 1.2 kcal/mol for the MeRhOEP complex photodecomposition barrier [39]. The degenerate S7 and S8 states are non-bonding or weakly bonded states, and their population will ultimately lead to radical homolysis of the Rh-CMe bond. The photophysical pathway between the relaxed states S1 and S2 and the products of axial bond rupture is marked with small green arrows in Figure 9 and described as path A. As a result of photohomolysis, the final product of the decomposition is a methyl radical and a RhPor complex in the D1 and D2 π/d → dz2 excited doublet states, denoted as 2[RhPor]LMCT/LF in Figure 9. The quenching of the D1 and D2 states in the RhPor complex is a thermal process occurring as vibrational energy dissipation and leads to the formation of a radical complex with a singly occupied dz2 rhodium orbital. Further processes, such as the formation of a two-centred [RhPor]2 dimer complex, radical recombination, and interaction with the external environment within the solvent cage are the result of the solvent character and the dynamics of the processes depending on temperature, medium viscosity, MeRhPor concentration, etc. Experimental findings suggest the existence of an additional state mediating the homolysis of the axial bond, S1 → X → RhPor + Me, in which the excited complex undergoes direct decomposition to radical products and in the presence of oxygen undergoes a direct radiationless deactivation to the ground-state MeRhPor [39]. In light of the computational results discussed above, the intermediate state X may be the degenerate states S7 and S8 available through IC during Rh-CMe bond elongation. It is probable that the delay in the bond-breaking process in these states is not directly related to their electronic structure but to the above-mentioned dynamics of the processes accompanying photohomolysis, considering the fact that the kinetic factor (the velocity of the decomposition products) is insufficient to quickly overcome the barrier associated with the relaxation of the solvent cavity during the rupture of the bond. Therefore, the structure/electronic state X formation seen in the experiment may be associated with an increased bond-breaking time in the intermediate states S7 and S8. In the case of ultrafast reactions like photochemical bond cleavage, where strong changes in the molecular structure occur, the solvent molecules directly interact with the molecular motion of the solute. The solute’s fragments, moving apart, push into the solvent cage, and their initial motion can be significantly hindered [53,54].
Absorption in the Soret band (exp. 395 nm, calc. 330 nm) involves excitation to higher singlet excited states of π → π*, S9 and S10 (generally SSoret). By analogy with the S1 and S2 states, it should be thought that also in this case, after vertical excitation, the geometry of the complex in the excited state does not undergo much change in order to reach the minimum on the potential energy surface. In contrast to the situation in the S1 and S2 states, where the energy minima occur at a greater distance from the intersections with the potential energy surfaces of other states, in the case of excited states in the Soret band, the minima are closer to such intersections. With a slight elongation of the Rh-CMe bond in the range of ~0.5 Å to ~1.0 Å, first of all, the PECs of the S9 and S10 states intersect with the curves of the S13 and S14 states with a π/d → σ* character. The possible intersection points are close enough to the minima so that the processes of internal conversion of S9,S10  S13,S14 are practically barrierless. This IC occurrence starts a cascade of successive geometry relaxations and a series of subsequent internal conversions, involving a sequence of increasingly lower-energy excited states: S9,S10  S13,S14  S6  S7,S8, as is shown by the small blue arrows in Figure 9. As a result, the S7 and S8 states are occupied, which leads directly to photolysis of the axial bond. The photophysical pathway discussed above, labelled path B in Figure 9, is virtually devoid of energy barriers. From the perspective of experimental results, it is said that the MeRhPor complex undergoes direct photochemical decomposition after excitation with a wave λ < 400 nm, which also includes the Soret band. Of course, in the context of the presented computational results, one can speak of rapid but rather not direct photodissociation. The experimental results also do not indicate the presence of intermediate stages in the case of photodissociation from SSoret states. Although the final stage of photolysis of the Rh-CMe bond, involving the S7 and S8 states, is common to both paths A and B, it can be assumed that the kinetic factor has a significant impact on the dynamics of the final stage of photodissociation. Due to the high energy of the SSoret, bond dissociation will occur with a much higher kinetic energy of decomposition fragments compared to the bond rupture from the SQ states, and this applies in particular to methyl radical formation. In this case, the cleavage rate of the Rh-CMe bond in states S7 and S8 may be fast enough that homolysis occurs rapidly before the geometry of the complex has time to relax to states S1 and S2 as a result of IC, and simultaneously, it could be said that there is no transition from the SSoret to the SQ states.
However, there are some indications that direct photolysis of the Rh-CMe bond may occur with involvement of the states S13, S14, and S17, as is shown in Figure 10. The direct path involving S9 and S10 ⇝ S17 internal conversion is rather unlikely because such IC occurs with a much more elongated axial bond than is the case for the S13 and S14 states and is associated with a higher energy barrier of about 5 kcal/mol. Concerning path B, the internal conversion S9,S10 ⇝ S13,S14 is more likely in the first stage of deactivation of SSoret states. Ultimately, the photohomolysis of the axial bond only through states S13 and S14 would have to lead to the formation of the RhPor product in a relatively highly excited state (states D4 and D5 in Figure 6), which is very unlikely. Regarding the TDDFT method in a closed-shell version, the potential energy curves of the excited states are heading toward incorrect dissociation limits. The analysis of the characters and energetics of singlet states based on the BS wave function at large Rh-CMe distances and at the dissociation limit indicates that the potential energy surface of the high-energy d → σ* state (S17) should intersect the surfaces of the π → σ* states (S13, S14) as the axial bond lengthens. Such intersections open an alternative path towards homolysis and lead to the formation of the RhPor final product in the excited state of D3 d → dz2. Thus, a photolysis path can be proposed in the form of the following sequence: S9,S10 ⇝ S13,S14 ⇝ S172[RhPor]LF + Me (Figure 10).
Although the energy of state D3 is relatively high, considering its character, it can be assumed that, like states D1 and D2, it is thermally quenched. In such a mechanism, partial quenching of the energy of the SSoret states would take place on the photolysis product. Unfortunately, based on the results discussed here, it is not possible to demonstrate a clear preference for the occurrence of a B or C path in the photodissociation process of the axial bond. This requires further, more detailed research, both theoretically and experimentally. Although the mechanism based on path C practically excludes the possibility of the creation of an X state, which is in good agreement with the interpretation of the experimental results, the creation of a photolysis product in a higher, excited state may be debatable. On path B, thermal quenching of the excitation energy takes place mainly as a relaxation of the states with an activated Rh-CMe bond, while on path C, ~30% quenching of energy should take place with the participation of the RhPor photolysis product.
According to the experimental findings, in the excitation wavelength range 410 < λ < 550 nm, the MeRhOEP molecule, after excitation, is directed to the triplet state or forms the photodecomposition product from excited singlet states. Therefore, from the perspective of the experimental findings, MeRhPor photophysics must also involve intersystem crossing (ISC) processes for the forbidden transitions between excited singlet and triplet states. The population of triplet states through the ISC channel results in the occurrence of phosphorescence due to the deactivation of the lowest π → π* triplet states. Experimentally, the phosphorescence band is observed at a wavelength of λ = 650 nm, which correlates very well with the calculated value of 653 nm determined on the basis of vertical deexcitation energy for the optimised T1 state geometry using the UKS formalism. Since the main singlet photophysical channel leads to the rupture of the Rh-CMe bond, primarily the possibilities of S/T crossing by changing the R(Rh-CMe) distance should be considered. By analysing the PECs as a function of the R(Rh-CMe) coordinate, for the SQ singlet states, two intersection points of the energy curves of electronic states S1 and S2 and T9, T10, and T34 can be indicated, as shown in Figure 11. To reach the first point, where the PECs of the SQ π → π* states intersect with the T9 and T10 π/d → σ* states, the system has to overcome a small energy barrier of about 3 kcal/mol. However, due to the low value of the calculated SOC coefficient, 4 cm−1, and the presence of a clearly unfavourable gradient factor, such an ISC channel of the triplet state population is rather unlikely. In the case of the second intersection point, where the PECs of the SQ states intersect the triplet energy curve of the T34 state of σ → σ* character, the calculated SOCC value is 69 cm−1; nevertheless, the importance of this ISC channel for occupying triplet states is rather minor. As emphasised in Section 2, this point occurs at a higher energy and a larger rhodium–carbon distance than the intersection of the singlet curves, where the main photophysical channel leading to the rupture of the Rh-CMe bond is opened. Even ignoring the importance of the gradient factor in the probability of this S/T transition, it can be assumed that this ISC pathway is most likely inactive. For Rh-CMe distances greater than 2.30 Å, the lowest-energy singlet states, S7 and S8, intersect the curves of the states of T1–T4, as well as the T34 non-bonding state. In general, the calculated value of the SOC coefficient for the coupling of the S7 and S8 states with the T1-T4 states is not high and is in the range of 14 cm−1–34 cm−1. Moreover, the very flat curves of the singlet states are intersected by the curves of the triplet state at a relatively large angle; hence, the gradient coefficient at the curves’ intersection points will significantly reduce the probability of the S T transition, and thus, the efficiency of the ISC process may be very limited. For the T34 state, the calculated SOCC is very large, equal to 533 cm−1. Such a high SOC coefficient value may indicate that the probability of occupying the triplet state is high and that a non-radiative transition to the triplet homolysis pathway would be possible during photodissociation along the singlet path. However, in the case of the S7 and S8 and T34 states, PEC intersections also occur at a large angle, which increases the gradient factor and thus limits the transition to the triplet state.
The interaction of the SSoret singlet states (S9, S10) with the triplet states is very similar to the SQ interactions with the T9, T10, and T34 states described above. The direct PEC intersection of the π → π* singlet states with the T34 σ → σ* state is particularly interesting. This intersection occurs at a slight elongation of the Rh-CMe bond from the singlet states’ energy minimum and is associated with a very small energy barrier, approximately 2 kcal/mol. The SOCC calculated around the intersection point is ~108 cm−1. Due to the SOC coefficient, the probability of triplet state occupancy could be significant, but analogously to the SQ states, the crossing of the S9 and S10 states with T34 occurs at a greater Rh-CMe distance than the intersection of the PECs of the S9, S10/S13, and S14 singlet states. Therefore, the S9,S10/T34 ISC channel requires a larger stretching of the rhodium–carbon bond than the singlet IC channel. Although the bond length difference is not very large (0.5–1.0 Å), taking into account that IC processes are generally faster than ISC, it can be assumed that the occupation of the T34 σ → σ* state in such a photophysical pathway is rather inefficient. Likewise, the almost vertical slope of the T34 triplet curve relative to the S9 and S10 singlet curves decreases the probability of S/T transition between these states. The T16 and T21 triplet states are π → σ* states, while the T19 state is of d → σ* type. These excited states have a common feature, i.e., the donation of electronic density to the anti-bonding orbital of the Rh-CMe bond. The PECs of these states cross the singlet state curves near their minimum, which means that in this case, the ISC process practically does not require overcoming an energy barrier. The calculated SOCCs for the interaction of these triplet states with SSoret states for a distance R(Rh-CMe) = 2.00 Å range from 21 cm−1 for the S9/T19 coupling to 75 cm−1 for the S9/T21 and S10/T21 couplings. Although for some of the couplings the SOC coefficient is relatively large, the gradient factor may have, in this case, an unfavourable influence on the probability of S/T transitions due to the significant angle at which the curves of the triplet and singlet states intersect.
The computational results reveal that the direct ISC process between the SQ and SSoret states and the non-bonding triplet σ → σ* state is ineffective, which basically confirms the results of experimental studies indicating the lack of participation of triplet states in the photodissociation of the rhodium–carbon bond. However, it is difficult to rule out an indirect channel for the σ → σ* state population via other triplet states. Therefore, while such a possibility certainly exists, experimentally, geometry relaxation and internal conversion within the triplet states according to the Kasha rule lead exclusively to the occupation of the lowest T1 and T2 states. The T1 and T2 states are phosphorescent states and, simultaneously, they are not the initial states for the triplet homolysis of the rhodium–carbon bond. Meanwhile, from the point of view of the computational results, one could assume the existence of a dissociative triplet pathway: T1,T2  T9,T10  T34   3[RhPor Me] RhPor + Me. The scope of the presented calculation results does not allow for the full solution to this problem; however, several reasons can be mentioned that may cause photophysical inactivity on the dissociative triplet path described above. Thus, it should be noted that in the case of photodissociation involving singlet states, the energies of the starting states SQ and the energy of the products in the dissociation limit are comparable (Figure 6, S1 and S2 at R(Rh-CMe) = 2.00 Å vs. D1 and D2 levels at R(Rh-CMe) = ). The high energy of deexcitation of the S1 and S2 states to the S0 state, which according to calculations is 2.65 eV (61 kcal/mol), would correspond to a relatively short emission wave by fluorescence, 468 nm, but experimentally, fluorescence is not observed. It can therefore be assumed that despite the existence of an energy barrier in the photodissociation process, the rate constant for fluorescence is probably lower than for IC, which favours a dissociative path. In the case of triplet states, the opposite is true, i.e., the energy of the lowest triplet states T1 and T2 is lower than the energy of the products in the dissociation limit, and the total energy effect on the path T1 and T2  RhPor + Me is associated with an increase in the system energy by ~6 kcal/mol. The computationally predicted maximum energy barrier on this path is 12 kcal/mol. Although this is not a high barrier for achieving the dissociated state, in the case of energetically unstable excited states, it may be the reason why the dissociative path is uncompetitive with respect to phosphorescence. It can, of course, be assumed that triplet photodissociation, despite everything, occurs to some very small extent, but the efficiency of this process is so low that it is difficult to observe it in an experiment.
The above discussion on singlet–triplet interactions involving the R(Rh-CMe) coordinate leads to the conclusion that most of the photophysical pathways related to ISC, which may occur via relaxation of the system in singlet states, are inactive or weakly active and do not contribute significantly to the triplet state population. It is worth noting, however, that the S/T interaction is possible for many closely lying triplet states around the photophysically active singlet states SQ and SSoret. This can be evidenced by the calculated SOCC values for R(Rh-CMe) = 2.00 Å, which are varied but sometimes relatively large, even above 100 cm−1. For the vast majority of states, both singlet and triplet, the R(Rh-CMe) = 2.00 Å distance is, if not exactly, then essentially close to, the equilibrium geometry. In this context, for example, the SOCC for the interaction of the SSoret states with the T17 triplet state of π/d → d character should be noted. The SOCC value for the coupling of the SSoret/T17 states is 200 cm−1 but quickly decreases to a level of several cm−1 with increasing rhodium–carbon distance. However, both in the case of the T17 state and many other triplet states located in the immediate vicinity of SQ and SSoret, an alternative active coordinate for the S/T interaction could be the deformation of the equatorial coordination sphere of the complex. Such deformation is associated with the vibrational motion of the central ion along the axial axis of the complex and causes changes in the electronic structure of excited states (usually, it involves mixing of π and d orbitals). Consequently, this deformation may cause an increase in the value of the SOC integral and increase the probability of S T conversion [55,56]. It should not be expected that even if such a mechanism of S/T interaction is taken into account, a dominant channel leading to triplet states will appear in the photophysics of MeRhPor, but probably for most pathways, ISC may be more effective in the case of coordination sphere deformation than in the case of rhodium–carbon bond elongation. Simultaneously, both the S T ISC processes and the internal conversion between triplet states would proceed primarily along active coordinates other than the Rh-CMe distance. The participation of different active coordinates in the ISC process may cause preference for specific photophysical paths, so in the case of triplet states, individual photophysical processes (ISC, IC, and geometric relaxation) that occur during the deformation of the coordination sphere may become more favourable for occupying the lowest triplet states and less advantageous for occupying potentially dissociative triplet states. The overall picture emerging from the computational results, regarding the contribution of triplet states to the photophysics of MeRhPor, seems to be that the population of the lowest phosphorescent triplet states is not the result of some dominant SQ  T1,T2 and SSoret  T1,T2 pathway but rather a composite of a larger number of individual and less effective S/T transitions. However, more elaborate consideration of this problem goes beyond the scope of this article and forms the basis for further theoretical studies and undoubtedly poses a challenge for experimental research.

4. Materials and Methods

The reported calculational results were obtained by applying the DFT [57,58,59] and TDDFT [60,61] level of theory with the use of the hybrid PBE0 functional [62,63], which mixes the Perdew–Burke–Ernzerhof (PBE) and HF exchange energy. The PBE0 hybrid functional is known as one of the effective functionals for describing the electronic structure of excited states of transition metal-containing systems and has been used for rhodoporphyrin complexes [33,64,65,66]. The def2-TZVP basis function [67] was applied for all atoms in the complex, albeit for rhodium, this basis utilises the effective core potential (ECP), replacing 28 core electrons [68]. In calculations, the RIJCOSX approximation [69,70] for the Coulomb and exchange parts of the Fock matrix was used, and the corresponding auxiliary basis sets were applied [71]. D3BJ dispersion corrections [72,73] were also employed in the calculations. To account for the interaction of the complex with the solvent environment, the continuous solvent model CPCM [74,75,76] with benzene as a solvent (ε = 2.28) was used in the calculations. The choice of solvent was due to the fact that the basic experimental data cited in the article [39] are from experiments for benzene solutions of methylrhodium(III)–octaethylporphyrin, MeRhOEP. For the calculation of the Rh-CMe bond energy, the B3LYP [77,78], PBE [63], and BP86 [77,79] functionals were additionally used.
In the calculations, the model structure of methylrhodium(III)–porphyrin (MeRhPor) was used, in which the valence of the carbon atoms on the outer side of the macrocyclic ring is saturated with hydrogen atoms. The molecular structure of MeRhPor is shown in Figure 12. The Rh+3 ion is coordinated by a porphyrin ligand with a total charge of −2 through four nitrogen atoms present in four pyrrole subunits. The methyl group, formally with a charge of −1, coordinates the central ion through a carbon atom in the axial position. Thus, the MeRhPor structural model under consideration is a five-coordinate complex with a total charge of zero. The electronic ground state of MeRhPor is a low-spin singlet state, denoted as S0. Simultaneously, rhodium(II)–porphyrin (RhPor), the product of homolytic photolysis of the Rh-CMe bond, is a four-coordinate complex with a total charge of 0 and a D0 doublet ground state. For all molecular structures considered, it was formally assumed that these systems had no symmetry, and no symmetry options were used in the calculations.
By applying the restricted Kohn–Sham (RKS) and unrestricted Kohn–Sham (UKS) formalisms, respectively, the geometry of the MeRhPor complex was fully optimised for the two lowest electronic states, S0 and T1. Full geometry optimisation of the lowest excited singlet state S1 was performed using the TDDFT method. For the RhPor photolysis product, the geometry optimisation of the doublet ground state D0 was performed within the framework of the UKS formalism, while the geometry optimisation of the five lowest excited doublet states, D1–D5, was performed with a time-dependent variant of DFT.
For MeRhPor, to determine a PEC as a function of the Rh-CMe distance, a scan of the S0 ground-state potential energy surface was performed in the range of 1.8 Å to 3.5 Å with a step of 0.05 Å. For individual PEC points, the Rh-CMe distance was frozen, and the remaining parameters were fully optimised. Based on the obtained relaxed S0 state geometries, TDDFT calculations were performed for 50 singlet and triplet excited states. For the lowest triplet state T1, the potential energy curve was determined using the unrestricted KS method, optimising the geometry of the complex at particular frozen Rh-CMe distances. In the calculations, both singlet and triplet pathways were examined to verify the experimental thesis. It is known that while the TDDFT method accurately captures the energies of singlet states, the triplet energies are somewhat less well reproduced [80]. A good description of triplet states is necessary in the calculations of many physical processes, including the TADF phenomenon [81,82,83]. Numerous benchmark calculations have been performed to verify this, showing that hybrid functionals perform better than non-hybrid functionals in this respect [81,82,83,84,85,86]. Using the Tamm–Dancoff approximation improves the quality of triplets [87]. However, the situation also strongly depends on the system being studied. Usually, the lowest triplet optimised by the UKS method provides the most reliable value for the energy of this state. Additionally, using the UKS formalism with a broken-symmetry wave function (BS WF), the PEC for the ground electronic state was determined in the full Rh-CMe distance range from 1.80 Å to 3.50 Å.
To estimate the interactions between singlet and triplet electronic states obtained from TD-DFT and determine the values of SOC integrals, the formalism of the Quasi-Degenerate Perturbation Theory (QDPT) [88,89] was used. Since, essentially, the reliable estimation of the degree of spin–orbit coupling based on scalar relativistic methods requires the use of a full basis function, in this part of the calculations, the ZORA-def2-TZVP [67,90] and SARC-ZORA-TZVP [91] bases were applied for N, C, H, and Rh atoms. The ZORA-def2-TZVP basis is a relativistically recontracted version of the all-electron def2-TZVP Ahlrichs basis set, whereas the SARC-ZORA-TZVP basis set used for palladium is the segmented all-electron relativistically contracted (SARC) basis set. Calculations were performed using the geometry of the complex for the S0 electronic ground state, at PEC points corresponding to Rh-CMe distances of 2.00 Å, 2.10 Å, 2.15 Å, and 2.30 Å. The geometry at these points was not re-optimised using the aforementioned basis sets. These points were chosen as characteristic of the interaction of singlet and triplet states in the photochemical process of elongation of the axial bond of rhodium–carbon.
Because MeRhPor is a structural model of the methylrhodium(III)–octaethylporphyrin complex (MeRhOEP), additional calculations were performed to check the influence of ethyl substituents on the structure and energetics of excited states. The results of the calculation are presented in Table S6 and Figure S8 and discussed in Supplementary Discussion S1 in the Supplementary Materials. Based on these results, it can be concluded that, according to the literature data [34,35], alkyl substituents at the β-pyrrolic position of the porphyrin ring have no significant effect on the energetics and electronic structure of the excited states of the rhodoporphyrin system.
All calculations were performed using the ORCA v. 5 package [92,93,94], with the SHARK package implemented for integral generation [95].

5. Conclusions

Under the influence of light with wavelengths in the Q and Soret band range, the MeRhPor complex molecule undergoes electronic excitation to singlet states, primarily S1 and S2 and S9 and S10 for the two absorption bands, respectively. All of these states are π → π* states, i.e., the electronic excitation involves the π orbitals of the porphyrin ligand. The excited states are deactivated in both singlet and triplet photophysical pathways. Singlet deactivation channels lead to homolytic scission of the axial rhodium–carbon bond. The excited states in the Soret band are deactivated via geometry relaxation and intersystem crossing (IC), which leads to a rapid scission of the Rh-CMe bond. In this case, the states of π → σ*, π/d → π*, d → σ*, and π/d → σ* type participate indirectly in the photodecomposition of the MeRhPor complex. On the other hand, an energy barrier must be overcome from the excited states belonging to the Q band to reach the states leading directly to bond dissociation. Typically, these photodissociation states are characterised by the donation of electronic density to the σ* anti-bonding orbital of the bond that is being broken. As a result of photolytic decomposition, a methyl radical Me and a RhPor radical are formed. The methyl radical can react with molecules in the environment but could also, depending on the photoreaction conditions, partially recombine with the deactivated radical RhPor. The lowest states of RhPor are thermally quenched via IC to the D0 state, in which the dz2 orbital of the rhodium ion is singly occupied. The RhPor radical in the electronic ground state D0 can recombine with the methyl radical and can react with molecules in the environment (e.g., O2) or form a two-centre dimer (RhPor)2. Ultimately, the reactions of the RhPor radical will be determined by the “external” conditions in which the photohomolysis reaction of the Rh-CMe bond takes place. The population of the lowest triplet states, T1 and T2, is mediated by a series of higher triplets due to ISC processes with Q and Soret singlet states. In singlet–triplet interactions, there is probably no dominant ISC photophysical channel; instead, there are many more or less efficient intersystem crossing processes between the Q and Soret states and a series of closely lying triplet states. The lowest triplet states, T1 and T2, are quenched by phosphorescence. The inactivity of the triplet photodissociation channel is related to (a) the uncompetitiveness of the ISC leading to the direct occupancy of the dissociative triplet state relative to the photophysical channel IC of singlet states and (b) inefficient IC in the thermal transition T1 and T2   3[RhPor Me] in relation to the radiative deactivation of the π → π* states T1 and T2.
For the MeRhPor molecule, quantum chemical calculations based on the DFT/TDDFT level of theory show a complex arrangement of closely spaced singlet as well as triplet states as a function of the Rh-CMe distance, which indicates the possibility of different competing photophysical pathways. This property is not unique but is somehow accumulated in the methylrhodoporphyrin molecule. This leads to the belief that this system, as well as its analogues containing an axial rhodium–carbon bond, is an extremely “flexible” system from the perspective of photophysical and photochemical processes. Both the structure change within the axial bond and changes in the complex’s surroundings will affect the quantum yield of a specific process from the point of view of the deactivation of excited states.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30193855/s1: Supplementary Discussion: Supplementary Discussion 1. Simulated UV/VIS spectrum of MeRhOEP complex and character of excited singlet electronic states. Table S1: Selected, optimised geometrical parameters of the coordination sphere for methylrhodium(III)–porphyrin complex (MeRhPor) and rhodium(II)–porphyrin complex (RhPor•). Table S2: Natural transition orbitals (NTOs) for singlet electronic states of the MeRhPor complex. Table S3: The lowest vertical triplet electronic transitions for the MeRhPor complex based on the TD-DFT/PBE0/def2-TZVP calculations with D3BJ dispersion correction and the CPCM/benzene solvent model. Table S4: The lowest vertical electronic transitions for the RhPor complex based on the TD-DFT/ PBE0/def2-TZVP calculations with D3BJ dispersion correction and the CPCM/benzene solvent model. Table S5: Calculated values of spin–orbit coupling constant (SOCC) for selected states of the MeRhPor complex at certain Rh-CMe distances. Table S6: The lowest vertical singlet electronic transitions for MeRhEOP (methylrhodium(III)–octaethylporphyrin) complex based on the TD-DFT/PBE0/def2-TZVP calculations with D3BJ dispersion correction and the CPCM/benzene solvent model. Figure S1: (a) Milliken spin population and (b) spin density isosurface (0.0035 a.u) of RhPor complex. Figure S2: Optimised structure of dimer (RhPor)2 and calculated bond energy (ΔEDE), bond dissociation energy (ΔEBDE), and bond dissociation free energy (ΔGBDFE) for Rh-Rh bond. Figure S3: Alpha and beta Kohn–Sham orbitals involved in the electronic excitations of the RhPor complex. Figure S4: Energy diagram of frontier Kohn–Sham orbitals for the RhPor complex. Figure S5: Energy diagram of Kohn–Sham frontier orbitals for two different Rh-CMe bond lengths in the MeRhPor complex. Figure S6: Potential energy curves (PECs) as a function of the Rh-CMe distance for the ground state (S0) and vertically excited singlet and triplet states of the MeRhPor complex. The potential energy curves for 21 singlet and 37 triplet states have been plotted on the basis of raw computational data obtained at the DFT and TDDFT levels of theory. Figure S7: Comparison of the potential energy curves of the ground state S0 and the lowest excited singlet state, obtained on the basis of calculations at the DFT, TDDFT, and CASSCF/NEVPT2 level of theory. Figure S8: Simulated UV/VIS spectra for the MeRhOEP complex and MeRhPor complex obtained on the basis of the TDDFT calculations [41,47,48,49,50,96,97].

Author Contributions

Conceptualisation, P.L.; methodology, P.L. and M.J.; validation, P.L. and M.J.; investigation, P.L.; writing—original draft preparation, P.L. and M.J.; writing—review and editing, P.L. and M.J.; visualisation, M.J. and P.L.; supervision, P.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

Calculations were carried out at the Wroclaw Centre for Networking and Supercomputing, https://wcss.pl (accessed on 1 September 2025), under grant No. 18.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
DFTDensity Functional Theory
TDDFTTime-Dependent Density Functional Theory
MeRhPorMethylrhodium(III)–porphyrin complex
RhPorRhodium(II)–porphyrin radical
PECPotential energy curve
SOCSpin–orbit coupling
SOCCSpin–orbit coupling coefficient
ICInternal conversion
ISCIntersystem crossing
Sii-th singlet electronic state
Tii-th triplet electronic state
SQS1 and S2 electronic state
SSoretS9 and S10 electronic state
NTOsNatural transition orbitals
CASSCF/NEVPT2Complete active space self-consistent field method in conjunction with the second order of perturbation theory in the N-Electron Valence State Perturbation Theory version.

Appendix A

The transition probability between singlet and triplet electronic states can be treated as a Landau–Zener transition probability [51,52], defined by the formula
P i j = 2 π H S T S O 2 F S T d R d t
where
F S T = d E S d R d E T d R
H S T S O is the spin–orbit coupling integral, F S T is the gradient factor, and R is the system active coordinate in the transition process between electronic states. F S T is defined as the difference in energy gradients at the intersection of the singlet and triplet electronic state surfaces. The S/T transition probability between the singlet and triplet states is proportional to the square of the modulus of the spin–orbital coupling integral (SOC coefficient, SOCC) and inversely proportional to the gradient factor and the change in the active coordinate over time.

References

  1. Fleischer, E.B. The Structure of Porphyrins and Metalloporphyrins. Acc. Chem. Res. 1970, 3, 105–112. [Google Scholar] [CrossRef]
  2. Smith, K.M. Porphyrins and Metalloporphyrins: A New Edition Based on the Original Volume by J. E. Falk, 2nd ed.; Elsevier Scientific Publishing Company: Amsterdam, The Netherlands; Oxford, UK; New York, NY, USA, 1975; pp. 1–910. [Google Scholar]
  3. Kadish, K.M.; Smith, K.M.; Guilard, R. The Porphyrin Handbook: Synthesis and Organic Chemistry; Academic Press INC: San Diego, CA, USA, 1999; Volume 1, pp. 1–399. [Google Scholar]
  4. Momenteau, M.; Reed, C.A. Synthetic Heme Dioxygen Complexes. Chem. Rev. 1994, 94, 659–698. [Google Scholar] [CrossRef]
  5. Kim, H.J.; Khalimonchuk, O.; Smith, P.M.; Winge, D.R. Structure, function, and assembly of heme centers in mitochondrial respiratory complexes. Biochim. Biophys. Acta 2012, 1823, 1604–1616. [Google Scholar] [CrossRef]
  6. Mukherjee, M. Heme Enzymes: Nature’s Versatile Catalysts. Am. J. Biomed. Sci. Res. 2022, 16, 406–408. [Google Scholar] [CrossRef]
  7. Gao, F.; Jiaxuan Guo, J.; Yuanyue Shen, Y. Advances from chlorophyll biosynthesis to photosynthetic adaptation, evolution and signaling. Plant Stress 2024, 12, 100470. [Google Scholar] [CrossRef]
  8. Denisov, I.G.; Makris, T.M.; Sligar, S.G.; Schlichting, I. Structure and Chemistry of Cytochrome P450. Chem. Rev. 2005, 105, 2253–2277. [Google Scholar] [CrossRef]
  9. Wasielewski, M.R. Photoinduced Electron Transfer in Supramolecular Systems for Artificial Photosynthesis. Chem. Rev. 1992, 92, 435–461. [Google Scholar] [CrossRef]
  10. Lomova, T.; Tsaplev, Y.; Klyueva, M.; Ovchenkova, E. Recent advances in the practical use of the redox properties of manganese porphyrins. J. Organomet. Chem. 2021, 945, 121880. [Google Scholar] [CrossRef]
  11. Cojocariu, I.; Carlotto, S.; Zamborlini, G.; Jugovac, M.; Schio, L.; Floreano, L.; Casarin, M.; Feyer, V.; Schneider, C.M. Reversible redox reactions in metal-supported porphyrin: The role of spin and oxidation state. J. Mater. Chem. C 2021, 9, 12559–12565. [Google Scholar] [CrossRef]
  12. Sheldon, R.A. Metalloporphyrins in Catalytic Oxidations, 1st ed.; CRC Press: Boca Raton, FL, USA, 1994; pp. 1–381. [Google Scholar]
  13. Kadish, K.M.; Smith, K.M.; Guilard, R. The Porphyrin Handbook: Applications: Past, Present, and Future; Academic Press INC: San Diego, CA, USA, 1999; Volume 6, pp. 1–346. [Google Scholar]
  14. Bonnett, R. Photosensitizers of the porphyrin and phthalocyanine series for photodynamic therapy. Chem. Soc. Rev. 1995, 24, 19–33. [Google Scholar] [CrossRef]
  15. Ogoshi, H.; Mizutani, T. Multifunctional and Chiral Porphyrins: Model Receptors for Chiral Recognition. Acc. Chem. Res. 1998, 31, 81–89. [Google Scholar] [CrossRef]
  16. Chandra, R.; Tiwari, M.; Kaur, P.; Sharma, M.; Jain, R.; Dass, S. Metalloporphyrins-Application and Clinical Significance. Indian J. Clin. Biochem. 2000, 15, 183–199. [Google Scholar] [CrossRef]
  17. Suslick, K.S.; Rakow, N.A.; Kosal, M.E.; Chou, J.-H. The materials chemistry of porphyrins and metalloporphyrins. J. Porphyr. Phthalocyanines 2000, 4, 407–413. [Google Scholar] [CrossRef]
  18. Takagi, S.; Miharu Eguchi, M.; Donald, A.; Tryk, D.A.; Inoue, H. Porphyrin photochemistry in inorganic/organic hybrid materials: Clays, layered semiconductors, nanotubes, and mesoporous materials. J. Photochem. Photobiol. C Photochem. Rev. 2006, 7, 104–126. [Google Scholar] [CrossRef]
  19. Sekhar, A.R.; Chitose, Y.; Janoš, J.; Dangoor, S.I.; Ramundo, A.; Satchi-Fainaro, R.; Slavíček, P.; Klán, P.; Weinstain, R. Porphyrin as a versatile visible-light-activatable organic/metal hybrid photoremovable protecting group. Nat. Commun. 2022, 13, 3614. [Google Scholar] [CrossRef]
  20. Ouyang, J.; Li, D.; Zhu, L.; Cai, X.; Liu, L.; Pan, H.; Ma, A. Application and Challenge of Metalloporphyrin Sensitizers in Noninvasive Dynamic Tumor Therapy. Molecules 2024, 29, 4828. [Google Scholar] [CrossRef]
  21. Imran, M.; Ramzan, M.; Qureshi, A.K.; Khan, M.A.; Tariq, M. Emerging Applications of Porphyrins and Metalloporphyrins in Biomedicine and Diagnostic Magnetic Resonance Imaging. Biosensors 2018, 8, 95. [Google Scholar] [CrossRef]
  22. Boscencu, R.; Radulea, N.; Manda, G.; Machado, I.F.; Socoteanu, R.P.; Lupuliasa, D.; Burloiu, A.M.; Mihai, D.P.; Ferreira, L.F.V. Porphyrin Macrocycles: General Properties and Theranostic Potential. Molecules 2023, 28, 1149. [Google Scholar] [CrossRef]
  23. Brothers, P.J.; Collman, J.P. The Organometallic Chemistry of Transition-Metal Porphyrin Complexes. Acc. Chem. Res. 1986, 19, 209–215. [Google Scholar] [CrossRef]
  24. Wayland, B.B.; Sherry, A.E.; Coffin, V.L. Homogeneous Transition Metal Catalyzed Reactions; American Chemical Society: Washington, DC, USA, 2009; Volume 230, pp. 249–259. [Google Scholar]
  25. Cui, W.; Wayland, B.B. Hydrocarbon C-H bond activation by rhodium porphyrins. J. Porphyr. Phthalocyanines 2004, 08, 103–110. [Google Scholar] [CrossRef]
  26. de Bruin, B.; Hetterscheid, D.G.H. Paramagnetic (Alkene)Rh and (Alkene)Ir Complexes: Metal or Ligand Radicals? Eur. J. Inorg. Chem. 2007, 2007, 211–230. [Google Scholar] [CrossRef]
  27. Thompson, S.J.; Brennan, M.R.; Lee, S.Y.; Dong, G. Synthesis and applications of rhodium porphyrin complexes. Chem. Soc. Rev. 2018, 47, 929–981. [Google Scholar] [CrossRef]
  28. Campagna, S.; Puntoriero, F.; Nastasi, F.; Bergamini, G.; Balzani, V. Photochemistry and Photophysics of Coordination Compounds: Ruthenium. In Photochemistry and Photophysics of Coordination Compounds I. Topics in Current Chemistry; Balzani, V., Campagna, S., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; Volume 280, pp. 117–214. [Google Scholar]
  29. Bosch, H.W.; Wayland, B.B. The role of rhodium porphyrins in the photoassisted formation of formaldehyde and methanol from hydrogen and carbon monoxide. Chem. Soc. Chem. Commun. 1986, 12, 900–901. [Google Scholar] [CrossRef]
  30. Zhou, J.; Gai, L.; Mack, J.; Zhou, Z.; Qiu, H.; Chan, K.S.; Shen, Z. Synthesis and photophysical properties of orthogonal rhodium(III)–carbon bonded porphyrin–aza-BODIPY conjugates. Mater. Chem. C 2016, 4, 8422. [Google Scholar] [CrossRef]
  31. Yu, M.; Fu, X. Visible Light Promoted Hydroxylation of a Si-C(sp3) Bond Catalyzed by Rhodium Porphyrins in Water. J. Am. Chem. Soc. 2011, 133, 15926–15929. [Google Scholar] [CrossRef]
  32. Vasil’ev, V.V.; Borisov, S.M.; Golovina, I.V. Luminescence of Water-Soluble Rh(III) Porphyrins. Opt. Spectrosc. 2003, 95, 29–34. [Google Scholar] [CrossRef]
  33. Li, H.; Boao Han, B.; Wang, R.; Li, W.; Zhang, W.; Fu, X.; Fang, H.; Ma, F.; Wang, Z.; Zhang, J. Photochemical conversion of CO to C1 and C2 products mediated by porphyrin rhodium(II) metallo-radical complexes. Nat. Commun. 2024, 15, 7724. [Google Scholar] [CrossRef]
  34. Kalyanasundaram, K. Luminescence and triplet—Triplet absorption spectra of rhodium (III) porphyrins. Chem. Phys. Lett. 1984, 104, 357–362. [Google Scholar] [CrossRef]
  35. Ogoshi, H.; Omura, T.; Yoshida, Z. A New Rhodium(I)-Porphyrin Complex. II. Synthesis and Oxidative Alkylation. J. Am. Chem. Soc. 1973, 95, 1666–1668. [Google Scholar] [CrossRef]
  36. Hanson, L.K.; Gouterman, M.; Hanson, J.C. Porphyrins. XXIX. The Crystal and Molecular Structure and Luminescence of Bis(dimethylamine)etio(I)porphinatorhodium(III) Chloride Dihydrate. J. Am. Chem. Soc. 1973, 95, 4822–4829. [Google Scholar] [CrossRef]
  37. Lever, A.B.P.; Ramaswamy, B.S.; Licoccia, S. Sensitized photoreduction of methyl viologen by metalloporphyrins. J. Photochem. 1982, 19, 173–182. [Google Scholar] [CrossRef]
  38. Hoshino, M.; Nagamori, T.; Seki, H.; Tase, T.; Chihara, T.; Lillis, J.P.; Wakatsuki, Y. Laser Photolysis Studies on Photodissociation of Axial Ligands from Isocyanide Complexes of Cobalt(III) and Rhodium(III) Porphyrins in Toluene Solutions. A Comparison with the Photochemistry of Carbonylrhodium(III) Porphyrin. J. Phys. Chem. A 1999, 103, 3672–3677. [Google Scholar] [CrossRef]
  39. Hoshino, M.; Yasufuku, K.; Seki, H.; Yamazaki, H. Wavelength-Dependent Photochemlcal Reaction of Methylrhodlum(III) Octaethylporphyrin. Studies on CH3-Rh Bond Cleavage. J. Phys. Chem. 1985, 89, 3080–3085. [Google Scholar] [CrossRef]
  40. Whang, D.; Kim, K. Structure of a new form of octaethylporphyrinato(methyl)rhodium(III). Acta Crystallogr. Sect. C Struct. Chem. 1991, C47, 2547–2550. [Google Scholar] [CrossRef]
  41. Wayland, B.B. Rh-Rh, Rh-H, Rh-C and Rh-O bond energies in (OEP)Rh complexes: Thermodynamic criteria for addition of M-H and M-M bonds to C-O and C-C multiple bonds. Polyhedron 1998, 7, 1545–1555. [Google Scholar] [CrossRef]
  42. Li, G.; Zhang, F.F.; Pi, N.; Chen, H.L.; Zhang, S.Y.; Chan, K.S. Determination of Rh–C Bond Dissociation Energy in Methyl(porphyrinato)rhodium(III) Complexes: A New Application of Photoacoustic Calorimetry. Chem. Lett. 2001, 30, 284–285. [Google Scholar] [CrossRef]
  43. Fu, X.; Wayland, B.B. Thermodynamics of Rhodium Hydride Reactions with CO, Aldehydes, and Olefins in Water:  Organo-Rhodium Porphyrin Bond Dissociation Free Energies. J. Am. Chem. Soc. 2005, 127, 16460–16467. [Google Scholar] [CrossRef]
  44. Gouterman, M. Optical Spectra and Electronic Structure of Porphyrins and Related Rings. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New York, USA, 1978; Volume III, pp. 1–165. [Google Scholar]
  45. Antipas, A.; Gouterman, M. Porphyrins. 44. Electronic States of Co, Ni, Rh, and Pd Complexes. J. Am. Chem. Soc. 1983, 105, 4896–4901. [Google Scholar] [CrossRef]
  46. Kuznetsov, A.E. Stacks of Metalloporphyrins: Comparison of Experimental and Computational Results. J. Phys. Chem. B 2019, 123, 10044–10060. [Google Scholar] [CrossRef]
  47. Roos, B.O. The Complete Active Space Self-Consistent Field Method and its Applications in Electronic Structure Calculations. In Advances in Chemical Physics: Ab Initio Methods in Quantum Chemistry Part 2; Lawley, K.P., Ed.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 1987; Volume 69, pp. 399–445. [Google Scholar]
  48. Angeli, C.; Cimiraglia, R.; Evangelisti, S.; Leininger, T.; Malrieu, J.-P. Introduction of n-electron valence states for multireference perturbation theory. J. Chem. Phys. 2001, 114, 10252–10264. [Google Scholar] [CrossRef]
  49. Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. N-electron valence state perturbation theory: A fast implementation of the strongly contracted variant. Chem. Phys. Lett. 2001, 350, 297–305. [Google Scholar] [CrossRef]
  50. Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. n-electron valence state perturbation theory: A spinless formulation and an efficient implementation of the strongly contracted and of the partially contracted variants. J. Chem. Phys. 2002, 117, 9138–9153. [Google Scholar] [CrossRef]
  51. Landau, L.D. Zur Theorie der Energieübertragung. II. Phys. Sov. Union 1932, 2, 46–51. [Google Scholar]
  52. Zener, C. Non-Adiabatic Crossing of Energy Levels. Proc. R. Soc. Lond. A. 1932, 137, 696–702. [Google Scholar]
  53. Thallmair, S.; Kowalewski, M.; Zauleck, J.P.P.; Roos, M.K.; de Vivie-Riedle, R. Quantum Dynamics of a Photochemical Bond Cleavage Influenced by the Solvent Environment: A Dynamic Continuum Approach. J. Phys. Chem. Lett. 2014, 5, 3480–3485. [Google Scholar] [CrossRef]
  54. Thallmair, S.; Zauleck, J.P.P.; de Vivie-Riedle, R. Quantum Dynamics in an Explicit Solvent Environment: A Photochemical Bond Cleavage Treated with a Combined QD/MD Approach. J. Chem. Theory Comput. 2015, 11, 1987–1995. [Google Scholar] [CrossRef]
  55. Szczepańska, M.; Lodowski, P.; Jaworska, M. Electronic excited states and luminescence properties of palladium(II)corrin complex. J. Photochem. Photobiol. A Chem. 2020, 389, 112226. [Google Scholar] [CrossRef]
  56. Jaworska, M.; Lodowski, P. Interaction of palladium porphyrin with dioxygen molecule. The perspective from theoretical calculation. In Proceedings of the 6th EuChemS Inorganic Chemistry Conference, Vienna, Austria, 3–7 September 2023. [Google Scholar]
  57. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B8713. [Google Scholar] [CrossRef]
  58. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  59. Hohenberg, P.C.; Kohn, W.; Sham, L.J. The Beginnings and Some Thoughts on the Future. Adv. Quantum Chem. 1990, 21, 7–26. [Google Scholar]
  60. Runge, E.; Gross, E.K.U. Density-Functional Theory for Time-Dependent Systems. Phys. Rev. Lett. 1984, 52, 997–1000. [Google Scholar] [CrossRef]
  61. Casida, M.E. Time-Dependent Density Functional Response Theory for Molecules. In Recent Advances in Density-Functional Methods; Chong, D.P., Ed.; World Scientific: Singapore, 1995; Volume 1, pp. 155–192. [Google Scholar]
  62. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  63. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  64. Munro, O.Q.; Camp, G.L.; Carlton, L. Structural, 103Rh NMR and DFT Studies of a Bis(phosphane)RhIII–Porphyrin Derivative. Eur. J. Inorg. Chem. 2009, 2009, 2512–2523. [Google Scholar] [CrossRef]
  65. Steinmetz, M.; Grimme, S. Benchmark Study of the Performance of Density Functional Theory for Bond Activations with (Ni,Pd)-Based Transition-Metal Catalysts. ChemistryOpen 2013, 2, 115–124. [Google Scholar] [CrossRef]
  66. Maity, B.; Scott, T.R.; Stroscio, G.D.; Gagliardi, L.; Cavallo, L. The Role of Excited States of LNiII/III(Aryl)(Halide) Complexes in Ni–Halide Bond Homolysis in the Arylation of Csp3–H Bonds. ACS Catal. 2022, 12, 13215–13224. [Google Scholar] [CrossRef]
  67. 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]
  68. Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjustedab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  69. Neese, F. An improvement of the resolution of the identity approximation for the formation of the Coulomb matrix. J. Comp. Chem. 2003, 24, 1740–1747. [Google Scholar] [CrossRef]
  70. Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys. 2009, 356, 98–109. [Google Scholar]
  71. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  72. 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]
  73. 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. [Google Scholar] [CrossRef] [PubMed]
  74. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  75. Garcia-Rates, M.; Neese, F. Efficient implementation of the analytical second derivatives of hartree-fock and hybrid DFT energies within the framework of the conductor-like polarizable continuum model. J. Comput. Chem. 2019, 40, 1816–1828. [Google Scholar] [CrossRef]
  76. Garcia-Rates, M.; Neese, F. Effect of the Solute Cavity on the Solvation Energy and Its Derivatives Within the Framework of the Gaussian Charge Scheme. J. Comput. Chem. 2020, 41, 922–939. [Google Scholar] [CrossRef]
  77. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic-behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  78. 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. [Google Scholar] [CrossRef]
  79. Perdew, J.P. Density-functional approximation for the correlation-energy of the inhomogeneous electron-gas. Phys. Rev. B Condens. Matter Mater. Phys. 1986, 33, 8822–8824. [Google Scholar] [CrossRef]
  80. Tozer, D.J.; Handy, N.C. On the determination of excitation energies using density functional theory. Phys. Chem. Chem. Phys. 2000, 2, 2117–2121. [Google Scholar] [CrossRef]
  81. Huang, S.; Zhang, Q.; Shiota, Y.; Nakagawa, T.; Kuwabara, K.; Yoshizawa, K.; Adachi, C. Computational Prediction for Singlet- and Triplet-Transition Energies of Charge-Transfer Compounds. J. Chem. Theory Comput. 2013, 9, 3872–3877. [Google Scholar] [CrossRef]
  82. Jacquemin, D.; Perpète, E.A.; Ciofini, I.; Adamo, C. Assessment of Functionals for TD-DFT Calculations of Singlet−Triplet Transitions. J. Chem. Theory Comput. 2010, 6, 1532–1537. [Google Scholar] [CrossRef]
  83. Wang, J.; Bai, F.Q.; Xia, B.H.; Zhang, H.X.; Cui, T. Accurate simulation of geometry, singlet-singlet and triplet-singlet excitation of cyclometalated iridium(III) complex. J. Mol. Model. 2014, 20, 2108. [Google Scholar] [CrossRef]
  84. Bousquet, D.; Fukuda, R.; Jacquemin, D.; Ciofini, I.; Adamo, C.; Ehara, M. Benchmark Study on the Triplet Excited-State Geometries and Phosphorescence Energies of Heterocyclic Compounds: Comparison Between TD-PBE0 and SAC-CI. J. Chem. Theory Comput. 2014, 10, 3969–3979. [Google Scholar] [CrossRef]
  85. Atkins, A.J.; Talotta, F.; Freitag, L.; Boggio-Pasqua, M.; González, L. Assessing Excited State Energy Gaps with Time-Dependent Density Functional Theory on Ru(II) Complexes. J. Chem. Theory Comput. 2017, 13, 4123–4145. [Google Scholar]
  86. Grotjahn, R.; Kaupp, M. Validation of Local Hybrid Functionals for Excited States: Structures, Fluorescence, Phosphorescence, and Vibronic Spectra. J. Chem. Theory Comput. 2020, 16, 5821–5834. [Google Scholar] [CrossRef]
  87. Rangel, T.; Hamed, S.M.; Bruneval, F.; J. Neaton, J.B. An assessment of low-lying excitation energies and triplet instabilities of organic molecules with an ab initio Bethe-Salpeter equation approach and the Tamm-Dancoff approximation. J. Chem. Phys. 2017, 146, 194108. [Google Scholar] [CrossRef]
  88. Roemelt, M.; Maganas, D.; DeBeer, S.; Neese, F. A combined DFT and restricted open-shell configuration interaction method including spin-orbit coupling: Application to transition metal L-edge X-ray absorption spectroscopy. J. Chem. Phys. 2013, 138, 204101. [Google Scholar] [CrossRef]
  89. de Souza, B.; Farias, G.; Neese, F.; Izsak, R. Predicting Phosphorescence Rates of Light Organic Molecules Using Time-Dependent Density Functional Theory and the Path Integral Approach to Dynamics. J. Chem. Theory Comput. 2019, 15, 1896. [Google Scholar] [CrossRef]
  90. ORCA Manual, Version 6.0; Max-Planck-Institut für Kohlenforschung: Mülheim a. d. Ruhr, Germany. 2025; Available online: https://www.faccts.de/docs/orca/6.0/manual/ (accessed on 1 August 2025).
  91. Rolfes, J.D.; Neese, F.; Pantazis, D.A. All-electron scalar relativistic basis sets for the elements Rb–Xe. J. Comput. Chem. 2010, 41, 1842–1849. [Google Scholar] [CrossRef]
  92. Neese, F. The ORCA program system. WIRES Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  93. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA quantum chemistry program package. J. Chem. Phys. 2020, 152, L224108. [Google Scholar] [CrossRef] [PubMed]
  94. Neese, F. Software update: The ORCA program system, version 5.0. WIRES Comput. Mol. Sci. 2022, 12, 1606. [Google Scholar] [CrossRef]
  95. Neese, F. The SHARK Integral Generation and Digestion System. J. Comput. Chem. 2023, 44, 381–396. [Google Scholar] [CrossRef]
  96. Wayland, B.B.; Ba, S.; Sherry, A.E. Activation of Methane and Toluene by Rhodium(II) Porphyrin Complexes. J. Am. Chem. Soc. 1991, 113, 5305–5311. [Google Scholar] [CrossRef]
  97. Wayland, B.B.; Coffin, V.L.; Farnos, M.D. Estimation of the Rh-Rh bond dissociation energy in the (octaethylporphyrina-to)rhodium(II) dimer by proton NMR line broadening. Inorg. Chem. 1988, 27, 2745–2747. [Google Scholar] [CrossRef]
Figure 1. Main photochemical deactivation pathways of singlet excited states SQ and SSoret for the MeRhOEP complex from the perspective of experimental studies [39].
Figure 1. Main photochemical deactivation pathways of singlet excited states SQ and SSoret for the MeRhOEP complex from the perspective of experimental studies [39].
Molecules 30 03855 g001
Figure 2. Superimposition of optimised geometries in the S0 and S1 electronic states of the MeRhPor complex: black colour—S0 state geometry; red colour—S1 state geometry. The superimposition of the geometries was performed relative to the three atoms N24, Rh, and N22.
Figure 2. Superimposition of optimised geometries in the S0 and S1 electronic states of the MeRhPor complex: black colour—S0 state geometry; red colour—S1 state geometry. The superimposition of the geometries was performed relative to the three atoms N24, Rh, and N22.
Molecules 30 03855 g002
Figure 3. Simulated UV/VIS spectrum of the MeRhPor complex based on the TDDFT calculations. Black solid line—simulated spectral line obtained using Lorentzian broadening with a half-width of 15 nm. Vertical red lines—calculated wavelengths for vertical excitations to singlet states; the height of the line corresponds to the calculated value of the oscillator strength f.
Figure 3. Simulated UV/VIS spectrum of the MeRhPor complex based on the TDDFT calculations. Black solid line—simulated spectral line obtained using Lorentzian broadening with a half-width of 15 nm. Vertical red lines—calculated wavelengths for vertical excitations to singlet states; the height of the line corresponds to the calculated value of the oscillator strength f.
Molecules 30 03855 g003
Figure 4. Energy diagram of frontier Kohn–Sham orbitals involved in electronic excitations for the MeRhPor complex. The shape and energetic placement of the orbitals correspond to the equilibrium geometry of the S0 state.
Figure 4. Energy diagram of frontier Kohn–Sham orbitals involved in electronic excitations for the MeRhPor complex. The shape and energetic placement of the orbitals correspond to the equilibrium geometry of the S0 state.
Molecules 30 03855 g004
Figure 5. Potential energy curves (PECs) as a function of the Rh-CMe distance for the ground state (S0) and selected vertically excited singlet and triplet states of the MeRhPor complex. Black line—PEC of S0 state obtained from restricted Kohn–Sham method (RKS); red lines—PECs of singlet excited states from TDDFT level of theory; blue lines—PECs of triplet excited states from TDDFT level of theory; black line with empty circles—PEC of ground state, S0,BS, obtained from broken-symmetry (BS) wave function; blue line with empty circles—PEC of lowest triplet state T1,UKS, obtained from unrestricted Kohn–Sham method (UKS); red line with empty circles—PEC of first excited singlet state S1,opt obtained for optimised geometry at TDDFT level of theory.
Figure 5. Potential energy curves (PECs) as a function of the Rh-CMe distance for the ground state (S0) and selected vertically excited singlet and triplet states of the MeRhPor complex. Black line—PEC of S0 state obtained from restricted Kohn–Sham method (RKS); red lines—PECs of singlet excited states from TDDFT level of theory; blue lines—PECs of triplet excited states from TDDFT level of theory; black line with empty circles—PEC of ground state, S0,BS, obtained from broken-symmetry (BS) wave function; blue line with empty circles—PEC of lowest triplet state T1,UKS, obtained from unrestricted Kohn–Sham method (UKS); red line with empty circles—PEC of first excited singlet state S1,opt obtained for optimised geometry at TDDFT level of theory.
Molecules 30 03855 g005
Figure 6. Energy diagram of selected electronic vertical states (S7, S8, S13, S14, S17) for different rhodium–carbon distances, R(Rh-CMe) = 2.00, 2.30, 2.80, and 3.50 Å, and comparison with the energies of the corresponding excited states determined by calculations with broken-symmetry wave function (BSWF) and energies of the lowest excited states of the RhPor complex (R(Rh-CMe) =   ). The numbering of the electronic states for distances R(Rh-CMe) > 2.00 Å corresponds to their order in the given geometry.
Figure 6. Energy diagram of selected electronic vertical states (S7, S8, S13, S14, S17) for different rhodium–carbon distances, R(Rh-CMe) = 2.00, 2.30, 2.80, and 3.50 Å, and comparison with the energies of the corresponding excited states determined by calculations with broken-symmetry wave function (BSWF) and energies of the lowest excited states of the RhPor complex (R(Rh-CMe) =   ). The numbering of the electronic states for distances R(Rh-CMe) > 2.00 Å corresponds to their order in the given geometry.
Molecules 30 03855 g006
Figure 7. Potential energy curves (PECs) in the vicinity of the Q band as a function of Rh-CMe distance for the vertically excited singlet and triplet states of the MeRhPor complex. Red lines—PECs of singlet excited states; blue lines—PECs of triplet excited states.
Figure 7. Potential energy curves (PECs) in the vicinity of the Q band as a function of Rh-CMe distance for the vertically excited singlet and triplet states of the MeRhPor complex. Red lines—PECs of singlet excited states; blue lines—PECs of triplet excited states.
Molecules 30 03855 g007
Figure 8. Potential energy curves (PECs) in the vicinity of the Soret band as a function of Rh-CMe distance for vertically excited singlet and triplet states of the MeRhPor complex. Red lines—PECs of singlet excited states; blue lines—PECs of triplet excited states.
Figure 8. Potential energy curves (PECs) in the vicinity of the Soret band as a function of Rh-CMe distance for vertically excited singlet and triplet states of the MeRhPor complex. Red lines—PECs of singlet excited states; blue lines—PECs of triplet excited states.
Molecules 30 03855 g008
Figure 9. PECs of the most important excited singlet states involved in the photophysics of the MeRhPor complex and the predicted pathways of IC and photohomolysis of the Rh-CMe axial bond.
Figure 9. PECs of the most important excited singlet states involved in the photophysics of the MeRhPor complex and the predicted pathways of IC and photohomolysis of the Rh-CMe axial bond.
Molecules 30 03855 g009
Figure 10. Energy diagram of singlet excited states of the MeRhPor complex involved in the photolysis of the Rh-CMe bond after excitation to SSoret states (S9 and S10).
Figure 10. Energy diagram of singlet excited states of the MeRhPor complex involved in the photolysis of the Rh-CMe bond after excitation to SSoret states (S9 and S10).
Molecules 30 03855 g010
Figure 11. PECs of the most important excited singlet and triplet states involved in the photophysics of the ISC process for the MeRhPor complex.
Figure 11. PECs of the most important excited singlet and triplet states involved in the photophysics of the ISC process for the MeRhPor complex.
Molecules 30 03855 g011
Figure 12. (a) Skeletal structure of MeRhPor (Me=CH3) and (b) structural model used in calculations; (c) skeletal structure of RhPor and (d) structural model used in calculations.
Figure 12. (a) Skeletal structure of MeRhPor (Me=CH3) and (b) structural model used in calculations; (c) skeletal structure of RhPor and (d) structural model used in calculations.
Molecules 30 03855 g012
Table 1. Most important geometrical parameters of the coordination sphere for methylrhodium(III)–porphyrin complex (MeRhPor) and rhodium(II)–porphyrin (RhPor) and calculated values of Rh-CMe binding energy (ΔEDE), dissociation energy (ΔEBDE), and bond dissociation free energy (ΔGBDFE) in kcal/mol.
Table 1. Most important geometrical parameters of the coordination sphere for methylrhodium(III)–porphyrin complex (MeRhPor) and rhodium(II)–porphyrin (RhPor) and calculated values of Rh-CMe binding energy (ΔEDE), dissociation energy (ΔEBDE), and bond dissociation free energy (ΔGBDFE) in kcal/mol.
MeRhPorRhPor
S0S1T1Exp. (a)S0S1S3S4
Bond length [Å]
Rh-CMe2.0022.0032.0001.974
Rh-N212.0232.0322.0412.0222.0292.0312.0262.017
Rh-N222.0232.0302.0412.0332.0282.0052.0262.017
Rh-N232.0222.0322.0402.0122.0292.0312.0262.017
Rh-N242.0232.0302.0422.0442.0282.0052.0262.017
Valence angle [°]
N21-Rh-N23177.3177.3178.2178.8180.0180.0180.0180.0
N22-Rh-N24177.2177.2178.2178.5180.0180.0180.0180.0
Dihedral angle [°]
N21-N22-N23-N240.00.10.00.20.00.00.00.0
N21-N22-N23-Rh−1.9−1.9−1.3−0.90.00.00.00.0
[kcal/mol]
With dispersion correction
ΔEDEΔEBDE (b)ΔGBDFE ΔEDEΔEBDE (b)ΔGBDFE
PBE049.945.835.2 54.750.539.9
B3LYP46.742.632.1 54.550.339.8
BP8655.952.340.2 63.860.248.1
PBE59.054.544.1 63.859.248.6
Exp. 58.0 (c)54.3 (d)41.0, 49.0 (e)
(a) Ref. [40]. (b) Bond dissociation energy (BDE, ΔEBDE) is defined as bond energy (ΔEDE) with ZPE correction and thermal corrections. (c) Ref. [41]. (d) Ref. [42]. (e) Bond dissociation free energies (BDFEs) for Rh-CMe bond: 41 kcal/mol and 49 kcal/mol are the BDFG values for the tetra(p-sulfonato-phenyl) porphyrin–rhodium complex in D2O and C6D6, respectively. Data from Ref. [43].
Table 2. Lowest vertical singlet electronic transitions for MeRhPor complex based on TDDFT/PBE0/def2-TZVP calculations with D3BJ dispersion correction and CPCM/benzene solvent model.
Table 2. Lowest vertical singlet electronic transitions for MeRhPor complex based on TDDFT/PBE0/def2-TZVP calculations with D3BJ dispersion correction and CPCM/benzene solvent model.
E (eV)λ (nm)f%Character Experimental (a)
S12.684630.01884193 → 94H → Lπ1 → πx*543 nm (2.28 eV)
3392 → 95H-1 → L+1π2 → πy*
1493 → 95H → L+1π1 → πy*
1192 → 94H-1 → Lπ2 → πx*
S22.684630.01884193 → 95H → L+1π1 → πy*
3392 → 94H-1 → Lπ2 → πx*
1493 → 94H → Lπ1 → πx*
1192 → 95H-1 → L+1π2 → πy*
S32.834380.00014691 → 95H-2 → L+1π/dxz → πy*
4490 → 94H-3 → Lπ/dyz → πx*
S42.904270.00004690 → 95H-3 → L+1π/dyz → πy*
4691 → 94H-2 → Lπ/dxz → πx*
S52.964190.00004790 → 94H-3 → Lπ/dyz → πx*
4691 → 95H-2 → L+1π/dxz → πy*
S63.333730.00534290 → 95H-3 → L+1π/dyz → πy*
4291 → 94H-2 → Lπ/dxz → πx*
S73.403650.09507491 → 96H-2 → L+2π/dxz → σ*
S83.403650.09677490 → 96H-3 → L+2π/dyz → σ*
S93.753301.59433792 → 94H-1 → Lπ2 → πx*395 nm (3.14 eV)
3293 → 95H → L+1π1 → πy*
S103.753301.62043892 → 95H-1 → L+1π2 → πy*
3293 → 94H → Lπ1 → πx*
S113.793270.01089289 → 95H-4 → L+1dx2-y2 → πy*
S123.793270.04039089 → 94H-4 → Ldx2-y2 → πx*
(a) Experimental absorption wavelengths from Ref. [39].
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.

Share and Cite

MDPI and ACS Style

Lodowski, P.; Jaworska, M. A Thorough Understanding of Methylrhodium(III)–Porphyrin Photophysics: A DFT/TDDFT Study. Molecules 2025, 30, 3855. https://doi.org/10.3390/molecules30193855

AMA Style

Lodowski P, Jaworska M. A Thorough Understanding of Methylrhodium(III)–Porphyrin Photophysics: A DFT/TDDFT Study. Molecules. 2025; 30(19):3855. https://doi.org/10.3390/molecules30193855

Chicago/Turabian Style

Lodowski, Piotr, and Maria Jaworska. 2025. "A Thorough Understanding of Methylrhodium(III)–Porphyrin Photophysics: A DFT/TDDFT Study" Molecules 30, no. 19: 3855. https://doi.org/10.3390/molecules30193855

APA Style

Lodowski, P., & Jaworska, M. (2025). A Thorough Understanding of Methylrhodium(III)–Porphyrin Photophysics: A DFT/TDDFT Study. Molecules, 30(19), 3855. https://doi.org/10.3390/molecules30193855

Article Metrics

Back to TopTop