Next Article in Journal
Large Electrocaloric Responsivity and Energy Storage Response in the Lead-Free Ba(GexTi1−x)O3 Ceramics
Previous Article in Journal
Experimental Investigation on the Potential Use of Magnetic Water as a Water Reducing Agent in High Strength Concrete
Previous Article in Special Issue
A Method to Realize Efficient Deep-Red Phosphorescent OLEDs with a Broad Spectral Profile and Low Operating Voltages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards Optimized Photoluminescent Copper(I) Phenanthroline-Functionalized Complexes: Control of the Photophysics by Symmetry-Breaking and Spin–Orbit Coupling

by
Christophe Gourlaouen
* and
Chantal Daniel
Laboratoire de Chimie Quantique Institut de Chimie UMR 7177, CNRS, Université de Strasbourg, 4, Rue Blaise Pascal CS 90032, F-67081 Strasbourg, France
*
Author to whom correspondence should be addressed.
Materials 2022, 15(15), 5222; https://doi.org/10.3390/ma15155222
Submission received: 22 June 2022 / Revised: 19 July 2022 / Accepted: 21 July 2022 / Published: 28 July 2022
(This article belongs to the Special Issue Transition Metal-Based Luminescent Materials and Their Applications)

Abstract

:
The electronic and structural alterations induced by the functionalization of the 1,10-phenanthroline (phen) ligand in [Cu(I) (phen-R2)2]+ complexes (R=H, CH3, tertio-butyl, alkyl-linkers) and their consequences on the luminescence properties and thermally activated delay fluorescence (TADF) activity are investigated using the density functional theory (DFT) and its time-dependent (TD) extension. It is shown that highly symmetric molecules with several potentially emissive nearly-degenerate conformers are not promising because of low S1/S0 oscillator strengths together with limited or no S1/T1 spin–orbit coupling (SOC). Furthermore, steric hindrance, which prevents the flattening of the complex upon irradiation, is a factor of instability. Alternatively, linking the phenanthroline ligands offers the possibility to block the flattening while maintaining remarkable photophysical properties. We propose here two promising complexes, with appropriate symmetry and enough rigidity to warrant stability in standard solvents. This original study paves the way for the supramolecular design of new emissive devices.

1. Introduction

The emissive properties of bis-phenanthroline substituted Cu(I) complexes have been intensively explored starting in the late 70s [1,2,3,4] because of intriguing luminescence properties extremely sensitive to structural effects driven by the surrounding ligands [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26]. The development of time-resolved spectroscopies and pump–probe X-ray experiments has allowed the mapping of ultrafast kinetics that precede the formation of potentially long-lived triplet T1 excited states [26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43]. The decay mechanism involves several intermediate singlet and triplet excited states coupled vibronically and by spin–orbit [44,45,46,47,48,49], the influence of the solvent being significant on the emission properties [50,51]. The relationship between the observed structural distortions, triggered by visible irradiation, and the excited state dynamics is not clear enough for a rigorous rationalization of the luminescent behavior of a variety of more or less constrained complexes proposed over the years. The structural reorganization in the active metal-to-ligand charge transfer (MLCT) state, from the pseudo tetrahedral D2d symmetry ground state to a flattened D2 symmetry excited state, has been investigated combining various spectroscopies with the functionalization of the phenanthrolines [31,33,39,40,52,53,54] (thereafter phen). The hope is to inhibit, by steric hindrance, the early pseudo-Jahn–Teller (PJT) distortion induced by the MLCT electronic transition and to prevent the formation of the flattened structure, critical to generate the appropriate long-lived 3MLCT state and high photoluminescence quantum yields. However, the synthesis of stable complexes with bulky ligands is still a challenge and the resulting photophysical properties do not always live up to expectations [55,56,57,58].
The present theoretical study is dedicated to a comparative investigation of the structural, optical and emissive properties of five [Cu(I) (phen-R2)2]+ complexes (R=H (1H), CH3 (1Met), tertio-butyl (1tBu) and alkyl-linker (2 and 3), Scheme 1) of increasing steric complexity associated with symmetry breaking. The potential energy profiles associated with the low-lying singlet and triplet excited states as well as their spin–orbit coupling (SOC) as a function of the structural deformations are scrutinized in order to rationalize the photophysical behavior within the series and to point to the consequence of symmetry-breaking on emission properties. In contrast to previous theoretical studies, the implication of upper triplet states and advanced spin–orbit effects are considered going beyond an oversimplified S1/T1 two-state model based on frontier molecular orbitals.

2. Computational Method

A first set of calculations has been performed with GAUSSIAN 09 (version D01) [59] at a density functional theory (DFT) level of theory with B3LYP functional [60] in CH2Cl2 through a PCM model [61]. All atoms were described by Pople’s 6–31+G** basis set [62]. Structures in the ground state and in triplet states were fully optimized and the nature of the encountered stationary point was checked by frequency analysis. Minima were characterized by a full set of real frequencies and the transition state by one imaginary frequency. Dispersion corrections were introduced through Grimme’s corrections [63].
A second set of calculations was performed with the ADF 2019 package [64]. The complexes (Scheme 1) electronic ground state and low-lying excited singlet and triplet state structures have been optimized by means of DFT using the B3LYP functional [60] and triple-ζ basis sets for all atoms [65]. Scalar relativistic effects were included using zeroth-order relativistic Hamiltonian [66]. Grimme’s corrections have been applied to consider dispersion effects [67]. Calculations have been performed in CH2Cl2 within the COSMO (conductor-like screening model) approximation [68]. The absorption spectra were computed at the TD-DFT level, including perturbative spin–orbit effects [69,70]. The Tamm–Dancoff approximation (TDA) was employed in order to avoid triplet instability problems [71].
Cuts of the potential energy surfaces along the key nuclear coordinates (Scheme 2), namely the Cu–phen distance (Cu–X1), the rocking angle (X1–Cu–X2) and the dihedral flattening angle (C1–X1–X2–C2), were obtained with remaining coordinates frozen at the ground state structure with ADF.
The excited state analysis was performed by TheoDORE, a package for theoretical density, orbital relaxation and exciton analysis [72].

3. Results

3.1. Ground State and Excited State Structural Properties

The structural data, namely a fully optimized electronic ground state and lowest T1 excited state optimized with GAUSSIAN are given in the SI section. The complexes are structurally flexible even in the ground state (GS) due to the weak inter-ligand interactions. This is illustrated by the variety of 1H GS crystal structures [73,74,75,76,77,78,79,80] (Table S1 in Supplementary Materials), for which the structural parameters depend on the nature of the counter-anion and of the solvent in which the compounds were crystallized. The Cu–N distances vary from 1.996 Å to 2.081 Å, the X1–Cu–X2 angle from 144.8° to 180.0° and the C1–X1–X2–C2 dihedral from 33.9° to 90° (Table S1). The situation is similar for 1Me [81,82,83,84,85,86] (Table S2). In both cases, the computed GS structure is within the range of experimental data. The situation is different for 1tBu, the bulkiness of the tertio-butyl group prevents the significant distortion of the geometry. However, this group introduces a new degree of freedom in the structure according to the orientation of the tertio-butyl groups. For each of them, there are two possibilities, a methyl group can point towards the copper cation or to the opposite of the cation (Figure S1). This generates six possible conformers (Scheme 3, Table S3), four of them (γ, δ, ε, ζ) exhibiting energy differences lower than 3 kcal mol−1 and separated by barriers lower than 7 kcal mol−1. This implies that in the solution there is a thermal distribution of the conformers in fast equilibrium. This is supported by the experimental structures of 1tbu [87]. Depending on the complex environment, the structure can adopt different conformations in the crystal (Table S4). According to frequency analysis, the GS structures of 1H and 1Me are of D2d symmetry point group, that of 1tBu most stable conformer (δ) is only C2. 2 exhibits C2v symmetry and 3 has Cs symmetry (Table S3).
The structural evolution is complex during excited state relaxation. In the optimized structures of the T1 excited state of 1H and 1Me, a significant flattening of the complex is observed leading to a D2 symmetry structure. The C1–X1–X2–C2 angle falls to 41.5° for 1H and to 65° for 1Me (Table S5). In contrast, for 1tBu (steric constraints, Table S5), 2 and 3 (ligand rigidity, Table S5), there is almost no flattening. The flattening generates two degenerated minima in 1H and 1Me, depending on the rotational direction (Scheme S1). These two structures are connected by a transition state (thereafter TS1) of 12.9 kcal mol−1 for 1H and 4.4 kcal mol−1 for 1Me.
The symmetry of 1Me and 1tBu is also reduced by the Cu–N bond breathing. One phen ligand is associated with a significant shortening of the Cu–N bonds, between 0.05 and 0.20 Å. The relaxation of the second phen ligand depends on the steric hindrance with a Cu–N bond shortening in 1Me (Table S5) and a lengthening in 1tBu (Table S5). A similar distortion operates in complexes 2 and 3. The bond breathing generates for each complex (except for 1H) two minima in T1 connected by a transition state (therafter TS2). These minima are strictly degenerate for 1Me and 1tBu, the two phen being identical in these molecules. In 2 and 3, the linkers induce constraints on the phen ligands and slightly lift the degeneracy of the two minima. The barrier associated with TS2 is very low, 2.4 kcal mol−1 for 1tBu. The flattening induces two minima depending on the phen orientations (left and right) in 1H and 1Me. Consequently, the T1 PES is characterized by four minima in 1Me (breathing and flattening, Scheme S1), two minima in 1H (flattening), two minima in 2 and 3 (breathing). The case of 1tBu is more complicated. The orientation of the tBu groups generates six possible conformers in the ground state (Scheme 3), number potentially doubled in the T1 excited state due to the bond breathing. Though some conformers (α, β) can be neglected due to their relative instability, the others may play a central role in the luminescence properties. More specifically, the most stable δ conformer (Scheme 3) will be used as reference for the 1tBu excited state properties described below.
The calculated values of TS1 and TS2 are in favor of non-negligible dynamical effects. The bond breathing (TS2) is associated with low barriers and the occurrence of true minima is uncertain. Furthermore, the barriers increase with the steric hindrance of the R group (none for 1H, and 1Me < 1tBu). The formal oxidation of the copper tends to shorten the Cu–N bonds. However, the ligand repulsion in T1 counterbalances this shortening leading to a pendulum motion: one ligand approaches the copper while the second one moves away from it. This effect increases with the size of the R group. The movement of interconversion of the phen (TS1) is more costly but possible for 1Me at the time-scale of excited state lifetime.

3.2. Low-Lying Singlet and Triplet Excited State Properties

3.2.1. Energetics, Absorption Spectra and Spin–Orbit Coupling

The calculated absorption spectra are depicted in Figure 1 on the basis of ADF optimized structure. The transition energies, absorption wavelengths and oscillator strengths associated with the 25 lowest singlet excited states of complexes 1H, 1Me, 1tBu, 2 and 3 are reported in Table 1 and Table 2, respectively. Complexes 1 absorb in the visible at ~490 nm via the 3rd singlet excited state (S3, f > 10−1) and in the UV domain (f > 10−2) at ~340 nm (S17) and at ~320 nm (S23). The orientation of the tBu groups does not affect the structure of the absorption spectrum of 1tBu, inducing unsignificant shifts of about 25 nm in the position of the absorption bands (Figure S2).
The absorption properties of complex 2 are very similar with three bands at 493 nm (S3), 350 (S18) and 320 nm (S22). The absorption of complex 3 starts at 465 nm (S3, f > 10−1), slightly blue-shifted as compared to the other complexes. As illustrated in Figure 2, the low-lying singlet excited states of the five complexes are mainly of metal-to-ligand charge transfer (MLCT) character with minor additions of ligand-centered (LC) and ligand-to-ligand charge transfer (LLCT). These states principally arise from HOMO and HOMO-1 orbitals towards LUMO to LUMO+3 (Figures S3 and S4).
From TheoDORE analysis (Figure 3), it appears that the low-lying triplet states are essentially MLCT with increasing LC contributions when moving to higher excitation energies. Including spin–orbit corrections does not modify the absorption spectra because the low-lying triplet states do not gain intensity by SOC, as illustrated by the data reported in Table 3 and Table 4, which describe the transition energies of the “spin–orbit” states together with associated oscillator strengths and state mixing. Table 5 reports spin–orbit mixing between the low-lying singlet and triplet states.
Whereas SOC has no effect on the absorption process, we may expect a large influence of SO effects on the early time photophysics. Indeed, within 0.28 eV, we find three singlet (S3, S2, S1) and four triplet (T4, T3, T2, T1) states, all MLCT and potentially activated by absorption, vibronic and spin–orbit coupling for driving the ultrafast decay observed experimentally [38] and tentatively analyzed by quantum dynamics simulations [45] for complex 1H. These seven states, significantly coupled by SOC, lead to a high density of spin–orbit states for all complexes (Table 3 and Table 4) between 2.24 eV (E1) and 2.53 eV (E15) in favor of an ultrafast S3 to T1 decay via an efficient spin-vibronic mechanism induced at Franck–Condon by Cu–N breathing mode activation within ~100 fs and at a longer time-scale (~400 fs) by the PJT distortion, in the case of D2d complexes 1H and 1Me [44,48,49,50]. In the highly symmetric molecules, singlets S3, S2 and S1 are strongly coupled with triplet T1, T4 and T2, respectively; T4 and T3 being strongly coupled to T1 and T2, respectively (Table 5). Breaking the symmetry in the sterically-hindered complex 1tBu, complexes 2 and 3 do not drastically modify the energetics but influence the SO interactions, potentially increasing the singlet–triplet interactions at Franck–Condon, in particular S1/T1 and S3/T3 (complex 1tBu; complex 2) or S2/T1, S3/T2, S3/T3 and T2/T1 in complex 3. However, inhibiting the PJT distortion in these complexes will decrease some of the vibronic interactions. Altogether and without experimental data and/or quantum dynamics simulations for these molecules it is difficult to conclude as to the consequences of the phen substitution on the ultrafast S3 to T1 decay.
From the above considerations we may expect concurrent elementary processes to occur upon irradiation in the MLCT band at about 490 nm. The branching ratio between the different scenario will depend on the experimental conditions and ligand substitutions. We have to distinguish between highly symmetric molecules with small steric constraints (complexes 1H, 1Me), for which flattening nearly inhibits luminescence and sterically-demanding molecules as complex 1tBu is characterized by significant quantum yields (ϕem > 10−2) and rather long lifetimes (a few hundred of ns), challenging the [Ru(bpy)3]2+ complex. These case studies, although well documented, are not fully rationalized and fail at explaining the differences in observed quantum yields and lifetimes.
Quantum dynamics simulations are too prohibitive to be performed systematically on the molecules described here, so we scrutinized the cuts of the PES associated with the seven low-lying active excited states discussed above as a function of the Cu–X distances (C2v symmetry constraint), the X1–Cu–X2 rocking angle (Cs symmetry constraint) and the C1–X1–X2–C2 dihedral angle (D2 symmetry constraint) (Scheme 2).

3.2.2. Calculated Potential Energy Curves Associated with the Low-Lying Excited States and Their Spin–Orbit Interactions

A spin-vibronic mechanism is controlled by the distortions at Franck–Condon that induce intrastate coupling leading to a shift of the excited state potentials in position and in energy. This generates critical geometries favorable to efficient non-adiabatic transitions induced by vibronic and SO couplings. Let us examine the case study of complex 1H and the cuts of PES depicted in Figure 4 for this D2d molecule. It should be kept in mind that in the D2d structure the HOMO is of e symmetry (doubly degenerate orbital) as well as the LUMO, generating degenerate S1 and T1 states. Upon geometry distortion, the symmetry is reduced and the degeneracy is lifted. This has no effect on the PES curve (Figure 4) but has some influence on the SOC curves (Figure 5). The S1–T1 SOC is nil for the breathing and rocking modes but not for the flattening due to the electronic reorganization in the excited states upon symmetry reduction.
The breathing of the phen ligands clearly induces the stabilization of S1 and T1 with the formation of two minima at Cu–X = 3.87 Å and 4.07 Å, corresponding to the exciton delocalization on one or the other phen (Figure 4a and Figure 5). Several crossings between S1 and T4, T3 and T2 will favorize an ultrafast population of these triple states. Moreover, the near-degeneracy between T2 and T1 at 3.97 Å is promising for an efficient population of T1 (Figure 3). The rocking angle X1–Cu–X2 deformation does not modify the ordering of the low-lying excited states and the associated PEC are nearly flat (Figure 4b). More interestingly, the dihedral deformation C1–X1–X2–C2 associated with the flattening of the molecule drastically destabilizes the T3 and T4 excited states and significantly stabilizes the S1, T2 and T1 states (Figure 4c). The large S1–T2 SOC at Frank–Condon and the increase of T1–T2 SOC as function of Cu–X with two maxima at 3.87 Å and 4.07 Å, combined with small S1–T1 and S1–T2 energy gaps (<0.15 eV) (Figure 6), favor a T2/T1 exchange of population at the early time (<1 ps), as observed experimentally [38]. These ultrafast processes have been rationalized theoretically by means of dynamics simulations [44,50].
It seems clear that only S1, T2 and T1 will be involved at the longer time-scales (>a few ps) that control the luminescence quantum yields and lifetimes. The parameters which govern the different mechanisms and the branching ratio between them, namely S1–S0 fluorescence, T1–S0 phosphorescence, indirect TADF and non-radiative decay, are the oscillator strengths; the S1/T1, S1/T2 and T1/T2 SOC; the S1–T1, S1–T2 and T1–T2 ΔE energy gaps and their evolution as functions of the structural deformations.
A TADF mechanism is excluded for 1H because S1–T1 SOC is null and the flattening induces both a large increase of the S1–T1 energy gap and a significant decrease of the S1–S0 oscillator strength. The only possibility of a back population of S1 would be via the T2 state driven by a small S1–T2 energy gap and a significant S1/T2 SOC at the equilibrium structure. However, the structural deformations, both the distortion and the flattening, do not induce important S1/T1 SOC, reduce S1/T2 SOC and increase the energy gaps (Figure 6).
At the T1 optimized structure, the spin–orbit state E1, E2 and E3 oscillator strengths vary drastically as functions of the key nuclear coordinates (Scheme 2), as illustrated in Figure 7. The flattening (Figure 7c) quenches the phosphorescence. Moreover, according to the data reported in Table 6 for 1H, the calculated emission wavelengths associated with S1 → S0 and T1 → S0 are too low to make these transitions radiative.
Let us focus now on the 1tBu complex, in which the flattening is nearly inhibited. The only allowed substantial structural deformation is the breathing, which generates S1 and T1 potentials characterized by two minima, in a manner similar to 1H but totally dissymmetric with one potential well at 3.90 Å, where the system can be trapped either in S1 or in T1 (Figure 8).
In 1tBu, ΔES1–T1 amounts to 0.12 eV on average with very small variations and the symmetry breaking activates S1/T1 SOC. In addition, the structural deformation generates a S1/S0 oscillator strength of ~10−3 in 1tBu as compared to ~10−13 in 1H. At Franck–Condon, the S1–T1 SOC is large, warranting the efficient population of T1. As soon as T1 is populated the system evolves to the T1 minimum at 3.90 Å, where S1 (Figure 9) can be back populated through S1–T1 SOC, assisted by large T1/T2 and S1/T2 SOC (Figure 8a). Consequently, the system may be the seat of T1 → S0 phosphorescence and of both direct (at Franck–Condon) and TADF (at 3.90 Å) S1 → S0 fluorescence. The contribution of T2 to the phosphorescence cannot be excluded. This explains the unique photophysical properties of 1tBu developed after absorption at 425 nm, namely a long-lived MLCT emission (λem, 599 nm; τ, 3260 ns) and the largest quantum yield (ϕ, 5.6%) of all [Cu(R2phen)2]+ complexes [53]. This mechanism, corroborated by the data reported in Table 6 for 1tBu, namely the deformation, emission and stabilization energies, emission wavelengths, singlet-triplet energy gaps, SOC and oscillator strengths at the excited state optimized structures, explains the occurrence of a superposition of phosphorescence and TADF contributions in the steady-state emission spectra as discussed experimentally for a number of new sterically-hindered complexes [54,88,89,90].
The proposed mechanism for the two case studies reported above, namely 1H and 1tBu, may be drastically modified by the experimental conditions (solvent, temperature, etc.). This points to the weakness of the two-state S1/T1 model based on frontier molecular orbitals defined for one minimum. The presence of six identified conformers in the ground state PES of 1tBu (α, β, γ, δ, ε and ζ, Scheme 3), multiplied by two in the S1 and T1 excited states (by breathing motion, Table S11) thwarts the above mechanism on a long time-scale. Based on the low calculated energy barriers (<7 kcal mol−1) the free rotation of the tBu groups in solution and the thermal distribution of the different conformers are ensured (Table 7).
The emission properties (deformation, emission, stabilization energies, emission wavelengths, oscillator strengths, S1–T1 energy splitting and SOC) computed at S1 and T1 minima are reported in Table S11. The C1 symmetry of the T1 state originated from the δ conformer warranties a significant S1–T1 SOC (10.1 cm−1) and a small S1–T1 splitting (0.143 eV) favorable to TADF associated with a highly emissive S1 state (f > 10−4). In contrast the T1 and S1 states generated by the conformer ζ are of C2 symmetry, resulting in the C2v symmetry reduction in the ground state structure. If the S1–T1 splitting (0.140 eV) is unaffected in the ζ conformer, the S1–T1 SOC is inhibited despite this small symmetry breaking (Table S11), which deactivates the TADF mechanism. Moreover, the low oscillator strength of S1 (f < 10−7) does not support fluorescence. Due to its stability, the ζ conformer is undoubtedly present in the solution and contributes to the emission properties either by phosphorescence or by non-radiative decay. Another possibility is that ζ plays the role of a reservoir: by tBu rotation it may evolve to the γ, δ and ε conformers reactivating the SOC and, thus, the TADF mechanism.
The co-existence of several conformers in the solution drastically complicates the mechanism due to the presence of shallow minima in the PES associated with S1, T1 and S0 accessible on a longer time-scale. Moreover, we may expect conformer specific ultrafast decay channels, as observed in some organic chromophores [91].
Whereas the ultrafast population of the low-lying S1, T2 and T1 excited states is well documented, both theoretically and experimentally for 1H, the longer time-scale non-adiabatic dynamics, including spin-vibronic effects and involving these three key states, has to be discovered. The present study paves the way to more sophisticated dynamical simulations in the ns time-scale.

3.3. Towards Supramolecular Design

Despite its remarkable photophysical properties, the tertio-butyl-substituted complex 1tBu suffers from serious drawbacks, namely instability in various solvents [53]. The steric congestion due to the tertio-butyl groups favors the decoordination of one of the substituted phen by a solvent molecule (as CH3CN). This can be illustrated by the calculated complexation energies of the second phen onto the [Cu(phen)]+ complex to give [Cu(phen)2]+. The ΔG of complexation amounts to −31.6 kcal mol−1 for 1H and −38.5 kcal mol−1 for 1Me but decreases to −19.8 kcal mol−1 for 1tBu (computed with GAUSSIAN, in CH3CN). In light of the performance of 1tBu, other functionalizations of the phen have been explored in position 2 and 9 of the phen to prevent flattening by steric congestion. However, none of the tested complexes exhibit performances similar to 1tBu. For instance, inter-ligand interactions may cause flattening even in the singlet ground state, as with 2,9-diphenylphenantroline [92,93]. To the best or our knowledge, no example of flattening blocked due to the linkage of the two ligands exists, while partial linkage has been explored [56,94,95]. We present here two promising structures illustrating this possibility. They derive from the structure ζ of 1tBu, a link being created between the alkyl groups of position 2 of the two phens. The same link is created with the groups in position 9 (Scheme 1). We propose complex 2 (Scheme 1), in which a C2v symmetry is retained, and complex 3, which is asymmetric due to the presence of an isopropyl group.

3.3.1. Complex 2

The linkage constrains the geometry and make the two phen core inequivalent, due to the different orientations of the methyl groups in the linker. Two methyl point towards the copper center and two methyl point outside the complex (Figure 10). The Cu–N distances are significantly larger than in 1H and 1Me, being close to those in 1tBu (Table S11). The X1–Cu–X2 and C1–X1–X2–C2 angles fit the ideal values, namely 180° and 90°, respectively. The absorption spectrum of 2 is very close to that of 1tBu (Figure 1) with similar transitions (Figure 2 and Figure 3).
Upon excitation, the structural rigidity imposed by the linker leads to the retention of the C2v symmetry. As for the previous complexes, the lowest excited states are mainly MLCT, and the electron transferred to the ligand is localized on one of the phen and generates two almost degenerate minima on the lowest S1 and T1 PESs (Table S13). All the structures belong to the A2 symmetry point group and their photophysical characteristics are presented in Table 8. They are very close to those of the ζ form of 1tBu with a small singlet–triplet gap (ΔEST = 0.135 eV) at T1 geometry. However, due to the A2 symmetry of T1 and S1, the SOC S1–T1 is strictly zero, deactivating the TADF mechanism. Furthermore, the oscillator strength associated with S1 is also zero. This first structure proves that the introduction of linkers between the two phen ligands induce structural rigidity preventing any flattening. This results in a small ΔEST similar to those computed with 1tBu. However, the C2v symmetry of the excited singlet and triplet state disfavors the TADF mechanism by cancelling the SOC between S1 and T1 and the oscillator strength of S1.

3.3.2. Complex 3

The structure of cage 2 is of C2v symmetry, and similarly to the ζ conformer of 1tBu, this disfavors emission. The emission properties of 1tBu are due to its asymmetric conformers. We modified the structure of 2 by deleting three of the methyl groups of the linker and replacing the fourth one with an isopropyl (Scheme 1) to introduce an asymmetry in complex 3. The two phen ligands are no longer equivalents (Table S9) and the Cu–N bonds especially are significantly different (2.020 to 2.299 Å) in 3 as compared to 2 (2.140 to 2.156 Å). Furthermore, the phen ligand facing the isopropyl moiety is no longer planar. The absorption spectrum of 3 (Figure 1) is similar in shape to that of 2 but blue-shifted by roughly 50 nm. The nature of the singlet transitions (Figure 2) is the same as in the other complexes, being almost exclusively dominated by MLCTs.
We optimized the lowest excited singlet and triplet states of 3 (Table 8, Table S13, Figure 11). We retrieved the two minima on the S1 and T1 PES, due to the localization of the exciton on each of the phen. The presence of the isopropyl group breaks the symmetry, and the two minima are no longer degenerate. Furthermore, both S1 and T1 minima do not have any symmetry. The consequence is an activation of the SOC between S1–T1 (7.7 cm−1). The linkers, by preventing any flattening of T1 structure (Table S12), retain a small ΔEST (0.146 eV, Table 8). These values are comparable to that of 1tBu for forms γ, δ or ε (Table S9), which are those contributing to the emission properties of 1tBu.

4. Discussion

The emission properties of the [Cu(phen)2]+ class of complexes is governed by the TADF mechanism, namely a back-population of S1 from T1. This process depends on two critical parameters, the spin–orbit coupling and the energy splitting between S1 and T1 in the simplest case. Upon excitation, the lowest excited state of these complexes is a MLCT from the copper cation towards the ligand, which induces nuclear relaxation. Two main motions (Scheme S1) are involved. First, the formal electronic state of the copper is Cu2+, with only nine electrons in the 3d shell, this leads to the rotation of the phen ligand to form a planar complex, with a half-filled 3dx2-y2 orbital (the “flattening”). This is the first motion. The second one is due to the localization of the excited electron on the ligand. In the Franck–Condon geometry, the electron is generally delocalized on the two phen. However, upon relaxation, there is a breathing of the Cu–N bonds due to the change of Cu formal oxidation state associated with the localization of the excited electron on only one of the phen (“the breathing”). This potentially generates two minima, as the excited electron could be on one or the other phen. The minima are degenerate if the two phen are identical, and if they are asymmetric, degeneracy is lifted. In addition to these two motions, a last point has to be considered. Indeed, with complex ligands, the conformational flexibility has to be taken into account. 1tBu is a good example with several conformers generated by the relative orientation of the tertio-buthyl groups, each conformer exhibiting its own emission properties. The experimental values are the results of the contribution of the conformers thermal distribution.
The flattening is associated with an increase in the ΔEST, which disfavors TADF and paves the way to phosphorescence or non-radiative decay. We also observe a decrease in the S1–T1 SOC. The changes associated with the breathing are less important, the variations of the ΔEST are negligible but the SOC decreases with the distortion of the structure. The prevention of the flattening and limitation of the breathing are mandatory to retain good emission properties. This is achieved with 1tBu, which bulky tertio-butyl groups makes flattening impossible and is associated with the best emission performances. Less bulky CH3 groups in 1Me lead to significant flattening with poor emission and 1H is not emissive. However, the bulkiness of the tertio-butyl groups is the source of inter-ligand repulsion, destabilizing the complex. In a coordinative solvent, such as CH3CN, one of the ligands may be substituted by solvent molecule. To overcome this problem, many complexes have been synthesized with less bulky substituents. However, if the complex stability increases, the less sterically hindered complex leads to significant flattening, weakening the emission performance. The reason for this flattening is that the substituent can adapt its position, as observed with isopropyl [96]. For the latter, the isopropyl rotate and allows a significant flattening. An extreme case can be seen in phenyl rings instead of tertio-butyl, in this complex the structure is flattened even in the ground state [92,93]. All the proposed structural modifications rely on the geometry constraints due to substituent repulsion to maintain a geometry close to Franck–Condon geometry in the excited state.
An alternative path is possible by linking the phen ligand together. Some structures have been synthesized but not with linkers between the phens on the same complex. Our detailed theoretical study led to promising molecules with emission characteristics tailored by symmetry breaking and spin–orbit coupling. Due to the structural rigidity, the complex should be stable in a standard solvent, with the departure of one of the phen requiring the departure of both of them. The flattening in excited states is prevented here by linkers instead of steric repulsion. Consequently, the S1–T1 energy gap remains small in complexes 2 and 3 and compares to the calculated value in 1tBu. However, as shown in 1tBu, structures that are too symmetric and close to C2V, are not in favor of efficient emission, as illustrated by complex 2 or by some conformers of 1tBu. Complex 3, with its asymmetric linkers, exhibits the most promising emission characteristics.

5. Conclusions

We report here a complete computational study of the structural, optical and photophysical properties of copper(I) phenanthroline-functionalized complexes, including solvent and spin–orbit coupling effects. The study focusses on the parameters which govern the different decay mechanisms and the branching ratio between them, namely S1–S0 fluorescence, T1–S0 phosphorescence, indirect TADF and non-radiative process. For this purpose, the oscillator strengths, the S1/T1, S1/T2 and T1/T2 spin–orbit interactions as well the S1–T1, S1–T2 and T1–T2 energy gaps are calculated as a function of the structural deformations of the complexes. More specifically, three modes are important: the Cu–X breathing, the X–Cu–X angular distortion (X being the center of the functionalized phen ligands) and the flattening mode. A TADF mechanism is excluded for [Cu(I)(phen-R2)2]+ (R=H, methyl) because of a null S1–T1 SOC, a large S1–T1 energy gap and a significant decrease in S1–S0 oscillator strength induced by flattening. Radiative processes are hindered as well. The inhibition of the flattening in [Cu(I)(phen-R2)2]+ (R = tertio-butyl) and the identification of six conformers in its electronic ground state, generating twelve local minima in the S1 and T1 potential energy surfaces along the breathing mode lead to a complex mechanism. At the early time, a large S1–T1 SOC activated by symmetry breaking, a small S1–T1 energy gap and a large S1/S0 oscillator strength spawned by structural deformation are in favor of both direct fluorescence (at Franck–Condon) and TADF. An efficient phosphorescence is expected from T1 with a potential contribution of T2. At longer time-scales the co-existence of several conformers in solution drastically complicates the mechanism. These features explain the unique photophysical properties of tertio-butyl complex as well as the occurrence of the superposition of phosphorescence and TADF contributions in newly synthesized sterically hindered complexes.
In order to overcome the drawback of instability in solution, the consequence of the decoordination of one phen by a solvent molecule, two complexes with promising photophysical properties are proposed. The new structures are derived from the C2v conformer ζ of the tertio-butyl substituted Cu(I) complex by linking the two phen in positions 2 and 9 (complex 2). The introduction of asymmetry (complex 3) by appropriate linkers leads to the most interesting molecular cage in terms of photophysical characteristics paving the way for a new supramolecular design.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15155222/s1, Figure S1: In (left) and out (right) orientations of a tBu group on a simplified complex holding only one tBu; Figure S2: ADF computed Absorption spectra for the different conformers of 1tBu; Figure S3: Frontier orbitals of 1H and energies in eV; Figure S4: Frontier orbitals of 3 and energies in eV; Scheme S1: Representation of the four possible Triplet minima after irradiation of 1Me complex. In bold blue are the long Cu-N distance, in bold red the short Cu-N distance and in purple the indication of how the phen ligands is out of plane.; Table S1: Experimental (from CSD reference) and computed (GAUSSIAN) geometrical parameters for complex 1H. Distances are in Angstroms, angles and dihedral angles in degrees; Table S2: Experimental (from CSD reference) and computed (GAUSSIAN) geometrical parameters for complex 1Me. Distances are in Angstroms, angles and dihedral angles in degrees; Table S3: Computed (GAUSSIAN) geometrical parameters for complexes 1tBu, 2 and 3 in the ground state. Distances are in Angstroms, angles and dihedral angles in degrees; Table S4: Experimental geometrical parameters for complex 1tBu. Distances are in Angstroms, angles and dihedral angles in degrees; Table S5: Computed (GAUSSIAN) geometrical parameters for complexes in the triplet state. The values are given for conformer δ for 1tBu. Distances are in Angstroms, angles and dihedral angles in degrees and energies in kcal mol−1; Table S6: Selected nuclear coordinates at the optimized ground state and low-lying excited states for 1H computed with ADF. Distances are in Angstroms, angles and dihedral angles in degrees; Table S7: Selected nuclear coordinates at the optimized ground state and low-lying excited states for 1Me computed with ADF. Distances are in Angstroms, angles and dihedral angles in degrees; Table S8: Selected nuclear coordinates at the optimized ground state and low-lying excited states for 1tBu for experimental δ structure computed with ADF. Distances are in Angstroms, angles and dihedral angles in degrees; Table S9: Computed ground state (ADF) geometrical parameters. Distances are in Angstroms, angles and dihedral angles in degrees; Table S10: Selected nuclear coordinates at the low-lying excited states for 1tBu computed with ADF. Distances are in Angstroms, and angles in degrees; Table S11: Emission data computed with ADF for S1 and T1 states of 1tBu. Edef, Eem, Estab, ΔE and ΔEST are in eV, λem is in nm, SOC is in cm−1. 1) Values for the 2nd minima due to oscillation of the exciton for γ, ε and ζ. Data are not given for α, β and δ because the second minima is strictly identical to the first one due to symmetry; Table S12: Spin-orbit coupling (in cm−1) between the low-lying excited states in 1Me computed with ADF; Table S13: Computed S1 and T1 (ADF) geometrical parameters. Distances are in Angstroms, angles and dihedral angles in degrees. References [73,74,75,76,77,78,79,80,81,82,83,84,85,86,87] are cited in the Supplementary Materials.

Author Contributions

Writing—original draft, C.G. and C.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The calculations were carried at the Strasbourg computing center (HPC).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Buckner, M.T.; McMillin, D.R. Photoluminescence from copper(I) complexes with low-lying metal-to-ligand charge transfer excited states. J. Chem. Soc. Chem. Commun. 1978, 759–761. [Google Scholar] [CrossRef]
  2. Blaskie, M.W.; McMillin, D.R. Photostudies of copper(I) systems. 6. Room-temperature emission and quenching studies of bis(2,9-dimethyl-1,10-phenanthroline)copper(I). Inorg. Chem. 1980, 19, 3519–3522. [Google Scholar] [CrossRef]
  3. Rader, R.A.; McMillin, D.R.; Buckner, M.T.; Matthews, T.G.; Casadonte, D.J.; Lengel, R.K.; Whittaker, S.B.; Darmon, L.M.; Lytle, F.E. Photostudies of [Cu(bpy)(PPh3)2]+, [Cu(phen)(PPh3)2]+, and [Cu(dmp)(PPh3)2]+ in Solution and in Rigid, Low-Temperature Glasses. Simultaneous Multiple Emissions from Intraligand and Charge-Transfer States. J. Am. Chem. Soc. 1981, 103, 5906–5912. [Google Scholar] [CrossRef]
  4. Kirchhoff, J.R.; Gamache, R.E., Jr.; Blaskie, M.W.; Del Paggio, A.A.; Lengel, R.K.; McMillin, D.R. Temperature dependence of luminescence from Cu(NN)2+ systems in fluid solution. Evidence for the participation of two excited states. Inorg. Chem. 1983, 22, 2380–2384. [Google Scholar] [CrossRef]
  5. Kutal, C. Spectroscopic and photochemical properties of d10 metal complexes. Coord. Chem. Rev. 1990, 99, 213–252. [Google Scholar] [CrossRef]
  6. Sabin, F.; Ryu, C.K.; Ford, P.C.; Vogler, A. Photophysical properties of hexanuclear copper(I) and silver(I) clusters. Inorg. Chem. 1992, 31, 1941–1945. [Google Scholar] [CrossRef] [Green Version]
  7. Li, D.; Che, C.-M.; Wong, W.T.; Shieh, S.J.; Peng, S.M. Luminescent three-co-ordinated dinuclear copper(I) complexes of bis(diphenylphosphino)methane: Photophysical properties and crystal structure. J. Chem. Soc. Dalton Trans. 1993, 653–654. [Google Scholar] [CrossRef]
  8. Yam, V.W.-W.; Fung, W.K.-M.; Cheung, K.-K. Synthesis, Structure, Photophysics, and Excited-State Redox Properties of the Novel Luminescent Tetranuclear Acetylidocopper(I) Complex [Cu4(μ-dppm)4412-C{C-)](BF4)2. Angew. Chem. Int. Ed. 1996, 35, 1100–1102. [Google Scholar] [CrossRef]
  9. Yam, V.W.-W.; Lo, K.K.-W.; Fung, W.K.-M.; Wang, C.R. Design of luminescent polynuclear copper(I) and silver(I) complexes with chalcogenides and acetylides as the bridging ligands. Coord. Chem. Rev. 1998, 171, 17–41. [Google Scholar] [CrossRef]
  10. Miller, M.T.; Karpishin, T.B. Phenylethynyl substituent effects on the photophysics and electrochemistry of [Cu(dpp2)]2+ (dpp = 2,9-Diphenyl-1,10-phenanthroline). Inorg. Chem. 1999, 38, 5246–5249. [Google Scholar] [CrossRef]
  11. Miller, M.T.; Gantzel, P.K.; Karpishin, T.B. A highly emissive heteroleptic copper(I) bis(phenanthroline) complex: [Cu(dbp)(dmp)]+ (dbp = 2, 9-Di-tert-butyl-1,10-phenanthroline; dmp = 2, 9-Dimethyl-1,10-phenanthroline). J. Am. Chem. Soc. 1999, 121, 4292–4293. [Google Scholar] [CrossRef]
  12. Yam, V.W.-K.; Lo, K.K.-W. Luminescent polynuclear d10 metal complexes. Chem. Soc. Rev. 1999, 28, 323–334. [Google Scholar] [CrossRef]
  13. Ford, P.C.; Cariati, E.; Bourassa, J. Photoluminescence properties of multinuclear copper(I) compounds. Chem. Rev. 1999, 99, 3625–3647. [Google Scholar] [CrossRef] [PubMed]
  14. Scaltrito, D.V.; Thompson, D.W.; O’Calaghan, J.A.; Meyer, G.J. MLCT excited states of cuprous bis-phenanthroline coordination compounds. Coord. Chem. Rev. 2000, 208, 243–266. [Google Scholar] [CrossRef]
  15. Che, C.M.; Mao, Z.; Miskowski, V.M.; Tse, M.C.; Chan, C.K.; Cheung, K.K.; Phillips, D.L.; Leung, K.H. Cuprophilicity: Spectroscopic and Structural Evidence for Cu−Cu Bonding Interactions in Luminescent Dinuclear Copper(I) Complexes with Bridging Diphosphane Ligands. Angew. Chem. Int. Ed. 2000, 112, 4250–4254. [Google Scholar] [CrossRef]
  16. Mao, Z.; Chao, H.Y.; Hui, Z.; Che, C.-M.; Fu, W.-F.; Cheung, K.-K.; Zhu, N. Excited States of Binuclear Copper(I) Phosphine Complexes: Effect of Copper–Ligand and Copper–Copper Interactions on Excited State Properties and Photocatalytic Reductions of the 4,4′-Dimethyl-2,2′-bipyridinium Ion in Alcohols. Chem. Eur. J. 2003, 9, 2885–2894. [Google Scholar] [CrossRef]
  17. Pawlowski, V.; Knoer, G.; Lennartz, C.; Vogler, A. Luminescence and Theoretical Studies of Cu(tripod)X [tripod = 1,1,1-Tris(diphenylphosphanylmethyl)ethane; X = Halide, Thiophenolate, Phenylacetylide]. Eur. J. Inorg. Chem. 2005, 15, 3167–3171. [Google Scholar] [CrossRef]
  18. Peng, R.; Li, D.; Wu, T.; Zhou, X.-P.; Ng, S.W. Increasing Structure Dimensionality of Copper(I) Complexes by Varying the Flexible Thioether Ligand Geometry and Counteranions. Inorg. Chem. 2006, 45, 4035–4046. [Google Scholar] [CrossRef]
  19. Armaroli, N.; Accorsi, G.; Cardinali, F.; Listorti, A. Photochemistry and Photophysics of Coordination Compounds: Copper. Top. Curr. Chem. 2007, 280, 69–115. [Google Scholar]
  20. Barbieri, A.; Accorsi, G.; Armaroli, N. Luminescent complexes beyond the platinumgroup: The d10 avenue. Chem. Commun. 2008, 2185–2193. [Google Scholar] [CrossRef]
  21. Tsubomura, T.; Tsukuda, T.; Matsumoto, K. Luminescent d10 transition metal complexes. Bull. Jpn. Soc. Coord. Chem. 2008, 52, 29–42. [Google Scholar] [CrossRef]
  22. Lavie-Combot, A.; Cantuel, M.; Leydet, Y.; Jonusauskas, G.; Bassani, D.M.; McClenaghan, N.D. Improving the photophysical properties of copper(I) bis(phenanthroline) complexes. Coord. Chem. Rev. 2008, 252, 2572–2584. [Google Scholar] [CrossRef]
  23. Tsuge, K. Luminescent Complexes Containing Halogeno-bridged Dicopper(I) Unit {Cu2(µ-X)2} (X = Cl, Br, and I). Chem. Lett. 2013, 42, 204–208. [Google Scholar]
  24. Tsubomura, T.; Kimura, K.; Nishikawa, M.; Tsukuda, T. Structures and photophysical properties of copper(I) complexes bearing diphenylphenanthroline and bis(diphenylphosphino)alkane: The effect of phenyl groups on the phenanthroline ligand. Dalton Trans. 2015, 44, 7554–7562. [Google Scholar] [CrossRef] [PubMed]
  25. Kakizoe, D.; Nishikawa, N.; Fuji, Y.; Tsubomura, T. Photophysicle properties of three coordinated copper(I) complexes bearing 1,10-phenanthroline and a monodentate phosphine ligand. Dalton Trans. 2017, 46, 14804–14811. [Google Scholar] [CrossRef] [PubMed]
  26. Wallesch, M.; Volz, D.; Zink, D.M.; Schepers, U.; Nieger, M.; Baumann, T.; Braese, S. Bright Coppertunities: Multinuclear CuI Complexes with N–P Ligands and Their Applications. Chem. Eur. J. 2014, 20, 6578–6590. [Google Scholar] [CrossRef]
  27. Giereth, R.; Mengele, A.K.; Frey, W.; Kloss, M.; Stuffen, A.; Karnahl, M.; Tschierlei, S. Copper(I) phosphinooxalozine complexes: Impact of the ligand substitution and steric demand on the electrochemical and photophysical properties. Chem. Eur. J. 2020, 26, 2675–2684. [Google Scholar] [CrossRef]
  28. Gunaratne, T.; Rodgers, M.A.J.; Felder, D.; Nierengarten, J.-F.; Accorsi, G.; Armaroli, N. Ultrafast dynamics of Cu(I)-phenanthrolines in dichloromethane. Chem. Commun. 2003, 3010–3011. [Google Scholar] [CrossRef]
  29. Samia, A.C.S.; Cody, J.; Fahrni, C.J.; Burda, C. The effects of ligand constraints on the metal-to-ligand charge transfer relaxation dynamics in copper (I)-phenanthroline complexes: A comparative study by femtosecond time-resolved spectroscopy. J. Phys. Chem. B 2000, 108, 563–569. [Google Scholar] [CrossRef]
  30. Cody, J.; Dennisson, J.; Gilmore, J.; VanDerveer, D.G.; Henary, M.M.; Gabrielli, A.; Sherrill, C.D.; Zhang, Y.; Pan, C.-P.; Burda, C.; et al. X-ray structures, photophysical characterization, and computational analysis of geometrically constrained copper(I)-phenanthroline complexes. Inorg. Chem. 2003, 42, 4918–4929. [Google Scholar] [CrossRef]
  31. Kohler, L.; Hayes, D.; Hong, J.Y.; Carter, T.J.; Shelby, M.L.; Fransted, K.A.; Chen, L.X.; Mulfort, K.L. Synthesis, structure, ultrafast kinetics, and light-induced dynamics of CuHETPHEN chromophores. Dalton Trans. 2016, 45, 9871–9883. [Google Scholar] [CrossRef]
  32. Kelley, M.S.; Shelby, M.L.; Mara, M.W.; Haldrup, K.; Hayes, D.; Hadt, R.G.; Zhang, X.; Stickrath, A.B.; Ruppert, R.; Sauvage, J.-P.; et al. Ultrafast dynamics of two copper bisphenanthroline complexes measured by X-ray transient absorption spectroscopy. J. Phys. B-Atom. Mol. Opt. Phys. 2017, 50, 154006. [Google Scholar] [CrossRef] [Green Version]
  33. Iwamura, M.; Takeuchi, S.; Tahara, T. Substituent effect on the photoinduced structural change of Cu(I) complexes observed by femtosecond emission spectroscopy. Phys. Chem. Chem. Phys. 2014, 16, 4143–4154. [Google Scholar] [CrossRef]
  34. Fransted, K.A.; Jackson, N.E.; Zong, R.; Mara, M.W.; Huang, J.; Harpham, M.R.; Shelby, M.L.; Thununel, R.P.; Chen, L.X. Ultrafast Structural Dynamics of Cu(I)-Bicinchoninic Acid and Their Implications for Solar Energy Applications. J. Physical Chem. A 2014, 118, 10497–10506. [Google Scholar] [CrossRef]
  35. Shaw, G.B.; Grant, C.D.; Shirota, H.; Castner, E.W.; Meyer, G.J.; Chen, L.X. Ultrafast structural rearrangements in the MLCT excited state for copper(I) bis-phenanthrolines in solution. J. Am. Chem. Soc. 2007, 129, 2147–2160. [Google Scholar] [CrossRef] [Green Version]
  36. Iwamura, M.; Takeuchi, S.; Tahara, T. Real-time observation of the photoinduced structural change of bis(2,9-dimethyl-1,10-phenanthroline)copper(I) by femtosecond fluorescence spectroscopy: A realistic potential curve of the Jahn-Teller distortion. J. Am. Chem. Soc. 2007, 129, 5248–5256. [Google Scholar] [CrossRef] [PubMed]
  37. Lockard, J.V.; Kabehie, S.; Zink, J.I.; Smolentsev, G.; Soldatov, A.; Chen, L.X. Influence of Ligand Substitution on Excited State Structural Dynamics in Cu(I) Bisphenanthroline Complexes. J. Phys. Chem. B 2010, 114, 14521–14527. [Google Scholar] [CrossRef] [PubMed]
  38. Iwamura, M.; Takeuchi, S.; Tahara, T. Ultrafast Excited-State Dynamics of Copper(I) Complexes. Acc. Chem. Res. 2015, 48, 782–791. [Google Scholar] [CrossRef]
  39. Hua, L.; Iwamura, M.; Takeuchi, S.; Tahara, T. The substituent effect on the MLCT excited state dynamics of Cu(I) complexes studied by femtosecond time-resolved absorption and observation of coherent nuclear wavepacket motion. Phys. Chem. Chem. Phys. 2015, 17, 2067–2077. [Google Scholar] [CrossRef]
  40. Gothard, N.A.; Mara, M.W.; Huang, J.; Szarko, J.M.; Rolczynski, B.; Lockard, J.V.; Chen, L.X. Strong Steric Hindrance Effect on Excited State Structural Dynamics of Cu(I) Diimine Complexes. J. Phys. Chem. A 2012, 116, 1984–1992. [Google Scholar] [CrossRef]
  41. Papanikolaou, P.A.; Tkachenko, N.V. Probing the excited state dynamics of a new family of Cu(I)-complexes with an enhanced light absorption capacity: Excitation-wavelength dependent population of states through branching. Phys. Chem. Chem. Phys. 2013, 15, 13128–13136. [Google Scholar] [CrossRef] [PubMed]
  42. Grupe, M.; Bappler, F.; Theiss, M.; Busch, J.M.; Dietrich, F.; Volz, D.; Gerhards, M.; Bräse, S.; Diller, R. Real-time observation of molecular flattening and intersystem crossing in [(DPEPhos)Cu(I)(PyrTet)] via ultrafast UV/Vis- and mid-IR spectroscopy on solution and solid samples. Phys. Chem. Chem. Phys. 2020, 22, 14187–14200. [Google Scholar] [CrossRef] [PubMed]
  43. Iwamura, M.; Watanabe, H.; Ishii, K.; Takeuchi, S.; Tahara, T. Coherent Nuclear Dynamics in Ultrafast Photoinduced Structural Change of Bis(diimine)copper(I) Complex. J. Am. Chem. Soc. 2011, 133, 7728–7736. [Google Scholar] [CrossRef] [PubMed]
  44. Capano, G.; Penfold, T.J.; Chergui, M.; Tavernelli, I. Photophysics of a copper phenanthroline elucidated by trajectory and wavepacket-based quantum dynamics: A synergetic approach. Phys. Chem. Chem. Phys. 2017, 19, 19590–19600. [Google Scholar] [CrossRef] [Green Version]
  45. Capano, G.; Chergui, M.; Rothlisberger, U.; Tavernelli, I.; Penfold, T.J. A Quantum Dynamics Study of the Ultrafast Relaxation in a Prototypical Cu(I)–Phenanthroline. J. Phys. Chem. A 2014, 118, 9861–9869. [Google Scholar] [CrossRef]
  46. Zgierski, M.Z. Cu(I)-2,9-dimethyl-1,10-phenanthroline: Density functional study of the structure, vibrational force-field, and excited electronic states. J. Chem. Phys. 2003, 118, 4045. [Google Scholar] [CrossRef]
  47. Stoïanov, A.; Gourlaouen, C.; Vela, S.; Daniel, C. Luminescent Dinuclear Copper(I) Complexes as Potential Thermally Activated Delayed Fluorescence (TADF) Emitters: A Theoretical Study. J. Phys. Chem. A 2018, 122, 1413–1421. [Google Scholar] [CrossRef]
  48. Levi, G.; Biasin, E.; Dohn, A.O.; Jonsson, H. On the interplay of solvent and conformational effects in simulated excited-state dynamics of a copper phenanthroline photosensitizer. Phys. Chem. Chem. Phys. 2020, 22, 748–757. [Google Scholar] [CrossRef]
  49. Du, L.; Lan, Z. Ultrafast structural flattening motion in photoinduced excited state dynamics of bis(diamine)copper(I) complex. Phys. Cem. Chem. Phys. 2016, 18, 7641–7650. [Google Scholar] [CrossRef]
  50. Capano, G.; Rothlisberger, U.; Tavernelli, I.; Penfold, T.J. Theoretical Rationalization of the Emission Properties of Photophysical Cu(I)-phenanthroline Complexes. J. Phys. Chem. A. 2015, 119, 7026–7037. [Google Scholar] [CrossRef]
  51. Guda, A.; Windisch, J.; Probst, B.; van Bokhoven, J.A.; Alberto, R.; Nachtegaal, M.; Chen, L.X.; Smolentsev, G. Excited-State Structure of Copper Phenanthroline-Based Photosensitizers. Phys. Chem. Chem. Phys. 2021, 23, 26729–26736. [Google Scholar] [CrossRef] [PubMed]
  52. Kalsani, V.; Schmittel, M.; Listorti, A.; Accorsi, G.; Armaroli, N. Novel Phenanthroline Ligands and Their Kinetically Locked Copper(I) Complexes with Unexpected Photophysical Properties. Inorg. Chem. 2006, 45, 2061–2067. [Google Scholar] [CrossRef] [PubMed]
  53. Green, O.; Gandhi, B.A.; Burstyn, J.N. Photophysical Characteristics and Reactivity of Bis(2,9-di-tert-butyl-1,10-phenanthroline) Copper(I). Inorg. Chem. 2009, 48, 5704–5714. [Google Scholar] [CrossRef]
  54. Garakyaraghi, S.; Koutnik, P.; Castellano, F.N. Photoinduced Structural Distortions and Singlet-Triplet Intersystem Crossing in Cu(I) MLCT Excited States Monitored by Optically Gated Fluorescence Spectroscopy. Phys. Chem. Chem. Phys. 2017, 19, 16662–16668. [Google Scholar] [CrossRef] [PubMed]
  55. Garakyaraghi, S.; McCusker, C.E.; Khan, S.; Koutnik, P.; Bui, A.T.; Castellano, F.N. Enhancing the Visible-Light Absorption and Excited-State Properties of Cu(I) MLCT Excited States. Inorg. Chem. 2018, 57, 2296–2307. [Google Scholar] [CrossRef]
  56. Brandl, T.; Kerzig, C.; Le Pleux, L.; Prescimone, A.; Wenger, O.S.; Mayor, M. Improved Photostability of a CuI Complex by Macrocyclization of the Phenanthroline Ligands. Chem. Eur. J. 2020, 26, 3119–3128. [Google Scholar] [CrossRef] [PubMed]
  57. Livshits, M.Y.; Reeves, B.J.; DeWeerd, J.; Strauss, S.H.; Boltalina, O.V.; Rack, J.J. Trifluoromethylated Phenanthroline Ligands Reduce Excited-State Distortion in Homoleptic Copper(I) Complexes. Inorg. Chem. 2020, 59, 2781–2790. [Google Scholar] [CrossRef] [PubMed]
  58. Rosko, M.C.; Wells, K.A.; Hauke, C.E.; Castellano, F.N. Next Generation Cuprous Phenanthroline MLCT Photosensitizer Featuring Cyclohexyl Substituents. Inorg. Chem. 2021, 60, 8394–8403. [Google Scholar] [CrossRef]
  59. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09; Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  60. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  61. Miertuš, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilisation of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  62. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724. [Google Scholar] [CrossRef]
  63. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. ADF, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. Available online: https://www.scm.com/doc/ADF/index.html (accessed on 1 June 2019).
  65. Van Lenthe, E.; Baerends, E.J. Optimized Slater-type basis sets for the elements 1−118. J. Comp. Chem. 2003, 24, 1142–1156. [Google Scholar] [CrossRef] [PubMed]
  66. Van Lenthe, E.; van Leeuwen, R.; Baerends, E.J.; Snijders, J.G. Relativistic regular two-component Hamiltonians. Int. J. Quant. Chem. 1996, 57, 281–293. [Google Scholar] [CrossRef]
  67. Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  68. Klamt, A. Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation Phenomena. J. Phys. Chem. 1995, 99, 2224–2235. [Google Scholar] [CrossRef]
  69. Wang, F.; Ziegler, T. A simplified relativistic time-dependent density-functional theory formalism for the calculations of excitation, energies including spin-orbit coupling effect. J. Chem. Phys. 2005, 123, 154102–154112. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, F.; Ziegler, T.; van Lenthe, E.; van Gisbergen, S.; Baerends, E.J. The calculation of excitation energies based on the relativistic two-component zeroth-order regular approximation and time-dependent density-functional with full use of symmetry. J. Chem. Phys. 2005, 122, 204103–204112. [Google Scholar] [CrossRef]
  71. Peach, M.J.G.; Tozer, D.J. Overcoming Low Orbital Overlap and Triplet Instability Problems in TDDFT. J. Phys. Chem. A 2012, 116, 9783–9789. [Google Scholar] [CrossRef]
  72. Plasser, F. TheoDORE: A toolbox for a detailed and automated analysis of electronic excited state computations. J. Chem. Phys. 2020, 152, 084108. [Google Scholar] [CrossRef] [Green Version]
  73. Wang, R.Z.; Xu, J.Q.; Yang, G.Y.; Bu, W.M.; Xing, Y.H.; Li, D.M.; Liu, S.Q.; Ye, L.; Fan, Y.G. Structure of [Cu(phen)2]2[{Cu(phen)}2Mo8O26].H2O. Pol. J. Chem. 1999, 73, 1909. [Google Scholar]
  74. Avdeeva, V.V.; Dziova, A.E.; Polyakova, I.N.; Malinina, E.A.; Goeva, L.V.; Kuznetsov, N.T. Copper(I), copper(II), and heterovalent copper(I,II) complexes with 1,10-phenanthroline and the closo-decaborate anion. Inorg. Chim. Acta 2015, 430, 74–81. [Google Scholar] [CrossRef]
  75. Rheingold, A.L.; Department of Chemistry, University of California, La Jolla, CA, USA; Maatta, E.; Department of Chemistry, Kansas State University, Manhattan, KS, USA. CSD Communication, 2015.
  76. Huang, Z.; Hartwig, J.F. Copper(I) Enolate complexes in α-Arylation Reactions: Synthesis, Reactivity and Mechanism. Angew. Chem. Int. Ed. 2012, 51, 1028–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Lin, X.; Hou, C.; Li, H.; Weng, Z. Decarboxylative Trifluoromethylating Reagent [Cu(O2CCF3)(phen)] and Difluorocarbene Precursor [Cu(phen)2][O2CCF2Cl]. Chem. Eur. J. 2016, 22, 2075–2084. [Google Scholar] [CrossRef] [PubMed]
  78. Hillhouse, G.; Searle Chemistry Laboratory, Department of Chemistry, The University of Chicago, Chicago, IL, USA; Rheingold, A.L.; Department of Chemistry, University of Delaware, Newark, DE, USA; Allen, M.B. CSD Communication, 1996.
  79. Huang, Y.; Ajitha, M.J.; Huang, K.-W.; Zhang, Z.; Weng, Z. A class of effective decarboxylative perfluoroalkylating reagents: [(phen)2Cu](O2CRF). Dalton Trans. 2016, 45, 8468–8474. [Google Scholar] [CrossRef] [Green Version]
  80. Yu, J.-H.; Lu, Z.-L.; Xu, J.-Q.; Bie, H.-Y.; Lu, J.; Zhang, X. Syntheses, characterization and optical properties of some copper(i) halides with 1,10-phenanthroline ligand. New J. Chem. 2004, 28, 940–945. [Google Scholar] [CrossRef]
  81. Hamalainen, R.; Department of Inorganic Chemistry, University of Helsinki, Helsinki, Finland; Ahlgren, M.; Department of Inorganic Chemistry, University of Helsinki, Helsinki, Finland; Turpeinen, U.; Department of Inorganic Chemistry, University of Helsinki, Helsinki, Finland; Raikas, T.; Department of Inorganic Chemistry, University of Helsinki, Helsinki, Finland. Crystal Structure Communications, 1979.
  82. Dessy, G.; Laboratorio di Teoria e Struttura Elettronica e Comportamento Spettrochimico dei Composti di Coordinazione del CNR, Monterotondo Stazione, Italy; Fares, V.; Laboratorio di Teoria e Struttura Elettronica e Comportamento Spettrochimico dei Composti di Coordinazione del CNR, Monterotondo Stazione, Italy. Crystal Structure Communications, 1979.
  83. Royappa, A.T.; Royappa, A.D.; Moral, R.F.; Rheingold, A.L.; Papoular, R.J.; Blum, D.M.; Duong, T.Q.; Stepherson, J.R.; Vu, O.D.; Chen, B.; et al. Copper(I) oxalate complexes: Synthesis, structures and surprises. Polyhedron 2016, 119, 563–574. [Google Scholar] [CrossRef] [Green Version]
  84. Zheng, S.-L.; Gembicky, M.; Messerschmidt, M.; Dominiak, P.M.; Coppens, P. Effect of the Environment on Molecular Properties:  Synthesis, Structure, and Photoluminescence of Cu(I) Bis(2,9-dimethyl-1,10-phenanthroline) Nanoclusters in Eight Different Supramolecular Frameworks. Inorg. Chem. 2006, 45, 9281–9289. [Google Scholar] [CrossRef]
  85. Kovalevsky, A.Y.; Gembicky, M.; Novozhilova, I.V.; Coppens, P. Solid-State Structure Dependence of the Molecular Distortion and Spectroscopic Properties of the Cu(I) Bis(2,9-dimethyl-1,10-phenanthroline) Ion. Inorg. Chem. 2003, 42, 8794–8802. [Google Scholar] [CrossRef] [PubMed]
  86. Blake, A.J.; Hill, S.J.; Hubberstey, P.; Li, W.-S. Copper(I) chelated by 2,9-dimethyl-1,10-phenanthroline and bridged by 4,4′-bipyridine or trans-1,2-bis(pyridin-4-yl)ethene to give discrete dinuclear and polymeric cations. J. Chem. Soc. Dalton Trans. 1998, 909–916. [Google Scholar] [CrossRef]
  87. Gandhi, B.A.; Green, O.; Burstyn, J.N. Facile oxidation-based of sterically encumbered four-coordinate bis(2,9-di-ter-butyl-1,10-phenanthroline)copper(I) and related three-coordinate copper(I) complexes. Inorg. Chem. 2007, 46, 3816–3825. [Google Scholar] [CrossRef] [PubMed]
  88. Asano, M.S.; Tomiduka, K.; Sekizawa, K.; Yamashita, K.; Sugiura, K. Temperature-dependent Emission of Copper(I) Phenanthroline Complexes with Bulky Substituents: Estimation of an Energy Gap between the Singlet and Triplet MLCT States. Chem. Lett. 2010, 39, 376–378. [Google Scholar] [CrossRef]
  89. Czerwieniec, R.; Yersin, H. Diversity of Copper(I) Complexes Showing Thermally Activated Delayed Fluorescence: Basis Photophysical Analysis. Inorg. Chem. 2015, 54, 4322–4327. [Google Scholar] [CrossRef] [PubMed]
  90. Farias, G.; Salla, C.A.M.; Heying, R.S.; Bortoluzzi, A.J.; Curcio, S.F.; Cazati, T.; dos Santos, P.L.; Monkman, A.P.; de Souza, B.; Bechtlod, I.H. Reducing lifetime in Cu(I) Complexes with Thermally Activated Delayed Fluorescence and Phosphorescence Promoted by Chalcogenolate-diimine ligands. J. Mater. Chem. C 2020, 8, 14595–14604. [Google Scholar] [CrossRef]
  91. Greenough, S.E.; Horbury, M.D.; Thompson, J.O.F.; Roberts, G.M.; Karsili, T.N.V.; Marchetti, B.; Townsend, D.; Stavros, V.G. Solvent induced conformer specific photochemistry of guaiacol. Phys. Chem. Chem. Phys. 2014, 16, 16187–16195. [Google Scholar] [CrossRef]
  92. Miller, M.T.; Gantzel, P.K.; Karpishin, T.B. Structures of the Copper(I) and Copper(II) Complexes of 2,9-Diphenyl-1,10-phenanthroline:  Implications for Excited-State Structural Distortion. Inorg. Chem. 1998, 37, 2285–2290. [Google Scholar] [CrossRef]
  93. Gimeno, L.; Blart, E.; Rebilly, J.-N.; Coupeau, M.; Allain, M.; Roisnel, T.; Quarré de Verneuil, A.; Gourlaouen, C.; Daniel, C.; Pellegrin, Y. Non-symmetrical sterically challenged phenanthroline ligands and their homoleptic copper(I) complexes with improved excited state properties. Chem. Eur. J. 2020, 26, 11887–11899. [Google Scholar] [CrossRef] [PubMed]
  94. Heuft, M.A.; Fallis, A.G. Template-directed synthesis of helical phenanthroline cyclophanes. Angew. Chem. Int. Ed. 2002, 41, 4520–4523. [Google Scholar] [CrossRef]
  95. Dietrich-Buchecker, C.O.; Guilhem, J.; Khemiss, A.K.; Kintzinger, J.-P.; Pascard, C.; Sauvage, J.-P. Molecular Structure of a I3I-Catenate: Curling Up of the Interlocked System by Interaction between the Two Copper Complex Subunits. Angew. Chem. Int. Ed. Engl. 1987, 26, 661–663. [Google Scholar] [CrossRef]
  96. Gimeno, L.; Phelan, B.T.; Sprague-Klein, E.A.; Roisnel, T.; Blart, E.; Gourlaouen, C.; Chen, L.X.; Pellegrin, Y. Bulky and Stable Copper(I)-Phenanthroline Complex: Impact of Steric Strain and Symmetry on the Excited-State Properties. Inorg. Chem. 2022, 61, 7296–7307. [Google Scholar] [CrossRef]
Scheme 1. Schematic structures of complexes 1 (R=H, 1H; CH3, 1Me; tertio-butyl, 1tBu) 2 and 3.
Scheme 1. Schematic structures of complexes 1 (R=H, 1H; CH3, 1Me; tertio-butyl, 1tBu) 2 and 3.
Materials 15 05222 sch001
Scheme 2. Schematic view of the molecules with the key nuclear coordinates, with X1 and X2 being the center of the phenyl core.
Scheme 2. Schematic view of the molecules with the key nuclear coordinates, with X1 and X2 being the center of the phenyl core.
Materials 15 05222 sch002
Scheme 3. Schematic structure (only the tBu groups are represented), name and energies computed with a GAUSSIAN of the different conformers of 1tBu in the ground state (black values) and in the lowest triplet state (red values).
Scheme 3. Schematic structure (only the tBu groups are represented), name and energies computed with a GAUSSIAN of the different conformers of 1tBu in the ground state (black values) and in the lowest triplet state (red values).
Materials 15 05222 sch003
Figure 1. Calculated TD-DFT absorption spectra without spin–orbit.
Figure 1. Calculated TD-DFT absorption spectra without spin–orbit.
Materials 15 05222 g001
Figure 2. TheoDORE analysis of the lowest absorbing singlet states of complexes 1H, 1Me, 1tBu(δ), 2 and 3. In black are metal-centered transitions; in red, ligand-centered transitions; in blue, metal-to-ligand charge transfer; in yellow, ligand-to-metal charge transfer and in green, ligand-to-ligand charge transfer. The state numbers corresponding to main transitions (Table 1 and Table 2) are in bold.
Figure 2. TheoDORE analysis of the lowest absorbing singlet states of complexes 1H, 1Me, 1tBu(δ), 2 and 3. In black are metal-centered transitions; in red, ligand-centered transitions; in blue, metal-to-ligand charge transfer; in yellow, ligand-to-metal charge transfer and in green, ligand-to-ligand charge transfer. The state numbers corresponding to main transitions (Table 1 and Table 2) are in bold.
Materials 15 05222 g002
Figure 3. TheoDORE analysis of the lowest triplet states at Franck-Condon geometry of complexes 1H, 1Me, 1tBu(δ), 2 and 3. In black are metal-centered transitions; in red, ligand-centered transitions; in blue, metal-to-ligand charge transfer; in yellow, ligand-to-metal charge transfer and in green, ligand-to-ligand charge transfer.
Figure 3. TheoDORE analysis of the lowest triplet states at Franck-Condon geometry of complexes 1H, 1Me, 1tBu(δ), 2 and 3. In black are metal-centered transitions; in red, ligand-centered transitions; in blue, metal-to-ligand charge transfer; in yellow, ligand-to-metal charge transfer and in green, ligand-to-ligand charge transfer.
Materials 15 05222 g003
Figure 4. One-dimensional cuts of the PES associated with the low-lying singlet and triplet excited states of 1H as a function of the breathing mode (Cu–X) (a), the angular distortion (X–Cu–X) (b) and the flattening mode (c), as defined in Scheme 2.
Figure 4. One-dimensional cuts of the PES associated with the low-lying singlet and triplet excited states of 1H as a function of the breathing mode (Cu–X) (a), the angular distortion (X–Cu–X) (b) and the flattening mode (c), as defined in Scheme 2.
Materials 15 05222 g004aMaterials 15 05222 g004b
Figure 5. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for 1H. Electron density-enriched and -depleted areas are in green and red, respectively.
Figure 5. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for 1H. Electron density-enriched and -depleted areas are in green and red, respectively.
Materials 15 05222 g005
Figure 6. Evolution of the singlet–triplet energy gaps (a) and SOC (b) of 1H as a function of the breathing mode (Cu–X), the angular distortion (X–Cu–X) and the flattening mode, as defined in Scheme 2. The discontinuity on the SOC values for the dihedral angles is due to the crossing between S1 and S2 around 25°.
Figure 6. Evolution of the singlet–triplet energy gaps (a) and SOC (b) of 1H as a function of the breathing mode (Cu–X), the angular distortion (X–Cu–X) and the flattening mode, as defined in Scheme 2. The discontinuity on the SOC values for the dihedral angles is due to the crossing between S1 and S2 around 25°.
Materials 15 05222 g006aMaterials 15 05222 g006b
Figure 7. Evolution of the E1, E2 and E3 spin–orbit state oscillator strengths of the three components of the T1 state of 1H as a function of the breathing mode (a), the angular distortion (b) and the flattening mode (c), as defined in Scheme 2.
Figure 7. Evolution of the E1, E2 and E3 spin–orbit state oscillator strengths of the three components of the T1 state of 1H as a function of the breathing mode (a), the angular distortion (b) and the flattening mode (c), as defined in Scheme 2.
Materials 15 05222 g007aMaterials 15 05222 g007b
Figure 8. One-dimensional cuts of the PES associated with the low-lying singlet and triplet excited states (a), evolution of the singlet–triplet energy gaps (b) and SOC (c) of 1tBu as a function of the breathing mode (Cu–X), as defined in Scheme 2.
Figure 8. One-dimensional cuts of the PES associated with the low-lying singlet and triplet excited states (a), evolution of the singlet–triplet energy gaps (b) and SOC (c) of 1tBu as a function of the breathing mode (Cu–X), as defined in Scheme 2.
Materials 15 05222 g008aMaterials 15 05222 g008b
Figure 9. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for 1tBu. Electron density enriched and depleted areas are in green and red, respectively.
Figure 9. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for 1tBu. Electron density enriched and depleted areas are in green and red, respectively.
Materials 15 05222 g009
Figure 10. Optimized structure of 2, hydrogen atoms are omitted for clarity.
Figure 10. Optimized structure of 2, hydrogen atoms are omitted for clarity.
Materials 15 05222 g010
Figure 11. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for complex 3. Electron density enriched and depleted areas are in green and red, respectively.
Figure 11. Electron density difference maps for the vertical transition S1 → S0, computed at the S1 minima for complex 3. Electron density enriched and depleted areas are in green and red, respectively.
Materials 15 05222 g011
Table 1. S0 to Sn transition energies (ADF) Eabs (eV), absorption wavelengths λabs (in nm) and oscillator strengths (f) of [Cu(I) (phen-R2)2]+ complexes 1 (1H, 1Me, and δ conformation of 1tbu).
Table 1. S0 to Sn transition energies (ADF) Eabs (eV), absorption wavelengths λabs (in nm) and oscillator strengths (f) of [Cu(I) (phen-R2)2]+ complexes 1 (1H, 1Me, and δ conformation of 1tbu).
State1H1Me1tBu (δ)
EabsλabsfEabsλabsfEabsλabsf
S12.3705237.91 × 10−132.3615251.17 × 10−132.4195131.30 × 10−3
S22.4275115.49 × 10−112.4215122.71 × 10−112.4655039.46 × 10−5
S32.5364893.35 × 10−12.5314903.26 × 10−12.5344891.69 × 10−1
S42.6994598.94 × 10−42.7754472.20 × 10−32.7784467.14 × 10−3
S52.6994598.77 × 10−42.7904442.31 × 10−32.8104412.34 × 10−3
S62.7894454.48 × 10−22.8734315.25 × 10−22.8644333.91 × 10−3
S72.8004434.24 × 10−22.8734315.25 × 10−22.8954283.24 × 10−2
S83.0314092.21 × 10−63.0294091.54 × 10−62.9124261.40 × 10−2
S93.1014001.19 × 10−33.1014001.16 × 10−32.9154252.67 × 10−3
S103.1014001.20 × 10−33.1213971.16 × 10−32.9484216.70 × 10−3
S113.2653802.96 × 10−53.2873776.67 × 10−53.1573939.42 × 10−6
S123.2763782.68 × 10−53.2983766.74 × 10−53.1703911.82 × 10−3
S133.3793671.27 × 10−43.4013651.26 × 10−43.2123866.36 × 10−4
S143.3793671.26 × 10−43.4013651.37 × 10−43.2133866.01 × 10−4
S153.4543591.51 × 10−143.5263527.98 × 10−133.3293723.68 × 10−4
S163.4623583.21 × 10−123.5543496.72 × 10−133.3423717.34 × 10−4
S173.6283424.70 × 10−23.7223335.69 × 10−23.5813462.71 × 10−2
S183.6803379.90 × 10−73.7933271.12 × 10−63.6243424.93 × 10−4
S193.7533304.65 × 10−93.8293243.72 × 10−113.6253421.04 × 10−3
S203.7713292.53 × 10−143.8573213.58 × 10−133.6423404.41 × 10−5
S213.8123251.14 × 10−123.8953182.63 × 10−133.8413232.03 × 10−3
S223.8183251.38 × 10−93.8953181.03 × 10−103.8613211.26 × 10−5
S233.8753204.99 × 10−23.9293164.04 × 10−23.9173173.12 × 10−2
S244.0123093.19 × 10−73.9583131.53 × 10−73.9663133.84 × 10−3
S254.0793041.48 × 10−24.0273082.27 × 10−24.0093093.50 × 10−2
Table 2. S0 to Sn transition energies (ADF) Eabs (eV), absorption wavelengths λabs (in nm) and oscillator strengths (f) of [Cu(I) (phen-R2)2]+ complexes 2 and 3.
Table 2. S0 to Sn transition energies (ADF) Eabs (eV), absorption wavelengths λabs (in nm) and oscillator strengths (f) of [Cu(I) (phen-R2)2]+ complexes 2 and 3.
State 2 3
EabsλabsfEabsλabsf
S12.3815211.63 × 10−62.4485061.69 × 10−4
S22.4345093.08 × 10−62.6594666.74 × 10−2
S32.5154931.65 × 10−12.6694651.32 × 10−1
S42.7304542.20 × 10−42.8484356.53 × 10−4
S52.7674489.72 × 10−52.9684189.83 × 10−3
S62.7824462.33 × 10−22.9904152.31 × 10−2
S72.8564346.03 × 10−33.0134117.92 × 10−3
S82.8774313.02 × 10−23.0294092.31 × 10−2
S92.8864302.08 × 10−43.1253971.21 × 10−3
S102.9944142.01 × 10−33.2303841.92 × 10−3
S113.0934011.99 × 10−43.2973763.07 × 10−3
S123.1163981.25 × 10−53.3203731.98 × 10−5
S133.1743911.66 × 10−33.4393608.82 × 10−4
S143.2123861.03 × 10−33.4693573.24 × 10−3
S153.2993762.67 × 10−63.5013541.62 × 10−3
S163.3383711.75 × 10−73.6053441.83 × 10−4
S173.5153531.08 × 10−63.7283336.97 × 10−4
S183.5443502.26 × 10−23.7593302.34 × 10−2
S193.5513492.97 × 10−73.8263242.27 × 10−5
S203.6183432.38 × 10−43.8443232.48 × 10−3
S213.7053356.04 × 10−73.9043181.04 × 10−4
S223.8793203.74 × 10−23.9363155.64 × 10−3
S233.8923198.98 × 10−63.9623133.21 × 10−4
S243.9303157.24 × 10−33.9923111.78 × 10−2
S253.9933101.31 × 10−24.0293085.14 × 10−2
Table 3. Spin–orbit excited states: transition energies (in eV), oscillator strengths, mixing in terms of Sn/Tn states of complexes 1H, 1Me, 1tBu (δ).
Table 3. Spin–orbit excited states: transition energies (in eV), oscillator strengths, mixing in terms of Sn/Tn states of complexes 1H, 1Me, 1tBu (δ).
State1H1Me1tBu
EabsfMixingEabsfMixingEabsfMixing
E12.2414.16 × 10−6T1 83%
T4 16%
2.2336.36 × 10−6T1 81%
T4 18%
2.2623.35 × 10−6T1 85%
T4 9%
T2 4%
E22.2414.26 × 10−6T1 83%
T4 16%
2.2336.31 × 10−6T1 81%
T4 18%
2.2632.38 × 10−5T1 85%
T4 9%
T2 4%
E32.2483.26 × 10−6T2 61%
T3 38%
2.2354.81 × 10−6T2 68%
T3 31%
2.2693.09 × 10−3T1 95%
S3 2%
E42.2483.29 × 10−6T2 61%
T3 38%
2.2354.79 × 10−6T2 68%
T3 31%
2.2901.30 × 10−5T2 65%
T3 28%
E52.2515.76 × 10−3T1 97%
S3 2%
2.2435.29 × 10−3T1 98%
S3 2%
2.2914.80 × 10−5T2 65%
T3 29%
T1 3%
E62.2651.56 × 10−8T2 89%
S1 10%
2.2491.29 × 10−9T2 91%
S1 9%
2.3102.11 × 10−5T2 98%
E72.2923.43 × 10−7T3 99%2.2862.29 × 10−9T3 99%2.3311.17 × 10−3T3 82%
S1 16%
E82.3031.35 × 10−11T4 90%
S2 9%
2.2891.09 × 10−11T4 92%
S2 8%
2.3521.25 × 10−5T4 86%
S2 12%
E92.3226.05 × 10−7T3 61%
T2 38%
2.3115.98 × 10−7T3 68%
T2 31%
2.3681.09 × 10−5T3 55%
T2 27%
T4 13%
T1 3%
E102.3226.37 × 10−7T3 61%
T2 38%
2.3116.40 × 10−7T3 68%
T2 31%
2.3681.13 × 10−5T3 58%
T2 27%
T4 5%
T1 3%
E112.3291.81 × 10−6T4 82%
T1 16%
2.3152.18 × 10−6T4 80%
T1 18%
2.3809.55 × 10−6T4 77%
T3 11%
T1 8%
T2 2%
E122.3291.84 × 10−6T4 82%
T1 16%
2.3152.24 × 10−6T4 80%
T1 18%
2.3811.57 × 10−5T4 75%
T3 13%
T1 7%
T2 3%
E132.3725.45 × 10−8S1 89%
T2 10%
2.3613.09 × 10−11S1 90%
T2 9%
2.4221.18 × 10−3S1 81%
T3 15%
T14 1%
E142.4285.05 × 10−11S2 90%
T4 9%
2.4216.16 × 10−11S2 91%
T4 7%
2.4658.10 × 10−5S2 85%
T4 12%
T13 1%
E152.5293.23 × 10−1S3 97%
T1 2%
2.5243.15 × 10−1S3 97%
T1 2%
2.5231.60 × 10−1S3 95%
T1 2%
E162.6175.87 × 10−6T5 71%
T7 26%
2.6595.21 × 10−9T5 49%
T6 49%
2.7186.82 × 10−6T5 72%
T7 15%
T6 9%
E172.6171.06 × 10−4T5 71%
T7 26%
2.6591.41 × 10−11T5 49%
T6 49%
2.7188.35 × 10−6T5 72%
T7 16%
T6 9%
E182.6245.84 × 10−4T5 80%
S4 11%
T6 4%
S5 4%
2.6794.66 × 10−4T6 81%
S4 10%
T5 8%
2.7281.46 × 10−3T5 90%
S7 4%
S5 3%
E192.6371.43 × 10−7T6 70%
T8 25%
T7 3%
2.6804.43 × 10−4T5 83%
T6 10%
S5 8%
2.7369.80 × 10−6T6 68%
T8 20%
T5 6%
T7 4%
E202.6371.46 × 10−4T6 70%
T8 25%
T7 3%
2.7001.40 × 10−11T7 35%
T8 35%
T6 14%
T5 14%
2.7362.76 × 10−4T6 68%
T8 20%
T7 5%
T5 5%
Table 4. Spin–orbit excited states: transition energies (in eV), oscillator strengths, mixing in terms of Sn/Tn states of complexes 2 and 3.
Table 4. Spin–orbit excited states: transition energies (in eV), oscillator strengths, mixing in terms of Sn/Tn states of complexes 2 and 3.
State 2 3
EabsfMixingEabsfMixing
E12.2231.26 × 10−6T1 56%
T2 38%
T3 3%
2.2902.03 × 10−6T1 92%
T2 5%
E22.2232.24 × 10−5T1 56%
T2 38%
T3 3%
2.2901.07 × 10−5T1 92%
T2 6%
E32.2431.26 × 10−7T1 92%
S1 6%
2.2978.58 × 10−4T1 98%
E42.2563.29 × 10−3T2 97%
S3 2%
2.3938.39 × 10−5T2 47%
S1 41%
T2 10%
E52.2702.80 × 10−6T3 60%
T2 27%
T4 6%
T1 6%
2.4272.14 × 10−6T2 82%
T4 6%
T1 4%
T3 4%
E62.2701.25 × 10−5T3 59%
T2 28%
T4 6%
T1 6%
2.4284.79 × 10−5T2 80%
T3 6%
T4 5%
T1 5%
E72.2832.17 × 10−4T3 98%2.4582.19 × 10−3T3 44%
S1 26%
T2 26%
E82.2971.90 × 10−6T1 35%
T2 31%
T3 29%
T12 1%
2.4612.37 × 10−5T3 71%
T4 16%
T2 9%
E92.2984.42 × 10−6T1 36%
T2 32%
T3 29%
T12 1%
2.4624.17 × 10−5T3 71%
T4 16%
T2 10%
E102.3631.25 × 10−6S1 50%
T4 33%
S2 13%
T1 2%
2.4771.66 × 10−3T3 50%
S1 31%
T2 15%
S2 1%
E112.3901.41 × 10−6S1 41%
T4 29%
S2 24%
T1 4%
2.5142.32 × 10−3T4 96%
S2 1%
E122.4121.48 × 10−6T4 91%
T3 6%
T14 2%
2.5399.12 × 10−7T4 75%
T3 20%
E132.4132.54 × 10−5T4 91%
T3 6%
S13 1%
2.5401.16 × 10−5T4 75%
T3 21%
E142.4471.91 × 10−6S2 60%
T4 36%
T14 2%
2.6536.59 × 10−2S2 90%
S3 4%
E152.5031.56 × 10−1S3 95%
T2 2%
2.6621.22 × 10−1S3 91%
S2 4%
E162.6661.56 × 10−4T5 74%
T7 15%
T6 6%
T8 4%
2.7641.29 × 10−4T5 87%
T6 11%
E172.6662.11 × 10−8T5 73%
T7 15%
T6 6%
T8 4%
2.7647.16 × 10−6T5 87%
T6 11%
E182.6715.13 × 10−4T5 88%
S5 6%
T6 3%
S8 2%
2.7716.64 × 10−4T5 96%
S6 3%
E192.6819.91 × 10−6T6 57%
T7 22%
T5 15%
T8 5%
2.8134.31 × 10−4S4 66%
T6 33%
E202.6811.34 × 10−8T6 57%
T7 22%
T5 15%
T8 5%
2.8751.62 × 10−4T6 86%
T5 12%
Table 5. Spin–orbit coupling (in cm−1) between the low-lying excited states in 1H, 1tBu (δ), 2 and 3.
Table 5. Spin–orbit coupling (in cm−1) between the low-lying excited states in 1H, 1tBu (δ), 2 and 3.
1HS1S2S3S4T1T2T3T4
T10.00.0307.43.00.00.00.0390.4
T2275.70.00.017.30.00.0428.60.0
T30.00.00.011.20.0428.60.00.0
T40.0300.00.09.6390.40.00.00.0
1tBuS1S2S3S4T1T2T3T4
T1114.90.5298.421.30.0107.11.1423.7
T21.357.90.8171.1107.10.0441.52.2
T3289.23.1118.558.91.1441.50.088.3
T40.8322.44.125.1423.72.288.30.0
2S1S2S3S4T1T2T3T4
T1282.2139.90.27.00.0392.7201.90.4
T20.30.4302.15.6392.70.00.5183.7
T30.1302.176.59.9201.90.50.0399.3
T4129.85.60.522.40.4183.7399.30.0
3S1S2S3S4T1T2T3T4
T17.9159.0221.620.50.0425.3176.521.4
T2316.865.348.923.0425.30.024.8206.9
T397.4248.0210.65.4176.524.80.0411.0
T48.6149.2124.810.221.4206.9411.00.0
Table 6. Deformation, emission and stabilization energies (in eV), emission wavelengths (in nm), singlet–triplet energy gaps (in eV), SOC (in cm−1) and oscillator strengths at the S1 and T1 optimized structures of 1H, 1Me and 1tBu (δ).
Table 6. Deformation, emission and stabilization energies (in eV), emission wavelengths (in nm), singlet–triplet energy gaps (in eV), SOC (in cm−1) and oscillator strengths at the S1 and T1 optimized structures of 1H, 1Me and 1tBu (δ).
State 1H1Me1tBu (δ)
S1Edef0.8200.4420.253
Eem0.7071.5691.813
Estab1.5272.0112.066
λem1753790684
fosc2.25 × 10−51.77 × 10−22.54 × 10−4
T1Edef0.6100.4430.307
Eem0.7041.2461.656
Estab1.3141.6891.964
λem1761995749
ΔEST0.4640.4110.143
SOC S1–T125.554.220.1
fosc2.28 × 10−92.01 × 10−96.85 × 10−6
7.01 × 10−95.38 × 10−82.77 × 10−5
2.34 × 10−61.17 × 10−41.77 × 10−4
Table 7. Symmetry, deformation (Edef), emission (Eem) and stabilization (Estab) of energies in eV and the relative stability of the different conformers (ΔE in eV), emission wavelength (λem in nm), S1–T1 splitting (ΔEST in eV), S1–T1 SOC (in cm−1) and oscillator strength at S1 and T1 optimized geometries for 1tBu.
Table 7. Symmetry, deformation (Edef), emission (Eem) and stabilization (Estab) of energies in eV and the relative stability of the different conformers (ΔE in eV), emission wavelength (λem in nm), S1–T1 splitting (ΔEST in eV), S1–T1 SOC (in cm−1) and oscillator strength at S1 and T1 optimized geometries for 1tBu.
StateConformerαβγδεζ
GS SymmetryD2dD2dC1C2C1C2v
S1S1 SymmetryC2C2C1C1C1C2
Edef0.5420.4970.2790.2860.3760.249
Eem1.9441.7791.8701.8131.8011.884
Estab2.4862.2752.1492.0992.1772.133
ΔE0.3870.1760.0500.0000.0780.033
λem638697663684688658
fosc1.06 × 10−51.57 × 10−61.53 × 10−52.54 × 10−42.53 × 10−51.59 × 10−8
T1T1 SymmetryC2C2C1C1C1C2
Edef0.5850.5130.2950.3410.3970.270
Eem1.7451.6331.7151.6561.6481.726
Estab2.3312.1462.0111.9972.0461.997
ΔE0.3340.1490.0140.0000.0490.000
λem710759723749752718
ΔEST0.1880.1330.1440.1430.1380.140
SOC S1–T173.06.03.620.18.30.0
3.04 × 10−62.10 × 10−97.05 × 10−76.85 × 10−64.25 × 10−61.09 × 10−13
fosc8.97 × 10−66.86 × 10−64.69 × 10−52.77 × 10−51.42 × 10−56.25 × 10−6
3.04 × 10−43.01 × 10−42.81 × 10−41.77 × 10−42.46 × 10−42.88 × 10−4
Table 8. Deformation, emission and stabilization energies (in eV), emission wavelengths (in nm), singlet-triplet energy gaps (in eV), SOC (in cm−1) and oscillator strengths at the S1- and T1-optimized structures of 2 and 3.
Table 8. Deformation, emission and stabilization energies (in eV), emission wavelengths (in nm), singlet-triplet energy gaps (in eV), SOC (in cm−1) and oscillator strengths at the S1- and T1-optimized structures of 2 and 3.
State 2233
S1Edef0.2270.2410.2570.343
Eem1.7831.7971.8811.893
Estab2.0092.0372.1372.235
λem695690659655
fosc0.000.005.43 × 10−55.73 × 10−6
T1Edef0.2430.260.2770.365
Eem1.6341.6431.7171.733
Estab1.8771.9031.9952.098
λem759754722715
ΔEST0.1350.1380.1460.143
SOC S1–T10.00.07.72.9
fosc9.96 × 10−81.04 × 10−72.37 × 10−71.28 × 10−8
4.81 × 10−61.07 × 10−55.68 × 10−63.87 × 10−6
3.03 × 10−41.39 × 10−43.10 × 10−41.51 × 10−4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gourlaouen, C.; Daniel, C. Towards Optimized Photoluminescent Copper(I) Phenanthroline-Functionalized Complexes: Control of the Photophysics by Symmetry-Breaking and Spin–Orbit Coupling. Materials 2022, 15, 5222. https://doi.org/10.3390/ma15155222

AMA Style

Gourlaouen C, Daniel C. Towards Optimized Photoluminescent Copper(I) Phenanthroline-Functionalized Complexes: Control of the Photophysics by Symmetry-Breaking and Spin–Orbit Coupling. Materials. 2022; 15(15):5222. https://doi.org/10.3390/ma15155222

Chicago/Turabian Style

Gourlaouen, Christophe, and Chantal Daniel. 2022. "Towards Optimized Photoluminescent Copper(I) Phenanthroline-Functionalized Complexes: Control of the Photophysics by Symmetry-Breaking and Spin–Orbit Coupling" Materials 15, no. 15: 5222. https://doi.org/10.3390/ma15155222

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop