Next Article in Journal
New Oxaliplatin-Pyrophosphato Analogs with Improved In Vitro Cytotoxicity
Next Article in Special Issue
Development of Amino Acids Functionalized SBA-15 for the Improvement of Protein Adsorption
Previous Article in Journal
Effect of Pyrolysis Temperature on the Characterisation of Dissolved Organic Matter from Pyroligneous Acid
Previous Article in Special Issue
Recent Advances in the Synthesis of Polymer-Grafted Low-K and High-K Nanoparticles for Dielectric and Electronic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

P∩N Bridged Cu(I) Dimers Featuring Both TADF and Phosphorescence. From Overview towards Detailed Case Study of the Excited Singlet and Triplet States †

1
Institut für Physikalische Chemie, Universität Regensburg, D-93053 Regensburg, Germany
2
Institut Lumière Matière, Univ Lyon, Université Claude Bernard Lyon 1, CNRS, F-69622 Villeurbanne, France
3
Institut für Mineralogie und Kristallographie, Geozentrum—Universität Wien, Althanstr. 9, A-1090 Wien, Austria
4
School of Education, Chemistry, Johannes Kepler University Linz, Altenbergerstr. 69, A-4040 Linz, Austria
*
Authors to whom correspondence should be addressed.
This publication is dedicated to Prof. Dr. Wolfgang Kaim on the occasion of his 70th birthday.
Molecules 2021, 26(11), 3415; https://doi.org/10.3390/molecules26113415
Submission received: 10 April 2021 / Revised: 23 May 2021 / Accepted: 25 May 2021 / Published: 4 June 2021
(This article belongs to the Special Issue 25th Anniversary of Molecules—Recent Advances in Applied Chemistry)

Abstract

:
We present an overview over eight brightly luminescent Cu(I) dimers of the type Cu2X2(P∩N)3 with X = Cl, Br, I and P∩N = 2-diphenylphosphino-pyridine (Ph2Ppy), 2-diphenylphosphino-pyrimidine (Ph2Ppym), 1-diphenylphosphino-isoquinoline (Ph2Piqn) including three new crystal structures (Cu2Br2(Ph2Ppy)3 1-Br, Cu2I2(Ph2Ppym)3 2-I and Cu2I2(Ph2Piqn)3 3-I). However, we mainly focus on their photo-luminescence properties. All compounds exhibit combined thermally activated delayed fluorescence (TADF) and phosphorescence at ambient temperature. Emission color, decay time and quantum yield vary over large ranges. For deeper characterization, we select Cu2I2(Ph2Ppy)3, 1-I, showing a quantum yield of 81%. DFT and SOC-TDDFT calculations provide insight into the electronic structures of the singlet S1 and triplet T1 states. Both stem from metal+iodide-to-ligand charge transfer transitions. Evaluation of the emission decay dynamics, measured from 1.2 ≤ T ≤ 300 K, gives ∆E(S1-T1) = 380 cm−1 (47 meV), a transition rate of k(S1→S0) = 2.25 × 106 s−1 (445 ns), T1 zero-field splittings, transition rates from the triplet substates and spin-lattice relaxation times. We also discuss the interplay of S1-TADF and T1-phosphorescence. The combined emission paths shorten the overall decay time. For OLED applications, utilization of both singlet and triplet harvesting can be highly favorable for improvement of the device performance.

1. Introduction

Potential applications of luminescent materials in organic light emitting diodes (OLEDs) have strongly stimulated the development of emitters that are suited for exploiting all singlet (25%) and triplet (75%) excitons [1] generated in the emission layer. Essentially, there are two different mechanisms that are already used for harvesting 100% of the excitons, namely the triplet harvesting mechanism [2,3,4,5,6] and the singlet harvesting mechanism [7,8,9,10,11,12]. For light generation based on the triplet harvesting mechanism, brightly and fast phosphorescent emitting compounds are required, such as Ir(III) or Pt(II) complexes [2,3,4,5,6,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28]. While for singlet harvesting, the molecules have to show efficient thermally activated delayed fluorescence (TADF) at ambient temperature. Examples are found among Cu(I), Ag(I), Au(III), W(VI) and Zn(II) complexes [7,9,10,11,12,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65] or specifically designed organic molecules [66,67,68,69,70,71,72,73,74,75,76,77,78,79]. For completeness, it is also referred to a very recently proposed mechanism, the direct singlet harvesting (DSH) mechanism [80,81]. This strategy, being based on compounds with very small energy gap of ∆E(S1-T1) ≪ kBT (kB = Boltzmann constant), allows also for 100% exciton harvesting and additionally for drastic reduction of the emission decay time. In this report, however, we want to focus on complexes that exhibit at ambient temperature both thermally activated delayed fluorescence (TADF) and phosphorescence. Thus, they may be regarded as singlet and triplet harvesting materials. This effect of combined TADF-phosphorescence emission, suited for decreasing the overall emission decay time, has already been addressed in the literature [50,51,82,83,84,85,86,87,88,89]. In particular, Cu(I) and Ag(I) dimers, in which the metal centers are linked by P∩N ligands can show this effect [34,50,87,90,91,92,93,94,95] and very probably also the complexes reported recently [96]. To evaluate the class of materials of Cu2X2(P∩N)3 complexes (with X = Cl, Br, I), we study eight compounds with respect to their crystal structures and, especially, their emission properties at T = 300 K and 77 K, respectively. These data allow us to select one prominent material, Cu2I2(P∩N)3, 1-I with P∩N = 2-diphenylphosphino-pyridine (Ph2Ppy), that shows distinct TADF and phosphorescence at ambient temperature even at high emission quantum yield of ΦPL = 81% at relatively short decay time. Therefore, we investigate the emission behavior of this material as a detailed case study over the large temperature range of 1.2 ≤ T ≤ 300 K. Thus, we obtain deep insight into properties of the lowest triplet state T1 including its zero-field splitting (ZFS), spin-lattice relaxation (SLR) dynamics and we determine the TADF activation energy gap ∆E(S1-T1) between the lowest excited singlet state S1 and the T1 state, as well as the S1→S0 fluorescence rate. The experimental results are largely supported by SOC-TDDFT computations. Indeed, as will be shown, the combined TADF-phosphorescence decay time is distinctly shorter than the TADF-only decay. This is due to the relatively fast phosphorescence rate and the small ∆E(S1-T1) gap. Potentially, such materials showing both singlet and triplet harvesting character are attractive for applications that require short photoluminescence decay.

2. Results and Discussion

2.1. Syntheses and Structural Characterization

Previously, we have reported on the preparation of complexes 1-X (X = Cl, Br, I) and 3-I (Scheme 1) [34]. In the present work, we additionally report on the new complexes 2-X (with X = Cl, Br, I) and 3-Br. They were prepared analogously by reactions of the respective ligands 2 and 3 [97] with copper(I) halides in dichloromethane. For ligand 3, the pure chloride complex could not be obtained. 1-X and 2-X give yellow, while 3-X red powders. Once precipitated, the complexes are only slightly soluble in standard, weakly or non-coordinating solvents, hence, no NMR spectra could be measured. Therefore, the complexes were characterized exclusively by elemental analysis. In addition, for complexes 1-Br, 2-I and 3-I the crystal structures could be determined.
Single crystals suitable for X-ray diffraction could be obtained by slow gas-phase diffusion of diethyl ether into the filtered reaction solution of 1-Br, 2-I and 3-I. To our knowledge, no bromide of this class of di-nuclear copper compounds has been structurally characterized. Therefore, the structure of 1-Br provides the first structural data and completes the series of accessible halides. Together with the crystal structures of 1-Cl and 1-I reported in reference [34], it is now possible to compare all three halides of a homologous series of complexes of the type Cu2X2(P∩N)3. All so far structurally characterized complexes of this type have in common a butterfly-shaped Cu2X2 core surrounded by three P∩N ligands (Figure 1). Two ligands coordinate exclusively via the phosphorus atom to Cu(I), while the third ligand is bound in a bridging manner with both the nitrogen and phosphorus atoms to two Cu(I) centers.
Interestingly, not all halides 1-X are isostructural (Table 6 below and [34]): 1-Br/Cl crystallize in the monoclinic space group P21/n, whereas 1-I is triclinic P 1 ¯ . Nevertheless, all structures are similar and show the expected trend for a series chloride → iodide, with only very small differences between 1-Cl and 1-Br (Table 1). For example, the Cu–Cu distances are almost identical for 1-Cl (2.878(1) Å) and 1-Br (2.883(1) Å), whereas 1-I features a shorter Cu–Cu distances of 2.7694(5) Å. On first sight, it seems a contradictory observation that the biggest anion causes the smallest intermetallic separation. However, this fact is due the most acute angles Cu1–X–Cu2 of around 62 °C for 1-I, whereas for 1-Cl, we find around 73°C. The respective angle for the bromide 1-Br lies between these values (≈69 °C). Together with a significant longer Cu–X bond distance (Cu–Cl: 2.39–2.44; Cu–Br: 2.51–2.57; Cu–I: 2.65–2.81 Å) this leads to the shortest Cu–Cu distance for 1-I. All Cu–Cu distances are around or slightly above the sum of the van-der-Waals radii (r(Cu) = 1.40 Å [98]) indicative of only neglectable cuprophilic interactions.
Although the isoelectronic pyridyl and pyrimidyl moieties are expected to have similar steric characteristics, the structure of 2-I (monoclinic) is not isostructural to 1-I (triclinic) [34]. 3-I crystallizes as the dichloromethane solvate. It should be noted that another form of complex 3-I has been reported previously which does not differ considerably in its structural parameters (compare Table 1) [34]. For all halides and P∩N ligands summarized in Table 1, the Cu–P and Cu–N distances lie in a very narrow range of 2.24–2.29 Å and 2.08–2.14 Å, respectively.

2.2. Computational Investigations

An overview of the emission data, as presented below, shows that the compound Cu2I2(Ph2Ppy)3, 1-I (Figure 2) exhibits particularly interesting properties with respect to a distinct combination of TADF and phosphorescence and shows high emission quantum yield at short emission decay time. Accordingly, we will discuss this compound’s photophysical properties in deeper detail below. In this section, we first investigate this material by computational methods to shed light on the electronic properties.
Initially, the compound was optimized in the singlet ground state. The calculated results reveal good agreement with the data from the X-ray structure (Table 2). Although the Cu–Cu bond length is slightly overestimated, the general structural motifs like the butterfly shape of the Cu2I2 core with an equilateral CuI2 triangle are nicely reproduced. Further geometry optimization was performed in the lowest triplet state using unrestricted DFT. In a next step, time-dependent DFT (TDDFT) calculations with and without self-consistent spin-orbit coupling using the ZORA Hamiltonian [99] (SOC-TDDFT [100]) were carried out for the S0 and T1 geometries with ADF2014 [101] solving for the lowest six spin-mixed excitations. Due to computational constraints, the basis set for these calculations was chosen to be DZP, a double-zeta plus polarization basis set [102]. The TDDFT calculations at the S0 geometry show that the S1 and T1 states correspond to a nearly pure excitation (98% and 97%) from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) as similarly reported in [34]. As depicted in Figure 3, the HOMO is localized on the Cu2I2 core with significant contributions of the iodides, while the LUMO is localized on the bridging organic ligand. This leads to a classification of the HOMO-LUMO transition as being of (I+M)LCT character, abbreviated shortly as metal-to-ligand charge transfer (MLCT) transition.
Taking spin-orbit coupling (SOC) into account, also the energetic separations between the three triplet substates, the zero field splittings (ZFSs) and the radiative lifetimes of these states can be assessed (Table 3). At the S0 geometry, a moderate agreement with the experimental data is observable. The calculated vertical excitation energies are somewhat underestimated and the small computed singlet-triplet splitting, although predicting 1-I as a TADF material, underestimate the experimental ∆E(S1-T1) gap. Clearly, several important aspects are not taken into account at this level of theory. First, the inclusion of the difference in zero-point energies of ground and excited state may alter transition energies by 0.1–0.2 eV, usually leading to red shifted emission energies. Second, one generally expects excited state geometry relaxation to occur [103,104,105]. To investigate its impact, simulations were also carried out at the optimized T1 state geometry. As Table 3 shows, this leads to a strongly red shifted emission, probably being significantly influenced by the strong shortening of the Cu–Cu bond length (Table 2). Experimentally, a strong dependence of the emission wavelength on the molecular environment was observed. For example, related Cu2I2(P∩N)3 [(P∩N) = (2-diphenylphosphino)-4-alkyl-pyridine] complexes, as presented in [34], show strong red shifts going from the powder material to PMMA (poly(methyl methacrylate)) matrices and, finally, to solvents (also compare [37]). In the latter environments, large scale atomic rearrangements can easily occur. We find that the computed bond length reduction at the T1 optimized geometry is accompanied by a 30 °C rotation of the P∩N ligand around the Cu1–P2 axis (Table 2). While the former geometry change could also be realized in a polycrystalline environment, the latter one would likely be sterically hindered. Another influence is the effect of the dielectric environment, which probably is important due to the charge transfer character of the emission. Since a realistic modeling of the environment of the complex within the powder material is challenging, we performed SOC-TDDFT calculations with the COSMO continuum [106] solvent model (parameters: dielectric constant of dichloromethane (DCM) ε = 8.9, radius of solvent molecules r(sol) = 2.94 Å) for the S0 geometry to obtain information about the general trends. Estimates of the radiative lifetimes are based on transition energies and transition dipole matrix elements as outlined in [107]. Corrections due to the refractive index of the molecular environment (n(DCM) = 1.42) were accounted for by the empty spherical cavity model [108]. Table 3 reveals a strong blue shift of ≈0.4 eV (≈3000 cm−1) compared to a gas phase environment. The observed blue shift can probably be explained by the considerable dipole moment change by the charge transfer from the Cu2I2 core to the bridging ligand.
The calculated data, presented in Table 3, show considerably different energies, depending on the model and the state geometry chosen [109]. For comparison to emission data, we focus on the T1 state geometry. Although the calculated transition energies are underestimated (in gas phase environment), it is usually accepted that energy differences display the experimental situation more realistically. The corresponding emission data are presented below in Section 2.4 and in Figure 8. It is seen that the energy gap ∆E(S1-T1) is very well reproduced in the T1 state geometry, while the ZFS value ∆E(III–II,I) is overestimated. This might be related to the model restricted to a too small number of spin-state-mixing of higher lying states. However, the very small splitting of ∆E(II–I) or the almost degenerate situation for the two low-lying triplet substates I and II is well reproduced by the SOC-TDDFT model (Table 3 and Figure 8, below).

2.3. Luminescence Properties of Cu2X2(P∩N)3 Complexes with X = Cl, Br, I. An Overview

In this section, we will discuss the emission behavior of eight Cu2X2(P∩N)3 complexes studied at 300 K and 77 K and give an overview over the large scale of varying emission properties.
The complexes studied are not soluble in common, non-coordinating organic solvents. Therefore, all measurements were performed with powders. Usually, investigations of solid-state samples are not convenient for studies of detailed emission properties, due to the influence of processes like triplet-triplet annihilation or energy transfer. However, for Cu(I) complexes, usually a self-trapping mechanism takes place that leads to quasi-isolated molecules embedded in the neat material without any significant excited state resonance interaction with the environment [37,110,111].
As summarized in Table 4, all studied di-nuclear Cu2X2(P∩N)3 complexes show relatively intense photoluminescence under UV excitation with broad emission bands in the green to red spectral range assigned to (iodide+metal)-to-ligand charge transfer, (I+M)LCT transitions, as predicted by DFT calculations presented in the previous section. The HOMO resides on the Cu(I)-halide core, while the LUMO is largely localized on the bridging (P∩N) ligand (Figure 3). Thus, modification of the ligand, in particular, of an extension of the ligand’s aromatic system leads to a red shift of the MLCT transition [34]. There is a clear trend of the emission maxima depending on the halides. In each series, the emission maxima are blue shifted from Cl to Br to I complexes. For example, Figure 4 shows the emission spectra of the 1-X series. For ambient temperature, the peak maxima are blue shifted from 1-Cl to 1-I by about 1200 cm−1. The flank of the excitation spectra in the region above about 400 nm shows a similar trend. This observation is rationalized by a reduction of the ligand field strength of the halides in the series Cl > Br > I (being contrary to the trend of electronegativity) and thus, by a larger HOMO-LUMO energy gap [34,112]. The photoluminescence quantum yields at T = 300 K lie between ΦPL = 9% and 81%. At ambient temperature, the emission decay times are found in the lower µs time regime between 1.2 to 8.8 µs. However, with respect to photophysical interpretations, it is usually better to compare radiative decay times τr = τ/ΦPL. In the series of compounds given in Table 4, they lie between 4.7 × 104 s−1 (21 µs) and 1.25 × 105 s−1 (8 µs).
Upon cooling to 77 K, the emission quantum yields increase remarkably, for example, for compound 2-I by a factor of more than five. Such a behavior is not unusual, since non-radiative deactivation processes frequently become less important with temperature reduction [113]. Moreover, the emission bands of all complexes are red shifted. Most affected is the 1-X series with shifts of about 440 cm−1 (54 meV, 1-Cl and 1-I) and about 710 cm−1 (88 meV, 1-Br), while for the compounds 2-X and 3-X, values below around 300 cm−1 (37 meV) are observed. These energy separations frequently display approximately the energy gap ∆E(S1-T1) that is responsible for thermal activation of the TADF emission. Interestingly, the value of 440 cm−1 (54 meV) found for Cu2I2(Ph2Ppy)3, 1-I corresponds well to the calculated one of 411 cm−1 (51 meV, Table 3). Concomitantly, the radiative rates of the emission decrease strongly (increasing decay times) with cooling, for example, by a factor of about 16 for compound 2-Cl. These effects, red shift and radiative rate decrease, occurring upon cooling, are consequences of freezing out the additional emission decay channel via the energetically higher lying S1 state. In other words, TADF emission is largely frozen out. Thus, at sufficiently low temperature, mostly at T = 77 K but sometimes only below T = 50 K [29], all compounds exhibit only phosphorescence.
The spectral changes of the emission bands with temperature variation are shown in Figure 5a for compound 1-I. At low temperature (1.3 K to almost 77 K), only T1→S0 phosphorescence is occurring, while with further temperature increase to T = 300 K, the emission band center is blue shifted, mainly observable by a blue-side flank growing in above around T = 70 K (Figure 5b). The corresponding additional band is assigned to the thermally activated fluorescence from the higher lying S1 state. Apparently, already the spectral changes displayed in Figure 5 indicate that at ambient temperature, the emission consists of overlapping TADF and phosphorescence. The appearance of a phosphorescence contribution even at ambient temperature is additionally based on the relatively fast radiative rate of kr(T1→S0, 77 K). Such a behavior of combined TADF and phosphorescence is not very frequently observed [48], but see refs. [50,51,83,86,114]. Mostly, however, the TADF channel dominates strongly [48]. For a more detailed discussion see Section 2.4.
Of the series of compounds studied (Table 4), especially, the iodide containing complexes 1-I, 2-I and 3-I show fast radiative phosphorescence rates lying above kr(T1→S0) = 2 × 104 s−1 (50 µs). This may be rationalized by two factors. First, significant SOC is induced by admixtures of higher lying singlet states Sn to the T1 state. Secondly, the large contributions of 5p-orbitals of iodide to the higher lying occupied orbitals (Figure 3) with the high SOC constant of iodide of ξ(I) = 5069 cm−1 [115] may also play a role in speeding up the T1→S0 transition rate [116]. This latter effect is supported by the distinctly lower T1→S0 rate found for the chloride compounds with ξ(Cl) = 587 cm−1 [115] (Table 4).
A deeper discussion of the data summarized in Table 4 will not give more detailed information, in particular, not the spectral features available. This is a consequence of the very broad emission bands of MLCT character with halfwidths of ≈3500 cm−1 (440 meV) even at T = 1.3 K. On the other hand, investigation of the emission decay behavior with temperature variation will lead to a detailed characterization of the compounds’ electronic structure, especially, of the lowest excited states, as will be shown in the next section.

2.4. Detailed Case Study of Cu2I2(Ph2Ppy)3, 1-I. the Lowest Excited Triplet and Singlet States

For a deeper case study, we selected Cu2I2(Ph2Ppy)3, 1-I (powder) due to four reasons. First, this material exhibits the highest emission quantum yield of ΦPL = 81% of all compounds summarized in Table 4. Second, 1-I emits with one of the fastest radiative phosphorescence rate of kr(77 K) = 2.88 × 104 s−1r(77 K) = 35 µs). Third, the emission decay time is mono-exponential over the whole temperature range (apart from very low temperature) and fourth, the emission quantum yield changes only slightly with temperature. These two latter properties are required for a detailed characterization based on the τ(T) fitting procedure as discussed below. With respect to these properties, 1-I represents a remarkable material and is expected to show clearly the interplay of phosphorescence and TADF. Moreover, according to the fast rate, efficient SOC experienced by the T1 state should lead to a well observable zero-field splitting (ZFS) of T1 into substates I, II and III, as already predicted by the SOC-TDDFT calculations (Section 2.2). This should lead to specific relaxation properties within the manifold of the substates, in particular, such as effects of spin-lattice relaxation (SLR) [117].
For the detailed characterization of the compound’s electronic structure and the related decay rates, we study the emission decay behavior of 1-I over the large temperature range of 1.2 ≤ T ≤ 300 K. Figure 6 displays selected emission decay curves measured at different temperatures. At T = 1.2 K, the emission decays bi-exponentially with two clearly different components of 10 µs and 174 µs, respectively. Such a behavior can be well rationalized, if the lowest triplet state exhibits distinct ZFS. Thus, the short component refers to spin-lattice-relaxation (SLR) processes within the triplet substate manifold [117], while the longer component is ascribed to the thermalized emission that is established after around 60 µs at T = 1.2 K (estimated from the 1.2 K decay curve reproduced in Figure 6). It is known that thermalization according to SLR processes is strongly temperature dependent and becomes faster with temperature increase [117]. Indeed, the observed short decay component decreases from 10 µs at T = 1.2 K to 1.7 µs at T = 10 K and is faster than detectable (with our equipment) at T = 15 K (Figure 6). This means that SLR processes are getting much faster than the thermalized emission decay time. Therefore, shortly after the excitation pulse, the population numbers of the excited states behave according to a Boltzmann distribution. We will come back to SLR properties below in this section.
The decay time of the long (thermalized) component is also drastically shortened with temperature increase (Figure 7). Due to the photophysical mechanisms involved, the temperature dependence can be classified into three ranges: (i) Up to T ≈ 20 K, the decay behavior is determined by the zero-field split T1 substates. At T = 1.2 K, we find a long component of τ = 174 µs. This emission dominantly stems from the thermally equilibrated two lower triplet substates I and II. By temperature increase the decay is strongly shortened (by a factor of more than five) when reaching a plateau near 20 K. The plateau decay time amounts to about 32 µs. This decrease is resulting from population of the higher lying triplet substate III. Because this transition III(T1)→S0 exhibits a much faster rate than the rates from the two lower lying substates I and II, an additional decay path is opened. (ii) Within the temperature range of the plateau, between about 20 K and about 70 K, the triplet substates I, II and III are thermally equilibrated and emit as T1→S0 phosphorescence with an average decay time of the three substates of 32 µs (see also Equation (4), below). (iii) With further temperature increase from T ≈ 70 K to 300 K, the decay time decreases to τ(300 K) = 6.5 µs. In this temperature range, the photoluminescence quantum yield changes only slightly (Table 4). Hence, the decay time decrease can (largely) be ascribed to a rate increase by a factor of more than four from kr(77 K) = 2.88 × 104 s−1 to kr(300 K) = 12.5 × 104 s−1 (Table 4). This increase is induced by opening the TADF decay path via the S1→S0 transition in addition to the still remaining phosphorescence decay path (see also below).
For the mono-exponential ranges of the decay curves (Figure 6), it can be concluded that the emitting states are in fast thermal equilibration with respect to the individual decay times of the states involved. This is not only valid for the SLR processes but also for down- and up-inter-system crossing (ISC) processes between the T1 and the S1 states. Down-ISC will probably occur within around 10 ps or even shorter [103,104,105,118,119,120,121], while the up-ISC time (also named RISC time) is strongly temperature dependent [122] and may be estimated very roughly to around 30 ns at 80 K and to about 0.2 ns at ambient temperature. For completeness, it is mentioned that very fast ISC processes are probably based on direct SOC of higher lying singlet states to T1 substates as is displayed in a high allowedness of the T1→S0 transition. This is in contrast to the situation of molecules with weak SOC with respect to the lowest triplet state, as it seems to be valid for most organic molecules. For these, spin-vibronic processes will probably dominate the ISC rate [123].
Thus, with respect to the emission decay times of many µs, fast thermalization is well realized for the Cu(I) compounds discussed here. Accordingly, it is justified to describe the temperature dependence of the emission decay time τ(T), as shown in Figure 7, by a modified Boltzmann distribution of the four thermally equilibrated excited states, the three T1 substates I, II and III and the S1 state [19,50,87,124].
τ T = 1 + exp Δ E II I k B T + exp Δ E III I k B T + exp Δ E S 1 I k B T k I + k II exp Δ E II I k B T + k III exp Δ E III I k B T + k S 1 exp Δ E S 1 I k B T
Herein, ∆E(II–I) and ∆E(III–I) are the ZFS values and ∆E(S1–I) is the energy gap between the S1 state and the T1 state (for ∆E(S1–T1) >> ∆E(III–I)). k(I), k(II), k(III) and k(S1) are the transition rates of the respective states to the electronic ground state S0. kB is the Boltzmann constant.
If we assume constant ΦPL over the whole temperature range, as approximately justified (Table 4), very good fit of Equation (1) to the measured decay times is realized (Figure 7). As photophysical fit parameters we find ∆E(II–I) ≈ 0 cm−1 (meaning < 1 cm−1) and ∆E(III–I) = 3 cm−1 (0.37 meV). These values are smaller than the ones obtained from our SOC-TDDFT computations, but the splitting pattern is well reproduced (Table 3). For the manifold of the three triplet substates we expect further that for T < 1 K, a low-temperature plateau of the emission decay time τ(T) will be adapted near 210 µs. Although, the individual decay times of the substates I and II cannot be determined directly, we will roughly set τ(I) ≈ τ(II) ≈ 210 µs. This is not unreasonable, since any larger deviation from this approach would not fit to the set of experimental data and the fitting procedure. Moreover, from the fit, we obtain k(III) = 8.3 × 104 s−1, corresponding to formally 12 µs (compare Figure 8, below and Table 5). Since the III→S0 decay path is much faster than the I,II→S0 decays, state III involvement is already of importance at T = 1.2 K by reducing the decay time from ≈ 210 µs to 174 µs as found experimentally (Figure 6).
The short decay component of 10 µs observed at T = 1.2 K (Figure 6) is assigned to the direct mechanism of SLR [117] from substate III to both substates I and II. Using the fit data presented above, we can determine the SLR rate k(SLR) for these processes by
k(SLR) = k(obs) − k(III→S0)
Herein, k(obs) corresponds to the observed decay component of 10 µs and k(III→S0) is the rate of the III→S0 transition with 8.3 × 104 s−1. Accordingly, we find k(SLR) = 1.7 × 104 s−1 corresponding to τ(SLR) = 59 µs. This means, the SLR process is relatively slow. Values of similar size have, for example, been reported for Ir(III)complexes [125,126].
The fitting procedure gives also ∆E(S1-T1) = 380 cm−1 (47 meV), the energy gap that is crucial for TADF properties. The computed value for the T1 state geometry fits relatively well (411 cm−1, 51 meV, Table 3). Moreover, the rate of the S1→S0 fluorescence, also resulting from the fit, amounts to k(S1→S0) = 2.25 × 106 s−1 (445 ns, Figure 7 and Figure 8). However, the corresponding prompt fluorescence decay cannot be measured directly, since the S1–T1 ISC is orders of magnitude faster. It amounts to around 10 ps [103,104,105,118,119,120,121] as already discussed above. The computations lead to a value of 1/k(S1–S0) = τ(S1–S0) = τ(S1) = 1.5 µs for the prompt fluorescence (Table 3), thus, being overestimated by a factor of about three. Even the experimental S1→S0 transition rate is relatively slow for an “allowed” fluorescence, but this is in line with the distinct CT character of the transition. At small HOMO-LUMO overlap, one obtains small exchange interaction and hence, a small ∆E(S1-T1) gap (as required for TADF materials), however, as well as a slow S1→S0 transition rate [37,48]. For completeness, it is remarked that design of materials with faster S1→S0 transition rates and small gaps ∆E(S1-T1) is highly attractive and indeed, became possible recently [40,41,42,43,44,52].
Interestingly, the emission intensity at ambient temperature represents a combined phosphorescence and TADF. According to the Supporting Information of ref. [87] and [82], we find for the ratio of phosphorescence intensity Int(T1) to the total emission intensity Int(tot)
Int T 1 Int tot = 1 1 + τ T 1 3 τ S 1 exp Δ E S 1 T 1 k B T
For the phosphorescence decay time τ(T1), we can insert the value given by the plateau displayed in Figure 7 with τ(plateau) = 32 µs or we can calculate it from the average decay time determined from the three triplet substates according to [19,117,127].
τ T 1 = 3 1 τ I + 1 τ II + 1 τ III 1
with τ(I) ≈ τ(I) ≈ 210 µs and τ(III) = 12 µs, we also obtain τ(T1) = 32 µs, as expected.
Inserting into Equation (3) the mean value of τ(T1) = 32 µs, τ(S1) = 0.445 µs and ∆E(S1-T1) = 380 cm−1, we find for ambient temperature the percentage of phosphorescence intensity relative to the total emission intensity of ≈20%. Thus, the fractional emission intensity of the TADF-only emission amounts to ≈80%. Accordingly, compound Cu2I2(Ph2Ppy)3 1-I shows two emission decay paths, as already indicated by the relatively fast phosphorescence rate and by the temperature dependent development of the emission spectra (Figure 5). The rate of the TADF-only process is expressed by
kr(TADF-only) = kr(com) − kr(T1)
With the values of the radiative combined rate of kr(com) = 1.25 × 105 s−1 (kr(com) = ΦPL/τ = 0.81/(6.5 µs), Table 4) and the radiative phosphorescence rate of kr(T1) = 2.88 × 104 s−1r = 35 µs, Table 4), we obtain for the radiative TADF-only process kr(TADF-only) = 9.62 × 104 s−1 corresponding to τr = 10.4 µs decay time. Hence, the radiative decay time of the combined process, phosphorescence and TADF, is significantly shorter than the TADF-only process. This is valid for the radiative as well as for the measured decays. The combined emission decays by 23 % faster than the TADF-only emission. This is a favorable result and may help in future design strategies to shorten the overall emission decay time of emitter materials to reduce OLED device stability problems [68].
Essential properties of Cu2I2(Ph2Ppy)3, 1-I worked out above are summarized in Table 5 and visualized by an energy level diagram in Figure 8.

3. Summarizing Conclusions

In this report, we present an overview over a series of P∩N linked Cu(I) dimers of the type of Cu2X2(P∩N)3 with respect to their X-ray structures and with particular focus on photo-luminescence data. Within the series of eight different compounds, the emission color varies from green to red, the ambient temperature radiative decay time τr covers a range from 8 to 21 µs and the emission quantum yield ΦPL lies between 9 and 81%. All compounds studied show combined TADF and phosphorescence at ambient temperature. This is due to relatively large spin-orbit coupling (SOC) experienced by the lowest lying triplet state T1. Around T = 70 K, the TADF component is largely frozen out and only phosphorescence is observed. The set of data described in the first part, leads us to select a specific compound, namely Cu2I2(Ph2Ppy)3 1-I, for deeper photophysical characterizations by DFT and SOC-TDDFT computations and for emission measurements over the large temperature range of 1.2 ≤ T ≤ 300 K. Thus, the interplay of phosphorescence and TADF is clarified and properties of the lowest excited singlet and triplet states can be revealed in detail. For example, we determine the S1→S0 transition rate, the ∆E(S1−T1) gap, detailed triplet state features, such as the T1→S0 transition rate as well as the rates from the substates, zero-field splitting (ZFS) of the T1 state and the spin-lattice relaxation (SLR) rate between triplet substates. Interestingly, the investigation of the decay time as function of temperature allows us to determine electronic splitting features of the order of less than 1 cm−1 (0.1 meV), although the spectral halfwidth of about 3500 cm−1 (0.43 eV) is by a factor of 3500 larger.
Thus, this presentation proceeds from an overview over properties of a series of related Cu(I) dimers to detailed photophysical characterization at the state of art. Moreover, it becomes evident that the occurrence of a combined emission process at ambient temperature, consisting of distinct phosphorescence and TADF, or if applied in an OLED, consisting of combined singlet and triplet harvesting, leads to significant shortening of the overall photoluminescent decay time. This is a favorable result and may help in future design to improve the performance of OLED devices with respect to decrease of roll-off and stability problems and even increase of the external quantum efficiency (EQE).

4. Materials and Methods

4.1. General

All commercially available solvents and starting materials were used without further purification. 2-(Diphenylphosphino)pyridine was purchased from Acros Organics. 2-(Diphenylphosphino)pyrimidine, 2,1-(diphenylphosphino)isoquinoline, 3, [97] complexes 1-X (X = Cl, Br, I) and 3-I were prepared according to described procedures [34]. Elemental analyses were carried out by the Center for Chemical Analysis of the Faculty of Natural Sciences of the University Regensburg. Single-crystal structure analysis was carried out on a Bruker Smart X2S (1-Br), Bruker X8 APEX-II (2-I) and STOE-IPDS (3-I) diffractometer with graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å). The structures were solved by direct methods (SHELXS-97 [128,129], SIR-92 [130]) and refined by full-matrix least-squares on F2 (SHELXL-97 [131,132] and SHELXL-2014/7 [133]). The H atoms were calculated geometrically and a riding model was applied in the refinement process. Crystallographic details can be found in Table 6.

4.2. Photophysical Measurements

The complexes were investigated as powders. Emission spectra and decay curves were measured by use of a Fluorolog 3 spectrometer (Horiba Jobin Yvon, Munich, Germany) equipped with a cooled photomultiplier tube. The spectra were corrected with respect to the wavelength dependence of the instrument. The decay behavior of the phosphorescence was recorded using a multichannel scaler card (P7887, Fast ComTec, Munich, Germany) with a time resolution of 250 ps. For excitation, the third harmonic of a pulsed Nd:YAG laser (355 nm, pulse width < 8 ns) was used. A Konti IT (CryoVac, Troisdorf, Germany) cryostat was applied for the variation of temperature between 1.2 K and 300 K. Quantum yield measurements at ambient temperature and at 77 K were carried out with an integrating sphere applying a C9920-02 system (Hamamatsu, Herrsching am Ammersee, Germany).

4.3. Computational Investigations

Density functional theory (DFT) geometry optimizations were performed with the Turbomole [134] package using a basis set of triple-zeta plus polarization (def2-TZVP) quality [135] and the hybrid B3LYP exchange-correlation functional [136]. For iodine the corresponding effective core potential [137] was employed. Time-dependent DFT (TDDFT) calculations were carried out with ADF2014 [101] as mentioned in the main text.

4.4. Syntheses

General Procedure for the syntheses of complexes 2-X and 3-Br according to a simplified literature method: The copper(I) halide salt (2 equivalents) and the ligand (3 equivalents) were suspended in dichloromethane (15 mL) and stirred 12 h under ambient conditions. To the filtered reaction solution diethyl ether was added. The formed solid was filtered off, washed with diethyl ether and dried in vacuum. As soon as the complexes were precipitated form the reaction mixture, they were not sufficiently soluble in common weakly or non-coordinating solvents to perform NMR-spectroscopy. Slow gas phase diffusion of diethyl ether into a solution which was obtained by filtration of the crude reaction mixture gave single crystal suitable for X-ray diffraction.
[(2-Diphenylphosphino)pyrimidine)3Cu2Cl2] (2-Cl). 2 (150 mg, 0.57 mmol), CuCl (38 mg, 0.38 mmol). Yield: 127 mg, 0.128 mmol, 68%, yellow powder. Anal. Calcd for C48H39Cu2Cl2N6P3 (990.79 g·mol−1): C, 58.19; H, 3.97; N, 8.48. Found: C, 58.45; H, 4.12; N, 8.43.
[(2-Diphenylphosphino)pyrimidine)3Cu2Br2] (2-Br): 2 (150 mg, 0.57 mmol) CuBr (54 mg, 0.38 mmol). Yield: 156 mg, 0.144 mmol, 74%, yellow powder. Anal. Calcd for C48H39Cu2Br2N6P3 (1079.70 g·mol−1): C, 53.40; H, 3.64; N, 7.78. Found: C, 53.24; H, 3.53; N, 7.76.
[(2-Diphenylphosphino)pyrimidine)3Cu2I2] (2-I): 2 (150 mg, 0.57 mmol) CuI (72 mg, 0.38 mmol). Yield: 176 mg, 0.150 mmol, 79%, yellow powder. Anal. Calcd for C48H39Cu2I2N6P3·½CH2Cl2 (1173.69 g·mol−1): C, 47.90; H, 3.32; N, 6.91. Found: C, 47.73; H, 3.44; N, 6.75.
[(2-Diphenylphosphino)isoquinoline)3Cu2Br2] (3-Br). 3 (75 mg, 0.24 mmol) CuBr (23 mg, 0.16 mmol). Yield: 72 mg, 0.059 mmol, 75%, yellow powder. Anal. Calcd for C63H48Cu2Br2N3P3·½CH2Cl2 (1205.84 g·mol−1): C, 60.08; H, 3.89; N, 3.31. Found: C, 60.29; H, 4.08; N, 3.23.

Author Contributions

Conceptualization, H.Y. and U.M.; synthesis, U.M.; photophysical measurements, T.H.; X-ray structures, M.F. and U.M.; Quantum-chemical calculations, T.A.N.; writing H.Y., T.A.N. and U.M.; visualization, T.A.N., U.M., T.H. and H.Y. 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.

Data Availability Statement

CCDC 2034780 (1-Br), 2034779 (2-I), 2034781 (3-I·CH2Cl2), contain supplementary crystallographic data for this paper. This information can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/.

Acknowledgments

The authors thank Manfred Zabel for the determination of the crystal structure of Cu2I2(Ph2Piqn)3, 3-I and Philipp Edelsbacher for help with formatting the manuscript.

Conflicts of Interest

The authors declare that they do not have any competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Helfrich, W.; Schneider, W.G. Transients of volume-controlled current and of recombination radiation in anthracene. J. Chem. Phys. 1966, 44, 2902–2909. [Google Scholar] [CrossRef]
  2. Yersin, H. Highly Efficient OLEDs with Phosphorescent Materials; Wiley-VCH Verlag: Weinheim, Germany, 2008. [Google Scholar] [CrossRef]
  3. Brütting, W.; Adachi, C. Physics of Organic Semiconductors; Wiley-VCH Verlag: Weinheim, Germany, 2012. [Google Scholar]
  4. Baldo, M.A.; O’Brien, D.F.; You, Y.; Shoustikov, A.; Sibley, S.; Thompson, M.E.; Forrest, S.R. Highly efficient phosphorescent emission from organic electroluminescent devices. Nature 1998, 395, 151–154. [Google Scholar] [CrossRef]
  5. Adachi, C.; Baldo, M.A.; Thompson, M.E.; Forrest, S.R. Nearly 100% internal phosphorescence efficiency in an organic light-emitting device. J. Appl. Phys. 2001, 90, 5048–5051. [Google Scholar] [CrossRef] [Green Version]
  6. Yersin, H. Triplet Emitters for OLED applications. Mechanisms of exciton trapping and control of emission properties. Top. Curr. Chem. 2004, 241, 1–26. [Google Scholar] [CrossRef]
  7. Yersin, H. Highly Efficient OLEDs: Materials Based on Thermally Activated Delayed Fluorescence; Wiley-VCH Verlag: Weinheim, Germany, 2019. [Google Scholar]
  8. Yersin, H.; Monkowius, U. Komplexe mit Kleinen Singulett-Triplett-Energie-Abständen zur Verwendung in Opto-Elektronischen Bauteilen (Singulett-Harvesting-Effekt). German Patent No. DE102008033563, 21 January 2010. [Google Scholar]
  9. Czerwieniec, R.; Yu, J.; Yersin, H. Blue-light emission of Cu(I) complexes and singlet harvesting. Inorg. Chem. 2011, 50, 8293–8301. [Google Scholar] [CrossRef]
  10. Deaton, J.C.; Switalski, S.C.; Kondakov, D.Y.; Young, R.H.; Pawlik, T.D.; Giesen, D.J.; Harkins, S.B.; Miller, A.J.M.; Mickenberg, S.F.; Peters, J.C. E-type delayed fluorescence of a phosphine-supported Cu2(μ-NAr2)2 diamond core: Harvesting singlet and triplet excitons in OLEDs. J. Am. Chem. Soc. 2010, 132, 9499–9508. [Google Scholar] [CrossRef] [Green Version]
  11. Zink, D.M.; Volz, D.; Baumann, T.; Mydlak, M.; Flügge, H.; Friedrichs, J.; Nieger, M.; Bräse, S. Heteroleptic, dinuclear copper (I) complexes for application in organic light-emitting diodes. Chem. Mater. 2013, 25, 4471–4486. [Google Scholar] [CrossRef]
  12. Dumur, F. Recent advances in organic light-emitting devices comprising copper complexes: A realistic approach for low-cost and highly emissive devices? Org. Electron. 2015, 21, 27–39. [Google Scholar] [CrossRef]
  13. Lamansky, S.; Djurovich, P.; Murphy, D.; Abdel-Razzaq, F.; Lee, H.E.; Adachi, C.; Burrows, P.E.; Forrest, S.R.; Thompson, M.E. Highly phosphorescent bis-cyclometalated iridium complexes: Synthesis, photophysical characterization, and use in organic light emitting diodes. J. Am. Chem. Soc. 2001, 123, 4304–4312. [Google Scholar] [CrossRef]
  14. Zysman-Colman, E. Iridium (III) in Optoelectronic and Photonics Applications; John Wiley & Sons: Chichester, UK, 2017. [Google Scholar]
  15. Deaton, J.C.; Castellano, F.N. Archetypal iridium (III) compounds for optoelectronic and photonic applications. In Iridium(III) in Optoelectronic and Photonics Applications; Zysman-Colman, E., Ed.; John Wiley & Sons: Chichester, UK, 2017; pp. 1–69. [Google Scholar] [CrossRef] [Green Version]
  16. Tang, M.-C.; Chan, A.K.-W.; Chan, M.-Y.; Yam, V.W.-W. Platinum and gold complexes for OLEDs. Top. Curr. Chem. 2016, 374, 1–43. [Google Scholar] [CrossRef]
  17. Li, K.; Ming Tong, G.S.; Wan, Q.; Cheng, G.; Tong, W.-Y.; Ang, W.-H.; Kwong, W.-L.; Che, C.-M. Highly phosphorescent platinum (II) emitters: Photophysics, materials and biological applications. Chem. Sci. 2016, 7, 1653–1673. [Google Scholar] [CrossRef] [Green Version]
  18. Armaroli, N.; Bolink, H.J. Photoluminescent Materials and Electroluminescent Devices; Springer: Cham, Switzerland, 2017. [Google Scholar] [CrossRef]
  19. Yersin, H.; Rausch, A.F.; Czerwieniec, R.; Hofbeck, T.; Fischer, T. The triplet state of organo-transition metal compounds. Triplet harvesting and singlet harvesting for efficient OLEDs. Coord. Chem. Rev. 2011, 255, 2622–2652. [Google Scholar] [CrossRef]
  20. Yersin, H.; Rausch, A.F.; Czerwieniec, R. Organometallic emitters for OLEDs: Triplet harvesting, singlet harvesting, case structures, and trends. In Physics of Organic Semiconductors; Brütting, W., Adachi, C., Eds.; Wiley-VCH Verlag: Weinheim, Germany, 2012; pp. 371–424. [Google Scholar] [CrossRef]
  21. Yersin, H.; Finkenzeller, W.J. Triplet emitters for organic light-emitting diodes: Basic properties. In Highly Efficient OLEDs with Phosphorescent Materials; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2007; pp. 1–97. [Google Scholar] [CrossRef]
  22. Che, C.-M.; Kwok, C.-C.; Lai, S.-W.; Rausch, A.F.; Finkenzeller, W.J.; Zhu, N.; Yersin, H. Photophysical properties and OLED applications of phosphorescent platinum (II) Schiff base complexes. Chem. Eur. J. 2010, 16, 233–247. [Google Scholar] [CrossRef]
  23. Lin, Y.-Y.; Chan, S.-C.; Chan, M.C.W.; Hou, Y.-J.; Zhu, N.; Che, C.-M.; Liu, Y.; Wang, Y. Structural, photophysical, and electrophosphorescent properties of platinum (II) complexes supported by tetradentate N2O2 chelates. Chem. Eur. J. 2003, 9, 1263–1272. [Google Scholar] [CrossRef]
  24. Cheng, G.; Kui, S.C.F.; Ang, W.-H.; Ko, M.-Y.; Chow, P.-K.; Kwong, C.-L.; Kwok, C.-C.; Ma, C.; Guan, X.; Low, K.-H.; et al. Structurally robust phosphorescent [Pt(O^N^C^N)] emitters for high performance organic light-emitting devices with power efficiency up to 126 lm W−1 and external quantum efficiency over 20%. Chem. Sci. 2014, 5, 4819–4830. [Google Scholar] [CrossRef]
  25. Mao, M.; Peng, J.; Lam, T.-L.; Ang, W.-H.; Li, H.; Cheng, G.; Che, C.-M. High-performance organic light-emitting diodes with low-efficiency roll-off using bulky tetradentate [Pt(O^N^C^N)] emitters. J. Mater. Chem. C 2019, 7, 7230–7236. [Google Scholar] [CrossRef]
  26. Cheng, G.; Kwak, Y.; To, W.-P.; Lam, T.-L.; Tong, G.S.M.; Sit, M.-K.; Gong, S.; Choi, B.; Choi, W.I.; Yang, C.; et al. High-efficiency solution-processed organic light-emitting diodes with tetradentate platinum(II) emitters. ACS Appl. Mater. Interfaces 2019, 11, 45161–45170. [Google Scholar] [CrossRef]
  27. Vezzu, D.A.K.; Deaton, J.C.; Jones, J.S.; Bartolotti, L.; Harris, C.F.; Marchetti, A.P.; Kondakova, M.; Pike, R.D.; Huo, S. Highly luminescent tetradentate bis-cyclometalated platinum complexes: Design, synthesis, structure, photophysics, and electroluminescence application. Inorg. Chem. 2010, 49, 5107–5119. [Google Scholar] [CrossRef]
  28. Fleetham, T.; Li, G.; Wen, L.; Li, J. Efficient “pure” blue OLEDs employing tetradentate Pt complexes with a narrow spectral bandwidth. Adv. Mater. 2014, 26, 7116–7121. [Google Scholar] [CrossRef]
  29. Czerwieniec, R.; Yersin, H. Diversity of copper(I) complexes showing thermally activated delayed fluorescence: Basic photophysical analysis. Inorg. Chem. 2015, 54, 4322–4327. [Google Scholar] [CrossRef]
  30. Osawa, M.; Hoshino, M.; Hashimoto, M.; Kawata, I.; Igawa, S.; Yashima, M. Application of three-coordinate copper(I) complexes with halide ligands in organic light-emitting diodes that exhibit delayed fluorescence. Dalton Trans. 2015, 44, 8369–8378. [Google Scholar] [CrossRef]
  31. Yu, R.; Lu, C.-Z. Ionic [Cu(NN)(PP)]+ TAD9727 F complexes with pyridine-based diimine chelating ligands and their use in OLEDs. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 177–198. [Google Scholar] [CrossRef]
  32. Hirtenlehner, C.; Monkowius, U. Syntheses, crystal structures and blue luminescence of Cu2X2(Ph3P)2[(−)-nicotine]2 (X. = Br, I). Inorg. Chem. Commun. 2012, 15, 109–112. [Google Scholar] [CrossRef]
  33. So, G.K.-M.; Cheng, G.; Wang, J.; Chang, X.; Kwok, C.-C.; Zhang, H.; Che, C.-M. Efficient color-tunable copper(I) complexes and their applications in solution-processed organic light-emitting diodes. Chem. Asian J. 2017, 12, 1490–1498. [Google Scholar] [CrossRef] [Green Version]
  34. Zink, D.M.; Bächle, M.; Baumann, T.; Nieger, M.; Kühn, M.; Wang, C.; Klopper, W.; Monkowius, U.; Hofbeck, T.; Yersin, H.; et al. Synthesis, structure, and characterization of dinuclear copper(I) halide complexes with P^N ligands featuring exciting photoluminescence properties. Inorg. Chem. 2013, 52, 2292–2305. [Google Scholar] [CrossRef]
  35. Wallesch, M.; Volz, D.; Zink, D.M.; Schepers, U.; Nieger, M.; Baumann, T.; Bräse, S. Bright coppertunities: Multinuclear Cu(I) complexes with N-P ligands and their applications. Chem. Eur. J. 2014, 20, 6578–6590. [Google Scholar] [CrossRef] [PubMed]
  36. Leitl, M.J.; Zink, D.M.; Schinabeck, A.; Baumann, T.; Volz, D.; Yersin, H. Copper(I) complexes for thermally activated delayed fluorescence: From photophysical to device properties. Top. Curr. Chem. 2016, 374, 25–65. [Google Scholar] [CrossRef]
  37. Czerwieniec, R.; Leitl, M.J.; Homeier, H.H.H.; Yersin, H. Cu(I) complexes–thermally activated delayed fluorescence. Photophysical approach and material design. Coord. Chem. Rev. 2016, 325, 2–28. [Google Scholar] [CrossRef]
  38. Hamze, R.; Peltier, J.L.; Sylvinson, D.; Jung, M.; Cardenas, J.; Haiges, R.; Soleilhavoup, M.; Jazzar, R.; Djurovich, P.I.; Bertrand, G.; et al. Eliminating nonradiative decay in Cu(I) emitters: 99% quantum efficiency and microsecond lifetime. Science 2019, 363, 601–606. [Google Scholar] [CrossRef]
  39. Hamze, R.; Shi, S.; Kapper, S.C.; Muthiah Ravinson, D.S.; Estergreen, L.; Jung, M.-C.; Tadle, A.C.; Haiges, R.; Djurovich, P.I.; Peltier, J.L.; et al. “Quick-silver>” from a systematic study of highly luminescent, two-coordinate, d10 coinage metal complexes. J. Am. Chem. Soc. 2019, 141, 8616–8626. [Google Scholar] [CrossRef] [PubMed]
  40. Shi, S.; Jung, M.C.; Coburn, C.; Tadle, A.; Sylvinson, M.R.D.; Djurovich, P.I.; Forrest, S.R.; Thompson, M.E. Highly efficient photo- and electroluminescence from two-coordinate Cu(I) complexes featuring nonconventional N-heterocyclic carbenes. J. Am. Chem. Soc. 2019, 141, 3576–3588. [Google Scholar] [CrossRef] [PubMed]
  41. Di, D.; Romanov, A.S.; Yang, L.; Richter, J.M.; Rivett, J.P.; Jones, S.; Thomas, T.H.; Jalebi, M.A.; Friend, R.H.; Linnolahti, M.; et al. High-performance light-emitting diodes based on carbene-metal-amides. Science 2017, 356, 159–163. [Google Scholar] [CrossRef] [Green Version]
  42. Conaghan, P.J.; Menke, S.M.; Romanov, A.S.; Jones, S.T.E.; Pearson, A.J.; Evans, E.W.; Bochmann, M.; Greenham, N.C.; Credgington, D. Efficient vacuum-processed light-emitting diodes based on carbene-metal-amides. Adv. Mater. 2018, 30, e1802285. [Google Scholar] [CrossRef]
  43. Romanov, A.S.; Jones, S.T.E.; Gu, Q.; Conaghan, P.J.; Drummond, B.H.; Feng, J.; Chotard, F.; Buizza, L.; Foley, M.; Linnolahti, M.; et al. Carbene metal amide photoemitters: Tailoring conformationally flexible amides for full color range emissions including white-emitting OLED. Chem. Sci. 2020, 11, 435–446. [Google Scholar] [CrossRef] [Green Version]
  44. Föller, J.; Ganter, C.; Steffen, A.; Marian, C.M. Computer-aided design of luminescent linear N-heterocyclic carbene copper(I) pyridine complexes. Inorg. Chem. 2019, 58, 5446–5456. [Google Scholar] [CrossRef]
  45. Osawa, M.; Hoshino, M. Molecular design and synthesis of metal complexes as emitters for TADF-type OLEDs. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 119–176. [Google Scholar] [CrossRef]
  46. Nozaki, K.; Iwamura, M. Highly emissive d10 metal complexes as TADF emitters with versatile structures and photophysical properties. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 61–91. [Google Scholar] [CrossRef]
  47. Tsuboyama, A. Luminescent dinuclear copper(I) complexes with short intramolecular Cu-Cu distances. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 93–118. [Google Scholar] [CrossRef]
  48. Yersin, H.; Czerwieniec, R.; Shafikov, M.Z.; Suleymanova, A.F. TADF material design: Photophysical background and case studies focusing on CuI and AgI complexes. Chemphyschem 2017, 18, 3508–3535. [Google Scholar] [CrossRef]
  49. Armaroli, N.; Accorsi, G.; Cardinali, F.; Listorti, A. Photochemistry and photophysics of coordination compounds: Copper. In Photochemistry and Photophysics of Coordination Compounds I; Balzani, V., Campagna, S., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; pp. 69–115. [Google Scholar] [CrossRef]
  50. Schinabeck, A.; Leitl, M.J.; Yersin, H. Dinuclear Cu(I) complex with combined bright TADF and phosphorescence. Zero-field splitting and spin-lattice relaxation effects of the triplet state. J. Phys. Chem. Lett. 2018, 9, 2848–2856. [Google Scholar] [CrossRef]
  51. Schinabeck, A.; Rau, N.; Klein, M.; Sundermeyer, J.; Yersin, H. Deep blue emitting Cu(I) tripod complexes. Design of high quantum yield materials showing TADF-assisted phosphorescence. Dalton Trans. 2018, 47, 17067–17076. [Google Scholar] [CrossRef]
  52. Schinabeck, A.; Chen, J.; Kang, L.; Teng, T.; Homeier, H.H.H.; Suleymanova, A.F.; Shafikov, M.Z.; Yu, R.; Lu, C.-Z.; Yersin, H. Symmetry-based design strategy for unprecedentedly fast decaying thermally activated delayed fluorescence (TADF). Application to dinuclear Cu(I) compounds. Chem. Mater. 2019, 31, 4392–4404. [Google Scholar] [CrossRef]
  53. Xu, K.; Chen, B.-L.; Yang, F.; Liu, L.; Zhong, X.-X.; Wang, L.; Zhu, X.-J.; Li, F.-B.; Wong, W.-Y.; Qin, H.-M. Largely color-tuning prompt and delayed fluorescence: Dinuclear Cu(I) halide complexes with tert -amines and phosphines. Inorg. Chem. 2021, 60, 4841–4851. [Google Scholar] [CrossRef]
  54. Gan, X.-M.; Yu, R.; Chen, X.-L.; Yang, M.; Lin, L.; Wu, X.-Y.; Lu, C.-Z. A unique tetranuclear Ag(I) complex emitting efficient thermally activated delayed fluorescence with a remarkably short decay time. Dalton Trans. 2018, 47, 5956–5960. [Google Scholar] [CrossRef]
  55. Shafikov, M.Z.; Suleymanova, A.F.; Czerwieniec, R.; Yersin, H. Design strategy for Ag(I)-based thermally activated delayed fluorescence reaching an efficiency breakthrough. Chem. Mater. 2017, 29, 1708–1715. [Google Scholar] [CrossRef]
  56. Shafikov, M.Z.; Suleymanova, A.F.; Schinabeck, A.; Yersin, H. Dinuclear Ag(I) complex designed for highly efficient thermally activated delayed fluorescence. J. Phys. Chem. Lett. 2018, 9, 702–709. [Google Scholar] [CrossRef] [PubMed]
  57. Osawa, M.; Kawata, I.; Ishii, R.; Igawa, S.; Hashimoto, M.; Hoshino, M. Application of neutral d10 coinage metal complexes with an anionic bidentate ligand in delayed fluorescence-type organic light-emitting diodes. J. Mater. Chem. C 2013, 1, 4375–4383. [Google Scholar] [CrossRef]
  58. Chen, J.; Teng, T.; Kang, L.; Chen, X.-L.; Wu, X.-Y.; Yu, R.; Lu, C.-Z. Highly efficient thermally activated delayed fluorescence in dinuclear Ag(I) complexes with a bis-bidentate tetraphosphane bridging ligand. Inorg. Chem. 2016, 55, 9528–9536. [Google Scholar] [CrossRef]
  59. Chan, K.-T.; Lam, T.-L.; Yu, D.; Du, L.; Phillips, D.L.; Kwong, C.-L.; Tong, G.S.M.; Cheng, G.; Che, C.-M. Strongly luminescent tungsten emitters with emission quantum yields of up to 84 %: TADF and high-efficiency molecular tungsten OLEDs. Angew. Chem. Int. Ed. Engl. 2019, 58, 14896–14900. [Google Scholar] [CrossRef]
  60. Fernandez-Cestau, J.; Bertrand, B.; Blaya, M.; Jones, G.A.; Penfold, T.J.; Bochmann, M. Synthesis and luminescence modulation of pyrazine-based gold(III) pincer complexes. Chem. Commun. 2015, 51, 16629–16632. [Google Scholar] [CrossRef] [Green Version]
  61. Zhou, D.; To, W.-P.; Kwak, Y.; Cho, Y.; Cheng, G.; Tong, G.S.M.; Che, C.-M. Thermally stable donor-acceptor type (Alkynyl)gold(III) TADF emitters achieved EQEs and luminance of up to 23.4% and 70 300 cd m-2 in vacuum-deposited OLEDs. Adv. Sci. 2019, 6, 1802297. [Google Scholar] [CrossRef] [Green Version]
  62. Zhou, D.; To, W.-P.; Tong, G.S.M.; Cheng, G.; Du, L.; Phillips, D.L.; Che, C.-M. Tetradentate gold(III) complexes as thermally activated delayed fluorescence (TADF) emitters: Microwave-assisted synthesis and high-performance OLEDs with long operational lifetime. Angew. Chem. Int. Ed. Engl. 2020, 59, 6375–6382. [Google Scholar] [CrossRef]
  63. Li, L.-K.; Tang, M.-C.; Lai, S.-L.; Ng, M.; Kwok, W.-K.; Chan, M.-Y.; Yam, V.W.-W. Strategies towards rational design of gold(III) complexes for high-performance organic light-emitting devices. Nat. Photonics 2019, 13, 185–191. [Google Scholar] [CrossRef]
  64. Lee, C.-H.; Tang, M.-C.; Kong, F.K.-W.; Cheung, W.-L.; Ng, M.; Chan, M.-Y.; Yam, V.W.-W. Isomeric tetradentate ligand-containing cyclometalated gold(III) complexes. J. Am. Chem. Soc. 2020, 142, 520–529. [Google Scholar] [CrossRef]
  65. Sakai, Y.; Sagara, Y.; Nomura, H.; Nakamura, N.; Suzuki, Y.; Miyazaki, H.; Adachi, C. Zinc complexes exhibiting highly efficient thermally activated delayed fluorescence and their application to organic light-emitting diodes. Chem. Commun. 2015, 51, 3181–3184. [Google Scholar] [CrossRef]
  66. Uoyama, H.; Goushi, K.; Shizu, K.; Nomura, H.; Adachi, C. Highly efficient organic light-emitting diodes from delayed fluorescence. Nature 2012, 492, 234–238. [Google Scholar] [CrossRef]
  67. Hosokai, T.; Matsuzaki, H.; Nakanotani, H.; Tokumaru, K.; Tsutsui, T.; Furube, A.; Nasu, K.; Nomura, H.; Yahiro, M.; Adachi, C. Evidence and mechanism of efficient thermally activated delayed fluorescence promoted by delocalized excited states. Sci. Adv. 2017, 3, e1603282. [Google Scholar] [CrossRef] [Green Version]
  68. Noda, H.; Nakanotani, H.; Adachi, C. Excited state engineering for efficient reverse intersystem crossing. Sci. Adv. 2018, 4, eaao6910. [Google Scholar] [CrossRef] [Green Version]
  69. Li, X.; Shi, Y.-Z.; Wang, K.; Zhang, M.; Zheng, C.-J.; Sun, D.-M.; Dai, G.-L.; Fan, X.-C.; Wang, D.-Q.; Liu, W.; et al. Thermally activated delayed fluorescence carbonyl derivatives for organic light-emitting diodes with extremely narrow full width at half-maximum. ACS Appl. Mater. Interfaces 2019, 11, 13472–13480. [Google Scholar] [CrossRef]
  70. Liu, Y.; Li, C.; Ren, Z.; Yan, S.; Bryce, M.R. All-organic thermally activated delayed fluorescence materials for organic light-emitting diodes. Nat. Rev. Mater. 2018, 3, 18020. [Google Scholar] [CrossRef]
  71. Komatsu, R.; Ohsawa, T.; Sasabe, H.; Nakao, K.; Hayasaka, Y.; Kido, J. Manipulating the electronic excited state energies of pyrimidine-based thermally activated delayed fluorescence emitters to realize efficient deep-blue emission. ACS Appl. Mater. Interfaces 2017, 9, 4742–4749. [Google Scholar] [CrossRef]
  72. Li, M.; Liu, Y.; Duan, R.; Wei, X.; Yi, Y.; Wang, Y.; Chen, C.-F. Aromatic-imide-based thermally activated delayed fluorescence materials for highly efficient organic light-emitting diodes. Angew. Chem. Int. Ed Engl. 2017, 56, 8818–8822. [Google Scholar] [CrossRef]
  73. Sharma, N.; Wong, M.Y.; Samuel, I.D.W.; Zysman-Colman, E. Solution-processed TADF materials and devices based on organic emitters. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 501–541. [Google Scholar] [CrossRef]
  74. Wong, M.Y.; Zysman-Colman, E. Purely organic thermally activated delayed fluorescence materials for organic light-emitting diodes. Adv. Mater. 2017, 29. [Google Scholar] [CrossRef] [Green Version]
  75. Chen, X.-K.; Kim, D.; Brédas, J.-L. Thermally activated delayed fluorescence (TADF) path toward efficient electroluminescence in purely organic materials: Molecular level insight. Acc. Chem. Res. 2018, 51, 2215–2224. [Google Scholar] [CrossRef]
  76. Sommer, G.A.; Mataranga-Popa, L.N.; Czerwieniec, R.; Hofbeck, T.; Homeier, H.H.H.; Müller, T.J.J.; Yersin, H. Design of conformationally distorted donor-acceptor dyads showing efficient thermally activated delayed fluorescence. J. Phys. Chem. Lett. 2018, 9, 3692–3697. [Google Scholar] [CrossRef] [PubMed]
  77. Schmidt, T.D.; Brütting, W. Efficiency enhancement of organic light-emitting diodes exhibiting delayed fluorescence and nonisotropic emitter orientation. In Highly Efficient OLEDs; Yersin, H., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2018; pp. 199–228. [Google Scholar] [CrossRef]
  78. Yang, Z.; Mao, Z.; Xie, Z.; Zhang, Y.; Liu, S.; Zhao, J.; Xu, J.; Chi, Z.; Aldred, M.P. Recent advances in organic thermally activated delayed fluorescence materials. Chem. Soc. Rev. 2017, 46, 915–1016. [Google Scholar] [CrossRef] [PubMed]
  79. Cai, X.; Su, S.-J. Marching toward highly efficient, pure-blue, and stable thermally activated delayed fluorescent organic light-emitting diodes. Adv. Funct. Mater. 2018, 28, 1802558. [Google Scholar] [CrossRef]
  80. Yersin, H.; Mataranga-Popa, L.; Czerwieniec, R.; Dovbii, Y. Design of a new mechanism beyond thermally activated delayed fluorescence toward fourth generation organic light emitting diodes. Chem. Mater. 2019, 31, 6110–6116. [Google Scholar] [CrossRef]
  81. Yersin, H.; Mataranga-Popa, L.; Li, S.-W.; Czerwieniec, R. Design strategies for materials showing thermally activated delayed fluorescence and beyond: Towards the fourth-generation OLED mechanism. J. Soc. Inf. Disp. 2018, 26, 194–199. [Google Scholar] [CrossRef]
  82. Leitl, M.J.; Küchle, F.-R.; Mayer, H.A.; Wesemann, L.; Yersin, H. Brightly blue and green emitting Cu(I) dimers for singlet harvesting in OLEDs. J. Phys. Chem. A 2013, 117, 11823–11836. [Google Scholar] [CrossRef] [PubMed]
  83. Gneuß, T.; Leitl, M.J.; Finger, L.H.; Rau, N.; Yersin, H.; Sundermeyer, J. A new class of luminescent Cu(I) complexes with tripodal ligands–TADF emitters for the yellow to red color range. Dalton Trans. 2015, 44, 8506–8520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Chen, X.-L.; Yu, R.; Wu, X.-Y.; Liang, D.; Jia, J.-H.; Lu, C.-Z. A strongly greenish-blue-emitting Cu4Cl4 cluster with an efficient spin-orbit coupling (SOC): Fast phosphorescence versus thermally activated delayed fluorescence. Chem. Commun. 2016, 52, 6288–6291. [Google Scholar] [CrossRef]
  85. Bizzarri, C.; Hundemer, F.; Busch, J.; Bräse, S. Triplet emitters versus TADF emitters in OLEDs: A comparative study. Polyhedron 2018, 140, 51–66. [Google Scholar] [CrossRef]
  86. Baranov, A.Y.; Berezin, A.S.; Samsonenko, D.G.; Mazur, A.S.; Tolstoy, P.M.; Plyusnin, V.F.; Kolesnikov, I.E.; Artem’ev, A.V. New Cu(I) halide complexes showing TADF combined with room temperature phosphorescence: The balance tuned by halogens. Dalton Trans. 2020, 49, 3155–3163. [Google Scholar] [CrossRef]
  87. Hofbeck, T.; Monkowius, U.; Yersin, H. Highly efficient luminescence of Cu(I) compounds: Thermally activated delayed fluorescence combined with short-lived phosphorescence. J. Am. Chem. Soc. 2015, 137, 399–404. [Google Scholar] [CrossRef] [PubMed]
  88. Li, C.; Li, W.; Henwood, A.F.; Hall, D.; Cordes, D.B.; Slawin, A.M.Z.; Lemaur, V.; Olivier, Y.; Samuel, I.D.W.; Zysman-Colman, E. Luminescent dinuclear Copper(I) complexes bearing an imidazolylpyrimidine bridging ligand. Inorg. Chem. 2020, 59, 14772–14784. [Google Scholar] [CrossRef] [PubMed]
  89. Hobbollahi, E.; List, M.; Hupp, B.; Mohr, F.; Berger, R.J.F.; Steffen, A.; Monkowius, U. Highly efficient cold-white light emission in a Au2CuCl2(P∩N)2PF6 type salt. Dalton Trans. 2017, 46, 3438–3442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Yersin, H.; Fischer, T.; Monkowius, U.; Hofbeck, T. Copper Complexes for Optoelectronic Applications. German Patent DE 102009030475 A1, 5 January 2011. [Google Scholar]
  91. Chen, K.; Shearer, J.; Catalano, V.J. Subtle modulation of Cu4X4L2 phosphine cluster cores leads to changes in luminescence. Inorg. Chem. 2015, 54, 6245–6256. [Google Scholar] [CrossRef]
  92. Artemev, A.V.; Shafikov, M.Z.; Schinabeck, A.; Antonova, O.V.; Berezin, A.S.; Bagryanskaya, I.Y.; Plusnin, P.E.; Yersin, H. Sky-blue thermally activated delayed fluorescence (TADF) based on Ag(I) complexes: Strong solvation-induced emission enhancement. Inorg. Chem. Front. 2019, 6, 3168–3176. [Google Scholar] [CrossRef]
  93. Artem’ev, A.V.; Davydova, M.P.; Berezin, A.S.; Ryzhikov, M.R.; Samsonenko, D.G. Dicopper(I) paddle-wheel complexes with thermally activated delayed fluorescence adjusted by ancillary ligands. Inorg. Chem. 2020, 59, 10699–10706. [Google Scholar] [CrossRef]
  94. Davydova, M.P.; Rakhmanova, M.I.; Bagryanskaya, I.Y.; Brylev, K.A.; Artemev, A.V. A 1D coordination polymer based on CuI and 2-(Diphenylphosphino)pyrimidine: Synthesis, structure and luminescent properties. J. Struct. Chem. 2020, 61, 894–898. [Google Scholar] [CrossRef]
  95. 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]
  96. Busch, J.M.; Koshelev, D.S.; Vashchenko, A.A.; Fuhr, O.; Nieger, M.; Utochnikova, V.V.; Bräse, S. Various structural design modifications: Para-substituted diphenylphosphinopyridine bridged Cu(I) complexes in organic light-emitting diodes. Inorg. Chem. 2021, 60, 2315–2332. [Google Scholar] [CrossRef]
  97. Monkowius, U.; Zabel, M.; Fleck, M.; Yersin, H. Gold(I) complexes bearing P∩N-ligands: An unprecedented twelve-membered ring structure stabilized by aurophilic interactions. Z. Naturforsch. B 2009, 64, 1513–1524. [Google Scholar] [CrossRef]
  98. Bondi, A. van der waals volumes and radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  99. van Lenthe, E.; Baerends, E.J.; Snijders, J.G. Relativistic regular two-component Hamiltonians. J. Chem. Phys. 1993, 99, 4597–4610. [Google Scholar] [CrossRef]
  100. 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. [Google Scholar] [CrossRef]
  101. ADF2014, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. Available online: http://www.scm.com (accessed on 26 May 2021).
  102. van Lenthe, E.; Baerends, E.J. Optimized Slater-type basis sets for the elements 1-118. J. Comput. Chem. 2003, 24, 1142–1156. [Google Scholar] [CrossRef]
  103. Chen, L.X.; Shaw, G.B.; Novozhilova, I.; Liu, T.; Jennings, G.; Attenkofer, K.; Meyer, G.J.; Coppens, P. MLCT state structure and dynamics of a Copper(I) diimine complex characterized by pump-probe X-ray and laser spectroscopies and DFT calculations. J. Am. Chem. Soc. 2003, 125, 7022–7034. [Google Scholar] [CrossRef]
  104. Siddique, Z.A.; Yamamoto, Y.; Ohno, T.; Nozaki, K. Structure-dependent photophysical properties of singlet and triplet metal-to-ligand charge transfer states in Copper(I) bis(diimine) compounds. Inorg. Chem. 2003, 42, 6366–6378. [Google Scholar] [CrossRef]
  105. 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]
  106. Pye, C.C.; Ziegler, T. An implementation of the conductor-like screening model of solvation within the Amsterdam density functional package. Theor. Chem. Acc. 1999, 101, 396–408. [Google Scholar] [CrossRef]
  107. Mori, K.; Goumans, T.P.M.; van Lenthe, E.; Wang, F. Predicting phosphorescent lifetimes and zero-field splitting of organometallic complexes with time-dependent density functional theory including spin-orbit coupling. Phys. Chem. Chem. Phys. 2014, 16, 14523–14530. [Google Scholar] [CrossRef]
  108. Toptygin, D. Effects of the solvent refractive index and its dispersion on the radiative decay rate and extinction coefficient of a fluorescent solute. J. Fluoresc. 2003, 13, 201–219. [Google Scholar] [CrossRef]
  109. Niehaus, T.A.; Hofbeck, T.; Yersin, H. Charge-transfer excited states in phosphorescent organo-transition metal compounds: A difficult case for time dependent density functional theory? RSC Adv. 2015, 5, 63318–63329. [Google Scholar] [CrossRef] [Green Version]
  110. Rössler, U.; Yersin, H. Destabilization of a self-trapped exciton in a quasi-one-dimensional semiconductor: Mg[Pt(CN)4]7 H2O with hydrostatic pressure. Phys. Rev. B 1982, 26, 3187–3191. [Google Scholar] [CrossRef] [Green Version]
  111. Gliemann, G.; Yersin, H. Spectroscopic properties of the quasi one-dimensional tetracyanoplatinate(II) compounds. In Clusters; Springer: Berlin/Heidelberg, Germany, 1985; pp. 87–153. [Google Scholar] [CrossRef]
  112. Tsuboyama, A.; Kuge, K.; Furugori, M.; Okada, S.; Hoshino, M.; Ueno, K. Photophysical properties of highly luminescent Copper(I) halide complexes chelated with 1,2-bis(diphenylphosphino)benzene. Inorg. Chem. 2007, 46, 1992–2001. [Google Scholar] [CrossRef]
  113. Turro, N.J. Modern Molecular Photochemistry of Organic Molecules; Benjamin/Cummings: Melon Park, CA, USA, 1978. [Google Scholar]
  114. Leitl, M.J.; Krylova, V.A.; Djurovich, P.I.; Thompson, M.E.; Yersin, H. Phosphorescence versus thermally activated delayed fluorescence. Controlling singlet-triplet splitting in brightly emitting and sublimable Cu(I) compounds. J. Am. Chem. Soc. 2014, 136, 16032–16038. [Google Scholar] [CrossRef]
  115. Murov, S.L.; Carmichael, J.; Hug, G.L. Handbook of Photochemistry, 2nd ed.; Marcel Dekker: New York, NY, USA, 1993; p. 340. [Google Scholar]
  116. Shafikov, M.Z.; Zaytsev, A.V.; Kozhevnikov, V.N. Halide-enhanced spin-orbit coupling and the phosphorescence rate in Ir(III) complexes. Inorg. Chem. 2021, 60, 642–650. [Google Scholar] [CrossRef]
  117. Yersin, H.; Strasser, J. Triplets in metal–organic compounds. Chemical tunability of relaxation dynamics. Coord. Chem. Rev. 2000, 208, 331–364. [Google Scholar] [CrossRef]
  118. Bergmann, L.; Hedley, G.J.; Baumann, T.; Bräse, S.; Samuel, I.D.W. Direct observation of intersystem crossing in a thermally activated delayed fluorescence Copper complex in the solid state. Sci. Adv. 2016, 2, e1500889. [Google Scholar] [CrossRef] [Green Version]
  119. Tschierlei, S.; Karnahl, M.; Rockstroh, N.; Junge, H.; Beller, M.; Lochbrunner, S. Substitution-controlled excited state processes in heteroleptic Copper(I) photosensitizers used in hydrogen evolving systems. Chem. Phys. Chem. 2014, 15, 3709–3713. [Google Scholar] [CrossRef]
  120. Garakyaraghi, S.; Danilov, E.O.; McCusker, C.E.; Castellano, F.N. Transient absorption dynamics of sterically congested Cu(I) MLCT excited states. J. Phys. Chem. A 2015, 119, 3181–3193. [Google Scholar] [CrossRef]
  121. Iwamura, M.; Takeuchi, S.; Tahara, T. Ultrafast excited-state dynamics of Copper(I) complexes. Acc. Chem. Res. 2015, 48, 782–791. [Google Scholar] [CrossRef] [PubMed]
  122. Eng, J.; Penfold, T.J. Understanding and designing thermally activated delayed fluorescence emitters: Beyond the energy gap approximation. Chem. Rec. 2020, 20, 831–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Penfold, T.J.; Gindensperger, E.; Daniel, C.; Marian, C.M. Spin-vibronic mechanism for intersystem crossing. Chem. Rev. 2018, 118, 6975–7025. [Google Scholar] [CrossRef] [Green Version]
  124. Striplin, D.R.; Crosby, G.A. Nature of the emitting 3MLCT manifold of rhenium(I)(diimine)(CO)3Cl complexes. Chem. Phys. Lett. 1994, 221, 426–430. [Google Scholar] [CrossRef]
  125. Finkenzeller, W.J.; Thompson, M.E.; Yersin, H. Phosphorescence dynamics and spin-lattice relaxation of the OLED emitter Ir(btp)2(acac). Chem. Phys. Lett. 2007, 444, 273–279. [Google Scholar] [CrossRef]
  126. Finkenzeller, W.J.; Hofbeck, T.; Thompson, M.E.; Yersin, H. Triplet state properties of the OLED emitter Ir(btp)2(acac): Characterization by site-selective spectroscopy and application of high magnetic fields. Inorg. Chem. 2007, 46, 5076–5083. [Google Scholar] [CrossRef]
  127. Tinti, D.S.; El-Sayed, M.A. New techniques in triplet state phosphorescence spectroscopy: Application to the emission of 2,3-dichloroquinoxaline. J. Chem. Phys. 1971, 54, 2529–2549. [Google Scholar] [CrossRef]
  128. Sheldrick, G.M. SHELXS-97, Program for the Solution of Crystal Structures; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  129. Sheldrick, G.M. Phase annealing in SHELX-90: Direct methods for larger structures. Acta Crystallogr. 1990, A46, 467–473. [Google Scholar] [CrossRef]
  130. Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A. Completion and refinement of crystal structures with SIR 92. J. Appl. Crystallogr. 1993, 26, 343–350. [Google Scholar] [CrossRef]
  131. Sheldrick, G.M. SHELXL-97, Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  132. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  133. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  134. Turbomole V6.4 2012. A Development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989-2007, TURBOMOLE GmbH, since 2007. Available online: http://www.turbomole.com (accessed on 26 May 2021).
  135. 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] [PubMed]
  136. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  137. Peterson, K.A.; Figgen, D.; Goll, E.; Stoll, H.; Dolg, M. Systematically convergent basis sets with relativistic pseudopotentials. II. Small-core pseudopotentials and correlation consistent basis sets for the post- d group 16–18 elements. J. Chem. Phys. 2003, 119, 11113–11123. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Synthesis of the di-nuclear copper complexes.
Scheme 1. Synthesis of the di-nuclear copper complexes.
Molecules 26 03415 sch001
Figure 1. Molecular structure of Cu2Br2(Ph2Ppy)3 1-Br, Cu2I2(Ph2Ppym)3 2-I and Cu2I2(Ph2Piqn)3 3-I (hydrogen atoms and solvent molecules are omitted for clarity; displacement parameters are drawn at 50% probability level).
Figure 1. Molecular structure of Cu2Br2(Ph2Ppy)3 1-Br, Cu2I2(Ph2Ppym)3 2-I and Cu2I2(Ph2Piqn)3 3-I (hydrogen atoms and solvent molecules are omitted for clarity; displacement parameters are drawn at 50% probability level).
Molecules 26 03415 g001
Figure 2. Chemical structure of Cu2I2(Ph2Ppy)3, 1-I.
Figure 2. Chemical structure of Cu2I2(Ph2Ppy)3, 1-I.
Molecules 26 03415 g002
Figure 3. HOMO (left) and LUMO (right) of Cu2I2(Ph2Ppy)3, 1-I at the B3LYP/DZP level for the ground state S0 geometry. Iso-contour values are set to 0.03 with blue/purple color representing the sign of the wave function. Color code: P (orange), Cu (brown), I (violet), N (blue), C (grey), H (white). In the right plot, the two iodine atoms are exactly on top of each other.
Figure 3. HOMO (left) and LUMO (right) of Cu2I2(Ph2Ppy)3, 1-I at the B3LYP/DZP level for the ground state S0 geometry. Iso-contour values are set to 0.03 with blue/purple color representing the sign of the wave function. Color code: P (orange), Cu (brown), I (violet), N (blue), C (grey), H (white). In the right plot, the two iodine atoms are exactly on top of each other.
Molecules 26 03415 g003
Figure 4. Emission and excitation spectra of Cu2Cl2(Ph2Ppy)3, 1-Cl, Cu2Cl2(Ph2Ppy)3, 1-Br and Cu2I2(Ph2Ppym)3, 1-I (powder, 300 K, λexc = 350 nm).
Figure 4. Emission and excitation spectra of Cu2Cl2(Ph2Ppy)3, 1-Cl, Cu2Cl2(Ph2Ppy)3, 1-Br and Cu2I2(Ph2Ppym)3, 1-I (powder, 300 K, λexc = 350 nm).
Molecules 26 03415 g004
Figure 5. Emission spectra of Cu2I2(Ph2Ppy)3, 1-I at different temperatures (powder, λexc = 355 nm). The spectra for different temperatures shown in (a) are super-imposed in (b) to visualize the intensity growing in at the blue side flank with temperature increase.
Figure 5. Emission spectra of Cu2I2(Ph2Ppy)3, 1-I at different temperatures (powder, λexc = 355 nm). The spectra for different temperatures shown in (a) are super-imposed in (b) to visualize the intensity growing in at the blue side flank with temperature increase.
Molecules 26 03415 g005
Figure 6. Emission decay behavior of Cu2I2(Ph2Ppy)3, 1-I at different temperatures (powder, λexc = 355 nm). At 1.2 K, a bi-exponentially decay can be observed showing a short component that is determined by spin-lattice relaxation (SLR) processes between the triplet substates. The long component represents the decay of the involved thermalized states. At higher temperatures, the decay is mono-exponential. Note the different scales.
Figure 6. Emission decay behavior of Cu2I2(Ph2Ppy)3, 1-I at different temperatures (powder, λexc = 355 nm). At 1.2 K, a bi-exponentially decay can be observed showing a short component that is determined by spin-lattice relaxation (SLR) processes between the triplet substates. The long component represents the decay of the involved thermalized states. At higher temperatures, the decay is mono-exponential. Note the different scales.
Molecules 26 03415 g006
Figure 7. Temperature dependence of the emission decay time of the thermally equilibrated excited states of Cu2I2(Ph2Ppy)3 1-I. The calculated fit is based on Equation (1). As inset, we summarize the fit parameters.
Figure 7. Temperature dependence of the emission decay time of the thermally equilibrated excited states of Cu2I2(Ph2Ppy)3 1-I. The calculated fit is based on Equation (1). As inset, we summarize the fit parameters.
Molecules 26 03415 g007
Figure 8. Energy level diagram, decay data and rates for Cu2I2(Ph2Ppy)3, 1-I powder. The T1 substates I and II emit independently at T = 1.2 K with τ(I) ≈ τ(II) ≈ 210 µs. The transition rate k(III→S0) amounts to 8.3 × 104 s−1, corresponding to τ(III) = 12 µs. The spin-lattice relaxation time at T = 1.2 K is determined to τ(SLR) = 59 µs. The T1→S0(0-0) energy is estimated from the blue energy flank of the phosphorescence band as displayed in Figure 5. At T = 300 K, the total emission is composed of phosphorescence (20%) and TADF (80%), leading to a combined decay time of 6.5 µs at ΦPL = 81%.
Figure 8. Energy level diagram, decay data and rates for Cu2I2(Ph2Ppy)3, 1-I powder. The T1 substates I and II emit independently at T = 1.2 K with τ(I) ≈ τ(II) ≈ 210 µs. The transition rate k(III→S0) amounts to 8.3 × 104 s−1, corresponding to τ(III) = 12 µs. The spin-lattice relaxation time at T = 1.2 K is determined to τ(SLR) = 59 µs. The T1→S0(0-0) energy is estimated from the blue energy flank of the phosphorescence band as displayed in Figure 5. At T = 300 K, the total emission is composed of phosphorescence (20%) and TADF (80%), leading to a combined decay time of 6.5 µs at ΦPL = 81%.
Molecules 26 03415 g008
Table 1. Selected bond lengths (Å) and bond angles (°) for Cu2X2(Ph2Ppy)3, 1-X (with X = Cl, Br, I), Cu2I2(Ph2Ppym)3, 2-I and Cu2I2(Ph2Piqn)3, 3-I.
Table 1. Selected bond lengths (Å) and bond angles (°) for Cu2X2(Ph2Ppy)3, 1-X (with X = Cl, Br, I), Cu2I2(Ph2Ppym)3, 2-I and Cu2I2(Ph2Piqn)3, 3-I.
Compound1-Cl a1-Br1-I a2-I3-I·CH2Cl23-I a
Cu1–Cu22.878(1)2.883(1)2.7694(5)2.693(1)2.799(1)2.7204(6)
Cu1–P12.242(1)2.240(1)2.2514(6)2.263(1)2.292(1)2.2555(8)
Cu1–P22.248(1)2.258(1)2.2522(7)2.249(1)2.291(1)2.2404(9)
Cu2–P32.224(1)2.237(1)2.2507(7)2.244(1)2.285(1)2.2468(9)
Cu2–N2.106(3)2.098(4)2.104(1)2.101(3)2.140(3)2.076(2)
Cu1–X12.395(1)2.570(1)2.6733(7)2.684(1)2.699(1)2.6930(5)
Cu1–X22.426 (1)2.522(1)2.6803(5)2.718(1)2.641(1)2.6954(5)
Cu2–X12.436(1)2.509(1)2.7280(6)2.647(1)2.714(1)2.6277(5)
Cu2–X22.390(1)2.543(1)2.6446(5)2.702(1)2.687(1)2.6802(5)
X1–Cu1–X298.38(4)101.89(3)107.63(2)106.85(2)107.02(2)108.86(2)
X1–Cu2–X298.25(3)103.00(3)107.07(1)108.41(2)105.31(2)111.32(2)
N–Cu2–P3123.23(8)123.35(9)117.47(2)115.9(1)111.8(1)120.35(2)
P1–Cu1–P2123.80(4)123.90(5)119.87(2)118.40(4)118.8(1)118.16(3)
Cu1–X1–Cu273.42(3)69.17(2)61.68(1)60.69(1)62.27(2)61.48(1)
Cu1–X2–Cu273.13(3)69.40(2)62.67(1)59.59(1)63.38(2)60.80(1)
a—values from Ref. [34].
Table 2. Selected bond lengths [Å] and angles [°] for Cu2I2(Ph2Ppy)3 1-I as obtained from DFT calculations (B3LYP/TZVP) for the S0 and T1 minimum compared to the values from the X-ray structure determination. The atom numbering scheme is analogous to Cu2Br2(Ph2Ppy)3 1-Br shown in Figure 1.
Table 2. Selected bond lengths [Å] and angles [°] for Cu2I2(Ph2Ppy)3 1-I as obtained from DFT calculations (B3LYP/TZVP) for the S0 and T1 minimum compared to the values from the X-ray structure determination. The atom numbering scheme is analogous to Cu2Br2(Ph2Ppy)3 1-Br shown in Figure 1.
Exp aCalculations Geometry
S0
T1
Cu1–Cu22.772.862.59
Cu2–I12.732.712.67
Cu1–I12.672.762.76
Cu1–P22.252.362.36
Cu1–P12.252.362.35
Cu2–P32.252.332.37
Cu2–N12.102.222.03
I1–Cu1–I2107.6106.8102.5
I1–Cu2–I2107.1108.9103.4
Cu1–I1–Cu261.763.256.9
Cu1–I2–Cu262.762.355.9
P1–Cu1–Cu287.185.585.0
P2–Cu1–P1119.9119.7117.3
a—taken from ref. [34].
Table 3. SOC-TDDFT (B3LYP/DZP) vertical excitation energies [eV] and radiative lifetime calculated for 100% quantum yield in µs (in brackets) for the S1 state and the three substates of the T1 state of Cu2I2(Ph2Ppy)3, 1-I. Singlet-triplet splittings Δ(S1−T1) [cm−1] obtained from the average triplet energy and zero field splittings ZFS [cm−1] are also given. Results are presented for calculations at the S0 optimized geometry with and without solvent effects and at the T1 geometry in gas phase.
Table 3. SOC-TDDFT (B3LYP/DZP) vertical excitation energies [eV] and radiative lifetime calculated for 100% quantum yield in µs (in brackets) for the S1 state and the three substates of the T1 state of Cu2I2(Ph2Ppy)3, 1-I. Singlet-triplet splittings Δ(S1−T1) [cm−1] obtained from the average triplet energy and zero field splittings ZFS [cm−1] are also given. Results are presented for calculations at the S0 optimized geometry with and without solvent effects and at the T1 geometry in gas phase.
GeometryI(T1)II(T1)III(T1)S1Δ(S1−T1)ZFS
∆E(III−I)
S02.418 (752.4)2.420 (95.1)2.424 (6.5)2.437 (1.6)12952
S0(solvent)2.796
(92.5)
2.797 (8.1)2.801 (1.3)2.816 (0.2)14439
T11.452
(>103)
1.453 (>103)1.456 (54.4)1.507 (2.8)41132
Exp a2.54 (≈210)2.54 (≈210)2.54
(12)
≈2.588 b
(0.445)
3803
a—compare Figure 8, below. b—estimated from the blue flank of the emission spectrum at T = 1.3 K + Δ(S1-T1) = 380 cm−1 (Figure 5, below).
Table 4. Luminescence data of a series of P∩N linked Cu(I) dimer complexes.
Table 4. Luminescence data of a series of P∩N linked Cu(I) dimer complexes.
Compoundλmax(300 K) a
[nm]
ΦPL(300 K) b
[%]
τ(300 K) a
[µs]
kr(300 K) c
[s−1]
knr(300K) d
[s−1]
λmax(77 K) a
[nm]
ΦPL(77 K) b
[%]
τ(77 K) a
[µs]
kr(77 K) c
[s−1]
knr(77 K) d
[s−1]
Cu2Cl2(Ph2Ppy)31-Cl577377.94.7 × 1048.0 × 10459271651.1 × 1044.5 × 103
Cu2Br2(Ph2Ppy)31-Br545538.86.0 × 1045.3 × 104567891108.1 × 1031.0 × 103
Cu2I2(Ph2Ppy)31-I539816.51.25 × 1052.92 × 10455292322.88 × 1042.5 × 103
Cu2Cl2(Ph2Ppym)32-Cl61691.27.5 × 1047.6 × 10562614304.7 × 1032.9 × 104
Cu2Br2(Ph2Ppym)32-Br583332.51.3 × 1052.7 × 10558456291.9 × 1041.5 × 104
Cu2I2(Ph2Ppym)32-I565131.7 e (2.7 f)7.6 × 1045.1 × 1055756717.4 e3.9 × 1041.9 × 104
Cu2Br2(Ph2Piqn)33-Br660112.0 e5.5 × 1044.5 × 1056682442 e5.7 × 1031.8 × 104
Cu2I2(Ph2Piqn)33-I636383.3 e1.2 × 1051.9 × 10564559222.7 × 1041.9 × 104
a—excitation wavelength λexc = 372 nm; b—accuracy at 77 K: ± 10%, at 300 K: ± 5 % (relative error), (excitation wavelength λexc = 400 nm); c—determined according to kr = ΦPL/τ; d—determined according to knr = (1-ΦPL)/τ; e—the decay curve deviates from mono-exponential behavior. The lifetime given represents the main component; f—long component.
Table 5. Energies, emission quantum yields and decay data of Cu2I2(Ph2Ppy)3, 1-I (powder). Note the different temperatures.
Table 5. Energies, emission quantum yields and decay data of Cu2I2(Ph2Ppy)3, 1-I (powder). Note the different temperatures.
PropertyValue
E0-0(T1–S0) a20,500 cm−1
2.541 eV
E0-0(S1–S0) b20,880 cm−1
2.588 eV
ΦPL(300 K)81%
ΦPL(77 K)92%
k(S1–S0)2.25 × 106 s−1
(445 ns)
k(T1–S0), plateau3.1 × 104 s−1
(32 µs)
kr(T1–S0)2.88 × 104 s−1
(35 µs)
k(TADF + phos), 300 K
observed
15.4 × 104 s−1
6.5 µs
k(TADF-only), 300 K11.9 × 104 s−1
(8.4 µs)
∆E(S1–T1)380 cm−1
(47 meV)
∆E(II–I)<1 cm−1
∆E(III–I,II)3 cm−1
(0.37 meV)
k(III–S0)8.3 × 104 s−1
(12 µs)
k(I–S0) ≈
k(II–S0)
≈4.76 × 103 s−1
(≈210 µs)
k(SLR) c
at 1.2 K
1.7 × 104 s−1
(59 μs)
a—estimated from the blue flank of the emission spectrum at T = 1.3 (Figure 5); b—E0-0(T1–S0) + ∆E(S1-T1); c—spin-lattice relaxation rate between the triplet substates III and I, II at T = 1.2 K.
Table 6. Crystal data and data collection and structure refinement details for 1-Br, 2-I and 3-I.
Table 6. Crystal data and data collection and structure refinement details for 1-Br, 2-I and 3-I.
1-Br2-I3-I·CH2Cl2
Empirical formulaC51H42Br2Cu2N3P3C48H39Cu2I2N6P3·CH2Cl2C63H48Cu2I2N3P3·CH2Cl2
Mr, g mol−11076.691258.571405.78
Size, mm30.40 × 0.25 × 0.110.28 × 0.18 × 0.080.20 × 0.18 × 0.15
Crystal systemmonoclinicmonoclinictriclinic
Space groupP21/nC2/c P 1 ¯
a, Å14.202(2)42.494(5)12.4000(13)
b, Å18.139(3)11.0172(14)15.5224(17)
c, Å17.628(3)22.311(3)16.3296(17)
α, deg909097.462(12)
β, deg98.223(5)103.076(3)106.923(12)
γ, deg909092.752(13)
V, Å34494.5(11)10174(2)2969.3(6)
ρcalcd., g cm−11.5911.6431.572
Z482
μ(Mo-Kα), mm−12.872.291.97
T, K300293123
Θ range, deg1.6–25.13.0–28.52.2–27.0
Measured reflections27,86514,02742,450
Independent reflections7896951611,932
Reflections with I > 2σ(I))515176117253
Absorption correctionmulti-scanmulti-scananalytical
Tmin/Tmax0.39, 0.740.894, 0.9140.725, 0.769
Restraints/refined param.0/5500/6170/685
R1 (I ≥ 2σ(I))0.0470.0310.034
wR20.0920.0830.074
ρfin (max/min), e Å−3 1.15/−0.650.60/−0.541.81/−1.45
CCDC no.2,034,7802,034,7792,034,781
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hofbeck, T.; Niehaus, T.A.; Fleck, M.; Monkowius, U.; Yersin, H. P∩N Bridged Cu(I) Dimers Featuring Both TADF and Phosphorescence. From Overview towards Detailed Case Study of the Excited Singlet and Triplet States. Molecules 2021, 26, 3415. https://doi.org/10.3390/molecules26113415

AMA Style

Hofbeck T, Niehaus TA, Fleck M, Monkowius U, Yersin H. P∩N Bridged Cu(I) Dimers Featuring Both TADF and Phosphorescence. From Overview towards Detailed Case Study of the Excited Singlet and Triplet States. Molecules. 2021; 26(11):3415. https://doi.org/10.3390/molecules26113415

Chicago/Turabian Style

Hofbeck, Thomas, Thomas A. Niehaus, Michel Fleck, Uwe Monkowius, and Hartmut Yersin. 2021. "P∩N Bridged Cu(I) Dimers Featuring Both TADF and Phosphorescence. From Overview towards Detailed Case Study of the Excited Singlet and Triplet States" Molecules 26, no. 11: 3415. https://doi.org/10.3390/molecules26113415

Article Metrics

Back to TopTop