Next Article in Journal
Recent Advances in the Synthesis and Antioxidant Activity of Low Molecular Mass Organoselenium Molecules
Previous Article in Journal
Structural Fluctuation, Relaxation, and Folding of Protein: An Approach Based on the Combined Generalized Langevin and RISM/3D-RISM Theories
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pt(II) Complexes with Tetradentate C^N*N^C Luminophores: From Supramolecular Interactions to Temperature-Sensing Materials with Memory and Optical Readouts

1
Institut für Anorganische und Analytische Chemie, Universität Münster, Corrensstraße 28/30, D-48149 Münster, Germany
2
Center of Nanotechnology (CeNTech), Center for Soft Nanosciences (SoN), Cells in Motion Interfaculty Cluster (CiMIC), Universität Münster, Heisenbergstraße 11, D-48149 Munster, Germany
3
Center for Nanointegration Duisburg-Essen (CENIDE), Faculty of Chemistry (Organic Chemistry), University of Duisburg-Essen, Universitätsstraße 7, D-45141 Essen, Germany
4
Center for Multiscale Theory and Computation, Institut für Festkörpertheorie, Universität Münster, Wilhelm-Klemm-Straße 10, D-48149 Münster, Germany
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(21), 7353; https://doi.org/10.3390/molecules28217353
Submission received: 14 September 2023 / Revised: 10 October 2023 / Accepted: 17 October 2023 / Published: 30 October 2023
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
A series of four regioisomeric Pt(II) complexes (PtLa-n and PtLb-n) bearing tetradentate luminophores as dianionic ligands were synthesized. Hence, both classes of cyclometallating chelators were decorated with three n-hexyl (n = 6) or n-dodecyl (n = 12) chains. The new compounds were unambiguously characterized by means of multiple NMR spectroscopies and mass spectrometry. Steady-state and time-resolved photoluminescence spectroscopy as well quantum chemical calculations show that the effect of the regioisomerism on the emission colour and on the deactivation rate constants can be correlated with the participation of the Pt atom on the excited state. The thermal properties of the complexes were studied by DSC, POM and temperature-dependent steady-state photoluminescence spectroscopy. Three of the four complexes (PtLa-12, PtLb-6 and PtLb-12) present an intriguing thermochromism resulting from the responsive metal–metal interactions involving adjacent monomeric units. Each material has different transition temperatures and memory capabilities, which can be tuned at the intermolecular level. Hence, dipole–dipole interactions between the luminophores and disruption of the crystalline packing by the alkyl groups are responsible for the final properties of the resulting materials.

Graphical Abstract

1. Introduction

Luminescent Pt(II) complexes have been widely studied as optical sensors for both physical (mechanic [1,2,3,4], temperature [5,6,7], polarity [6]) and chemical (vapours [2,8,9], ions [10,11], small molecules [12], molecular oxygen [5,13,14]) stimuli. Moreover, several patents on their use in optoelectronic devices and as sensors have been applied for [15,16,17,18]. The phosphorescence of Pt(II) complexes, which is widely discussed in the bibliographic literature, generally stems from metal-perturbed ligand-centred triplet states (3MP-LC), where the emissive species possesses a mixed 3LC (ligand-centred, i.e., π-π*) and 3MLCT (metal-to-ligand charge transfer, i.e., d-π*) character [19,20,21,22]. There are numerous examples of Pt(II) complexes with both bidentate and tridentate ligands that are flexible enough to allow for geometrical distortion out of the coordination plane, which upon photoexcitation facilitates non-radiative decay pathways through conical intersections with the ground state [23,24]. Hence, the use of tetradentate ligands can increase the rigidity and, in consequence, enhance the photoluminescence efficiency of the complexes [25,26]. Moreover, cyclometallation can be further used to enhance the ligand-field splitting by the introduction of strong σ-donors, raising the antibonding dx2-y2 orbital to higher energies and, in consequence, destabilizing dissociative states (including d-d* and π-d* configurations), which in turn become less thermally accessible [26,27,28,29,30].
Several Pt(II)-based species with sensing applications rely on their capability to form aggregates, both in solution and in the solid state. The formation of Pt(II)-based aggregates is possible due to their planar coordination geometry and is mainly driven by van der Waals interactions. When the distance between the Pt atoms is approximately 3.5 Å or shorter (i.e., below the sum of their van der Waals radii), coupling between the dz2 orbital lobes protruding out of the coordination plane becomes feasible. The resulting aggregates show a red-shifted emission if compared with the monomeric species, which arises from triplet metal–metal-to-ligand charge transfer (3MMLCT) states with a certain degree of excimeric character [14,21,31,32]. Hence, we have now prepared a series of Pt(II) complexes where the modification in the length and position of peripheral alkyl chains allows us to control the photophysical properties, the characteristics of the solid phase and the transition temperatures of the resulting solids. These concepts can lead to temperature-sensing materials with memory based on Pt-Pt interactions and providing optical readouts.
The complexes presented in this work have been developed as a natural progression of our previous research [4]. The herein-reported replacement of the methoxyphenyl group at the bridging N-atom by alkyl chains aims at promoting the aggregation of the complexes. This stacking was partially hampered in our previous complexes, due to the 90° rotation of the phenyl ring with respect to the luminophoric plane [4,14]. In addition, modulation of the length and position of the alkyl chains affects the packing and supramolecular interactions between the molecular units in the solid state, leading to changes in the thermal properties of the complexes.

2. Synthesis

The synthesis of the ligands (Scheme 1) started with the reaction between 2,6-dibromopyridine and 2-amino-6-bromopyridine promoted by NaH to give the precursor 1. The intermediate 1 was reacted with 1-bromohexane or 1-bromododecane to yield the precursors 2-6 and 2-12, respectively. A further Suzuki–Miyaura cross-coupling with an excess of the corresponding hydroxyphenylboronic acid allowed us to prepare the intermediates 3a-n and 3b-n (n = 6 or 12). The phenolic species were not isolated and the ligands La-n and Lb-n were finally obtained by Williamson’s ether synthesis using the corresponding alkylbromide [4]. The Pt(II) complexes were prepared by a standard cycloplatination reaction with K2PtCl4 in acetic acid at reflux with yields between 30 and 44%, which are in agreement with previously reported syntheses [4,14]. The comparable yields obtained for both regioisomers suggest that, in the case of PtLb-n complexes, only one of the three potential cyclometallation products is formed. This result may be attributed to the steric hindrance imposed by the alkyl chains during the cycloplatination reaction. The complexes show a good solubility in polar organic solvents, such as DCM and THF, mostly due to the presence of the peripheral alkyl chains [4]. All the precursors and complexes were unambiguously characterized by means of 2D-NMR (including a full assignment of 1H and 13C signals) as well as by high-resolution mass spectrometry (Figures S1–S61). Further experimental details can be found in the ESI.

3. Photophysics in Diluted Conditions

The photophysical characterization of the complexes was initially performed in diluted DCM solutions (complex concentration c ≈ 10−5 M) at room temperature. The absorption bands in the region between 235 and 300 nm can be assigned to transitions into 1LC states, while the bands in the region between 350 and 450 nm are related to mixed singlet states with variable degrees of LC and MLCT character, as was previously reported for comparable complexes (Figure 1 left, Figure S62 and Table S1) [14,21,33,34,35]. Interestingly, the PtLb-n complexes present a stronger absorption band in the 1MLCT (≈420 nm) region if compared with the PtLa-n complexes. This can be attributed to the higher contribution of the alkoxy group in para position (with respect to the Pt atom) to the HOMO than in case of the moiety in meta position, and therefore to a higher destabilization of the HOMO (as previously observed for Ir(III) complexes and [36,37] in phenylpyridine Pt(II)-based complexes [38]).
The emission spectra of the PtLa-n complexes (Figure 1 right, Figure S63) show a maximum at 517 nm with vibrational shoulders at 553 nm, which arise mainly from 3MP-LC states [14,33,34,39]. The spectra of PtLb-n are red-shifted with a maximum at 555 nm (Figure 1 right, Figure S64). The red shift in the emission can be also ascribed to a destabilization of the HOMO and consequently to a reduction in the gap between the excited and ground state [37,38]. To support and interpret the experimental results, (TD-)DFT calculations were performed for the methoxy-substituted complexes (PtLa-1 and PtLb-1) as model compounds. After the geometry optimization of the ground state (S0) and the lowest triplet state (T1) was achieved, the emission maxima were calculated (Figure S77) [40,41,42,43,44,45,46,47].
A detailed characterization of the emissive state can be achieved with the aid of a correlated electron-hole pair analysis using the software package for Theoretical Density, Orbital Relaxation and Exciton analysis (TheoDORE) [48]. The result of the decomposition into contributions from MLCT, LMCT (ligand-to-metal charge transfer), LC and MC is presented in Figure 2. The higher MLCT character in the excited state of PtLb-n, particularly if compared with PtLa-n (25.0% vs 21.1%, respectively), may be responsible for the lack of vibronic progression in PtLb-n complexes, due the stronger geometrical distortion in the excited state. The distortion of the excited state (ΔQ) can be estimated by the so-called Huang–Rhys factor (S) according to Equation (1), where I0-0 is the emission intensity of the 0-0 transition and I1-0 the one corresponding to the first vibronic peak (Table 1) [26]. Consistently, S is higher for PtLb-n than for PtLa-n complexes.
S = I 0 1 I 0 0
The photophysical properties were also studied in diluted glassy matrices of 2-methyltetrahydrofuran (complex concentration, c ≈ 10−5 M) at 77 K. All the complexes show a blue shift in the emission spectra of approximately 10 nm, if compared with the solutions at r.t. This is a consequence of the weaker charge-transfer stabilization associated with the limited solvent reorientation in the rigid matrices, thereby decreasing the 3MLCT character of the emissive state [49,50]. Another consequence of the restricted solvent reorientation is the reduced density of solvent-related roto-vibrational states, thus leading to a well-defined vibrational progression [51]. Interestingly, the reduction in S is more significant for PtLb-n than for PtLa-n when the fluid solution is changed to the glassy matrix, which is in agreement with a higher loss of MLCT character from the lack of solvent reorientation.
The excited state of Pt(II) complexes can be deactivated in fluid solution by 3O2-mediated diffusional quenching, which is reflected in the decrease (by a factor > 50) in the knr after Ar purging of the samples. The PtLa-n series shows longer lifetimes (Figures S65–S76) and higher quantum yields than their PtLb-n isomers (Table 2); a similar effect was previously reported for Ir(III) complexes [37].
When the complexes are studied in dilute fluid solutions, different conformations are thermally accessible and possess comparable excited state characters yet different deactivation rates; therefore, multiexponential decays can be observed in some cases. This effect is also noticeable in glassy matrices, yielding multiexponential decays as previously reported for other Pt(II) complexes [33,51,52]. Compared to the analogous complex with a methoxyphenyl group (which has been previously reported by our group [4]), PLa-6 presents somewhat higher kr and knr rates. This can be attributed to the reduced steric strain associated with the n-alkyl substitution, as opposed to a phenyl moiety at the bridging N-atom. The kr and knr values of PtLa-n and PtLb-n show two interesting trends. The kr are in the same range for both luminophores; however, PtLb-n complexes present knr values between 2 and 3 times higher than the PtLa-n compounds. In this case, the faster non-radiative intersystem crossing rate constant between the T1 and S0 states can be ascribed to the higher geometrical distortion in the excited state of PtLb-n, if compared with the PtLa-n complexes. It would be expected that a lower MLCT character of the excited state leads to a higher kr at 77 K than at r.t. However, the four complexes show a counterintuitive tendency, with higher kr at lower temperatures. This was already observed for other Pt(II) complexes with tridentate luminophores, and (TD-)DFT calculations at a higher level of sophistication (including a temperature-dependent Boltzmann analysis of the radiative rate constants) will be needed to elucidate the nature of this observation [53].
Table 2. Photophysical parameters of the complexes.
Table 2. Photophysical parameters of the complexes.
PtLa-6PtLa-12PtLb-6PtLb-12
τav a
(µs)
ΦL bkr c/104 (s−1) knr d/104 (s−1)τav a
(µs)
ΦL bkr c/104 (s−1)knr d/104 (s−1)τav a
(µs)
ΦL bkr c/104 (s−1)knr d/104 (s−1)τav a
(µs)
ΦL bkr c/104 (s−1)knr d/104 (s−1)
Air-equilibrated DCM solution (r.t.)0.203 ± 0.001 e<0.02<10480
< knr < 490
0.174 ± 0.001 e<0.02<12560
< knr < 570
0.163 ± 0.001 e<0.02<12600
< knr < 610
0.209 ± 0.001 f<0.02<10470
< knr < 480
Ar-purged DCM solution (r.t.)11.22 ± 0.02 e0.49 ± 0.024.4 ± 0.24.5 ± 0.210.60 ± 0.02 f0.57 ± 0.025.4 ± 0.24.1 ± 0.26.102 ± 0.003 e0.26 ± 0.024.3 ±
0.3
12.1 ± 0.38.55 ± 0.03 f0.27 ± 0.023.2 ± 0.28.5 ± 0.3
2-MeTHF glassy matrix (77 K)16.8 ±
0.1 e
0.99 ± 0.016.0 ± 0.2<611.4 ±
0.8 f
0.99 ± 0.018.8 ± 0.8<97.60 ± 0.03 f0.99 ± 0.0113.2 ± 0.3<138.7 ±
0.1 f
0.99 ± 0.0111.6 ±
0.4
<11
a Amplitude-weighted average lifetime. b Photoluminescence quantum yield. c Average radiative rate constant k r = Φ L τ a v ; k r = Φ L τ a v + Φ L τ a v 2   τ a v [54]. d Average radiationless deactivation rate constant k n r = 1 Φ L τ a v ; k n r = Φ L τ a v + τ a v τ a v 2   + Φ L τ a v 2   τ a v [54]. e Monoexponential decay. f Biexponential decay.

4. Thermochromic Properties

The thermochromic properties of the compounds were studied by differential scanning calorimetry (DSC), polarized optic microscopy (POM) and temperature-dependent steady-state luminescence spectroscopy.
In DSC experiments (Table 3, Figures S78–S81), PtLa-6 melts at 245 °C with decomposition. No changes below this temperature were observed in POM. PtLa-12 shows, within a first cycle, a crystal-to-crystal transition at 74 °C, while melting at 197 °C and crystallizing again at 186 °C (Figure S82). Changes in the melting temperature within cycles for PtLa-12 indicate a small decomposition during the phase transition. PtLb-6 and PtLb-12 both melt at around 130 °C, returning not to a crystalline phase, but rather to a glassy phase in a reversible way (Figures S83 and S84). In the case of the “open” series PtLb-n, the increasing chain length produces an enhancement of the van der Waals interactions between the alkyl groups while causing a slight raise of the melting temperature and a significant increment of the melting enthalpy.
Interestingly, the transitions undergone by PtLa-12 involve a change in the luminescence colour, with a decrease of the monomeric emission and the growth of an unstructured yet red-shifted band. This new luminescence can be ascribed to the formation of dimers and aggregates where the Pt···Pt distances are shorter than 3.5 Å. [3,14,34,55]. The assignment of this emission band was carried out by comparing the calculated emission spectra of the dimers with the experimental bands measured upon heating of the samples. The calculated emission spectra of PtLa-1 and PtLb-1 dimers are in very good agreement with the experimental data (Figure 3).
Powder X-ray diffractometry of PtLa-12 was performed before and after heating the sample above the transition temperature. The diffraction pattern before the thermal treatment shows typical Bragg reflections of a polycrystalline sample. Upon heating, no Bragg reflections are observed while showing the formation of a mostly amorphous solid (Figure 4).
In the case of complexes PtLb-6 and PtLb-12, the monomeric emission intensities decrease with increasing temperature (Figures S86 and S87), as expected for the progressive population of roto-vibrational states while leading to faster non-radiative deactivation rates. On the other hand, PtLa-12 shows a very stable emission intensity at 512 nm as a function of temperature below the transition temperature. In the range between 65 °C and 75 °C, the emission intensity at 512 nm decreases by roughly 80%.
The emission spectra of PtLa-12, PtLb-6 and PtLb-12 were measured 48 h after melting (Figure 5). Notably, each complex exhibits a different behaviour. PtLa-12 keeps the emission of the aggregates kinetically blocked and PtLb-6 shows partial recovery, whereas for PtLb-12 the original emission (i.e., before heating) is fully recovered. These observations can be explained in terms of the intermolecular interactions within the solids: the “closed” complexes PtLa-n possess a higher dipolar moment than the “open” PtLb-n series. Consequently, the interaction between the aromatic cores within the aggregates is more pronounced. Conversely, the longer alkyl moieties disrupt the interactions between the aromatic planes [56,57]. As a result of these combined effects, PtLb-12 is able to return to a monomer-dominated phase, whereas PtLb-6 achieves only a partial reversion. Ultimately, the more robust dipole interactions sustain the aggregates in the case of PtLa-12.

5. Conclusions

Four new Pt(II) complexes with C^N*N^C ligands were synthesized and characterized. The position of the alkoxy groups has a crucial effect on the photophysical properties of the monomeric complexes by varying the metal participation on the excited state. The PtLb-n complexes with an extended metal contribution display emission maxima that are red-shifted by approximately 45 nm while having higher knr values in comparison with their PtLa-n isomers; this demonstrates that small changes in the ligand structure can significantly affect the photophysical properties. The alkyl chain length affects the capability of the complexes to form crystalline phases and their transition temperatures. Interestingly, the phase transitions facilitate the Pt-Pt coupling, which causes a red-shifted luminescence due to the formation of dimers. Depending on the substitution pattern and the alkyl chain length, the thermal properties of the materials can be tuned. Moreover, the balance between the different classes of intermolecular interactions affects the memory of thermal exposure above the transition temperatures. Hence, variable alkyl chain lengths can be exploited to design materials with different sensitivity ranges, in order to cover appropriated windows for their use as sensors with multiple optical readouts (i.e., based on photoluminescence wavelengths, intensities and lifetimes).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28217353/s1. Figure S1. 1H-NMR spectrum (500 MHz, CD2Cl2) of 1. Figure S2. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of 1. Figure S3. ESI-MS (MeOH) of 1. Figure S4. 1H-NMR spectrum (500 MHz, CD2Cl2) of 2-6. Figure S5. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of 2-6. Figure S6. ESI-MS (MeOH) of 2-6. Figure S7. 1H-NMR spectrum (500 MHz, CD2Cl2) of 2-12. Figure S8. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of 2-12. Figure S9. ESI-MS (MeOH) of 2-12. Figure S10. 1H-NMR spectrum (500 MHz, CD2Cl2) of La-6. Figure S11. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of La-6. Figure S12. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of La-6. Figure S13. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of La-6. Figure S14. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of La-6. Figure S15. ESI-MS (MeOH) of La-6. Figure S16. 1H-NMR spectrum (500 MHz, CD2Cl2) of La-12. Figure S17. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of La-12. Figure S18. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of La-12. Figure S19. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of La-12. Figure S20. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of La-12. Figure S21. ESI-MS (MeOH) of La-12. Figure S22. 1H-NMR spectrum (500 MHz, CD2Cl2) of Lb-6. Figure S23. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of Lb-6. Figure S24. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of Lb-6. Figure S25. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of Lb-6. Figure S26. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of Lb-6. Figure S27. ESI-MS (MeOH) of Lb-6. Figure S28. 1H-NMR spectrum (400 MHz, CD2Cl2) of Lb-12. Figure S29. 13C-{1H}-NMR spectrum (101 MHz, CD2Cl2) of Lb-12. Figure S30. 1H/1H-COSY-NMR spectrum (400 MHz/400 MHz, CD2Cl2) of Lb-12. Figure S31. 1H/13C-gHSQC-NMR spectrum (400 MHz/101 MHZ, CD2Cl2) of Lb-12. Figure S32. 1H/13C-gHMBC-NMR spectrum (400 MHz/101 MHz, CD2Cl2) of Lb-12. Figure S33. ESI-MS (MeOH) of Lb-12. Figure S34. 1H-NMR spectrum (400 MHz, CD2Cl2) of PtLa-6. Figure S35. 13C-{1H}-NMR spectrum (101 MHz, CD2Cl2) of PtLa-6. Figure S36. 195Pt-NMR spectrum (86 MHz, CD2Cl2) of PtLa-6. Figure S37. 1H/1H-COSY-NMR spectrum (400 MHz/400 MHz, CD2Cl2) of PtLa-6. Figure S38. 1H/13C-gHSQC-NMR spectrum (400 MHz/101 MHZ, CD2Cl2) of PtLa-6. Figure S39. 1H/13C-gHMBC-NMR spectrum (400 MHz/101 MHz, CD2Cl2) of PtLa-6. Figure S40. ESI-MS (MeOH) of PtLa-16. Figure S41. 1H-NMR spectrum (500 MHz, CD2Cl2) of PtLa-12. Figure S42. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of PtLa-12. Figure S43. 195Pt-NMR spectrum (107 MHz, CD2Cl2) of PtLa-12. Figure S44. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of PtLa-12. Figure S45. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of PtLa-12. Figure S46. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of PtLa-12. Figure S47. ESI-MS (MeOH) of PtLa-12. Figure S48. 1H-NMR spectrum (500 MHz, CD2Cl2) of PtLb-6. Figure S49. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of PtLb-6. Figure S50. 195Pt-NMR spectrum (107 MHz, CD2Cl2) of PtLb-6. Figure S51. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of PtLb-6. Figure S52. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of PtLb-6. Figure S53. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of PtLb-6. Figure S54. ESI-MS (MeOH) of PtLb-6. Figure S55. 1H-NMR spectrum (500 MHz, CD2Cl2) of PtLb-12. Figure S56. 13C-{1H}-NMR spectrum (126 MHz, CD2Cl2) of PtLb-12. Figure S57. 195Pt-NMR spectrum (107 MHz, CD2Cl2) of PtLa-12. Figure S58. 1H/1H-COSY-NMR spectrum (500 MHz/500 MHz, CD2Cl2) of PtLb-12. Figure S59. 1H/13C-gHSQC-NMR spectrum (500 MHz/126 MHZ, CD2Cl2) of PtLb-12. Figure S60. 1H/13C-gHMBC-NMR spectrum (500 MHz/126 MHz, CD2Cl2) of PtLb-12. Figure S61. ESI-MS (MeOH) of PtLb-12. Figure S62. Absorption spectra of the four complexes in DCM at room temperature. Table S1. Absorption coefficients of the four complexes at selected wavelengths. Figure S63. Excitation (left, at the corresponding emission maximum) and emission (right, λexc = 350 nm) spectra of PtLa-6 and PtLa-12. Excitation (left, at the corresponding emission maximum) and emission (right, λexc = 350 nm) spectra of PtLb-6 and PtLb-12. Figure S65. Left: Raw time-resolved photoluminescence decay of PtLa-6 (c ≈ 10−5 M) in fluid air-equilibrated DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 515 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S66. Left: Raw time-resolved photoluminescence decay of PtLa-6 (c ≈ 10−5 M) in fluid Ar-purged DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 515 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S67. Left: Raw time-resolved photoluminescence decay of PtLa-6 (c ≈ 10−5 M) in a frozen glassy matrix of 2Me-THF at 77 K, including the residuals (λexc = 376.7 nm, λem = 505 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S68. Left: Raw time-resolved photoluminescence decay of PtLa-12 (c ≈ 10−5 M) in fluid air-equilibrated DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 515 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S69. Left: Raw time-resolved photoluminescence decay of PtLa-12 (c ≈ 10−5 M) in fluid Ar-purged DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 515 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S70. Left: Raw time-resolved photoluminescence decay of PtLa-12 (c ≈ 10−5 M) in a frozen glassy matrix of 2Me-THF at 77 K, including the residuals (λexc = 376.7 nm, λem = 505 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S71. Left: Raw time-resolved photoluminescence decay of PtLb-6 (c ≈ 10−5 M) in fluid air-equilibrated DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 555 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S72. Left: Raw time-resolved photoluminescence decay of PtLb-6 (c ≈ 10−5 M) in fluid Ar-purged DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 555 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S73. Left: Raw time-resolved photoluminescence decay of PtLb-6 (c ≈ 10−5 M) in a frozen glassy matrix of 2Me-THF at 77 K, including the residuals (λexc = 376.7 nm, λem = 540 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S74. Left: Raw time-resolved photoluminescence decay of PtLb-12 (c ≈ 10−5 M) in fluid air-equilibrated DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 555 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S75. Left: Raw time-resolved photoluminescence decay of PtLb-12 (c ≈ 10−5 M) in fluid Ar-purged DCM at r.t., including the residuals (λexc = 376.7 nm, λem = 555 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S76. Left: Raw time-resolved photoluminescence decay of PtLb-12 (c ≈ 10−5 M) in a frozen glassy matrix of 2Me-THF at 77 K, including the residuals (λexc = 376.7 nm, λem = 540 nm). Right: Fitting parameters including pre-exponential factors and confidence limits. Figure S77. Calculated emission spectra at 77 K in THF. Figure S78. DSC profile of PtLa-6 on heating with a heating/cooling rate of 10 °C/min. First cycle (black and red curves), second cycle (blue and green curves), third cycle (purple and brown curves). Figure S79. DSC profile of PtLa-12 on heating with a heating/cooling rate of 10 °C/min. First cycle (black and red curves), second cycle (blue and green curves), third cycle (purple and brown curves). Figure S80. DSC profile of PtLb-6 on heating with a heating/cooling rate of 10 °C/min. First cycle (black and red curves), second cycle (blue and green curves), third cycle (purple and brown curves). Figure S81. DSC profile of PtLb-12 on heating with a heating/cooling rate of 10 °C/min. First cycle (black and red curves), second cycle (blue and green curves), third cycle (purple and brown curves). Figure S82. POM images of PtLa-12 in the crystalline phase (left) and in the isotropic phase (right) after melting on cooling. Figure S83. POM images of PtLb-6 in the glassy phase (left) and in the isotropic phase (right) after melting on cooling. Figure S84. POM images of PtLb-12 in the glassy phase (left) and in the isotropic phase (right) after melting on cooling. Figure S85. Left: Emission spectra of solid PtLa-12 upon heating (λexc = 470 nm). Centre: Emission intensity at 512 nm and 661 nm as a function of the temperature. Right: Photographs of solid PtLa-12 at different temperatures under UV excitation, λexc = 365 nm. Figure S86. Left: Emission spectra of solid PtLb-6 upon heating (λexc = 470 nm). Centre: Emission intensity at 512 nm and 661 nm as a function of the temperature. Right: Photographs of solid PtLa-12 at different temperatures under UV excitation, λexc = 365 nm. Figure S87. Left: Emission spectra of solid PtLb-12 upon heating (λexc = 470 nm). Centre: Emission intensity at 512 nm and 661 nm as a function of the temperature. Right: Photographs of solid PtLa-12 at different temperatures under UV excitation, λexc = 365 nm.

Author Contributions

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

Funding

M.E.G.S. gratefully acknowledges a doctoral fellowship from the Deutscher Akademischer Austauschdienst (DAAD). The authors gratefully acknowledge funding from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)—Project-ID 433682494–SFB 1459. 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”).

Data Availability Statement

Further data supporting the results can be found in the electronic supporting information.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, A.; Du, P.; Sun, Z.; Wu, H.; Jia, H.; Zhang, R.; Liang, Z.; Cao, R.; Eisenberg, R. Reversible Mechanochromic Luminescence at Room Temperature in Cationic Platinum(II) Terpyridyl Complexes. Inorg. Chem. 2014, 53, 3338–3344. [Google Scholar] [CrossRef]
  2. Zhang, X.; Wang, J.Y.; Ni, J.; Zhang, L.Y.; Chen, Z.N. Vapochromic and Mechanochromic Phosphorescence Materials Based on a Platinum(II) Complex with 4-Trifluoromethylphenylacetylide. Inorg. Chem. 2012, 51, 5569–5579. [Google Scholar] [CrossRef] [PubMed]
  3. Theiss, T.; Buss, S.; Maisuls, I.; López-Arteaga, R.; Brünink, D.; Kösters, J.; Hepp, A.; Doltsinis, N.L.; Weiss, E.A.; Strassert, C.A. Room-Temperature Phosphorescence from Pd(II) and Pt(II) Complexes as Supramolecular Luminophores: The Role of Self-Assembly, Metal-Metal Interactions, Spin-Orbit Coupling, and Ligand-Field Splitting. J. Am. Chem. Soc. 2023, 145, 3937–3951. [Google Scholar] [CrossRef]
  4. Gutierrez Suburu, M.E.; Maisuls, I.; Kösters, J.; Strassert, C.A. Room-Temperature Luminescence from Pd(II) and Pt(II) Complexes: From Mechanochromic Crystals to Flexible Polymer Matrices. Dalton Trans. 2022, 51, 13342–13350. [Google Scholar] [CrossRef] [PubMed]
  5. Köse, M.E.; Carroll, B.F.; Schanze, K.S. Preparation and Spectroscopic Properties of Multiluminophore Luminescent Oxygen and Temperature Sensor Films. Langmuir 2005, 21, 9121–9129. [Google Scholar] [CrossRef] [PubMed]
  6. Zheng, Q.; Borsley, S.; Tu, T.; Cockroft, S.L. Reversible Stimuli-Responsive Chromism of a Cyclometallated Platinum(Ii) Complex. Chem. Commun. 2020, 56, 14705–14708. [Google Scholar] [CrossRef]
  7. Ma, Y.; Chen, K.; Lu, J.; Shen, J.; Ma, C.; Liu, S.; Zhao, Q.; Wong, W.Y. Phosphorescent Soft Salt Based on Platinum(II) Complexes: Photophysics, Self-Assembly, Thermochromism, and Anti-Counterfeiting Application. Inorg. Chem. 2021, 60, 7510–7518. [Google Scholar] [CrossRef]
  8. Kobayashi, A.; Kato, M. Vapochromic Platinum(II) Complexes: Crystal Engineering toward Intelligent Sensing Devices. Eur. J. Inorg. Chem. 2014, 27, 4469–4483. [Google Scholar] [CrossRef]
  9. Grove, L.J.; Rennekamp, J.M.; Jude, H.; Connick, W.B. A New Class of Platinum(II) Vapochromic Salts. J. Am. Chem. Soc. 2004, 126, 1594–1595. [Google Scholar] [CrossRef]
  10. Wong, Y.S.; Ng, M.; Yeung, M.C.L.; Yam, V.W.W. Platinum(II)-Based Host-Guest Coordination-Driven Supramolecular Co-Assembly Assisted by Pt···Pt and π-ΠStacking Interactions: A Dual-Selective Luminescence Sensor for Cations and Anions. J. Am. Chem. Soc. 2021, 143, 973–982. [Google Scholar] [CrossRef]
  11. Soto, M.A.; Carta, V.; Cano, M.T.; Andrews, R.J.; Patrick, B.O.; MacLachlan, M.J. Multiresponsive Cyclometalated Crown Ether Bearing a Platinum(II) Metal Center. Inorg. Chem. 2022, 61, 2999–3006. [Google Scholar] [CrossRef] [PubMed]
  12. Sinn, S.; Biedermann, F.; De Cola, L. Platinum Complex Assemblies as Luminescent Probes and Tags for Drugs and Toxins in Water. Chem. Eur. J. 2017, 23, 1965–1971. [Google Scholar] [CrossRef] [PubMed]
  13. Yeh, T.S.; Chu, C.S.; Lo, Y.L. Highly Sensitive Optical Fiber Oxygen Sensor Using Pt(II) Complex Embedded in Sol-Gel Matrices. Sens. Actuators B Chem. 2006, 119, 701–707. [Google Scholar] [CrossRef]
  14. 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] [PubMed]
  15. De Cola, L.; Strassert, C.A.; Mydlak, M.; Felicetti, M.; Diener, G.; Leonhardt, J. Neu Platin(II)Komplexe Als Triplett-Emitter Für OLED Anwendungen. German Patent DE102012108129A1, 31 August 2012. [Google Scholar]
  16. Kostic, N.M.; Zhou, X.-Y. Thermochromic Platinum Complexes 1989. U.S. Patent No. 4,857,231, 6 June 1989. [Google Scholar]
  17. Resch-Genger, U.; Wang, C.; Strassert, C.A.; Gutierrez Suburu, M.E.; Maisuls, I. Verwendung von D8-Metallkomplexverbindungen Mit Liganden-Kontrollierten Aggregations- Und Lumineszenzeigenschaften. European Patent EP 3896138A1, 20 October 2021. [Google Scholar]
  18. Arzhakova, O.V.; Bakeev, N.F.; Volynskii, A.L.; Dolgova, A.A.; Yarusheva, L.M.; Vasilievich, P.G. Optochemical Sensor for Sensing O2, and Method of Its Preparation. U.S. Patent No. 8,834,795, 16 September 2014. [Google Scholar]
  19. Paw, W.; Cummings, S.D.; Mansour, M.A.; Connick, W.B.; Geiger, D.K.; Eisenberg, R. Luminescent Platinum Complexes: Tuning and Using the Excited State. Coord. Chem. Rev. 1998, 171, 125–150. [Google Scholar] [CrossRef]
  20. Yam, V.W.W.; Lo, K.K.W. Luminescent Polynuclear D10 Metal Complexes. Chem. Soc. Rev. 1999, 28, 323–334. [Google Scholar] [CrossRef]
  21. Mydlak, M.; Mauro, M.; Polo, F.; Felicetti, M.; Leonhardt, J.; Diener, G.; De Cola, L.; Strassert, C.A. Controlling Aggregation in Highly Emissive Pt(II) Complexes Bearing Tridentate Dianionic N^N^N Ligands. Synthesis, Photophysics, and Electroluminescence. Chem. Mater. 2011, 23, 3659–3667. [Google Scholar] [CrossRef]
  22. Costa, R.D.; Ortí, E.; Bolink, H.J.; Monti, F.; Accorsi, G.; Armaroli, N. Luminescent Ionic Transition-Metal Complexes for Light-Emitting. Angew. Chem. Int. Ed. 2012, 51, 8178–8211. [Google Scholar] [CrossRef]
  23. Yoshida, M.; Kato, M. Cation-Controlled Luminescence Behavior of Anionic Cyclometalated Platinum(II) Complexes. Coord. Chem. Rev. 2020, 408, 213194. [Google Scholar] [CrossRef]
  24. Galstyan, A.; Naziruddin, A.R.; Cebrián, C.; Iordache, A.; Daniliuc, C.G.; De Cola, L.; Strassert, C.A. Correlating the Structural and Photophysical Features of Pincer Luminophores and Monodentate Ancillary Ligands in PtII Phosphors. Eur. J. Inorg. Chem. 2015, 2015, 5822–5831. [Google Scholar] [CrossRef]
  25. 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] [PubMed]
  26. 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] [PubMed]
  27. 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]
  28. Kui, S.C.F.; Keong Chow, P.; Cheng, G.; Kwok, C.C.; Lam Kwong, C.; Low, K.H.; Che, C.M. Robust Phosphorescent Platinum(Ii) Complexes with Tetradentate O^N^C^N Ligands: High Efficiency OLEDs with Excellent Efficiency Stability. Chem. Commun. 2013, 49, 1497–1499. [Google Scholar] [CrossRef] [PubMed]
  29. Li, G.; Wolfe, A.; Brooks, J.; Zhu, Z.Q.; Li, J. Modifying Emission Spectral Bandwidth of Phosphorescent Platinum(II) Complexes Through Synthetic Control. Inorg. Chem. 2017, 56, 8244–8256. [Google Scholar] [CrossRef] [PubMed]
  30. Williams, J.A.G.; Develay, S.; Rochester, D.L.; Murphy, L. Optimising the Luminescence of Platinum(II) Complexes and Their Application in Organic Light Emitting Devices (OLEDs). Coord. Chem. Rev. 2008, 252, 2596–2611. [Google Scholar] [CrossRef]
  31. Aliprandi, A.; Genovese, D.; Mauro, M.; De Cola, L. Recent Advances in Phosphorescent Pt (II) Complexes Featuring Metallophilic Interactions: Properties and Applications. Chem. Lett. 2015, 44, 1152–1169. [Google Scholar] [CrossRef]
  32. Carrara, S.; Aliprandi, A.; Hogan, C.F.; De Cola, L. Aggregation-Induced Electrochemiluminescence of Platinum(II) Complexes. J. Am. Chem. Soc. 2017, 139, 14605–14610. [Google Scholar] [CrossRef]
  33. Wilde, S.; González-Abradelo, D.; Daniliuc, C.G.; Böckmann, M.; Doltsinis, N.L.; Strassert, C.A. Fluorination-Controlled Aggregation and Intermolecular Interactions in Pt(II) Complexes with Tetradentate Luminophores. Isr. J. Chem. 2018, 58, 932–943. [Google Scholar] [CrossRef]
  34. Wilde, S.; Ma, D.; Koch, T.; Bakker, A.; Gonzalez-Abradelo, D.; Stegemann, L.; Daniliuc, C.G.; Fuchs, H.; Gao, H.; Doltsinis, N.L.; et al. Toward Tunable Electroluminescent Devices by Correlating Function and Submolecular Structure in 3D Crystals, 2D-Confined Monolayers, and Dimers. ACS Appl. Mater. Interfaces 2018, 10, 22460–22473. [Google Scholar] [CrossRef]
  35. Ren, J.; Cnudde, M.; Brünink, D.; Buss, S.; Daniliuc, C.G.; Liu, L.; Fuchs, H.; Strassert, C.A.; Gao, H.Y.; Doltsinis, N.L. On-Surface Reactive Planarization of Pt(II) Complexes. Angew. Chem. Int. Ed. 2019, 58, 15396–15400. [Google Scholar] [CrossRef] [PubMed]
  36. Grushin, V.V.; Herron, N.; Le Cloux, D.D.; Marshall, W.J.; Petrov, V.A.; Wang, Y. New, Efficient Electroluminescent Materials Based on Organometallic Ir Complexes. Chem. Commun. 2001, 1, 1494–1495. [Google Scholar] [CrossRef]
  37. Hasan, K.; Bansal, A.K.; Samuel, I.D.W.; Roldán-Carmona, C.; Bolink, H.J.; Zysman-Colman, E. Tuning the Emission of Cationic Iridium (III) Complexes towards the Red through Methoxy Substitution of the Cyclometalating Ligand. Sci. Rep. 2015, 5, 12325. [Google Scholar] [CrossRef] [PubMed]
  38. Brooks, J.; Babayan, Y.; Lamansky, S.; Djurovich, P.I.; Tsyba, I.; Bau, R.; Thompson, M.E. Synthesis and Characterization of Phosphorescent Cyclometalated Platinum Complexes. Inorg. Chem. 2002, 41, 3055–3066. [Google Scholar] [CrossRef] [PubMed]
  39. 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]
  40. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09 Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  41. Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  42. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-Adjusted Ab Initio Pseudopotentials for the Second and Third Row Transition Elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  43. Dunning, T.H.; Hay, P.J. Modern Theoretical Chemistry; Schaefer, H.F., III, Ed.; Plenum: New York, NY, USA, 1977; ISBN 9781475708899. [Google Scholar]
  44. Grimme, S.; Ehrlich, S.; Georigk, L. Software News and Updates Gabedit—A Graphical User Interface for Computational Chemistry Softwares. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  45. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef]
  46. Rappé, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A., III; Skiff, 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]
  47. Hanwell, M.D.; Curits, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An Advanced Semantic Chemical Editor, Visualization, and Analysis Platform. J. Cheminfomatics 2012, 4, 1–17. [Google Scholar] [CrossRef] [PubMed]
  48. 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] [PubMed]
  49. 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]
  50. Maisuls, I.; Boisten, F.; Hebenbrock, M.; Alfke, J.; Schürmann, L.; Jasper-Peter, B.; Hepp, A.; Esselen, M.; Müller, J.; Strassert, C.A. Monoanionic C^N^N Luminophores and Monodentate C-Donor Co-Ligands for Phosphorescent Pt(II) Complexes: A Case Study Involving Their Photophysics and Cytotoxicity. Inorg. Chem. 2022, 61, 9195–9204. [Google Scholar] [CrossRef] [PubMed]
  51. 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] [PubMed]
  52. Sanning, J.; Ewen, P.R.; Stegemann, L.; Schmidt, J.; Daniliuc, C.G.; Koch, T.; Doltsinis, N.L.; Wegner, D.; Strassert, C.A. Scanning-Tunneling-Spectroscopy-Directed Design of Tailored Deep-Blue Emitters. Angew. Chem. Int. Ed. 2015, 54, 786–791. [Google Scholar] [CrossRef] [PubMed]
  53. Krause, M.; Kourkoulos, D.; González-Abradelo, D.; Meerholz, K.; Strassert, C.A.; Klein, A. Luminescent PtII Complexes of Tridentate Cyclometalating 2,5-Bis(Aryl)-Pyridine Ligands. Eur. J. Inorg. Chem. 2017, 5215–5223. [Google Scholar] [CrossRef]
  54. Sillen, A.; Engelborghs, Y. The Correct Use of “Average” Fluorescence Parameters. Photochem. Photobiol. 1998, 67, 475–486. [Google Scholar] [CrossRef]
  55. Lin, C.J.; Liu, Y.H.; Peng, S.M.; Shinmyozu, T.; Yang, J.S. Excimer-Monomer Photoluminescence Mechanochromism and Vapochromism of Pentiptycene-Containing Cyclometalated Platinum(II) Complexes. Inorg. Chem. 2017, 56, 4978–4989. [Google Scholar] [CrossRef]
  56. Suarez, S.A.; Muller, F.; Gutierrez Suburu, M.E.; Fonrouge, A.; Baggio, R.F.; Cukiernik, F.D. Br⋯Br and van Der Waals Interactions along a Homologous Series: Crystal Packing of 1,2-Dibromo-4,5-Dialkoxybenzenes. Acta Crystallogr. B Struct. Sci. Cryst. Eng. Mater. 2016, 72, 693–701. [Google Scholar] [CrossRef]
  57. Fonrouge, A.; Cecchi, F.; Alborés, P.; Baggio, R.; Cukiernik, F.D. Relative Influence of Noncovalent Interactions on the Melting Points of a Homologous Series of 1,2-Dibromo-4,5-Dialkoxybenzenes. Acta Crystallogr. Sect. C Struct. Chem. 2013, 69, 204–208. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of the complexes. (i) NaH/DMF; (ii) alkylbromide, NaH /DMF; (iii) 4-hydroxyphenylboronic acid, K2CO3, Pd(PPh3)4/THF; (iv) 3-hydroxyphenylboronic acid, K2CO3, Pd(PPh3)4/THF; (v) alkylbromide, K2CO3/butanone; (vi) K2PtCl4/acetic acid.
Scheme 1. Synthesis of the complexes. (i) NaH/DMF; (ii) alkylbromide, NaH /DMF; (iii) 4-hydroxyphenylboronic acid, K2CO3, Pd(PPh3)4/THF; (iv) 3-hydroxyphenylboronic acid, K2CO3, Pd(PPh3)4/THF; (v) alkylbromide, K2CO3/butanone; (vi) K2PtCl4/acetic acid.
Molecules 28 07353 sch001
Figure 1. Absorption (left) and emission spectra (λexc = 350 nm, right) of Pta-6 (green) and Ptb-6 (orange) in fluid DCM solution at r.t. (solid lines), in 2-MeTHF glassy matrices at 77 K (dashed lines) and in PMMA films at r.t. (dotted lines).
Figure 1. Absorption (left) and emission spectra (λexc = 350 nm, right) of Pta-6 (green) and Ptb-6 (orange) in fluid DCM solution at r.t. (solid lines), in 2-MeTHF glassy matrices at 77 K (dashed lines) and in PMMA films at r.t. (dotted lines).
Molecules 28 07353 g001
Figure 2. Molecular orbitals dominating the T1 state of PtLa-1 and PtLb-1 at the optimized T1 geometry (left). Characterization of the emissive T1 state by correlated electron–hole pair analysis (right).
Figure 2. Molecular orbitals dominating the T1 state of PtLa-1 and PtLb-1 at the optimized T1 geometry (left). Characterization of the emissive T1 state by correlated electron–hole pair analysis (right).
Molecules 28 07353 g002
Figure 3. Molecular orbitals dominating the description of the T1 state for the dimers of PtLa-1 A (left) and PtLb-1 B (right) at the optimized T1 geometry (left). Calculated (solid lines) and experimental (dashed lines) emission spectra of the aggregates (right).
Figure 3. Molecular orbitals dominating the description of the T1 state for the dimers of PtLa-1 A (left) and PtLb-1 B (right) at the optimized T1 geometry (left). Calculated (solid lines) and experimental (dashed lines) emission spectra of the aggregates (right).
Molecules 28 07353 g003
Figure 4. (a) Emission spectra of solid PtLa-12 upon heating (λexc = 470 nm). (b) Emission intensity at 512 nm and 661 nm as a function of the temperature. (c) X-ray diffraction pattern of PtLa-12 before and after the phase transition. Inset: detail of the X-ray diffraction pattern after the phase transition. (d) Photographs of solid PtLa-12 at different temperatures under UV-light excitation, λexc = 365 nm.
Figure 4. (a) Emission spectra of solid PtLa-12 upon heating (λexc = 470 nm). (b) Emission intensity at 512 nm and 661 nm as a function of the temperature. (c) X-ray diffraction pattern of PtLa-12 before and after the phase transition. Inset: detail of the X-ray diffraction pattern after the phase transition. (d) Photographs of solid PtLa-12 at different temperatures under UV-light excitation, λexc = 365 nm.
Molecules 28 07353 g004
Figure 5. Emission spectra of PtLa-12 (red), PtLb-12 (green) and PtLb-6 (orange) after melting and 48 h storage under ambient conditions (left). Photographs of the complexes PtLa-12 (A), PtLb-12 (B) and PtLb-6 (C) after melting and 48 h storage under ambient conditions as observed under room light and under UV-light irradiation, λexc = 365 nm (right).
Figure 5. Emission spectra of PtLa-12 (red), PtLb-12 (green) and PtLb-6 (orange) after melting and 48 h storage under ambient conditions (left). Photographs of the complexes PtLa-12 (A), PtLb-12 (B) and PtLb-6 (C) after melting and 48 h storage under ambient conditions as observed under room light and under UV-light irradiation, λexc = 365 nm (right).
Molecules 28 07353 g005
Table 1. Huang–Rhys factor in fluid solution at r.t. and in glassy matrices at 77 K for the four complexes.
Table 1. Huang–Rhys factor in fluid solution at r.t. and in glassy matrices at 77 K for the four complexes.
ComplexHuang–Rhys Factor S
Fluid Solution at r.t.Glassy Matrix at 77 K
PtLa-60.610.57
PtLa-120.660.46
PtLb-60.830.61
PtLb-120.800.54
Table 3. Transition temperatures and enthalpies of the complexes.
Table 3. Transition temperatures and enthalpies of the complexes.
Phase Transitions
n612
PtLa-nCr 245 IdecompCr 197 (22.3) I
PtLb-nCr 127 (30.8) ICr 132 (45.4) I
Temperatures given in °C; enthalpies in kJ·mol−1 are indicated in parentheses. Cr: crystalline phase; I: isotropic liquid. Heating rates: 10 °C/s. Values are reported in the second heating cycle except for PLa-6.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gutierrez Suburu, M.E.; Blanke, M.; Hepp, A.; Maus, O.; Schwab, D.; Doltsinis, N.L.; Zeier, W.G.; Giese, M.; Voskuhl, J.; Strassert, C.A. Pt(II) Complexes with Tetradentate C^N*N^C Luminophores: From Supramolecular Interactions to Temperature-Sensing Materials with Memory and Optical Readouts. Molecules 2023, 28, 7353. https://doi.org/10.3390/molecules28217353

AMA Style

Gutierrez Suburu ME, Blanke M, Hepp A, Maus O, Schwab D, Doltsinis NL, Zeier WG, Giese M, Voskuhl J, Strassert CA. Pt(II) Complexes with Tetradentate C^N*N^C Luminophores: From Supramolecular Interactions to Temperature-Sensing Materials with Memory and Optical Readouts. Molecules. 2023; 28(21):7353. https://doi.org/10.3390/molecules28217353

Chicago/Turabian Style

Gutierrez Suburu, Matias E., Meik Blanke, Alexander Hepp, Oliver Maus, Dominik Schwab, Nikos L. Doltsinis, Wolfgang G. Zeier, Michael Giese, Jens Voskuhl, and Cristian A. Strassert. 2023. "Pt(II) Complexes with Tetradentate C^N*N^C Luminophores: From Supramolecular Interactions to Temperature-Sensing Materials with Memory and Optical Readouts" Molecules 28, no. 21: 7353. https://doi.org/10.3390/molecules28217353

Article Metrics

Back to TopTop