DOI:
10.1039/D5QI01086A
(Research Article)
Inorg. Chem. Front., 2025,
12, 4769-4775
Controlling circularly polarized luminescence of a pyrene modified chiral Zn(II) complex based on a temperature-dependent diastereomer equilibrium and solid-state excimer formation†
Received
7th May 2025
, Accepted 14th June 2025
First published on 17th June 2025
Abstract
A pyrene modified chiral Schiff-base Zn(II) complex was synthesized and its circularly polarized luminescence (CPL) properties were examined. Single crystal X-ray diffraction analysis of the (S,S)-isomer revealed its distorted tetrahedral coordination geometry with Λ coordination chirality. An equilibrium of the two diastereomers with different coordination chiralities, Λ-(S,S) and Δ-(S,S), was observed in solution, and the ratio of the isomers was found to change with temperature. The complex showed negligible CPL in CHCl3 at room temperature, whereas a clear mirror-image signal was obtained when the solution was cooled to 253 K. Interestingly, excimer-derived long wavelength emission was also observed in the solid state, resulting in CPL sign inversion compared to the solution state. The relationship between their photophysical properties and coordination geometries was theoretically investigated using density functional theory (DFT) calculations.
Introduction
Circularly polarized luminescence (CPL) is a phenomenon in which chiral luminescent materials exhibit left- or right-handed biased CP light.1 In the last decade, various CPL materials such as liquid crystals,2 polymers3 and small organic molecules4 have been developed for application in three-dimensional displays,5 biological probes,6 and anti-counterfeiting materials.7 In particular, recently there has been interest in developing methodologies to control CPL properties using various external stimuli.8 Metal complexes are promising stimuli-responsive materials because various coordination conformations can be designed by adjusting metal ions and chelating ligands.9 A number of chiral metal complexes have been developed and methods to control their CPL properties have been broadly explored.10–14 Zinc(II) complexes are known to adopt various coordination geometries, such as tetrahedral,15–18 trigonal bipyramidal,19–21 square pyramidal,22 and octahedral,23,24 depending on the design of the chelating ligands, enabling unique CPL properties based on their three-dimensional structures. More recently, application of chiral Zn(II) complexes to circularly polarized organic light-emitting diodes (CP-OLEDs) has also been achieved.25
Pyrene is an aromatic hydrocarbon with four fused benzene rings, and its derivatives have a wide range of applications in organic electronics,26 biological probes,27 and optical materials.28 One of the most important characteristics of pyrene derivatives is the formation of excimers in the excited state. It is known that when two pyrene rings are in a chiral orientation, they exhibit CPL properties that are distinct from those of the monomeric state.29 Therefore, the pyrene backbone is considerably useful as a CPL switching component, and a variety of derivatives responsive to various external environments and external stimuli have been reported.21,30–34
Our research group is interested in chiral metal complexes with Schiff-base ligands. In the course of our studies, we have succeeded in controlling their CPL properties by inducing chiral distortions in the coordination structure through appropriate ligand design.35–38 Janiak's group has investigated the Zn(II) complex of the salicylaldiminato ligand derived from chiral 1-phenylethylamine [Zn(O-N)2] and established an equilibrium of diastereomers, differing only in the configuration at the Zn center.39 We prepared and characterized the corresponding naphthalene and phenanthrene complexes, in which CPL inversion was successfully achieved by controlling Λ and Δ coordination chirality.17 In the present work, we studied the pyrene-modified Zn(II) complex 1 (Fig. 1), introducing the new aspect of excimer formation through stacking of the pyrene rings. The complex showed a diastereomer equilibrium between Λ-(S,S) and Δ-(S,S) in solution, and the ratio of the diastereomers was found to change depending on temperature. Temperature-dependent CPL measurements strongly reflect the diastereomer ratio changes, and CPL enhancement was observed at low temperature. Moreover, excimer-derived CPL was observed in the solid state, and the signal was inverted compared with that of the solution state. Density functional theory (DFT) calculations were performed to further understand their chiroptical properties.
 |
| Fig. 1 Chemical structures and the diastereomer equilibrium of chiral zinc(II) complexes (S,S)-1. | |
Results and discussion
Synthesis and structures
The synthetic route to the pyrene modified chiral Zn(II) complex 1 is shown in Scheme 1. 1-Hydroxy-2-pyrenecarboxaldehyde was prepared according to a published procedure.40 Schiff-base ligand (S)-L1 was synthesized by condensation of 1-hydroxy-2-pyrenecarboxaldehyde with commercially available (S)-1-phenylethylamine. The pyrene-modified Zn(II) complex (S,S)-1 was prepared by the reaction of ZnCl2 with (S)-L1 in the presence of tBuOK. The enantiomer, (R,R)-1, was also synthesized by the same method using (R)-1-phenylethylamine as the starting material. Complex 1 was characterized by nuclear magnetic resonance (NMR) spectroscopy (Fig. S1 and S2, ESI†) and high-resolution mass spectrometry (HRMS).
 |
| Scheme 1 Synthesis of the chiral Zn(II) complex (S,S)-1. | |
Single crystals of (S)-L1, (S,S)-1 and (R,R)-1 were obtained by recrystallization from CHCl3/EtOH, and the molecular structures were elucidated by single crystal X-ray diffraction (XRD) analysis. The crystallographic data are presented in Table S1.† ORTEP drawings of (S)-L1 and (S,S)-1 are shown in Fig. 2. The ligand (S)-L1 crystallized in the tetragonal, chiral space group P43212. Because ligand (S)-L1 does not contain any heavy atoms, the Flack parameter (−0.5(10)) is not sufficient to determine the absolute configuration. However, since optically pure (S)-1-phenylethylamine was used as the starting material, the absolute configuration of S is considered correct (Fig. 2a). In contrast, complex (S,S)-1 crystallized in the orthorhombic, chiral space group P212121 with a sufficiently low Flack parameter (−0.006(5)). The chirality angle of O(1)–N(1)–O(2)–N(2) (φ) of (S,S)-1 has the positive value 72.02° (Fig. 2b). Hence, (S,S)-1 has the Λ configuration in the crystal form.17,41 In addition, XRD measurements confirm that the (R,R)-1 configuration is a perfect mirror image of (S,S)-1 (Fig. S4, ESI†). The absolute configuration was assigned as Δ-(R,R) from the negative value of φ (−67.98°). The packing structure in the crystal of (S,S)-1 is shown in Fig. S5,† where non-covalent interactions such as CH–π and π–π stacking were observed between the pyrene rings.
 |
| Fig. 2 ORTEP drawings of (a) (S)-L1 and (b) Λ-(S,S)-1. Thermal ellipsoids are shown at the 50% probability level. Hydrogen atoms and solvent molecules are omitted for clarity. The dihedral angle of the tetrahedral geometry O(1)–N(1)–O(2)–N(2) (φ) is given under the structure in (b). | |
Diastereomer equilibrium in solution
Although the complexes crystallized with single coordination chirality (Λ or Δ) in the crystalline state, a diastereomer equilibrium was observed in solution. The 1H NMR spectra of (S,S)-1 in CDCl3 showed a diastereomer equilibrium based on different signals associated with Λ-(S,S)-1 and Δ-(S,S)-1 (Fig. 3). All proton signals of each diastereomer were successfully assigned using gCOSY and NOESY spectra (Fig. S6 and S7, ESI†). The diastereomer ratio was estimated to be Λ
:
Δ = 1
:
0.38 (diastereomeric excess (de) is 45%) at 293 K from integration values of the 1H NMR spectrum. Upon cooling, the signals of the Δ-isomer gradually increased, and the de reached 36% at 253 K. Such temperature-dependent changes were similar to those of a previous report of our group.17
 |
| Fig. 3 Variable-temperature (VT)-1H NMR spectra (500 MHz) of (S,S)-1 in CDCl3. The peak with an asterisk symbol corresponds to the free ligand in the sample. | |
Photophysical properties
The photophysical properties of complexes (S,S)- and (R,R)-1 in CHCl3 and in the KBr-dispersed pellet state were investigated, and the data of (S,S)-1 are summarized in Table 1. The UV-vis spectrum of (S,S)-1 showed a broad π–π* transition band at approximately 400–550 nm in CHCl3 (Fig. 4 bottom). Compared to previously reported Zn(II) complexes with naphthalene and phenanthrene frameworks,17 a bathochromic shift was observed due to the extension of the π-conjugated system. The circular dichroism (CD) spectra of the two enantiomers, (S,S)-1 and (R,R)-1, were recorded under the same conditions (Fig. 4 top). The two enantiomers showed completely mirror-image signals, with positive Cotton effects for (S,S)-1 and negative Cotton effects for (R,R)-1 in the low energy region attributed to the S0-to-S1 transition. The gabs (=Δε/ε) value of the transition was calculated to be 4.5 × 10−4.
 |
| Fig. 4 CD (upper plot) and UV-vis absorption (lower plot) spectra of (R,R)- and (S,S)-1 in CHCl3 at 293 K. | |
Table 1 Photophysical data for complex (S,S)-1
Medium |
λ
abs [nm] |
λ
max
[nm] |
Φ
,
|
τ
[ns] |
k
r × 107 e [s−1] |
k
nr × 108 f [s−1] |
g
lum
|
CIE [x, y]b |
Data were obtained from 2.0 × 10−4 M solutions at 293 K.
λ
ex = 450 nm.
Luminescence quantum efficiencies measured using the absolute method with an integrating sphere.
λ
ex = 366 nm.
k
r = Φ298 K/τ.
k
nr = (1 − Φ298 K)/τ.
The 2ΔI/I values around emission peak maxima (λmax) are listed.
The data measured at 253 K.
|
CHCl3a |
288, 360, 381, 472 |
549 (547)h |
0.08 |
2.8 |
2.9 |
3.3 |
(2 × 10−3)h |
0.45, 0.55 |
KBr pellet |
— |
565, 621 |
0.03 |
82 (65%), 3.7 (35%) |
0.055 |
0.18 |
−2 × 10−3 |
0.49, 0.51 |
Complex 1 exhibited yellow and orange photoluminescence in CHCl3 and in the KBr-dispersed state, respectively (Fig. S8, ESI†). The red-shifted emission observed in the KBr-dispersed state is attributed to excimer formation induced by molecular aggregation. This interpretation is further supported by the emission lifetime measurements in the KBr pellet, which revealed a long-lived component characteristic of excimer emission (Fig. S10, ESI†). Although well-defined π–π stacking interactions were not observed in the X-ray crystallographic analysis, it is considered that the mechanical grinding during the preparation of the KBr pellet facilitated excimer formation by promoting intermolecular interactions. In the excitation spectra, a red-shift was observed in the KBr-dispersed pellet compared to the solution, suggesting the formation of static excimers due to enhanced intermolecular interactions in the solid state (Fig. S11, ESI†).42,43 Emission quantum efficiencies (Φ) were determined as 8% for solution and 3% for solid samples. The CPL and total emission spectra were measured for CHCl3 solution and KBr-dispersed pellet samples (Fig. 5). While complex 1 showed only weak CPL signals in solution at room temperature (Fig. S12, ESI†), clear signals were observed when the solutions were cooled to 253 K (Fig. 5a, top). Such temperature-dependent CPL enhancement can be attributed to the variation in the diastereomer ratio of complex 1 and will be discussed based on theoretical calculations in a later section. The sign of CPL in the solution state corresponded to the S0-to-S1 transition in CD, which was positive in (S,S)-1 and negative in (R,R)-1 (Fig. 4, top). More interestingly, sign-inverted CPL signals were observed in the KBr-dispersed pellet state compared to the solution state (Fig. 5b, top). Pyrene derivatives are known to exhibit CPL sign inversion depending on the stacking orientation during excimer formation.29 Therefore, the interaction between the pyrene rings of complex 1 could have caused the excimer-derived CPL in the solid state. To investigate excimer formation in the aggregated state of complex 1, polymer-dispersed films with varying concentrations were prepared, and their photoluminescence properties were examined (Fig. S13, ESI†). Green luminescence was observed in a polymethyl methacrylate (PMMA) film doped with 5 wt% of complex 1. At higher concentrations (≥10 wt%), an additional orange emission band appeared in the low-energy region (>600 nm), which is attributed to excimer emission. The emission peaks around 540–570 nm are assigned to monomer emission arising from different conformations of the complex. In the excitation spectra, a red shift was observed with increasing doping concentration of the complex (Fig. S14, ESI†), suggesting the formation of static excimers, similar to the behavior seen in the KBr-dispersed pellet as shown in Fig. S11, ESI.†
 |
| Fig. 5 CPL (upper plot) and total emission (lower plot) spectra of (R,R)- and (S,S)-1. (a) In CHCl3 (2.0 × 10−4 M) at 253 K. (b) In the KBr-dispersed pellet state at 293 K. | |
DFT calculations
DFT calculations were performed for (S,S)-1 at the B3LYP/6-31+G(d,p) level of theory to reveal the correlation between coordination chirality and photophysical properties. Since NMR measurements confirmed that the two diastereomers, Λ-(S,S)-1 and Δ-(S,S)-1, are mixed in solution, structural optimization and estimation of their energies were performed for each diastereomer. Fig. 6 shows the optimized structures of Λ-(S,S)-1 and Δ-(S,S)-1 in the S0 state. The energy difference between the two diastereomers was estimated to be 0.65 kcal mol−1, with the Λ-(S,S)-form being slightly more stable than the Δ-(S,S)-form. This result is consistent with the fact that Λ-(S,S)-1 was the main product in solution as confirmed by 1H NMR spectra (Fig. 3). The frontier orbitals of Λ-(S,S)-1 and Δ-(S,S)-1 and their eigenvalues are given in Fig. S16† (S0) and S17 (S1). The HOMOs of both diastereomers are principally π orbitals of the pyrene framework, whereas the LUMOs are the π* orbitals of both the S0 and S1 states.
 |
| Fig. 6 Optimized structures of Λ-(S,S)-1 and Δ-(S,S)-1 in the S0 ground state estimated by DFT calculations (B3LYP/6-31+G(d,p)). The values in parentheses are relative Gibbs free energies in kcal mol−1. Hydrogen atoms are omitted for clarity. | |
The energy levels and electronic configurations of the singlet states of 1 were estimated from time-dependent (TD)-DFT calculations (B3LYP/6-31+G(d,p)) (Tables S2 and S3†). The major electronic configurations of the upward S0-to-S1 and downward S1-to-S0 transitions are determined to be the HOMO-to-LUMO transition, which suggests that emission from 1 is mainly attributable to π–π* transitions. The transition energies of the upward S0-to-S1 transition were calculated to be 2.65 eV (467 nm) for Λ-(S,S)-1 and 2.60 eV (477 nm) for Δ-(S,S)-1, which are consistent with the absorption wavelength of the π–π* transition band observed in CHCl3 (472 nm; Table 1). The downward S1-to-S0 transition energies were calculated to be 2.28 eV (544 nm) for Λ-(S,S)-1 and 2.21 eV (560 nm) for Δ-(S,S)-1, which also satisfactory reproduced the experimental emission spectrum in CHCl3 (549 nm; Table 1).
To gain further insight into the chiroptical properties of (S,S)-1, we subsequently analysed the transition dipole moments for the S0-to-S1 and S1-to-S0 transitions by using TD-DFT calculations (Fig. 7 and S18†). The magnitude and sign of CD and CPL can be estimated from the theoretical g value defined as g = 4(|μe||μm|cos
θe,m)/(|μe|2 + |μm|2). Here, |μe| and |μm| are the electric and magnetic transition dipole moments, respectively, and θe,m is the angle between the two vectors.44 For the upward S0-to-S1 transition, the two diastereomers were found to exhibit opposite CD signs, positive for Λ-(S,S)-1 and negative for Δ-(S,S)-1, with gabs values on the order of 10−2 (Fig. S16†). In solution, these two transitions cancel each other, resulting in a weak Cotton effect of the order of gabs = 10−4 (Fig. 4 top). In the downward S1-to-S0 transition, both diastereomers were estimated to have a positive CPL (Fig. 7). Furthermore, the glum values were found to differ six-fold between the diastereomers, and a comparably low value of the order of 10−4 was calculated for Λ-(S,S)-1, which was the main product in CHCl3. Combined with the VT-1H NMR results, the enhancement of CPL at low temperatures can be attributed to a shift of the diastereomer equilibrium to Δ-(S,S)-1, which is estimated to have a large glum value. Because both diastereomers were expected to show the same sign of CPL as monomers, the inversion of the CPL signal in the KBr pellet state is considered to be due to excimer formation caused by intermolecular stacking interactions between the pyrene rings.
 |
| Fig. 7 Electric (μe, orange) and magnetic (μm, green) dipole moments of the S1-to-S0 transition for (a) Λ-(S,S)-1 and (b) Δ-(S,S)-1 calculated at the B3LYP/6-31+G (d,p) level. Calculated values of transition dipole moments (|μe|, |μm| and θe,m) and glum are given under each structure. | |
Conclusions
In summary, we have synthesized the pyrene-modified Schiff-base zinc(II) complex 1. The complex crystallized in a single configuration, Λ-(S,S)-1, whereas it exhibits a diastereomer equilibrium in solution based on coordination chirality rearrangement. The diastereomer ratio in solution depended on the temperature, and it was a key factor in improving the CPL properties of the complexes at low temperatures. Interestingly, complex 1 exhibits excimer-derived CPL in the KBr pellet state, and the opposite sign of the CPL signal was obtained compared to that in solution. DFT and TD-DFT calculations of the structures and electronic configurations of (S,S)-1 support the experimental photophysical properties. We believe that the CPL control based on the coordination chirality rearrangements demonstrated in this study will provide important insights for the design of future CPL switching materials.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the ESI.†
Acknowledgements
This work was supported by JSPS KAKENHI (grant numbers JP23K13768 (M. I.), JP23H02040 (Y. I.) and JP21K05234 (T. T.)), Nihon University Industrial Technology Fund for Supporting Young Scholars (M. I.) and Iketani Science and Technology Foundation (grant number 0361227-A (M. I.)). We gratefully acknowledge Prof. Dr Henri Brunner (Universität Regensburg) for helpful discussion and comments. We also acknowledge Prof. Dr Shuichi Suzuki (Osaka University) for HRMS measurements. Elemental analysis was performed at the Center for Creative Materials Research, CST, Nihon University.
References
- J. P. Riehl and F. S. Richardson, Circularly polarized luminescence spectroscopy, Chem. Rev., 1986, 86, 1–16 CrossRef CAS
.
- J. Liu, Z.-P. Song, L.-Y. Sun, B.-X. Li, Y.-Q. Lu and Q. Li, Circularly polarized luminescence in chiral orientationally ordered soft matter systems, Responsive Mater., 2023, 1, e20230005 CrossRef
.
- S. Yang, S. Zhang, F. Hu, J. Han and F. Li, Circularly polarized luminescence polymers: From design to applications, Coord. Chem. Rev., 2023, 485, 215116 CrossRef CAS
.
- T. Zhang, Y. Zhang, Z. He, T. Yang, X. Hu, T. Zhu, Y. Zhang, Y. Tang and J. Jiao, Recent Advances of Chiral Isolated and Small Organic Molecules: Structure and Properties for Circularly Polarized Luminescence, Chem. – Asian J., 2024, 19, e202400049 CrossRef CAS PubMed
.
- D.-Y. Kim, Potential Application of Spintronic Light-Emitting Diode to Binocular Vision for Three-Dimensional Display Technology, J. Korean Phys. Soc., 2006, 49, S505–S508 Search PubMed
.
- G. Muller, Luminescent chiral lanthanide (III) complexes as potential molecular probes, Dalton Trans., 2009, 9692–9707 RSC
.
- F. Zhang, Q. Li, C. Wang, D. Wang, M. Song, Z. Li, X. Xue, G. Zhang and G. Qing, Multimodal, Convertible, and Chiral Optical Films for Anti-Counterfeiting Labels, Adv. Funct. Mater., 2022, 32, 2204487 CrossRef CAS
.
- J.-L. Ma, Q. Peng and C.-H. Zhao, Chem. – Eur. J., 2019, 25, 15441–15454 CrossRef CAS PubMed
.
- S.-Y. Chen, Z.-H. Zhao, C.-H. Li and Q. Li, Responsive materials based on coordination bonds, Responsive Mater., 2023, 1, e20240017 Search PubMed
.
- J. Gong and X. Zhang, Coordination-based circularly polarized luminescence emitters: Design strategy and application in sensing, Coord. Chem. Rev., 2022, 453, 214329 CrossRef CAS
.
- Z.-L. Gong, Z.-Q. Li and Y.-W. Zhong, Circularly polarized luminescence of coordination aggregates, Aggregate, 2022, 3, e177 CrossRef CAS
.
- M. Ikeshita, Y. Imai and T. Tsuno, Recent Advances in Circularly Polarized Luminescent Materials Based on Phosphorescent Platinum(II) Complexes, Chem. Lett., 2025, 54, upaf066 CrossRef
.
- M. Ikeshita, K. Takahashi, N. Hara, S. Kawamorita, N. Komiya, Y. Imai and T. Naota, Ultrasound-induced circularly polarized luminescence based on homochiral aggregation of clothespin-shaped Pt(II) complexes, Responsive Mater., 2024, 2, e20230011 Search PubMed
.
- J.-G. Yang, K. Li, J. Wang, S. Sun, W. Chi, C. Wang, X. Chang, C. Zou, W.-P. To, M.-D. Li, X. Liu, W. Lu, H.-X. Zhang, C.-M. Che and Y. Chen, Controlling Metallophilic Interactions in Chiral Gold(I) Double Salts towards Excitation Wavelength-Tunable Circularly Polarized Luminescence, Angew. Chem., Int. Ed., 2020, 59, 6915–6922 CrossRef CAS PubMed
.
- J. F. Kögel, S. Kusaka, R. Sakamoto, T. Iwashima, M. Tsuchiya, R. Toyoda, R. Matsuoka, T. Tsukamoto, J. Yuasa, Y. Kitagawa, T. Kawai and H. Nishihara, Heteroleptic [Bis(oxazoline)](dipyrrinato)zinc(II) Complexes: Bright and Circularly Polarized Luminescence from an Originally Achiral Dipyrrinato Ligand, Angew. Chem., Int. Ed., 2016, 55, 1377–1381 CrossRef PubMed
.
- Y. Imai and J. Yuasa, Off–off–on chiroptical property switching of a pyrene luminophore by stepwise helicate formation, Chem. Commun., 2019, 55, 4095–4098 RSC
.
- M. Ikeshita, M. Mizugaki, T. Ishikawa, K. Matsudaira, M. Kitahara, Y. Imai and T. Tsuno, Sign control of circularly polarized luminescence of chiral Schiff-base Zn(II) complexes through coordination geometry changes, Chem. Commun., 2022, 58, 7503–7506 RSC
.
- D. Tauchi, K. Kanno, M. Hasegawa, Y. Mazaki, K. Tsubaki, K. Sugiura, T. Shiga, S. Mori and H. Nishikawa, Aggregation-induced enhanced fluorescence emission of chiral Zn(II) complexes coordinated by Schiff-base type binaphthyl ligands, Dalton Trans., 2024, 53, 8926–8933 RSC
.
- H. Isla, M. Srebro-Hooper, M. Jean, N. Vanthuyne, T. Roisnel, J. L. Lunkley, G. Muller, J. A. G. Williams, J. Autschbach and J. Crassous, Conformational changes and chiroptical switching of enantiopure bis-helicenic terpyridine upon Zn2+ binding, Chem. Commun., 2016, 52, 5932–5935 RSC
.
- Y. Tanaka, T. Murayama, A. Muranaka, E. Imai and M. Uchiyama, Ring-Opened Hemiporphyrazines: Helical Molecules Exhibiting Circularly Polarized Luminescence, Chem. – Eur. J., 2020, 26, 1768–1771 CrossRef CAS PubMed
.
- D. Niu, Y. Jiang, L. Ji, G. Ouyang and M. Liu, Self-Assembly through Coordination and p-Stacking: Controlled Switching of Circularly Polarized Luminescence, Angew. Chem., Int. Ed., 2019, 58, 5946–5950 CrossRef CAS PubMed
.
- M. Savchuk, S. Vertueux, T. Cauchy, M. Loumagine, F. Zinna, L. Di Bari, N. Zigon and N. Avarvari, Schiff-base [4]helicene Zn(ii) complexes as chiral emitters, Dalton Trans., 2021, 50, 10533–10539 RSC
.
- P. Reiné, A. M. Ortuño, S. Resa, L. Álvarez de Cienfuegos, V. Blanco, M. J. Ruedas-Rama, G. Mazzeo, S. Abbate, A. Lucotti, M. Tommasini, S. Guisán-Ceinos, M. Ribagorda, A. G. Campaña, A. Mota, G. Longhi, D. Miguel and J. M. Cuerva, OFF/ON switching of circularly polarized luminescence by oxophilic interaction of homochiral sulfoxide-containing o-OPEs with metal cations, Chem. Commun., 2018, 54, 13985–13988 RSC
.
- J. Liu, Y. Zhao, Z. Zhang, M. Li, W. Song, W. Li and Z. Miao, Circularly polarized blue fluorescence based on chiral heteroleptic six-coordinate bis-pyrazolonate-Zn2+ complexes, Dalton Trans., 2024, 53, 6625–6630 RSC
.
- Y. Chen, X. Li, Y. Quan, Y. Cheng and Y. Tang, Strong circularly polarized electroluminescence based on chiral salen-Zn(II) complex monomer chromophores, Mater. Chem. Front., 2019, 3, 867–873 RSC
.
- T. M. Figueira-Duarte and K. Müllen, Pyrene-Based Materials for Organic Electronics, Chem. Rev., 2011, 111, 7260–7314 CrossRef CAS PubMed
.
- O. A. Krasheninina, D. S. Novopashina, E. K. Apartsin and A. G. Venyaminova, Recent Advances in Nucleic Acid Targeting Probes and Supramolecular Constructs Based on Pyrene-Modified Oligonucleotides, Molecules, 2017, 22, 2108 CrossRef PubMed
.
- X. Feng, X. Wang, C. Redshaw and B. Z. Tang, Aggregation behaviour of pyrene-based luminescent materials, from molecular design and optical properties to application, Chem. Soc. Rev., 2023, 52, 6715–6753 RSC
.
- Y. Ohishi and M. Inouye, Circularly polarized luminescence from pyrene excimers, Tetrahedron Lett., 2019, 60, 151232 CrossRef CAS
.
- K. Nakabayashi, T. Amako, N. Tajima, M. Fujiki and Y. Imai, Nonclassical dual control of circularly polarized luminescence modes of binaphthyl–pyrene organic fluorophores in fluidic and glassy media, Chem. Commun., 2014, 50, 13228–13230 RSC
.
- S. Ito, K. Ikeda, S. Nakanishi, Y. Imai and M. Asami, Chem. Commun., 2017, 53, 6323–6326 RSC
.
- K. Takaishi, K. Iwachido and T. Ema, Solvent-Induced Sign Inversion of Circularly Polarized Luminescence: Control of Excimer Chirality by Hydrogen Bonding, J. Am. Chem. Soc., 2020, 142, 1774–1779 CrossRef CAS PubMed
.
- Y. Yang, C. Yang, X. Zhu, L. Zhang and M. Liu, Interfacial self-assembly of a chiral pyrene exciplex into a superhelix with enhanced circularly polarized luminescence, Chem. Commun., 2024, 60, 6631–6634 RSC
.
- S. Ito, S. Wakiyama, H. Chen, M. Abekura, H. Uekusa, R. Ikemura and Y. Imai, Contrasting Mechanochromic Luminescence of Enantiopure and Racemic Pyrenylprolinamides: Elucidating Solid-State Excimer Orientation by Circularly Polarized Luminescence, Angew. Chem., Int. Ed., 2025, 64, e202422913 CrossRef CAS PubMed
.
- M. Ikeshita, S. Furukawa, T. Ishikawa, K. Matsudaira, Y. Imai and T. Tsuno, Enhancement of Chiroptical Responses of trans-Bis[(β-iminomethyl)naphthoxy]platinum(II) Complexes with Distorted Square Planar Coordination Geometry, ChemistryOpen, 2022, 11, e202100277 CrossRef CAS PubMed
.
- M. Ikeshita, N. Hara, Y. Imai and T. Naota, Chiroptical Response Control of Planar and Axially Chiral Polymethylene-Vaulted Platinum(II) Complexes Bearing 1,1′-Binaphthyl Frameworks, Inorg. Chem., 2023, 62, 13964–13976 CrossRef CAS PubMed
.
- M. Ikeshita, S. C. Ma, G. Muller and T. Naota, Linker-dependent control of the chiroptical properties of polymethylene-vaulted trans-bis[(β-iminomethyl)naphthoxy]platinum(II) complexes, Dalton Trans., 2024, 53, 7775–7787 RSC
.
- M. Ikeshita, S. Watanabe, S. Suzuki, S. Tanaka, S. Hattori, K. Shinozaki, Y. Imai and T. Tsuno, Circularly polarized phosphorescence with a large dissymmetry factor from a helical platinum(II) complex, Chem. Commun., 2024, 60, 2413–2416 RSC
.
- M. Enamullah, V. Vasylyeva and C. Janiak, Chirality and diastereoselection of Δ/Λ-configured tetrahedral zinc(II) complexes with enantiopure or racemic Schiff base ligands, Inorg. Chim. Acta., 2013, 408, 109–119 CrossRef CAS
.
- P. Demerseman, J. Einhorn, J. F. Gourvest and R. Royer, Synthèse d'analogues furanniques du benzo[a]pyrène, J. Heterocycl. Chem., 1985, 22, 39–43 CrossRef CAS
.
- H. Sakiyama, H. Okawa, N. Matsumoto and S. Kida, A Tetrahedral Zinc(II) Complex of N-(R)-1-Phenylethylsalicylideneimine. Structural and Circular Dichroism Spectral Investigations on Stereoselectivity, J. Chem. Soc., Dalton Trans., 1990, 2935–2939 RSC
.
- F. M. Winnik, Photophysics of Preassociated Pyrenes in Aqueous Polymer Solutions and in Other Organized Media, Chem. Rev., 1993, 93, 587–614 CrossRef CAS
.
- X. He, S. Zhang, S. Qi, P. Xu, B. Dong and B. Song, Enhanced excimer fluorescence emission of pyrene derivatives: Applications in artificial light-harvesting systems, Dyes Pigm., 2023, 209, 110933 CrossRef CAS
.
- H. Tanaka, Y. Inoue and T. Mori, Circularly Polarized Luminescence and Circular Dichroisms in Small Organic Molecules: Correlation between Excitation and Emission Dissymmetry Factors, ChemPhotoChem, 2018, 2, 386–402 CrossRef CAS
.
Footnotes |
† Electronic supplementary information (ESI) available. CCDC 2426213–2426215. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5qi01086a |
‡ These authors contributed equally to this work. |
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.