Next Article in Journal
Highly Luminescent Crystalline Sponge: Sensing Properties and Direct X-ray Visualization of the Substrates
Next Article in Special Issue
Fluorescence vs. Phosphorescence: Which Scenario Is Preferable in Au(I) Complexes with Benzothiadiazoles?
Previous Article in Journal
Ginsenoside Rb1 from Panax notoginseng Suppressed TNF-α-Induced Matrix Metalloproteinase-9 via the Suppression of Double-Strand RNA-Dependent Protein Kinase (PKR)/NF-κB Pathway
Previous Article in Special Issue
Photophysical Study on the Rigid Pt(II) Complex [Pt(naphen)(Cl)] (Hnaphen = Naphtho[1,2-b][1,10]Phenanthroline and Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Triplet Emitting C^N^C Cyclometalated Dibenzo[c,h]Acridine Pt(II) Complexes

1
Universität zu Köln, Fakultät für Mathematik und Naturwissenschaften, Department für Chemie, Institut für Anorganische Chemie, Greinstraße 6, 50939 Köln, Germany
2
Institut für Anorganische und Analytische Chemie, CiMIC, CeNTech, Westfälische Wilhelms-Universität Münster, Heisenbergstraße 11, 48149 Münster, Germany
3
Institut für Festkörpertheorie and Center for Multiscale Theory and Computation, Westfälische Wilhelms-Universität Münster, Wilhelm-Klemm-Straße 10, 48149 Münster, Germany
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(22), 8054; https://doi.org/10.3390/molecules27228054
Submission received: 28 October 2022 / Revised: 11 November 2022 / Accepted: 15 November 2022 / Published: 19 November 2022
(This article belongs to the Special Issue Photochemistry and Photophysics of Metal Complexes)

Abstract

:
In a series of Pt(II) complexes [Pt(dba)(L)] containing the very rigid, dianionic, bis-cyclometalating, tridentate C^N^C2− heterocyclic ligand dba2– (H2dba = dibenzo[c,h]acridine), the coligand (ancillary ligand) L = dmso, PPh3, CNtBu and Me2Imd (N,N’-dimethylimidazolydene) was varied in order to improve its luminescence properties. Beginning with the previously reported dmso complex, we synthesized the PPh3, CNtBu and Me2Imd derivatives and characterized them by elemental analysis, 1H (and 31P) NMR spectroscopy and MS. Cyclic voltammetry showed partially reversible reduction waves ranging between −1.89 and −2.10 V and increasing along the series Me2Imd < dmso ≈ PPh3 < CNtBu. With irreversible oxidation waves ranging between 0.55 (L = Me2Imd) and 1.00 V (dmso), the electrochemical gaps range between 2.65 and 2.91 eV while increasing along the series Me2Imd < CNtBu < PPh3 < dmso. All four complexes show in part vibrationally structured long-wavelength absorption bands peaking at around 530 nm. TD-DFT calculated spectra agree quite well with the experimental spectra, with only a slight redshift. The photoluminescence spectra of all four compounds are very similar. In fluid solution at 298 K, they show broad, only partially structured bands, with maxima at around 590 nm, while in frozen glassy matrices at 77 K, slightly blue-shifted (~580 nm) bands with clear vibronic progressions were found. The photoluminescence quantum yields ΦL ranged between 0.04 and 0.24, at 298 K, and between 0.80 and 0.90 at 77 K. The lifetimes τ at 298 K ranged between 60 and 14040 ns in Ar-purged solutions and increased from 17 to 43 µs at 77 K. The TD-DFT calculated emission spectra are in excellent agreement with the experimental findings. In terms of high ΦL and long τ, the dmso and PPh3 complexes outperform the CNtBu and Me2Imd derivatives. This is remarkable in view of the higher ligand strength of Me2Imd, compared with all other coligands, as concluded from the electrochemical data.

Graphical Abstract

1. Introduction

Studies on luminescent transition metal complexes are motivated by their wide range of potential applications, in fields such as photocatalysis [1,2], sensing [3,4,5], optoelectronic devices [5,6,7,8,9,10,11], and biomedicine [4,5,12,13,14,15]. Phosphorescent metal complexes are of particular interest in the field of OLED (Organic Light Emitting Diodes) applications, due to the ability of these materials to harvest all generated excitons in electroluminescent devices, by means of efficient spin-orbit coupling (SOC) promoting intersystem crossing and phosphorescence, which is associated with the heavy metal centers [5,7]. Frequently, metal-containing phosphorescent emitters involve Ir(III), Ru(II) or Pt(II) centers [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. While the d6 species, such as Ir(III) and Ru(II), often adopt an octahedral coordination environment, metals with a d8 electron configuration (such as Pt(II), Pd(II) and Au(III), among others) often display square-planar geometries with open coordination flanks in the axial positions [4,16,17]. These open axial positions can lead to metal-metal (M···M) and/or π-stacking interactions upon aggregation with red-shifted emission from MMLCT states (metal-metal-to-ligand charge transfer, eventually with excimeric M...M shortening), as opposed to the LC (ligand-centered) character found in monomeric species [5,11,18,19,20,21,22].
Two main structural features have been discovered to be beneficial for efficient triplet luminescence in Pt(II) complexes. The first is rigid coordination environments, which prevent radiationless decay from the triplet excited states; the second is the use of polydentate heteroaromatic cyclometalated chromophores [5,9,16,17,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]. Such ligands energetically disfavor the thermal population of “dark” (dissociative) d-d* excited states, which are responsible for non-radiative decay paths via conical intersections with the ground state, and they also provide multiple options for the character of the excited triplet states (such as the metal(d8)-to-ligand(π*) charge transfer (MLCT) configurations, intraligand (π–π*) or charge-transfer excitations involving donor and acceptor parts of the ligand(s) called LL’CT states as well as mixtures thereof) [5,8,10,11,16,17,21]. Thus, a number of Pt(II) complexes containing tri- or tetra-dentate multifunctional (C-bonding, N-bonding, further substituents) ligands were studied previously [4,5,17,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43]
Amongst the simplest tridentate ligands, variations of the C^N^N, N^C^N and C^N^C cores based on phenyl(C) and pyridyl(N) donors have been reported with the double cyclometalation C^N^C having no marked benefit over the C^N^N or N^C^N coordination in terms of the efficiency of the triplet photoluminescence [33,45,46,47]. On the other hand, the doubly anionic C^N^C2− ligands allow researchers to choose from a variety of neutral L coligands (ancillary ligands), whereas the C^N^N and N^C^N derivatives require anionic coligands to yield neutral cyclometalated Pt(II) complexes. The parent 2,6-diphenyl-pyridine (dppy) type of C^N^C ligands (see Scheme 1, i.e., the derivatives [Pt(dppy)(L)] (H2dppy = 2,6-diphenyl-pyridine, L = dimethyl sulfoxide (dmso), pyridine (Py), 4-tBu-Py, MeCN, PPh3, PCy3, CNXyl (Xyl = 2,6-xylyl), Me-4,4′-bipyridinium, and the PPh2-crown ether-based species) showed no photoluminescence (PL) at ambient T in the solution [33,43,45,46]. In contrast to this finding, PL at ambient T in the solution was reported for L = pyridine derivatives with pending crown ether moieties [46] On the other hand, the expansion of one phenyl group to naphthyl and the central pyridine to 4-phenyl-pyridine did not yield PL at 298 K in the solution for L = dmso or CNXyl, while the replacement of the peripheral phenyl groups by thiophenyl or N-alkyl carbazolyl moieties reached efficient PL intensities under these conditions [47]. In line with the rather poor performance of the simple dppy ligand, very recently, the complexes [Pt(L)(dmso)] with the same type of flexible C^N^C ligands (H2L = 2,6-di(phenyl)-4-(3,4,5-X-phenyl)-pyridine, X = H or F), were reported as being non-emissive while showing very interesting four-photon absorption (4PA), which can be used for medical diagnosis [48]. A very recent theoretical work on these complexes focussed on the 4PA singlet-singlet excitation, followed by ISC to the triplet states T2 and T1 [49].
In contrast to the dppy system, in which the C–C single bonds between the phenyl and the pyridyl groups provides some flexibility for deformations (Scheme 1, left), the condensation of the three aromatic rings in the dba2− (H2dba = dibenzo[c,h]acridine) ligand (Scheme 1, right) led to a far more rigid coordination environment. In our previous study, we found markedly increased PL quantum yields ΦL for the complex [Pt(dba)(dmso)] (0.23 at 298 K in CH2Cl2 solution), compared with derivatives of [Pt(dppy)(dmso)] [33] (no emission under these conditions), which is in line with the idea that molecular rigidity helps to reduce or circumvent radiationless decay paths. This makes these Pt dba complexes interesting candidates for various applications.
We recently reported a similar strategy for C^N^N and N^C^N cyclometalated Pt(II) complexes, including the complexes [Pt(naphen)(X)] (Hnaphen = naphtho[1,2-b][1,10]phenanthroline, X = Cl or C≡CPh) with the rigid tridentate C^N^N-coordinating pericyclic naphen ligand, the slightly more flexible tetrahydro-derivative [Pt(thnaphen)(X)] (Hthnaphen = 5,6,8,9-tetrahydro-naphtho[1,2-b][1,10]phenanthroline), and the N^C^N-coordinated complex [Pt(bdq)(Cl)] (Hbdq = benzo[1,2-h:5,4-h’]diquinoline [55]. The rigid ligand backbones in these complexes, compared with the more flexible parent 6-phenyl-2,2′-bipyridyl (C^N^N) or dipyridyl-benzene (N^C^N) derivatives, leads to red-shifted photoluminescence with longer lifetimes and higher ΦL.
In a number of previous studies, the coligand (ancillary ligand) L in Pt(II) complexes with tridentate C^N^N, N^C^N, or C^N^C chromophores showed a marked impact on the luminescence properties [17,28,45,46,47,49,50,51,52,53,54,55,56]. This represents one of the benefits of the tridenate coordination approach for luminescent cyclometalated Pt(II) complexes, in comparison with the use of one tetradentate [17,22,23,24,25,42,43,52,53] or two bidentate ligands [5,11,39,51,57].
Beginning with the previously reported complex [Pt(dba)(dmso)] (1) [33], we sought to improve the photoluminescence properties by exchanging the dmso coligand for PPh3, CNtBu (tert-butyl-isocyanide), or Me2Imd (N,N’-dimethylimidazolydene) (complexes 2, 3, and 4; Scheme 2). All three of the new coligands are expected to produce a stronger ligand field splitting (or ligand field stabilization energy) due to their superior σ-donating and π-accepting capabilities. This should destabilize the dark d-d* excited states further, and thus lead to superior luminescence quantum yields for the corresponding complexes. In addition to UV-vis absorption and time-resolved photoluminescence spectroscopy, we also studied the electrochemistry of the complexes, using cyclic voltammetry and spectroelectrochemical UV-vis spectroscopy, as well as modelling the excited states and electronic transitions employing (TD)-DFT calculations. The 7-phenyl substituted derivatives [Pt(db(ph)a)(L)] (H2db(ph)a = 7-phenyldibenzo[c,h]acridine) were also included in this study for comparison, as the experimental data for L = dmso (complex 5) was already available [33], while further derivatives with L = PPh3, CNtBu, and Me2Imd (6, 7, and 8) were calculated to probe if the 7-phenyl derivative of dba is potentially more interesting for phosphorescent Pt(II) complexes.

2. Results and Discussion

2.1. Synthesis, Analytical Characterization and Molecular Structures

The two-step synthesis of the complex [Pt(dba)(dmso)] from the H2dba protoligand and K2PtCl4 has previously been described [33]. The [Pt(dba)(dmso)] (1) was reacted with PPh3, CNtBu and Me2Imd (N,N’-dimethylimidazolydene) to yield the three new complexes [Pt(dba)(L)] in excellent yields (93 to 98%) (Scheme 3, further details are found in the Section 3).
All four compounds gave single crystals suitable for XRD structure analysis, and the crystallographic data is shown in Table S1 of the Supplementary Material (SM). All of the structures, with the exception of [Pt(dba)(Me2Imd)] (4), are characterized by pronounced head-to-tail π-stacks of two molecules each (Figure S1 to Figure S3, SM), which is typical for square planar complexes with planar tridentate or tetradentate ligands [19,21,25,26,27,28,29,33,34,35,36,39]. For 4, the carbene ligand interrupts this interaction and a complex packing motive is observed (Figure 1 and Figure S4).
A closer inspection of the π-stacks showed that the centroid...centroid distance of the central pyridyl ring of [Pt(dba)(dmso)] constitutes approximately 3.62 Å, with an offset of approximately 1.25 Å, which lies in the range of typical π-interactions [45,55,56,58,59]. In contrast, for the PPh3 and the CNtBu derivatives 2 and 3, the bulky coligands enlarge these centroid...centroid distances, to 4.53 Å (0.93 Å offset) for 2 and to 5.59 Å (1.12 Å offset) for 3. Therefore, dimerization through stacking in solution might have an impact on the photophysical properties only for the dmso complex (1). On the other hand, our previous study did not give evidence for this phenomenon [33].
The molecular structures show the expected square planar coordination of the d8 configured Pt(II) (Figure 1). The dba ligand is also completely planar, and the S, P, or C donor atoms of the ancillary ligands complete the planar coordination with N–Pt–X angles (α angle, Figure 1) of around 180°, while the residual atoms of the coligands fill the space at this fourth coordination site. The planar Me2Imd ligand in 4 shows a tilt angle towards the Pt,C,C,N coordination plane of about 58° (δ angle). The Pt–C–N unit (γ angle) in the CNtBu complex 3 is almost linear (Table 1).
The optimized geometries of the singlet ground state S0 (Figure 1 and Figure S5) and the lowest excited triplet state T1 (Figure S6) were calculated by DFT methods. The S0 geometries are in good agreement with the experimental data from the X-ray diffractometry (Table 1). The S0 and T1 geometries hardly differ between all of the complexes; the largest differences occurred for the dmso and PPh3 derivatives, where the N–Pt–X and C–Pt–X angles change by a few degrees upon relaxation in the T1 state. In the case of PPh3, the planarity of the complex is broken in the T1 state, with the Pt(II) center clearly protruding out of the coordination plane. The Pt1–C1–C6–C10 dihedral angle (Figure 1) measures 7°.

2.2. Electrochemistry and DFT-Calculated Orbitals

On the first view, the complexes show irreversible oxidation processes ranging between around 0.55 and 1 V (Figure 2 and Figure S7, data in Table 2); based on the DFT calculations, they can be assigned to an essentially Pt-centered Pt(II)/Pt(III) couple. The DFT-calculated highest occupied molecular orbitals (HOMO) are energetically ordered in growing order as Me2Imd > PPh3 > CNtBu > dmso (Figure S8). The electrochemical potentials increase along the same series for the dba complexes, which roughly match the expected increasing σ-donating strength of the coligand.
A first reduction is observed for all of the complexes at around −2 V. This process is partly reversible for the complexes 2 to 4, but irreversible for the dmso complex 1 (Figure S7). Regardless of the slightly different reversibility, the very similar potentials point to a largely ligand π*(dba) centered reduction processes. This is in line with the DFT-calculated lowest unoccupied molecular orbitals (LUMO), which show the growing energetic sequence Me2Imd > PPh3 > CNtBu > dmso (Figure S8); this correlates with the expected higher π-backbonding ability of the newly introduced coligands compared to dmso.
The irreversibility of both the oxidation and reduction processes were explained by rapid chemical reactions (C) following the electrochemical processes (E). If the two steps are timely correlated, this is called an EC mechanism or EC process [31,33,36,60,61]. Thus, secondary yet small, reduction waves found at −2.8 to −3.0 V represent the EC products. The introduction of the pending phenyl group led to markedly higher reduction and oxidation potentials for [Pt(db(ph)a)(dmso)], which is in line with the contribution of the dba core to the HOMO.
The electrochemical gaps increase along the series of dba complexes Me2Imd < CNtBu < PPh3 < dmso following the trend of decreasing ligand strength (Table 2).

2.3. UV-Vis Absorption Spectroscopy

The experimental UV-vis absorption spectra of the four [Pt(dba)(L)] complexes are very similar (Figure 3 and Figure S9). Four series of absorption maxima are discernible with very intense bands between 250 and 300 nm, along with two vibrationally structured medium-intense groups in the range between 300 and 420 nm, as well as a fourth structured band system between 450 and 550 nm that tails down to cut-offs (zero absorption) between 570 and 580 nm (Table 3).
The long-wavelength absorption and cut-off energies are only slightly smaller (by 0.03 to 0.05 eV) for the complexes with the new coligands, 2, 3, and 4, compared with the parent dmso complex 1 (Table 3) and vary only slightly for these three complexes, in contrast to the marked differences in the electrochemical potentials and gaps (Table 2). The corresponding intensities are very different and the CNtBu complex reaches not even half the intensities of the PPh3 derivative. Nevertheless, the similar band shapes and energies point to comparable excited states; we assume transitions into π–π* configurations for the two high-energy band systems (200 to 350 nm) and transitions into mixed π–π*(LC)/metal(d)-to-ligand(π*)(MLCT) states for the two long-wavelength absorption processes (350 to 550 nm), in line with previous reports [10,11,16,17,18,19,20,21,45,46,47,49,50,51,52,53,54,55,56,62]. We infer a relatively small metal participation from the rather low impact of the coligand exchange.
The recently reported complexes [Pt(naphen)(X)] (Hnaphen = naphtho[1,2-b][1,10]phenanthroline, X = Cl or C≡CPh) containing the rigid tridentate C^N^N-coordinating pericyclic naphen ligand and the derivatives [Pt(thnaphen)(X)], bearing the slightly more flexible thnaphen unit (Hthnaphen = 5,6,8,9-tetrahydro-naphtho[1,2-b][1,10]phenanthroline), show long-wavelength absorption bands between 523 and 534 nm [55], and thus appear comparable to the dba complexes. In contrast to this, the N^C^N-coordinated complex [Pt(bdq)(Cl)] containing the rigid bdq ligand (Hbdq = benzo[1,2-h:5,4-h’]diquinoline) absorbs markedly blue-shifted (493 nm), which is reasonable for the comparison between the C^N^N and the N^C^N coordination pattern [16,36,55,63].

2.4. TD-DFT Calculated Absorption Spectra

The most intense calculated transitions are found in the range between 225 and 325 nm (Figure 3, left). The maxima for the complexes with L = dmso and Me2Imd show very good agreement with the experimental data, while deviating by only 4 to 9 nm, whereas for PPh3 and CNtBu, the deviation is somewhat more pronounced (ca. 25 nm).
The main peak of the [Pt(dba)(dmso)] complex has a maximum of approximately 283 nm and corresponds to a transition into a state with dominant contributions from HOMO−1→LUMO+1 (32%), HOMO−3→LUMO+3 (10%) and HOMO−7→LUMO excitations (10%). The [Pt(dba)(PPh3)] complex has its highest peak at 266 nm, leading to an excited state with contributions from HOMO−1→LUMO+1 (48%), HOMO−5→LUMO (13%), and HOMO−7→LUMO (13%) excitations. For [Pt(dba)(Me2Imd)], this peak is located at 277 nm and is associated with a state described with contributions of HOMO−2→LUMO+2 (23%), HOMO−7→LUMO (23%), HOMO−3→LUMO+2 (14%), and HOMO-6→LUMO (12%) excitations. For [Pt(dba)(CNtBu)], the maximum can be found at 282 nm, and it is dominated by a transition into an excited state with contributions from HOMO−1→LUMO+2 (51%), HOMO→LUMO+3 (25%) and HOMO−7→LUMO (14%) excitations.
When comparing the TD-DFT-calculated UV-vis absorption spectra of the complexes [Pt(dba)(L)] and [Pt(db(ph)a)(L)] (5 to 8), we found a red-shift of the main maximum when going from PPh3 (6) via CNtBu (7) to Me2Imd (8) and dmso (5) and a coincidence of the main peaks for the Me2Imd and DMSO complexes in both cases. In addition, for the long-wavelength absorption bands, there is only a very small difference between the spectra of the [Pt(dba)(L)] and [Pt(db(ph)a)(L)] complexes (Figure 3, right).

2.5. Time-Resolved Photoluminescence Spectroscopy

At room temperature and in diluted CH2Cl2 solutions, the four dba Pt(II) complexes 1 to 4 show a red luminescence with a broad band centered at around 600 nm (Figure S10), most likely arising from the mixed 3LC/3MLCT excited states; this is in line with previous reports [10,11,16,17,18,19,20,21,45,46,47,49,50,51,52,53,54,55,56,62], as well as with our TD-DFT calculations (vide infra). The change of the monodentate coligand does not substantially influence the emission profile of the complexes; only the vibrational progression is marginally affected. The absence of 3O2 in the Ar-purged solutions generally increased the lifetime of the complexes, compared to the air-equilibrated solutions, in line with a triplet character of the luminescence (Table 4). The most remarkable finding was that the replacement of L = dmso by the stronger ligands PPh3, CNtBu, and Me2Imd had almost no effect on the emission energies and a rather detrimental effect on ΦL, which decreases along the series of coligands dmso >> CNtBu >> Me2Imd > PPh3 and is generally lower for the complexes with stronger coligands L than for the parent complex [Pt(dba)(dmso)]. In the more concentrated solutions, we observed red-shifted emissions which were likely a result of the aggregation of the molecules. However, in this report we wanted to focus on the luminescence properties of the isolated molecules and avoided aggregation by using low concentrations (c = 10−5 M). In future studies, we will investigate the aggregation of these complexes in solution and in solid materials.
The short-lived component of the adjusted biexponential decays for τ in air-equilibrated CH2Cl2 solutions are due to traces of the free H2dba ligand (τdba ≈ 5 ns) (Table 4). However, in most cases, this effect was not seen in Ar-purged solution due to the vast increment of ΦL for the complexes under these conditions. The fluorescence quantum yield of dba is not negligible in CH2Cl2 at 298 K (ΦL = 0.44); thus, even traces of the ligand are detected in case of the poorly phosphorescent complexes in the presence of oxygen. Therefore, a trace amount of the ligand below 0.01% might escape the analysis by NMR, elemental analysis and other methods, while still being detectable through intrinsic fluorescence. Importantly, the studied dba Pt(II) complexes were purified by column chromatography, and NMR spectra show no signals for the H2dba precursor (with the exception of the PPh3 complex, see Figures S31–S35 in the Supporting Material). We assume that, in addition to the complexes as molecular species, small stacks (see crystal structures) of complexes with the ligand precursor H2dba are also eluted. Our findings also emphasize that special care has to be taken when measuring weakly phosphorescent complexes stemming from a brightly fluorescent ligand precursor, in order to avoid potential confusion with dual emission [55].
In frozen glassy matrices at 77 K, the charge-transfer states are destabilized, which lowers the MLCT character of the emissive T1 state, while causing a blue-shifted photoluminescence (576 nm for the dmso complex). Therefore, the observed emission at 77 K (Figure 4) predominantly originates from ligand-centered triplet states (3LC), as reflected by the enhanced vibrational progression, compared with 298 K (Figure S10). Under these conditions, all of the complexes show prolonged excited state lifetimes and quantum yields between 0.8 and 1.0 due to a diminished knr (Table 4). In addition, it is worth noting that at 77 K, diverse comparably stable conformations can be locked in the glassy matrix, leading to biexponential decays, as was previously observed for other Pt(II) complexes [22,31,32,55,56,62].
In addition, ΦL and τ at 77 K were used to determine the radiative (kr) and radiationless (knr) deactivation rate constants (Table 4), according to equations 1 to 4, and assuming a unitary intersystem crossing efficiency for S1TnT1, the following relations are valid
k r = Φ L τ
k nr = k ISC
τ = 1 k r + k nr  
k nr = 1 Φ L τ  
where kIC is the internal conversion rate constant and kISC′ is the intersystem crossing rate constant (T1S0). For biexponential decays, amplitude-weighted average lifetimes (τav_amp) were used to estimate the rate constants [64].
Depending on the nature of the coligand, minor changes in both kr and knr are found (Table 4). In a frozen glassy matrix, at 77 K, the complexes are trapped in their most stable conformation(s) within the solid and the differences in solvation environments and rotovibrational relaxation effects are diminished; therefore, only minor differences in kr and knr can be found.
For the complexes [Pt(naphen)(X)] (X = Cl or C≡CPh) with the rigid C^N^N-coordinating naphen ligand, or the slightly more flexible tetrahydro-derivative [Pt(thnaphen)(X)], similar broad PL spectra were found at 298 K in the CH2Cl2 solution, peaking at 630 nm for the naphen and 570 nm for the thnaphen derivatives [55], i.e., slightly red-shifted with respect to the dba complexes. The N^C^N-coordinated complex [Pt(bdq)(Cl)] containing the rigid bdq ligand shows a slightly more pronounced vibronic structure and a maximum at 568 nm [55]. The quantum yields ΦL ranged between 0.08 and 0.32, and are similar to the dba complexes.
At low T (77 K, frozen glassy matrices), vibrationally structured emission profiles were found for these complexes, and they appear similar to those observed for the dba complexes, but with slightly higher emission energies for the naphen (600 nm) and thnaphen (530 nm) complexes, yet with almost no shift for the bdq complex. The ΦL values were close to 1 for these complexes [55]. Taking into account the photoluminescence lifetime of the [Pt(naphen)(Cl)] complex, we assign the phosphorescence to mixed 3LC/3MLCT excited states with only a small metal contribution and dominant 3LC character [55].

2.6. DFT-Calculated Emission Spectra

The calculated vibrationally resolved emission spectra (red lines) and the zero-point corrected 0-0 emission energies (green dashed lines) are in very good agreement with the experimental data (blue lines). For all of the complexes (Figure 5, left), the main peak and the most intense shoulder deviate by less than 10 nm (except for PPh3). Nonetheless, for the calculated 0-0 energy of the PPh3, the agreement is still reasonable. A vibrationally resolved emission spectrum could not be computed for PPh3 due to numerical issues (see Computational Details).
When comparing the complexes [Pt(dba)(L)] and [Pt(db(ph)a)(L)] for the different coligands (Figure 5, right), the main peak shifts to shorter wavelengths (for both cases) when the dmso ligand is replaced by Me2Imd. In the case of [Pt(db(ph)a)(L)], the main emission peak for PPh3 lies inbetween the other two. Interestingly, the vibronic coupling pattern is very different for the two C^N^C chelate ligands and the change from dba to db(ph)a causes a general, yet slight, red-shift.
The calculated orbitals contributing to the description of the emissive T1 state show that the HOMO and LUMO are partially localized on the metal ion, as well as on the ligand scaffold (Figure 6). For Me2Imd, partial contributions from the coligand can also be identified. In general, the occupied orbitals have higher metal contributions than their unoccupied counterparts. For the dmso complex, exclusively, the T1 state can be described as a predominant HOMO→LUMO excitation, which involves a significant amount of charge transfer between the metal and the ligand (Figure 6).
However, for all of the complexes, the T1 state is composed of a number of monoelectronic excitations. For the dmso complex, there are two main contributions to the T1 state, namely, HOMO→LUMO (44.1%) and HOMO−1→LUMO (37.6%). In addition, there is a small contribution from the HOMO→LUMO+1 excitation (5.3%). The PPh3 derivative shows a major contribution (65.9%) from the HOMO−1→LUMO excitation. Other smaller contributions come from HOMO−3→LUMO (7.6%) and HOMO→LUMO+1 (5.7%). The Me2Imd complex exhibits two approximately equal contributions from HOMO−2→LUMO (32.8%) and HOMO−1→LUMO (29.5%), as well as a significant contribution from the HOMO→LUMO+1 (14.1%) excitation. In the case of CNtBu, the main contributions stem from the HOMO−1→LUMO (47.1%), HOMO→LUMO (33.3%), and HOMO→LUMO+1 (8.5%) excitations. Generally speaking, for all of the complexes, the character of the T1 state can be described as a mixture of intraligand LC and MLCT character; this is in line with our initial assumptions.
A quantitative description of the T1 character was obtained through a correlated electron-hole pair analysis [65]. The resulting decomposition into LC (π−π*), MLCT (d−π*), coligand-to-dba charge transfer (L’LCT), coligand-to-metal charge transfer (L’MCT), and MC (d−d*) contributions (Figure 7) is based on a partitioning of the complexes into the dba ligand, the metal and the coligand (as shown for 3 in Figure 7).
Overall, the analysis shows that all of the complexes exhibit a similar character for the T1 state. The dmso, PPh3, Me2Imd, and CNtBu complexes all predominantly show a LC character (86.8%, 76.3%, 76.6%, and 91.1%, respectively), and only a small degree of CT character. The proportion of the MLCT character increases from dmso (4.8%) via CNtBu (5.4%) and PPh3 (6.8%) to Me2Imd (7.9%).
The high LC character of the excited states explains our spectroscopic findings and helps us to understand the marginal influence of the coligands L on the PL properties. However, it cannot explain why the change from dmso to the stronger σ-donating and π-accepting ligands PPh3, CNtBu and Me2Imd drastically reduces the ΦL at 298 K, namely, from a remarkable 0.23 (dmso) to almost zero for PPh3. At the same time, at 77 K, the introduction of the new ligands led to increased ΦL values, with the highest performance observed for the PPh3 complex. In future work, we will seek insights into the dynamics of the excited states by dedicated calculations. In fact, the results depicted in Figure 7 only show the composition concerning the initial excited states [65], which might change over time, as recent work on related luminescent complexes has revealed [66,67].

2.7. Spectroelectrochemistry (SEC)

The UV-vis spectroscopic response upon cathodic reduction and anodic oxidation (SEC) was studied for all of the for complexes in the THF/n-Bu4NPF6 solution (Figure 8, more figures and data in the SM).
First of all, the SEC experiments show that the initial reductions and oxidations of all four compounds occur, irreversibly, on the timescale of minutes. In the CV experiments, the first reduction showed some degree of reversibility, but the CV experiment is on a timescale of seconds. Nevertheless, the spectral responses of the compounds upon reduction or oxidation leave the impression that defined EC (electrochemical + subsequent chemical reaction) [61,68,69,70] reactions occur. This assumption is derived from the almost isosbestic character of the produced spectral traces (Figure 8).
Upon reduction, a pronounced band, at around 435 nm, was observed for the set of reduced species, while long-wavelength bands at around 650 nm were only observed for the dmso (1) and Me2Imd (4) complexes. Assuming a ligand π*-centered reduction, we were surprised to see no long-wavelength transitions into π*−π* states, which are usually observed for complexes with reduced heteroaromatic ligands [27,36,68,69,70,71,72]. They most likely occur for the dba system at wavelengths below 1100 nm, beyond the detection limit of our spectrometer.
The oxidation of the four complexes leaves UV-vis absorption traces that are very dissimilar, thus confirming the mixed metal/coligand character of the oxidation process. Importantly, no evidence for stable or transient Pt(III) species was observed, in line with a rapid chemical reaction after oxidation (EC). Bands for Pt(III) species are expected in the range 500 to 1800 nm [72].

3. Experimental Section

3.1. General Information

Commercially available chemicals were purchased from Sigma-Aldrich, Acros, ABCR or Fisher-Scientific and were used without further purification. Dry THF was obtained from distillation over sodium/potassium alloy. All reactions and measurements involving metal complexes were conducted under argon, using standard Schlenk techniques. The complex [Pt(dba)(dmso)] (1) was synthesized and characterized as previously described [33]. Finally, 1,3-dimethylimidazolium iodide was prepared from commercially purchased methylimidazole and methyl iodide as described in the literature [73].

3.2. Syntheses

3.2.1. Synthesis of [Pt(dba)(PPh3)] (2)

A mixture of 40.14 mg [Pt(dba)(dmso)] (0.073 mmol, 1 eq.) and PPh3 (20.65 mg, 0.079 mmol, 1.1 eq.) was dissolved in 30 mL CH2Cl2 and the reaction mixture was stirred for 96 h at ambient temperature. The solvent was evaporated under reduced pressure and a red solid was obtained. The purification was conducted by column chromatography on silica (CH2Cl2/n-hexane 1:1). The product was obtained as a red solid (56.25 mg, 0.071 mmol, 97%). Elemental analysis found (calculated for C39H26NPPt, MW = 734.68 g mol−1): C, 63.78 (63.76); H, 3.55 (3.57); N, 1.92 (1.91). 1H NMR (600 MHz, CD2Cl2): δ (ppm) = 8.50 (s, 1H), 7.96 (t, J = 6.3 Hz, 6H), 7.63 (d, J = 6.3 Hz, 2H), 7.58 (d, J = 8.9 Hz, 2H), 7.50 (t, J = 7.3 Hz, 3H), 7.44 (t, J = 7.5 Hz, 6H), 7.38 (d, J = 7.7 Hz, 2H), 7.04 (t, J = 7.5 Hz, 2H), 6.29 (d, 2H). 31P NMR (234 MHz, CD2Cl2): δ (ppm) = 25.66 (s, 1JPt-P = 8209 Hz). HR-ESI-MS(+) (MeOH) calculated for [M]+ m/z = 734.1454, found 734.15015; calculated for [dba+H]+ m/z = 280.112624, found: 280.113714. UV-vis absorption (CH2Cl2): λ (nm) ε (103 M−1cm−1) = 296 (50.2), 347 (12.7), 409 (6.2), 532 (2.1).

3.2.2. Synthesis of [Pt(dba)(CNtBu)] (3)

A mixture of 40.14 mg [Pt(dba)(dmso)] (0.073 mmol, 1 eq.) and CNtBu (8.3 μL, 0.073 mmol, 1 eq.) was dissolved in 20 mL CH2Cl2 and stirred for 24 h at 50 °C. The solvent was evaporated under reduced pressure. The product was obtained as a red solid (39.7 mg, 0.072 mmol, 98%). Elemental analysis found (calculated for C26H20N2Pt, MW = 555.53 g mol−1): C, 56.22 (56.21); H, 3.65 (3.63); N, 5.08 (5.04). 1H NMR (600 MHz, CD2Cl2): δ (ppm) = 8.36 (s, 1H), 7.68 (dd, J = 2.6 Hz, 2H), 7.57 (d, J = 8.9 Hz, 2H), 7.47 (d, J = 8.9 Hz, 2H), 7.45–7.41 (m, 4H), 1.67 (s, 9H). HR-ESI-MS(+) (MeOH) calculated for [M]+ m/z = 555.127440, found 555.13287. UV-vis absorption (CH2Cl2): λ (nm) ε (103 M−1cm−1) = 294 (27.0), 343 (5.2), 408 (2.4), 533 (1.0).

3.2.3. Synthesis of [Pt(dba)(Me2Imd)] (4)

A solution of 40.14 mg [Pt(dba)(dmso)] (0.073 mmol, 1 eq.) was dissolved in 20 mL MeCN. Freshly synthesized 1,3-dimethylimidazolium iodide (17.2 mg, 0.077 mmol, 1.05 eq.) and KOtBu (9.83 mg, 0.088 mmol, 1.2 eq.) were added. The solution was heated under reflux at 90 °C for 16 h. After the solvent was reduced to the half volume, the product was precipitated by addition of 20 mL water. After filtration, the crude solid was washed with toluene to obtain the product as a red solid (38.61 mg, 0.068 mmol, 93%). Elemental analysis found (calculated for C26H19N3Pt, MW = 568.53 g mol−1): C, 54.95 (54.93); H, 3.38 (3.37); N, 7.39 (7.39). 1H NMR (600 MHz, CD2Cl2): δ (ppm) = 8.35 (s, 1H), 7.55 (d, J = 8.9 Hz, 2H), 7.49 (d, J = 8.9 Hz, 2H), 7.40 (d, J = 7.8 Hz, 2H), 7.30 (t, J = 7.3 Hz, 2H), 7.23 (d, J = 6.8 Hz, 2H), 7.01 (t, J = 5.4 Hz, 2H), 5.78 (s, 6H). EI-MS(+) (MeOH) calculated for [M]+ m/z = 568.122689, found: 568; calculated for [Ptdba]+ m/z = 472.053940, found 472; calculated for [dba]+ 279.104799, found 279. UV-vis absorption (CH2Cl2): λ (nm) ε (103 M−1cm−1) = 300 (24.8), 352 (8.9), 406 (3.2), 536 (1.3).

3.3. Instrumentation

1H and 31P NMR spectra were recorded on a Bruker Avance II 600 MHz spectrometer (1H: 600.13 MHz, 31P: 242.88 MHz) with a triple resonance (TBI) 5 mm inverse probehead with z-gradient coil using a triple resonance. Chemical shifts are relative to TMS, respectively. UV-vis absorption spectra were recorded with Varian Cary 05E or Cary 50 scan spectrophotometers. Photoluminescence spectra at room temperature and at 77 K were recorded with a PicoQuant FluoTime 300. The same equipment was used for lifetime measurements. Lifetime analysis was performed using the commercial EasyTau 2.2 software. The quality of the fit was assessed by minimising the reduced chi-squared function. Photoluminescence quantum yields were determined with a Hamamatsu Photonics absolute PL quantum yield measurement system (C9920–02), equipped with a L9799-01 CW Xenon light source, monochromator, photonic multichannel analyser and integrating sphere (error of maximum ± 2% for ΦL is estimated). All of the solvents were of spectroscopic grade and were degassed prior to use. Elemental analyses were obtained using a HEKAtech CHNS EuroEA 3000 analyzer. EI-MS spectra were measured with a Finnigan MAT 95, and HR-ESI-MS using a THERMO Scientific LTQ Orbitrap XL. MS Simulations were performed using ISOPRO 3.0. Electrochemical measurements were carried out in 0.1 M n-Bu4NPF6/THF solution using a three-electrode configuration (glassy carbon working electrode, Pt counter electrode, Ag/AgCl reference electrode) and a Metrohm Autolab PGSTAT30 potentiostat and function generator. The ferrocene/ferrocenium couple served as internal reference.

3.4. X-ray Diffractometric Analysis of Single Crystals for Structure Determination

The crystals of 1 to 4 were obtained via isothermic evaporation at 0 °C from acetone/Et2O mixtures (v:v = 2:1). The measurement of [Pt(dba)(Me2Imd)] was performed using graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å) with an IPDS 2T diffractometer (STOE and Cie.) at 100 K. [Pt(dba)(dmso)] and [Pt(dba)(CNtBu)] were measured on a D8 Advance diffractometer (Bruker), with the same wavelength, at 100 K. Diffraction data of [Pt(dba)(PPh3)] were collected at the beamline P24 of PETRA III at the German Electron Synchrotron facility DESY using a wavelength of λ = 0.56076 Å and at a temperature of 100 K. The data were recorded on a Pilatus 1M CdTe detector. All of the structures were solved by dual space methods (SHELXT-2015) [74] and refined by full-matrix least-squares techniques against F2 (SHELXL-2017/1) [75,76] The non-hydrogen atoms were refined with anisotropic displacement parameters, without any constraints. The hydrogen atoms were included by using appropriate riding models. The numerical absorption correction of the data of [Pt(dba)(Me2Imd)] (X-RED V1.31; STOE and Cie, 2005, Darmstadt, Germany) were performed after optimising the crystal shapes using the X-SHAPE V1.06 (STOE & Cie, 1999, Darmstadt, Germany) [77,78]. For the data set acquired with synchrotron radiation, no absorption correction was omitted. Residual electron density in the pores alongside the crystallographic b-axis of [Pt(dba)(PPh3)] was removed. The crystallographic details are summarized in Table S1. Data of the structure solutions and refinements can be obtained for [Pt(dba)(dmso)] (CCDC 2141888 (100 K) and 2141889 (298 K)), [Pt(dba)(PPh3)] (CCDC 2070867), [Pt(dba)(CNtBu)] (CCDC 2070866), and [Pt(dba)(Me2Imd)] (CCDC 2070868) free of charge at https://www.ccdc.cam.ac.uk/structures/ or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge, CB2 1EZ UK (fax: +44-1223 336033 or e-mail: [email protected]).

3.5. Computational Details

3.5.1. Vibrationally Resolved Emission Spectra

The quantum-chemical calculations of the vibrational Franck-Condon spectra and optimized geometries were carried out with the quantum chemistry package, Gaussian 09 Rev. D.01 [79]. The CAM-B3LYP functional [80] was used in the density functional theory (DFT) calculations and the SDD basis was set, which applies an effective core potential for the Pt atoms [81] and the D95 basis set for H, C, N and O atoms [82]. The vibrational Franck-Condon spectra were calculated according to the method of Barone et al. [83,84,85], with Kohn-Sham DFT-based geometry optimizations in the S0 and T1 states, followed by frequency analysis calculations.
The energies were corrected by zero-point vibrational energies and thermal free energy contributions. In order to calculate the overlap integrals for the vibronic spectra, the transitions are divided into classes Cn, where n is the number of the excited normal modes in the final electronic state. The maximum number of quanta per mode was set to 100 and the maximum number of quanta for combinations of two modes to 65. The number of integrals calculated was limited to 1.5 × 108. A maximum of 20 classes were computed. The line spectrum was broadened using Gaussian functions with a half-width at half-maximum of 500 cm−1. The solvent CH2Cl2 was taken into account by the polarizable continuum model (PCM) in an integral equation formalism framework [86] with atomic radii from the universal force field model (UFF) [87].

3.5.2. Character of the T1 State

The character of the emissive T1 state is determined by TD-DFT calculations, including 20 excited singlet and triplet states at the T1 geometry, optimized with Kohn-Sham DFT with multiplicity 3.

3.5.3. Absorption Spectra

To obtain the UV-vis absorption spectrum of the complexes, TD-DFT calculations of the 40 lowest excited singlet states were performed with the PBE0 functional [88] and the SDD basis set. A Lorentzian broadening with a half-width at a half-maximum (HWHM) of 10 nm was used for each transition.

4. Conclusions

Starting from the previously reported complex [Pt(dba)(dmso)], containing the rigid dianionic, bis-cyclometalating, tridentate C^N^C ligand dba2– (H2dba = dibenzo[c,h]acridine), we synthesized further derivatives [Pt(dba)L] with L = PPh3, CNtBu and Me2Imd (N,N’-dimethylimidazolydene) as coligands (ancillary ligands) to improve their luminescence properties through the introduction of these stronger σ-donating and π-accepting ligands as compared to dmso.
Cyclic voltammetry showed partially reversible reduction waves ranging between −1.89 and −2.10 V and irreversible oxidation processes between 0.55 (L = Me2Imd) and 1.00 V (dmso), depending on the ancillary ligand L. The strongly variable oxidation potentials raised the electrochemical gaps from 2.65 to 2.91 V along the series Me2Imd < CNtBu < PPh3 < dmso along the trend of decreasing ligand strength.
All four dba complexes partially show vibrationally structured long-wavelength absorption bands peaking at around 530 nm. TD-DFT calculated absorption spectra are in good agreement with the experimental findings, showing that they are slightly red-shifted, but perfectly reproduce the observed trends. The stronger coligands only cause a minor red-shift of the long-wavelength absorption bands (0.03 to 0.05 eV). The photoluminescence spectra of all four compounds are very similar and strongly depend on the presence or absence of 3O2, in line with the triplet character of the emissive state. In a solution at 298, they show broad bands with a minor vibronic structure and maxima at around 590 nm; in frozen glassy matrices, slightly blue-shifted (~580 nm) bands with vibrational progression were found. In contrast to the comparable emission energies, the photoluminescence quantum yields ΦL differ between the four complexes and the lifetimes increase along the series Me2Imd < CNtBu < PPh3 < dmso from 0.06 to 14 µs; surprisingly, the dmso complex constitutes the most efficient emitter, at 298 K (ΦL(Ar) = 0.24). This is highly counterintuitive, as the ligand strength is considered to decrease from Me2Imd to dmso, as supported by the electrochemical results. In frozen glassy matrices at 77 K, the lifetimes are comparable and increase from 17 µs for the Me2Imd complex, to approximately 30 µs for the other three derivatives. The TD-DFT calculated emission spectra agree well with the experimental data for all of the complexes. Decomposition of the excited T1 states showed mostly ligand-centered (LC) contributions (80 to 90%) to the description of the emissive states, which is in line with the vibrationally structured emission bands, a minor metal-to-ligand charge transfer (MLCT) character (5 to 8%), marginal coligand-to-metal (L’MCT) or coligand-to-chelate ligand charge transfer (L’LCT) contributions and virtually no metal-centered (MC) character.
The exclusively theoretical study on the dba-7-phenyl substituted derivatives [Pt(db(ph)a)(L)] (H2db(ph)a = 7-phenyldibenzo[c,h]acridine) with L = PPh3, CNtBu, and Me2Imd showed that these complexes were not superior to the dba congeners, in line with the previous experimental comparison of [Pt(db(ph)a)(dmso)] with the dba derivative.
The far lower ΦL for the complexes containing the stronger coligands Me2Imd, CNtBu, and PPh3 (in comparison with the parent dmso complex [Pt(dba)(dmso)] in solution at 298 K) is an unexpected result as most of the previous studies showed that the exchange of the coligand by a stronger σ-donor and/or π-acceptor improved the photoluminescence (i.e., higher quantum yields). However, at 77 K in frozen glassy matrix, the “ligand effect” was found as expected (higher ΦL for stronger ligands). The elucidation of the cause for these surprising observation at 298 K is the object of ongoing research. At the same time, the assessment of their suitability for optoelectronic devices will require in-depth photophysical studies in the solid state, e.g., in polymeric matrices, which will be carried out in the near future. This will also include a study of the aggregation of these complexes in solution and in solid materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27228054/s1, Figure S1: Crystal structure of [Pt(dba)(CNtBu)] viewed along the crystallographic c axis, the stacking of two molecules in the unit cell, and the complex molecule. Figure S2: Crystal structure of [Pt(dba)(PPh3)] viewed along the crystallographic b axis, the stacking of two molecules in the unit cell, and the complex molecule. Figure S3: Crystal structure of [Pt(dba)(DMSO)] viewed along the crystallographic a axis, the stacking of two molecules in the unit cell, and the complex molecule. Figure S4: Crystal structure of [Pt(dba)(Me2Imd)] viewed along the crystallographic b axis (left). Figure S5: DFT-optimized S0 ground state geometries of the complex [Pt(dba)(L)] with L = DMSO, PPh3, Me2Imd, and CNtBu. Figure S6: DFT-optimized T1 excited state geometries of the complex [Pt(dba)(L)] with L = DMSO, PPh3, Me2Imd, and CNtBu. Figure S7: Cyclic voltammogramms of [Pt(dba)(dmso)] and [Pt(db(ph)a)(dmso)] in 0.1 M n-Bu4NPF6/THF. Figure S8: DFT-calculated energies and composition of selected frontier molecular orbitals for [Pt(dba)(L)] (L = dmso, PPh3, Me2Imd, and CNtBu) at the optimized ground state S0 geometry. Figure S9: UV-vis absorption spectra of [Pt(dba)(L)] (L = dmso, PPh3, CNtBu and Me2Imd) in CH2Cl2. Figure S10: Normalized photoluminescence spectra of the [Pt(dba)L] complexes in CH2Cl2 at 298 K. Figure S11: Normalized excitation (dotted line) and emission (continuous line) spectra of [Pt(dba)(dmso)] in a CH2Cl2/MeOH 1:1 glassy matrix at 77 K. Figure S12: Normalized excitation (dotted line) and emission (continuous line) spectra of [Pt(dba)(PPh3)] in a CH2Cl2/MeOH 1:1 glassy matrix at 77 K. Figure S13: Normalized excitation (dotted line) and emission (continuous line) spectra of [Pt(dba)(CNtBu)] in a CH2Cl2/MeOH 1:1 glassy matrix at 77 K. Figure S14: Normalized excitation (dotted line) and emission (continuous line) spectra of [Pt(dba)(Me2Imd)] in a CH2Cl2/MeOH 1:1 glassy matrix at 77 K. Figure S15: Left: Time-resolved photoluminescence of [Pt(dba)(dmso)] in an air-equilibrated CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S16: Left: Time-resolved photoluminescence of [Pt(dba)(dmso)] in an Ar-purged CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S17: Left: Time-resolved photoluminescence of [Pt(dba)(dmso)] in a frozen CH2Cl2/MeOH glassy matrix at 77 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S18: Left: Time-resolved photoluminescence of [Pt(dba)(PPh3)] in an air-equilibrated CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S19: Left: Time-resolved photoluminescence of [Pt(dba)(PPh3)] in an Ar-purged CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S20: Left: Time-resolved photoluminescence of [Pt(dba)(PPh3)] in a frozen CH2Cl2/MeOH glassy matrix at 77 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S21: Left: Time-resolved photoluminescence of [Pt(dba)(CNtBu)] in an air-equilibrated CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S22: Left: Time-resolved photoluminescence of [Pt(dba)(CNtBu)] in an Ar-purged CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S23: Left: Time-resolved photoluminescence of [Pt(dba)(CNtBu)] in a frozen CH2Cl2/MeOH glassy matrix at 77 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S24: Left: Time-resolved photoluminescence of [Pt(dba)(Me2Imd)] in an air-equilibrated CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S25: Left: Time-resolved photoluminescence of [Pt(dba)(Me2Imd)] in an Ar-purged CH2Cl2 solution at 298 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S26: Left: Time-resolved photoluminescence of [Pt(dba)(Me2Imd)] in a frozen CH2Cl2/MeOH glassy matrix at 77 K (c = 10−5 M), including the residuals (λex = 376 nm, λem = 600 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S27: UV-vis absorption spectra of [Pt(dba)(dmso)] in THF/n-Bu4NPF6, recorded before and after cathodic reduction and before and after anodic oxidation. Figure S28: UV-vis absorption spectra of [Pt(dba)(PPh3)] in THF/n-Bu4NPF6, recorded before and after cathodic reduction and before and after anodic oxidation. Figure S29: UV-vis absorption spectra of [Pt(dba)(CNtBu)] in THF/n-Bu4NPF6, recorded before and after cathodic reduction and before and after anodic oxidation. Figure S30: UV-vis absorption spectra of [Pt(dba)(Me2Imd)] in THF/n-Bu4NPF6, recorded before and after cathodic reduction and before and after anodic oxidation. Table S1: Selected structure solution and refinement data for [Pt(dba)(Me2Imd)], [Pt(dba)(CNtBu)], [Pt(dba)(PPh3)] and [Pt(dba)(dmso)]. Table S2: Selected experimental structural data of [Pt(dba)(Me2Imd)], [Pt(dba)(CNtBu)], [Pt(dba)(PPh3)], and [Pt(dba)(dmso)]. Table S3: Electrochemical data of the Pt(II) complexes. Table S4: Experimental long-wavelength absorption maxima of the Pt(II) complexes. Table S5: Selected absorption maxima of the oxidized, parent, and reduced [Pt(dba)(L)]+/0/− complexes. Figure S31: 600 MHz 1H NMR spectra of [Pt(dba)(PPh3)] in CD2Cl2. Figure S32: 600 MHz 31P NMR spectrum of [Pt(dba)(PPh3)] in CD2Cl2. Figure S33: 600 MHz H,H COSY NMR spectrum of [Pt(dba)(PPh3)] in CD2Cl2. Figure S34: 600 MHz H,H NOESY NMR spectrum of [Pt(dba)(PPh3)] in CD2Cl2. Figure S35: 600 MHz 1H NMR spectrum of [Pt(dba)(CNtBu)] in CD2Cl2. Figure S36: 300 MHz 13C-135-DEPTQ-[1H] NMR spectra of [Pt(dba)(CNtBu)] in CD2Cl2. Figure S37: 600 MHz H,H COSY NMR spectrum of [Pt(dba)(CNtBu)] in CD2Cl2. Figure S38: 600 MHz 1H NMR spectrum of [Pt(dba)(Me2Imd)] in CD2Cl2. Figure S39: 300 MHz 13C-135-DEPTQ-[1H] NMR spectra of [Pt(dba)(Me2Imd)] in CD2Cl2. Figure S40: 600 MHz H,H COSY NMR spectrum of [Pt(dba)(Me2Imd)] in CD2Cl2. Figure S41: 300 MHz 1H-195Pt-HMBC NMR spectrum of [Pt(dba)(Me2Imd)] in CD2Cl2. Figure S42: ESI-MS(+) of [Pt(dba)(PPh3)]. Figure S43: ESI-MS(+) of [Pt(dba)(CNtBu)]. Figure S44: EI-MS(+) of [Pt(dba)(Me2Imd)].

Author Contributions

Conceptualization, A.K., C.A.S. and N.L.D.; methodology, A.K., C.A.S., I.M., and N.L.D.; software, N.L.D.; validation, M.K., I.M., and D.B.; formal analysis, M.K., R.J., I.M., and D.S.; investigation, J.F., M.K., R.J., I.M., and D.B.; resources, A.K., C.A.S., and N.L.D.; data curation, A.K., R.J., and D.S.; writing—original draft preparation, A.K., C.A.S., I.M., and N.L.D.; writing—review and editing, A.K. and C.A.S.; visualization, J.F., I.M., and D.S.; supervision, A.K.; C.A.S., and N.L.D.; project administration, A.K.; C.A.S., and N.L.D.; funding acquisition, A.K.; C.A.S., and N.L.D.; All authors have read and agreed to the published version of the manuscript.

Funding

The Deutsche Forschungsgemeinschaft [DFG Priority Programme 2102 “Light-controlled Reactivity of Metal Complexes” STR-1186/6-1 and/6-2 (C.A.S.) KL-1194/16-1 and/16-2 (A.K.) and DO-768/5-1 and/5-2 (N.L.D.) is acknowledged for the funding of this project. We would also like to thank the Regional Computing Center of the University of Cologne (RRZK) for providing computing time on the DFG-funded High Performance Computing (HPC) system CHEOPS as well as for their support. C.A.S gratefully acknowledges the generous financial support for the acquisition of an “Integrated Confocal Luminescence Spectrometer with Spatiotemporal Resolution and Multiphoton Excitation” (DFG/Land NRW: INST 211/915-1 FUGG; DFG EXC1003: “Berufungsmittel”).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Available from the authors on request.

Acknowledgments

Simon Schmitz, and Alexander Haseloer, University of Cologne, are acknowledged for measurements at the beamline P24 at Petra III and for the re-refinement of the crystal structures.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Mori, K.; Yamashita, H. Metal Complexes Supported on Solid Matrices for Visible-Light-Driven Molecular Transformations. Chem.-Eur. J. 2016, 22, 11122–11137. [Google Scholar] [CrossRef]
  2. Parasram, M.; Gevorgyan, V. Visible light-induced transition metal-catalyzed transformations: Beyond conventional photosensitizers. Chem. Soc. Rev. 2017, 46, 6227–6240. [Google Scholar] [CrossRef]
  3. Zhao, Q.; Li, F.; Huang, C. Phosphorescent chemosensors based on heavy-metal complexes. Chem. Soc. Rev. 2010, 39, 3007–3030. [Google Scholar] [CrossRef]
  4. Law, A.S.-Y.; Lee, L.C.-C.; Lo, K.K.-W.; Yam, V.W.-W. Aggregation and Supramolecular Self-Assembly of Low-Energy Red Luminescent Alkynylplatinum(II) Complexes for RNA Detection, Nucleolus Imaging, and RNA Synthesis Inhibitor Screening. J. Am. Chem. Soc. 2021, 143, 5396–5405. [Google Scholar] [CrossRef]
  5. Li, K.; Tong, G.S.M.; 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]
  6. Fantacci, S.; De Angelis, F. A computational approach to the electronic and optical properties of Ru(II) and Ir(III) polypyridyl complexes: Applications to DSC, OLED and NLO. Coord. Chem. Rev. 2011, 255, 2704–2726. [Google Scholar] [CrossRef]
  7. Housecroft, C.E.; Constable, E.C. Over the LEC rainbow: Colour and stability tuning of cyclometallated iridium(III) complexes in light-emitting electrochemical cells. Coord. Chem. Rev. 2017, 350, 155–177. [Google Scholar] [CrossRef] [Green Version]
  8. Yersin, H. Highly Efficient OLEDs with Phosphorescent Materials; Wiley-VCH: Weinheim, Germany, 2008. [Google Scholar] [CrossRef]
  9. Bizzarri, C.; Spuling, E.; Knoll, D.M.; Volz, D.; Bräse, S. Sustainable metal complexes for organic light-emitting diodes (OLEDs). Coord. Chem. Rev. 2018, 373, 49–82. [Google Scholar] [CrossRef]
  10. Ravotto, L.; Ceroni, P. Aggregation induced phosphorescence of metal complexes: From principles to applications. Coord. Chem. Rev. 2017, 346, 62–76. [Google Scholar] [CrossRef]
  11. 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]
  12. Zhen, X.; Qu, R.; Chen, W.; Wu, W.; Jiang, X. The development of phosphorescent probes for in vitro and in vivo bioimaging. Biomater. Sci. 2021, 9, 285–300. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, Y.; Guan, R.; Zhang, C.; Huang, J.; Ji, L.; Chao, H. Two-photon luminescent metal complexes for bioimaging and cancer phototherapy. Coord. Chem. Rev. 2016, 310, 16–40. [Google Scholar] [CrossRef]
  14. Balachandran, C.; Yokoi, K.; Naito, K.; Haribabu, J.; Tamura, Y.; Umezawa, M.; Tsuchiya, K.; Yoshihara, T.; Tobita, S.; Aoki, S. Cyclometalated Iridium(III) Complex–Cationic Peptide Hybrids Trigger Paraptosis in Cancer Cells via an Intracellular Ca2+ Overload from the Endoplasmic Reticulum and a Decrease in Mitochondrial Membrane Potential. Molecules 2021, 26, 7028. [Google Scholar] [CrossRef] [PubMed]
  15. Haribabu, J.; Tamura, Y.; Yokoi, K.; Balachandran, C.; Umezawa, M.; Tsuchiya, K.; Yamada, Y.; Karvembu, R.; Aoki, S. Synthesis and Anticancer Properties of Bis- and Mono(cationic peptide) Hybrids of Cyclometalated Iridium(III) Complexes: Effect of the Number of Peptide Units on Anticancer Activity. Eur. J. Inorg. Chem. 2021, 2021, 1796–1814. [Google Scholar] [CrossRef]
  16. Kalinowski, J.; Fattori, V.; Cocchi, M.; Williams, J.A.G. Light-emitting devices based on organometallic platinum complexes as emitters. Coord. Chem. Rev. 2011, 255, 2401–2425. [Google Scholar] [CrossRef]
  17. Koshevoy, I.O.; Krause, M.; Klein, A. Non-Covalent Intramolecular Interactions through Ligand-Design Promoting Efficient Luminescence from Transition Metal Complexes. Coord. Chem. Rev. 2020, 405, 213094. [Google Scholar] [CrossRef]
  18. Gray, H.B.; Záliš, S.; Vlček, A. Electronic structures and photophysics of d8-d8 complexes. Coord. Chem. Rev. 2017, 345, 297–317. [Google Scholar] [CrossRef]
  19. Ganesan, P.; Hung, W.; Tso, J.; Ko, C.; Wang, T.; Chen, P.; Hsu, H.; Liu, S.; Lee, G.; Chou, P.; et al. Functional Pyrimidinyl Pyrazolate Pt(II) Complexes: Role of Nitrogen Atom in Tuning the Solid-State Stacking and Photophysics. Adv. Funct. Mater. 2019, 29, 1900923. [Google Scholar] [CrossRef]
  20. Cinninger, L.M.; Bastatas, L.D.; Shen, Y.; Holliday, B.J.; Slinker, J.D. Luminescent properties of a 3,5-diphenylpyrazole bridged Pt(II) dimer. Dalton Trans. 2019, 48, 9684–9691. [Google Scholar] [CrossRef]
  21. Yam, V.W.-W.; Au, V.K.-M.; Leung, S.Y.-L. Light-Emitting Self-Assembled Materials Based on d8 and d10 Transition Metal Complexes. Chem. Rev. 2015, 115, 7589–7728. [Google Scholar] [CrossRef]
  22. Cnudde, M.; Brünink, D.; Doltsinis, N.L.; Strassert, C.A. Tetradentate N^N°N^N-type luminophores for Pt(II) complexes: Synthesis, photophysical and quantum-chemical investigation. Inorg. Chim. Acta 2021, 518, 120090. [Google Scholar] [CrossRef]
  23. Li, G.; Ameri, L.; Fleetham, T.; Zhu, Z.-Q.; Li, J. Stable and efficient blue and green organic light emitting diodes employing tetradentate Pt(II) complexes. Appl. Phys. Lett. 2020, 117, 253301. [Google Scholar] [CrossRef]
  24. Yu, F.; Sheng, Y.; Wu, D.; Qin, K.; Li, H.; Xie, G.; Xue, Q.; Sun, Z.; Lu, Z.; Ma, H.; et al. Blue-Phosphorescent Pt(II) Complexes of Tetradentate Pyridyl−Carbolinyl Ligands: Synthesis, Structure, Photophysics, and Electroluminescence. Inorg. Chem. 2020, 59, 14493–14500. [Google Scholar] [CrossRef] [PubMed]
  25. Cheng, G.; Kwak, Y.; To, W.-P.; Lam, T.-L.; Tong, G.S.M.; Sit, M.-K.; Gong, S.; Choi, B.; Choi, W.il.; 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]
  26. Fleetham, T.; Li, G.; Li, J. Phosphorescent Pt(II) and Pd(II) Complexes for Efficient, High-Color-Quality, and Stable OLEDs. Adv. Mater. 2017, 29, 1601861. [Google Scholar] [CrossRef]
  27. Krause, M.; von der Stück, R.; Brünink, D.; Buss, S.; Doltsinis, N.L.; Strassert, C.A.; Klein, A. Platinum and palladium complexes of tridentate -C^N^N (phen-ide)-pyridine-thiazol ligands—A case study involving spectroelectrochemistry, photoluminescence spectroscopy and TD-DFT calculations. Inorg. Chim. Acta 2021, 518, 120093. [Google Scholar] [CrossRef]
  28. Hebenbrock, M.; González-Abradelo, D.; Hepp, A.; Meadowcroft, J.; Lefringhausen, N.; Strassert, C.A.; Müller, J. Influence of the ancillary ligands on the luminescence of platinum(II) complexes with a triazole-based tridentate C^N^N luminophore. Inorg. Chim. Acta 2021, 516, 119988. [Google Scholar] [CrossRef]
  29. Sivchik, V.; Kochetov, A.; Eskelinen, T.; Kisel, K.S.; Solomatina, A.I.; Grachova, E.V.; Tunik, S.P.; Hirva, P.; Koshevoy, I.O. Modulation of Metallophilic and π–π Interactions in Platinum Cyclometalated Luminophores with Halogen Bonding. Chem.-Eur. J. 2021, 27, 1787–1794. [Google Scholar] [CrossRef]
  30. Puttock, E.V.; Sturala, J.; Kistemaker, J.C.M.; Williams, J.A.G. Platinum(II) Complexes of Tridentate-Coordinating Ligands Based on Imides, Amides, and Hydrazides: Synthesis and Luminescence Properties. Eur. J. Inorg. Chem. 2021, 2021, 335–347. [Google Scholar] [CrossRef]
  31. Eskelinen, T.; Buss, S.; Petrovskii, S.K.; Grachova, E.V.; Krause, M.; Kletsch, L.; Klein, A.; Strassert, C.A.; Koshevoy, I.O.; Hirva, P. Photophysics and Excited State Dynamics of Cyclometalated [M(C^N^N)(CN)] (M = Ni, Pd, Pt) Complexes: A Theoretical and Experimental Study. Inorg. Chem. 2021, 60, 8777–8789. [Google Scholar] [CrossRef]
  32. Gangadharappa, S.C.; Maisuls, I.; Schwab, D.A.; Kösters, J.; Doltsinis, N.L.; Strassert, C.A. Compensation of Hybridization Defects in Phosphorescent Complexes with Pnictogen-Based Ligands—A Structural, Photophysical, and Theoretical Case-Study with Predictive Character. J. Am. Chem. Soc. 2020, 142, 21353–21367. [Google Scholar] [CrossRef]
  33. Garbe, S.; Krause, M.; Klimpel, A.; Neundorf, I.; Lippmann, P.; Ott, I.; Brünink, D.; Strassert, C.A.; Doltsinis, N.L.; Klein, A. Cyclometalated Pt Complexes of CNC Pincer Ligands: Luminescence and Cytotoxic Evaluation. Organometallics 2020, 39, 746–756. [Google Scholar] [CrossRef]
  34. Iwakiri, A.; Konno, Y.; Shinozaki, K. Determination of excimer emission quantum yield of Pt(dpb)Cl (dpbH = 1,3-di(2-pyridyl)benzene and its analogues in solution. J. Lumin. 2019, 207, 482–490. [Google Scholar] [CrossRef]
  35. Garoni, E.; Boixel, J.; Dorcet, V.; Roisnel, T.; Roberto, D.; Jacquemin, D.; Guerchais, V. Controlling the emission in flexibly-linked (N^C^N)platinum dyads. Dalton Trans. 2018, 47, 224–232. [Google Scholar] [CrossRef]
  36. Kletsch, L.; Jordan, R.; Köcher, A.S.; Buss, S.; Strassert, C.A.; Klein, A. Photoluminescence of Ni(II), Pd(II), and Pt(II) Complexes [M(Me2dpb)Cl] Obtained from C-H Activation of 1,5-Di(2-pyridyl)-2,4-dimethylbenzene (Me2dpbH). Molecules 2021, 26, 5051. [Google Scholar] [CrossRef] [PubMed]
  37. Schulze, B.; Friebe, C.; Jäger, M.; Görls, H.; Birckner, E.; Winter, A.; Schubert, U.S. PtII Phosphors with Click-Derived 1,2,3-Triazole-Containing Tridentate Chelates. Organometallics 2018, 37, 145–155. [Google Scholar] [CrossRef]
  38. Hebenbrock, M.; Stegemann, L.; Kösters, J.; Doltsinis, N.L.; Müller, J.; Strassert, C.A. Phosphorescent Pt(II) complexes bearing a monoanionic C^N^N luminophore and tunable ancillary ligands. Dalton Trans. 2017, 46, 3160–3169. [Google Scholar] [CrossRef] [PubMed]
  39. Rodrigue-Witchel, A.; Rochester, D.L.; Zhao, S.-B.; Lavelle, K.B.; Williams, J.A.G.; Wang, S.; Connick, W.B.; Reber, C. Pressure-induced variations of MLCT and ligand-centered luminescence spectra in square-planar platinum(II) complexes. Polyhedron 2016, 108, 151–155. [Google Scholar] [CrossRef] [Green Version]
  40. Rausch, A.F.; Murphy, L.; Williams, J.A.G.; Yersin, H. Improving the Performance of Pt(II) Complexes for Blue Light Emission by Enhancing the Molecular Rigidity. Inorg. Chem. 2012, 51, 312–319. [Google Scholar] [CrossRef]
  41. Pander, P.; Zaytsev, A.V.; Sil, A.; Williams, J.A.G.; Kozhevnikov, V.N.; Dias, F.B. Enhancement of thermally activated delayed fluorescence properties by substitution of ancillary halogen in a multiple resonance-like diplatinum(II) complex. J. Mater. Chem. C 2022, 10, 4851–4860. [Google Scholar] [CrossRef]
  42. Li, G.; Wen, J.; Zhan, F.; Lou, W.; Yang, Y.-F.; Hu, Y.; She, Y. Fused 6/5/6 Metallocycle-Based Tetradentate Pt(II) Emitters for Efficient Green Phosphorescent OLEDs. Inorg. Chem. 2022, 61, 11218–11231. [Google Scholar] [CrossRef]
  43. Zhu, L.; Sha, C.; Lv, A.; Xie, W.; Shen, K.; Chen, Y.; Xie, G.; Ma, H.; Li, H.; Hang, X.-C. Tetradentate Pt(II) Complexes with Peripheral Hindrances for Highly Efficient Solution-Processed Blue Phosphorescent OLEDs. Inorg. Chem. 2022, 61, 10402–10409. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, X.; Ao, L.; Han, Y.; Gao, Z.; Wang, F. Modulating Pt...Pt metal–metal interactions through conformationally switchable molecular tweezer/guest complexation. Chem. Commun. 2018, 54, 1754–1757. [Google Scholar] [CrossRef]
  45. Lu, W.; Chan, M.C.W.; Cheung, K.-K.; Che, C.-M. π–π Interactions in Organometallic Systems. Crystal Structures and Spectroscopic Properties of Luminescent Mono-, Bi-, and Trinuclear Trans-cyclometalated Platinum(II) Complexes Derived from 2,6-Diphenylpyridine. Organometallics 2001, 20, 2477–2486. [Google Scholar] [CrossRef]
  46. Yam, V.W.-W.; Tang, R.P.-L.; Wong, K.M.-C.; Lu, X.-X.; Cheung, K.-K.; Zhu, N. Syntheses, Electronic Absorption, Emission, and Ion-Binding Studies of Platinum(II) C^N^C and Terpyridyl Complexes Containing Crown Ether Pendants. Chem.-Eur. J. 2002, 8, 4066–4076. [Google Scholar] [CrossRef]
  47. Kui, S.C.F.; Hung, F.-F.; Lai, S.-L.; Yuen, M.-Y.; Kwok, C.-C.; Low, K.-H.; Chui, S.S.-Y.; Che, C.-M. Luminescent Organoplatinum(II) Complexes with Functionalized Cyclometalated C^N^C Ligands: Structures, Photophysical Properties, and Material Applications. Chem.-Eur. J. 2012, 18, 96–109. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, Q.; Wang, S.; Zhu, Y.; Zhang, C.; Cao, H.; Ma, W.; Tian, X.; Wu, J.; Zhou, H.; Tian, Y. Functional Platinum(II) Complexes with Four-Photon Absorption Activity, Lysosome Specificity, and Precise Cancer Therapy. Inorg. Chem. 2021, 60, 2362–2371. [Google Scholar] [CrossRef]
  49. Zhu, R.; Chen, X.; Shu, N.; Shang, Y.; Wang, Y.; Yang, P.; Tang, Y.; Wang, F.; Xu, J. Computational Study of Photochemical Relaxation Pathways of Platinum(II) Complexes. J. Phys. Chem. A 2021, 125, 10144–10154. [Google Scholar] [CrossRef] [PubMed]
  50. Sanning, J.; Stegemann, L.; Ewen, P.R.; Schwermann, C.; Daniliuc, C.G.; Zhang, D.; Lin, N.; Duan, L.; Wegner, D.; Doltsinis, N.L.; et al. Colour-tunable asymmetric cyclometalated Pt(II) complexes and STM-assisted stability assessment of ancillary ligands for OLEDs. J. Mater. Chem. C 2016, 4, 2560–2565. [Google Scholar] [CrossRef]
  51. Yam, V.W.-W.; Wong, K.M.-C. Luminescent metal complexes of d6, d8 and d10 transition metal centres. Chem. Commun. 2011, 47, 11579–11592. [Google Scholar] [CrossRef]
  52. Chow, P.-K.; Cheng, G.; Tong, G.S.M.; To, W.-P.; Kwong, W.-L.; Low, K.-H.; Kwok, C.-C.; Ma, C.; Che, C.-M. Luminescent Pincer Platinum(II) Complexes with Emission Quantum Yields up to Almost Unity: Photophysics, Photoreductive C-C Bond Formation, and Materials Applications. Angew. Chem. Int. Ed. 2015, 54, 2084–2089. [Google Scholar] [CrossRef]
  53. Wang, L.; Zhang, Y.; Li, J.; He, H.; Zhang, J. Influence of primary and auxiliary ligand on spectroscopic properties and luminescent efficiency of organoplatinum(II) complexes bearing functionalized cyclometalated C^N^C ligands. Dalton Trans. 2014, 43, 14029–14038. [Google Scholar] [CrossRef]
  54. Qiu, D.; Wu, J.; Xie, Z.; Cheng, Y.; Wang, L. Synthesis, photophysical and electrophosphorescent properties of mononuclearPt(II) complexes with arylamine functionalized cyclometalating ligands. J. Organomet. Chem. 2009, 694, 737–746. [Google Scholar] [CrossRef]
  55. Krause, M.; Maisuls, I.; Buss, S.; Strassert, C.A.; Winter, A.; Schubert, U.S.; Nair, S.S.; Dietzek-Ivanšic, B.; Klein, A. Photophysical Study on the Rigid Pt(II) Complex [Pt(naphen)(Cl)] (Hnaphen = Naphtho[1,2-b][1,10]Phenanthroline and De-rivatives. Molecules 2022, 27, 7022. [Google Scholar] [CrossRef]
  56. Krause, M.; Friedel, J.; Buss, S.; Brünink, D.; Berger, A.; Strassert, C.A.; Doltsinis, N.L.; Axel Klein, A. Isoelectronic Pt(II) complexes of cyclometalating C^N^N ligands with phenyl/(benzo)thiophenyl and pyridyl/(benzo)thiazolyl moieties. Dalton Trans. 2022, 51, 16181–16194. [Google Scholar] [CrossRef]
  57. Föller, J.; Friese, D.H.; Riese, S.; Kaminski, J.M.; Metz, S.; Schmidt, D.; Würthner, F.; Lambert, C.; Marian, C.M. On the photophysical properties of IrIII, PtII, and PdII (phenylpyrazole) (phenyldipyrrin) complexes. Phys. Chem. Chem. Phys. 2020, 22, 3217–3233. [Google Scholar] [CrossRef] [PubMed]
  58. Martinez, C.R.; Iverson, B.L. Rethinking the term “pi-stacking”. Chem. Sci. 2012, 3, 2191–2201. [Google Scholar] [CrossRef] [Green Version]
  59. Janiak, C. A critical account on π–π stacking in metal complexes with aromatic nitrogen-containing ligands. J. Chem. Soc. Dalton Trans. 2000, 2000, 3885–3896. [Google Scholar] [CrossRef]
  60. Jain, R.; Al Mamun, A.; Buchanan, R.M.; Kozlowski, P.M.; Grapperhaus, C.A. Ligand-Assisted Metal-Centered Electrocatalytic Hydrogen Evolution upon Reduction of a Bis(thiosemicarbazonato)Ni(II) Complex. Inorg. Chem. 2018, 57, 13486–13493. [Google Scholar] [CrossRef]
  61. Klein, A.; Kaiser, A.; Wielandt, W.; Belaj, F.; Wendel, E.; Bertagnolli, H.; Zalis, S. Halide Ligands-More Than Just σ-Donors? A Structural and Spectroscopic Study of Homologous Organonickel Complexes. Inorg. Chem. 2008, 47, 11324–11333. [Google Scholar] [CrossRef]
  62. Maisuls, I.; Wang, C.; Gutierrez Suburu, M.E.; Wilde, S.; Daniliuc, C.-G.; Brünink, D.; Doltsinis, N.L.; Ostendorp, S.; Wilde, G.; Kösters, J.; et al. Ligand-controlled and nanoconfinement-boosted luminescence employing Pt(II) and Pd(II) complexes: From color-tunable aggregation-enhanced dual emitters towards self-referenced oxygen reporters. Chem. Sci. 2021, 12, 3270–3281. [Google Scholar] [CrossRef]
  63. Williams, J.A.G. The coordination chemistry of dipyridylbenzene: N-deficient terpyridine or panacea for brightly luminescent metal complexes? Chem. Soc. Rev. 2009, 38, 1783–1801. [Google Scholar] [CrossRef]
  64. Sillen, A.; Engelborghs, Y. The Correct Use of “Average” Fluorescence Parameters. Photochem. Photobiol. 1998, 67, 475–486. [Google Scholar] [CrossRef]
  65. 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]
  66. Bregenholt Zederkof, D.; Møller, K.B.; Nielsen, M.M.; Haldrup, K.; González, L.; Mai, S. Resolving Femtosecond Solvent Reorganization Dynamics in an Iron Complex by Nonadiabatic Dynamics Simulations. J. Am. Chem. Soc. 2022, 144, 12861–12873. [Google Scholar] [CrossRef] [PubMed]
  67. Mai, S.; González, L. Unconventional two-step spin relaxation dynamics of [Re(CO)3(im)(phen)]+ in aqueous solution. Chem. Sci. 2019, 10, 10405–10411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Von der Stück, R.; Krause, M.; Brünink, D.; Buss, S.; Doltsinis, N.L.; Strassert, C.A.; Klein, A. Luminescent Pd(II) Complexes with Tridentate -C^N^N Arylpyridine-(benzo)thiazole Ligands. Z. Anorg. Allg. Chem. 2022, 648, e202100278. [Google Scholar] [CrossRef]
  69. Vogt, N.; Sandleben, A.; Kletsch, L.; Schäfer, S.; Chin, M.T.; Vicic, D.A.; Hörner, G.; Klein, A. Role of the X Coligands in Cyclometalated [Ni(Phbpy)X] Complexes (HPhbpy = 6-Phenyl-2,2′-bipyridine). Organometallics 2021, 40, 1776–1785. [Google Scholar] [CrossRef]
  70. Hamacher, C.; Hurkes, N.; Kaiser, A.; Klein, A.; Schüren, A. Electrochemistry and Spectroscopy of Organometallic Terpyridine Nickel Complexes. Inorg. Chem. 2009, 48, 9947–9951. [Google Scholar] [CrossRef]
  71. Braterman, P.S.; Song, J.-I.; Wimmer, F.M.; Wimmer, S.; Kaim, W.; Klein, A.; Peacock, R.D. Electrochemistry and Spectroelectrochemistry (EPR, UV-Vis-Near-IR) of Platinum(II) 2,2′-Bipyridine and Ring-Metalated Bipyridine Complexes: PtII(L) and PtI(L) but Not PtI(L). Inorg. Chem. 1992, 31, 5084–5088. [Google Scholar] [CrossRef]
  72. Klein, A.; Kaim, W. Axial Shielding of 5d8 and 5d7 Metal Centers in Dimesitylplatinum Complexes with Unsaturated Chelate Ligands: Spectroscopic and Spectroelectrochemical Studies of Four Different Oxidation States. Organometallics 1995, 14, 1176–1186. [Google Scholar] [CrossRef]
  73. Gardner, S.S.; Kawamoto, T.; Curran, D.P. Synthesis of 1,3-Dialkylimidazol-2-ylidene Boranes from 1,3-Dialkylimidazolium Iodides and Sodium Borohydride. J. Org. Chem. 2015, 80, 9794–9797. [Google Scholar] [CrossRef] [PubMed]
  74. Sheldrick, G.M. ShelXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Sheldrick, G.M. SHELXL-2017/1, Program for the Solution of Crystal Structures; University of Göttingen: Göttingen, Germany, 2017. [Google Scholar]
  76. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  77. STOE X-RED. Data Reduction Program, Version 1.31/Windows; STOE & Cie: Darmstadt, Germany, 2005. [Google Scholar]
  78. STOE X-SHAPE. Crystal Optimisation for Numerical Absorption Correction, Version 1.06/Windows; STOE & Cie: Darmstadt, Germany, 1999. [Google Scholar]
  79. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09 Revision D.01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  80. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  81. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comp. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  82. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chem. Acc. 1990, 77, 123–141. [Google Scholar] [CrossRef]
  83. Dunning, T.H.; Hay, P.J. Modern Theoretical Chemistry; Schaefer, H.F., Ed.; Plenum: New York, NY, USA, 1977; Volume 3. [Google Scholar]
  84. Barone, V.; Bloino, J.; Biczysko, M. Vibrationally-Resolved Electronic Spectra in GAUSSIAN 09; GAUSSIAN 09, Revision A.02; Gaussian Inc.: Wallingford, CT, USA, 2009; pp. 1–20. [Google Scholar]
  85. Barone, V.; Bloino, J.; Biczysko, M.; Santoro, F. Fully Integrated Approach to Compute Vibrationally Resolved Optical Spectra: From Small Molecules to Macrosystems. J. Chem. Theory Comput. 2009, 5, 540–554. [Google Scholar] [CrossRef]
  86. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  87. Skid, W.M. UFF, a Full Periodic Table Force Field for Molecular Mechanics and Molecular Dynamics Simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. [Google Scholar] [CrossRef]
  88. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
Scheme 1. Schematic depiction of C^N^C Pt(II) complexes with the relatively flexible dppy ligand (left) or the rather rigid dba ligand (right).
Scheme 1. Schematic depiction of C^N^C Pt(II) complexes with the relatively flexible dppy ligand (left) or the rather rigid dba ligand (right).
Molecules 27 08054 sch001
Scheme 2. Schematic representation of the complexes [Pt(dba)(L)] (1 to 4) and [Pt(db(ph)a)(L)] (5 to 8).
Scheme 2. Schematic representation of the complexes [Pt(dba)(L)] (1 to 4) and [Pt(db(ph)a)(L)] (5 to 8).
Molecules 27 08054 sch002
Scheme 3. Synthesis of the complexes [Pt(dba)(L)] 2 to 4 from 1.
Scheme 3. Synthesis of the complexes [Pt(dba)(L)] 2 to 4 from 1.
Molecules 27 08054 sch003
Figure 1. Crystal structure of [Pt(dba)(Me2Imd)] (4) viewed along the crystallographic a axis (A) and the molecular structure in the crystal (B) with thermal displacement ellipsoids at 50% probability and including the numbering; H atoms were omitted for clarity. DFT-optimized S0 structure of 4 (C,D) displaying the specific angles listed in Table 1.
Figure 1. Crystal structure of [Pt(dba)(Me2Imd)] (4) viewed along the crystallographic a axis (A) and the molecular structure in the crystal (B) with thermal displacement ellipsoids at 50% probability and including the numbering; H atoms were omitted for clarity. DFT-optimized S0 structure of 4 (C,D) displaying the specific angles listed in Table 1.
Molecules 27 08054 g001
Figure 2. Cyclic voltammogramms of [Pt(dba)(L)] with L = PPh3 (left), CNtBu (middle), and Me2Imd (right) in 0.1 M n-Bu4NPF6/THF.
Figure 2. Cyclic voltammogramms of [Pt(dba)(L)] with L = PPh3 (left), CNtBu (middle), and Me2Imd (right) in 0.1 M n-Bu4NPF6/THF.
Molecules 27 08054 g002
Figure 3. Left: TD-DFT-calculated UV-vis absorption spectra (red lines, Theor.) of [Pt(dba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu in CH2Cl2 compared with the experimental spectra (black lines, Exp.). Right: Comparison of calculated UV-vis absorption spectra of [Pt(dba)(L)] and [Pt(d(ph)ba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu.
Figure 3. Left: TD-DFT-calculated UV-vis absorption spectra (red lines, Theor.) of [Pt(dba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu in CH2Cl2 compared with the experimental spectra (black lines, Exp.). Right: Comparison of calculated UV-vis absorption spectra of [Pt(dba)(L)] and [Pt(d(ph)ba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu.
Molecules 27 08054 g003
Figure 4. Normalized photoluminescence spectra of the [Pt(dba)(L)] complexes (1 to 4) in a frozen CH2Cl2:MeOH 1:1 glassy matrix at 77 K (λex = 350 nm, c = 10−5 M). The corresponding excitation spectra as well as the emission of the dba ligand can be found in the SM, Figures S11–S15.
Figure 4. Normalized photoluminescence spectra of the [Pt(dba)(L)] complexes (1 to 4) in a frozen CH2Cl2:MeOH 1:1 glassy matrix at 77 K (λex = 350 nm, c = 10−5 M). The corresponding excitation spectra as well as the emission of the dba ligand can be found in the SM, Figures S11–S15.
Molecules 27 08054 g004
Figure 5. Left: DFT-calculated emission spectra (red lines, Theor.) of [Pt(dba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu in CH2Cl2 compared with the respective experimental data at 77 K (blue lines, Exp.). The calculated 0-0 transitions are shown as vertical green dashed lines. Right: Comparison of calculated emission spectra of [Pt(dba)(L)] and [Pt(d(ph)ba)(L)].
Figure 5. Left: DFT-calculated emission spectra (red lines, Theor.) of [Pt(dba)(L)] for L = dmso, PPh3, Me2Imd, and CNtBu in CH2Cl2 compared with the respective experimental data at 77 K (blue lines, Exp.). The calculated 0-0 transitions are shown as vertical green dashed lines. Right: Comparison of calculated emission spectra of [Pt(dba)(L)] and [Pt(d(ph)ba)(L)].
Molecules 27 08054 g005
Figure 6. Frontier orbitals of complex [Pt(dba)(L)] with L = dmso, PPh3, Me2Imd, CNtBu at the optimized T1 geometry together with the highest contributions of monoelectronic excitations to the T1 state (in %).
Figure 6. Frontier orbitals of complex [Pt(dba)(L)] with L = dmso, PPh3, Me2Imd, CNtBu at the optimized T1 geometry together with the highest contributions of monoelectronic excitations to the T1 state (in %).
Molecules 27 08054 g006
Figure 7. Decomposition of the T1 state for [Pt(dba)(L)] with L = dmso (red), PPh3 (green), Me2Imd (black), CNtBu (blue) into contributions originating from MLCT, L’MCT, LC, L’LCT and MC configurations determined from a correlated electron-hole pair analysis [48]. Right: Partitioning used for the example of [Pt(dba)(CNtBu)].
Figure 7. Decomposition of the T1 state for [Pt(dba)(L)] with L = dmso (red), PPh3 (green), Me2Imd (black), CNtBu (blue) into contributions originating from MLCT, L’MCT, LC, L’LCT and MC configurations determined from a correlated electron-hole pair analysis [48]. Right: Partitioning used for the example of [Pt(dba)(CNtBu)].
Molecules 27 08054 g007
Figure 8. UV-vis absorption spectra of [Pt(dba)(CNtBu)] (3) in THF/n-Bu4NPF6, recorded during cathodic reduction (left, traces of reduced species in red) and during anodic oxidation (right, traces of oxidized species in blue)-spectroelectrochemistry.
Figure 8. UV-vis absorption spectra of [Pt(dba)(CNtBu)] (3) in THF/n-Bu4NPF6, recorded during cathodic reduction (left, traces of reduced species in red) and during anodic oxidation (right, traces of oxidized species in blue)-spectroelectrochemistry.
Molecules 27 08054 g008
Table 1. Selected data of experimental and DFT-optimized geometries in the S0 and T1 states for [Pt(dba)(L)] a.
Table 1. Selected data of experimental and DFT-optimized geometries in the S0 and T1 states for [Pt(dba)(L)] a.
Statedmso(1)PPh3(2)CNtBu(3)Me2Imd(4)
d Pt-X [Å]exp.2.19(7)2.23(2)1.89(3)1.99(1)
calc. S02.342.351.921.99
calc. T12.352.391.921.99
α = N-Pt-X [°]exp.173.2(6)176.2(1)177.6(1)177.8(4)
calc.S0174.1176.7179.8179.7
calc. T1179.1177.4179.8179.8
β = C-Pt-X [°]exp.94.47(7)96.8(1)102.9(1)101.9(4)
calc. S094.797.1100.4100.3
calc. T199.499.2100.4100.4
γ = Pt-X-Y [°]exp.115.7(9)113.4(2)177.4(3)128.7(9)
calc. S0115.1114.5179.7127.8
calc. T1115.2114.2179.9127.8
δ = C-Pt-X-Y [°]exp.129.9(1)68.8(2)17.8(5)57.9(1)
calc. S0111.165.55.164.2
calc. T1107.7101.9−33.064.2
a The X atom is the S, P, or C donor atom of the coligand, Y represents the following atom.
Table 2. Selected electrochemical data of the [Pt(dba)L] complexes (1 to 4) a.
Table 2. Selected electrochemical data of the [Pt(dba)L] complexes (1 to 4) a.
ReductionReduction OxidationEchem Gap
L =E1/2 Red2E1/2 Red1ΔRed1−Red2Epa Ox1ΔOx1−Red1
dmso (1)−2.70 (irr)−1.98 (irr)0.720.932.91
PPh3 (2)−2.70 (irr)−1.960.740.752.71
CNtBu (3)−3.26−1.891.370.802.69
Me2Imd (4)−3.02 (irr)−2.100.920.552.65
a From cyclic voltammetry in n-Bu4NPF6/THF. Potentials in V vs. ferrocene/ferrocenium; half-wave potentials (E1/2) for reversible processes; anodic peak potentials Epa for irreversible oxidations; cathodic peak potentials (Epc) for irreversible reductions (irr); accuracy of potentials: ±0.01 V.
Table 3. Experimental long-wavelength absorption maxima of the [Pt(dba)(L)] complexes (1 to 4) a.
Table 3. Experimental long-wavelength absorption maxima of the [Pt(dba)(L)] complexes (1 to 4) a.
L =λmax (ε)λmax (ε)λmax (ε)λmax (ε)λmax/eVλcut-off b/nmλcut-off b/eV
dmso (1)292 (47.8)340 (8.3)409 (4.1)525 (1.9)2.365662.19
PPh3 (2)296 (50.2)347 (12.7)409 (6.2)532 (2.1)2.335772.15
CNtBu (3)294 (27.0)343 (5.2)408 (2.4)533 (1.0)2.335782.15
Me2Imd (4)300 (24.8)352 (8.9)406 (3.2)536 (1.3)2.315802.14
a Measured in CH2Cl2, absorption maxima λmax in nm, molar absorption coefficients ε in 103 M−1 cm−1. b The λcut-off is defined as the wavelength of zero absorption and is obtained through extrapolation of the low-energy slope of the lowest energy absorption band.
Table 4. Photophysical properties of the [Pt(dba)(L)] complexes (1 to 4) and H2dba a.
Table 4. Photophysical properties of the [Pt(dba)(L)] complexes (1 to 4) and H2dba a.
L =τ(air)/nsτ(Ar)/µsτ(77 K)/µsΦL (air) ± 0.02ΦL (Ar) ± 0.02ΦL (77 K) ± 0.02kr (77 K)/
104 s−1
dmso (1)τ = 336.0 ± 0.5τ = 14.040 ± 0.013τ1 = 27.9 ± 1.2 (24%)
τ2 = 47.4 ± 0.4 (76%)
τav_amp = 42.79 ± 0.06
<0.020.240.821.90 ± 0.05
PPh3 (2)τ1 = 157.1 ± 0.3 (27%)
τ2 = 4.1± 0.3 (73%)
τav_amp = 46 ± 4
τ = 9.279 ± 0.008τ1 = 40 ± 3 (14%)
τ2 = 26.4 ± 0.4 (86%)
τav_amp = 28.40 ± 0.03
<0.02<0.020.903.16 ± 0.08
CNtBu (3)τ1 = 259.2 ± 0.7 (52%)
τ2 = 6.9 ± 0.3 (48%)
τav_amp = 139.0 ± 1.6
τ = 3.167 ± 0.002τ1 = 38.4 ± 0.6 (14%)
τ2 = 27.08 ± 0.12 (86%)
τav_amp = 28.640 ± 0.016
0.040.110.802.80 ± 0.08
Me2Imd (4)τ1 = 340 ± 20 (1%)
τ2 = 7.0 ± 0.3 (24%)
τ3 = 21.0 ± 1.2 (75%)
τav_amp = 22.4 ± 0.6
τ1 = 2.13 ± 0.07 (2%)
τ2 = 0.022 ± 0.001 (98%)
τav_amp = 0.061 ± 0.004
τ1 = 60 ± 2 (2%)
τ2 = 15.59 ± 0.05 (98%)
τav_amp = 16.7 ± 0.04
<0.020.040.865.16 ± 0.13
H2dbaτ/10−3= 6.404 ± 0.009τ = 6.81 ± 0.03n.d.0.420.420.53n.d.
a Photoluminescence quantum yields ΦL (±0.02), lifetimes τ, radiative and non-radiative deactivation rate constants (kr and knr) under different conditions: air-equilibrated (air) or argon-purged samples (Ar) at 298 K or in frozen CH2Cl2:MeOH 1:1 glassy matrices at 77 K. For the bi-exponential decays, the amplitude-weighted average lifetimes (τav_amp) [64] are shown. Raw time-resolved photoluminescence decays including the individual fitting components and their relative amplitudes (for biexponential decays) are available in the SM, Figures S16 to S29. Uncertainties for kr were calculated as follows: k r = Φ L τ a v + Φ L τ a v 2   τ a v whereas for knr we used k n r = Φ L τ a v + τ a v τ a v 2   + Φ L τ a v 2   τ a v . Uncertainties of τ are given by the decay fitting. n.d. = not determined, due to bad signal-to-noise ratios.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Friedel, J.; Krause, M.; Jordan, R.; Maisuls, I.; Brünink, D.; Schwab, D.; Doltsinis, N.L.; Strassert, C.A.; Klein, A. Triplet Emitting C^N^C Cyclometalated Dibenzo[c,h]Acridine Pt(II) Complexes. Molecules 2022, 27, 8054. https://doi.org/10.3390/molecules27228054

AMA Style

Friedel J, Krause M, Jordan R, Maisuls I, Brünink D, Schwab D, Doltsinis NL, Strassert CA, Klein A. Triplet Emitting C^N^C Cyclometalated Dibenzo[c,h]Acridine Pt(II) Complexes. Molecules. 2022; 27(22):8054. https://doi.org/10.3390/molecules27228054

Chicago/Turabian Style

Friedel, Joshua, Maren Krause, Rose Jordan, Iván Maisuls, Dana Brünink, Dominik Schwab, Nikos L. Doltsinis, Cristian A. Strassert, and Axel Klein. 2022. "Triplet Emitting C^N^C Cyclometalated Dibenzo[c,h]Acridine Pt(II) Complexes" Molecules 27, no. 22: 8054. https://doi.org/10.3390/molecules27228054

Article Metrics

Back to TopTop