Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Chiral tether-guided selective synthesis of Dn-symmetric chiral conjugated nanorings

Tai An ab, Jiayao Yao d, Zuo Xiao *a, Qi Yu ab, Yu Wang a, Yueyue Gao f, Yixiao Song a, Zuoxin Huang *d, Zheng Ding a, Xinyue Zhang ab, Yuanpeng Xie e, Menglan Lv *e, Chuantian Zuo *a, Junqiao Ding g and Liming Ding *c
aKey Laboratory of Nanosystem and Hierarchical Fabrication (CAS), National Center for Nanoscience and Technology, Beijing 100190, China. E-mail: xiaoz@nanoctr.cn; zuocht@nanoctr.cn
bUniversity of Chinese Academy of Sciences, Beijing 100049, China
cSchool of Chemical Engineering and Light Industry, Guangdong University of Technology, Guangzhou 510006, China. E-mail: ding@gdut.edu.cn
dSinopec Research Institute of Petroleum Processing Co., Ltd, Beijing 100083, China. E-mail: huangzx.ripp@sinopec.com
eSchool of Chemistry and Chemical Engineering, Guizhou University, Guiyang 550025, China. E-mail: mllv@gzu.edu.cn
fSchool of Future Technology, Henan University, Zhengzhou 450046, China
gSchool of Chemical Science and Technology, Yunnan University, Kunming 650091, China

Received 22nd August 2025 , Accepted 9th October 2025

First published on 10th October 2025


Abstract

Chiral conjugated nanorings with Dn symmetry exhibit extraordinary circularly polarized luminescence (CPL) properties due to their unique cylindrical helical conjugated system. However, their synthesis faces challenges such as numerous atropisomers and tedious separation and chiral resolution processes, which severely hinder their development. In this work, we report a chiral tether-guided synthesis strategy. By introducing a strained planar chiral alkyl chain tether into the fused-ring building unit of the nanoring, the energy differences between the various atropisomers of the nanoring are significantly increased. This guides the formation of the thermodynamically most stable Dn-symmetric isomer during synthesis, thus greatly enhancing the selectivity. Four chiral nanorings, D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2, and D4-(M)-NR2, as well as their enantiomers, were facilely obtained through this method. All molecules have shown remarkable and stable CPL capability, with a luminescence dissymmetry factor up to 0.076.


Introduction

In recent years, organic luminescent materials have achieved widespread applications in fields such as electroluminescence, biological probes, information encryption, sensing and detection.1–7 Chirality further endows them with circularly polarized luminescence (CPL) properties, significantly expanding their application potential and making them highly promising for 3D displays, optical storage, polarized imaging, chiral sensing, and spintronics.8–15 Two key metrics for evaluating the performance of chiral luminescent materials are the photoluminescence quantum yield (φ) and the luminescence dissymmetry factor (glum). For organic luminescent molecules, the φ can reach 100% through molecular design and excited-state modulation. However, the intrinsic glum values of most organic chiral molecules typically fall within the range of 10−5 to 10−3,16 far below the theoretical extremes (±2). Developing organic chiral molecules with intrinsically large glum values holds significant scientific importance.

Theoretically, the glum of chiral luminescent molecules can be described by the following equation:

image file: d5sc06445g-t1.tif
where μij and mij represent the electric transition dipole moment and magnetic transition dipole moment, respectively, during the transition from excited state ‘i’ to ground state ‘j’, and θ is the angle between the two transition dipole moments.17,18 Since the luminescence of most organic molecules follows Kasha's rule, the transition of interest is typically from the S1 to S0 state.19 The equation indicates that to maximize the g value while maintaining high luminescence intensity, both μ and m must be large and comparable in magnitude, while their dipole moments must be aligned parallel or antiparallel (i.e., θ = 0 or 180). However, for most organic chiral molecules, m is much smaller than μ, making it difficult to enhance glum.20 Therefore, the key to improving glum lies in molecular design strategies that enhance m while ensuring μ and m remain parallel or antiparallel. In recent years, significant efforts have been devoted to exploring chiral organic luminescent molecules with high glum values.21–47 Among all reported chiral organic emitters, conjugated nanorings with Dn symmetry exhibit exceptionally high intrinsic glum. For example, Isobe's D4-symmetric (P)-(12,8)-[4]CC45 and Du's D4-symmetric (+)-[4]CAn2,6 ref. 46 achieved |glum| of 0.152 and 0.103, respectively (Fig. 1a), representing the only two purely organic chiral emitters with |glum| on the order of 10−1 at the molecular level. Notably, (P)-(12,8)-[4]CC also exhibited a high fluorescence quantum yield of 80%. The extraordinary |glum| in these molecules arises from their unique cylindrical helical conjugated systems. On one hand, this structure arranges the local electric transition dipole moments of the building units in a circular fashion, generating a large induced magnetic transition dipole moment along the helical axis.48 On the other hand, their high Dn symmetry ensures that μ and m remain parallel or antiparallel, maximizing |cos[thin space (1/6-em)]θ| and thus achieving a large |glum|.44,49 Long et al. further demonstrated through theoretical calculations that increasing the number of building units to synthesize higher-order Dn-symmetric nanorings could significantly enhance the magnetic transition dipole moment along the helical axis, enabling giant g values (Fig. 1b).50 For instance, starting from Isobe's [4]CC framework, expanding to a D8-symmetric [8]CC (8 units) increases g to 0.71, while a D20-symmetric [20]CC (20 units) could further elevate g to 1.47, all while maintaining high transition oscillator strengths.


image file: d5sc06445g-f1.tif
Fig. 1 (a) Previously reported Dn-symmetric chiral conjugated nanorings. (b) Theoretically predicted high-order Dn-nanorings with giant g-values. (c) Chiral tether-guided selective synthesis of Dn-nanorings in this work.

The above experimental and theoretical studies demonstrate that Dn-symmetric chiral conjugated nanorings represent a highly promising class of molecules, offering potential breakthroughs in organic luminescent materials with intrinsically giant glum values. However, their synthesis faces a critical challenge: the formation of atropisomers (conformational isomers arising from flipping of building units) during preparation. The number of atropisomers increases rapidly with higher symmetry order (n). Tanaka et al. reported a D2-symmetric nanoring (Sp,Sp)-(M,M,M,M)-1a, which theoretically has 3 atropisomers (Fig. 1a). While an asymmetric catalytic cyclization enabled the selective synthesis of the D2-symmetric isomer, its |glum| was only 0.005.51 Isobe et al. observed all 4 possible atropisomers in the preparation of D3-symmetric nanoring (P)-(9,6)-[3]CdbC. The optically pure (P)-(9,6)-[3]CdbC was obtained via chiral HPLC purification.48 For D4-symmetric nanorings, the number of atropisomers rises to 6. Aforementioned (P)-(12,8)-[4]CC and (+)-[4]CAn2,6 both required laborious chiral HPLC purification.45,46 For higher-order systems like D8-symmetric [8]CC and D20-symmetric [20]CC proposed by Long et al., the atropisomer count escalates to 30 and 27012, respectively, making their synthesis and isolation extremely challenging.50 Thus, developing a method to selectively access Dn-symmetric nanorings is crucial to advance their applications in chiral organic luminescent materials. In this work, we report a novel chiral tether-guided synthesis strategy for Dn-symmetric chiral conjugated nanorings (Fig. 1c). By incorporating a strained planar-chiral alkyl tether into the fused-ring building block of the nanoring, we significantly amplify the energy differences between various atropisomers. This thermodynamic control drives the formation of the most stable Dn-symmetric isomer with exceptional selectivity during synthesis. Using this innovative approach, we successfully synthesized four chiral nanorings: D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2, and D4-(M)-NR2, along with their respective enantiomers. The helical chirality of these molecules is uniquely determined by the planar chirality of their tether moieties. These compounds exhibit remarkable CPL properties, with all derivatives demonstrating |glum| exceeding 10−2.

Results and discussion

The synthetic routes for the nanorings are illustrated in Scheme 1. For D3-(P)-NR1 and D4-(P)-NR1 (Scheme 1a), starting from the previously reported planar chiral cyclophane (Sp)-1,52 the synthesis proceeded via acylation to afford the dibenzyloxy-terminated amide (Sp)-2. Subsequent intramolecular Yamamoto cyclization yielded the pentacyclic fused lactam (Sp)-3.53–55 Deprotection of the benzyl groups followed by acylation furnished the triflate-terminated intermediate (Sp)-5, which was further converted into the boronic ester (Sp)-6 via Pd-catalyzed borylation. Equimolar reaction of (Sp)-6 with Pt(COD)Cl2 generated the Pt-macrocyclic complex intermediates, and final reductive elimination promoted by PPh3 afforded the chiral trimeric ring D3-(P)-NR1 and the chiral tetrameric ring D4-(P)-NR1 in yields of 40% and 3% (starting from (Sp)-6), respectively. The preferential formation of the trimer suggests that the highly curved backbone of the fused-ring unit (Sp)-6 favors smaller macrocycles in this Pt-mediated nanoring synthesis.48 For D3-(M)-NR2 and D4-(M)-NR2 (Scheme 1b), (Sp)-1 first underwent Suzuki coupling with 3-chloro-2-fluorobenzeneboronic acid to afford (Sp)-7, which was cyclized via a potassium tert-butoxide-mediated intramolecular nucleophilic substitution to form the pentacyclic carbazole (Sp)-8.52,56 Pd-catalyzed borylation then replaced the terminal chlorine with a boronic ester, yielding (Sp)-9. Subsequent Pt-mediated macrocyclization produced the chiral trimer D3-(M)-NR2 and the chiral tetramer D4-(M)-NR2 in 10% and 27% yields (starting from (Sp)-9), respectively. Unlike the fused-lactam-based system, the fused-carbazole-derived nanoring predominantly formed the tetrameric product. Using the enantiomeric planar chiral cyclophane (Rp)-1 as the starting material, we also synthesized the enantiomers of the above four nanorings—D3-(M)-NR1, D4-(M)-NR1, D3-(P)-NR2 and D4-(P)-NR2 (Scheme 1c). Detailed synthetic procedures are provided in the SI (Schemes S3 and S4). It is worth mentioning that all the nanorings were obtained through conventional column chromatography separation, without the need for chiral HPLC purification. All nanoring structures were characterized by NMR and mass spectroscopic methods. Both 1H and 13C NMR spectra confirmed that each nanoring was obtained as a single isomer (see SI). Additionally, single-crystal X-ray diffraction (SCXRD) analysis unambiguously determined the structures of D3-(P)-NR1, D3-(M)-NR2, and the enantiomer D3-(M)-NR1 (vide infra).
image file: d5sc06445g-s1.tif
Scheme 1 Synthetic routes for (a) D3-(P)-NR1 and D4-(P)-NR1, (b) D3-(M)-NR2 and D4-(M)-NR2, and (c) the enantiomers. Reaction conditions: (i) Et3N, 70 °C; (ii) Ni(COD)2/2,2′-bipyridine, 70 °C; (iii) Pd/C, NH4OOCH, 80 °C; (iv) Tf2O/pyridine; (v) (Bpin)2, Pd(OAc)2/S-Phos, 80 °C; (vi) (COD)PtCl2, CsF; (vii) PPh3, 110 °C; (viii) Pd(PPh3)4, K2CO3, 70 °C; (ix) tBuOK, 80 °C; (x) PPh3, 150 °C.

The single-crystal structures of D3-(P)-NR1 and D3-(M)-NR2 are shown in Fig. 2a, and that for D3-(M)-NR1 is presented in Fig. S64. Crystallographic data reveal that all molecules adopt a single configuration without other atropisomers, and their ring skeletons exhibit D3 symmetry. Although both D3-(P)-NR1 and D3-(M)-NR2 were synthesized from the same starting material, (Sp)-1, their macrocyclic structures differ significantly. In D3-(P)-NR1, all the tethers are positioned on the outer periphery of the nanoring, and the three fused-ring lactam building blocks (Sp)-M1 adopt a clockwise P-helical arrangement. In contrast, in D3-(M)-NR2, all the tethers are oriented inside the nanoring, and the three fused-ring carbazole units (Sp)-M2 form a counterclockwise M-helical arrangement. D3-(M)-NR1, synthesized from (Rp)-1, shows the mirror-image structure of D3-(P)-NR1, exhibiting M-helicity (Fig. S64). From the top view (Fig. 2b), the diameter of the lactam nanoring D3-(P)-NR1 is larger (∼12.7 Å), whereas that of the carbazole nanoring D3-(M)-NR2 is slightly smaller (∼12.1 Å). Regarding packing structures, D3-(P)-NR1 molecules form a 2D layered arrangement, where adjacent nanorings within each layer are nearly perpendicular, resembling a herringbone pattern—similar to some known cycloparaphenylene (CPP) nanorings (e.g., [6]CPP, [7]CPP).57,58 In contrast, D3-(M)-NR2 crystals exhibit a triangular packing motif: three nanoring molecules assemble into a circular trimer, with disordered solvent molecules filling the triangular pores. This arrangement resembles the triangular stacking observed in Itami's previously reported carbon nanobelts (CNB) (e.g., [24]CNB).55 To understand why D3-(P)-NR1 has all tethers outward while D3-(M)-NR2 has them inward, we examined the conformations of their building blocks, (Sp)-M1 and (Sp)-M2 (Fig. 2c). DFT-optimized structures show that the fused-ring backbone of (Sp)-M1 bends outward due to tether-induced strain, favoring an exterior tether placement, whereas (Sp)-M2 bends inward under the strain, directing its tethers inside the nanoring.


image file: d5sc06445g-f2.tif
Fig. 2 (a) Single crystal structures for D3-(P)-NR1 and D3-(M)-NR2; (b) top view and packing structures for D3-(P)-NR1 and D3-(M)-NR2; (c) the fused-ring building blocks (Sp)-M1 and (Sp)-M2, and their DFT-optimized configurations. Note: the solvent systems used for single-crystal growth are hexane/dichloromethane and methanol/chloroform for D3-(P)-NR1 and D3-(M)-NR2, respectively; the solvent molecules in the crystals of D3-(P)-NR1 and D3-(M)-NR2 are dichloromethane and chloroform, respectively.

To elucidate the high selectivity in obtaining Dn-symmetric nanorings without observable formation of other atropisomeric products, we calculated the energies of all possible atropisomers. For comparison, we also calculated the energies of hypothetical tether-free nanorings (where one alkyl tether is replaced by two methyl groups). Fig. 3 shows the relative Gibbs free energies of all four possible atropisomers for the D3-symmetric nanorings (D3-(P)-NR1 and D3-(M)-NR2) and their tether-free analogues. For D3-(P)-NR1 and D3-(M)-NR2, the lowest-energy isomers are Aooo and Biii (‘o’ (outside) and ‘i’ (inside) describe the tether's position with respect to the nanoring), respectively. These structures match the single-crystal configurations. Flipping one fused-ring unit to form Aooi and Biio raises the energy by 8.4 kcal mol−1 and 24.6 kcal mol−1, respectively. Obviously, the D3-symmetric Aooo and Biii are much favored energetically over other isomers. On the other hand, the tether-free analogues exhibit low energy differences (a few kcal mol−1) among atropisomers (‘a’ and ‘b’ denote the syn-facial and anti-facial arrangement of the fused-ring building units) (Fig. 3c and d), consistent with prior reports.59 Similar energies imply the lack of a dominant product, as all are likely to form during the synthesis. The above results indicate that by introducing strained chiral tethers into the nanoring significantly increases the energy differences among atropisomers, thereby enabling one Dn-symmetric isomer to become the dominant product and greatly enhancing selectivity. We also conducted energy analyses on all atropisomers of the D4-symmetric nanorings D4-(P)-NR1, D4-(M)-NR2, and their tether-free analogues, arriving at the same conclusion (Fig. S65 and S66). Notably, the tether does not significantly affect the rotational barriers between the atropisomers. Similar to the tether-free nanorings, D3-(P)-NR1, D3-(M)-NR2, D4-(P)-NR1, and D4-(M)-NR2 all exhibit high atropisomerization barriers (21.4–67.5 kcal mol−1), indicating that they can maintain stable helical configurations with high chiral stability (Fig. S65 and S66).


image file: d5sc06445g-f3.tif
Fig. 3 The relative Gibbs free energies for different atropisomers of (a) D3-(P)-NR1 (A series), (b) D3-(M)-NR2 (B series) and the tether-free analogues of (c) D3-(P)-NR1 (A′ series) and (d) D3-(M)-NR2 (B′ series).

The striking structural difference between the fused-lactam nanorings (D3-(P)-NR1 and D4-(P)-NR1) and fused-carbazole nanorings (D3-(M)-NR2 and D4-(M)-NR2), where the tethers are positioned exteriorly and interiorly respectively, results in significantly distinct chemical shifts of the tether protons in NMR spectra. Particularly, the most pronounced shift is observed for proton H1, which is closest to the central benzene ring of the building block (Fig. 4a). For the fused-lactam building unit (Sp)-M1, its H1 appears at −0.29 ppm due to its location above the central benzene ring, where it experiences shielding effects and resonates in a relatively upfield region. However, upon incorporation into nanorings D3-(P)-NR1 and D4-(P)-NR1, the chemical shift of H1 moves to 0.68 ppm and 0.57 ppm respectively, indicating markedly reduced shielding in the exterior regions of the nanorings. Conversely, for the fused-carbazole building unit (Sp)-M2, H1 initially appears at −1.86 ppm. After being integrated into nanorings D3-(M)-NR2 and D4-(M)-NR2, its chemical shift moves to −3.90 ppm and −3.11 ppm, respectively, demonstrating enhanced shielding within the interior cavity of the nanorings. These experimental observations align well with the trends predicted by DFT calculations. Previous theoretical studies have proposed that the exterior and interior of certain cyclic nanocarbons correspond to less-shielded and shielded regions, respectively.60 In this study, H1 worked as an NMR probe, providing direct experimental evidence. To further investigate the magnetic shielding environments, we computed the nucleus-independent chemical shift (NICS)61–63 of the central benzene rings (Fig. 4b). The results reveal that for all nanorings, the NICS(−1) values (interior) are consistently more negative than the NICS(1) values (exterior), confirming stronger shielding inside the rings, in agreement with our NMR observations. Moreover, compared to the building blocks, all nanorings exhibit less negative NICS(1) values, while the NICS(−1) values tend to become more negative (except for D4-(M)-NR2, which shows minimal change). This suggests that upon nanoring formation, the increased curvature of the building units leads to reduced shielding on the convex side (exterior) and enhanced shielding on the concave side (interior), consistent with prior literature reports.64,65 The average values (NICSavg) of NICS(1) and NICS(−1) show a clear trend toward less negative values from monomers to nanorings, indicating reduced aromaticity of the central benzene ring due to heightened curvature in the nanoring structures.


image file: d5sc06445g-f4.tif
Fig. 4 (a) The change of the NMR chemical shifts (the values in parentheses are DFT-calculated chemical shifts) of proton H1 from building blocks to nanorings; (b) the calculated NICS(1), NICS(−1) and the averages of NICS(1) and NICS(−1) for the building blocks and nanorings.

Next, we investigated the optical properties of the nanorings. Fig. 5a displays the UV-Vis absorption spectra of D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2. Their respective enantiomers, D3-(M)-NR1, D4-(M)-NR1, D3-(P)-NR2 and D4-(P)-NR2, show identical spectra (Fig. S67). It can be observed that the spectra of the fused-lactam nanorings D3-(P)-NR1 and D4-(P)-NR1 primarily exhibit two absorption bands (one at 320–400 nm and another at 400–500 nm), whereas the fused-carbazole nanorings D3-(M)-NR2 and D4-(M)-NR2 show only one main absorption band at 320–420 nm. This difference is mainly attributed to the fact that some lower-energy electronic transitions (e.g., S0 → S2, S0 → S3) in the fused-lactam nanorings exhibit higher oscillator strengths (Table S1). The oscillator strengths of the S0 → S1 transitions are very low for all nanorings, likely due to the high molecular symmetry, which makes the S0 → S1 transition forbidden. Consequently, the absorption coefficients near the bandgap are relatively low for all nanorings. From the onset absorption, the optical bandgaps of D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2 are calculated to be 2.40, 2.51, 2.54, and 2.62 eV, respectively. It is evident that, for both fused-lactam and fused-carbazole nanorings, the smaller-sized nanorings possess narrower bandgaps than their larger-sized counterparts, consistent with previously reported trends for CPP and CNB nanorings.22,55,66 DFT calculations further confirm this observation, showing that smaller-sized nanorings have smaller HOMO–LUMO gaps (Fig. S69). Additionally, theoretical calculations indicate that the fused-carbazole nanorings have higher HOMO and LUMO energy levels compared to the fused-lactam nanorings, which aligns with the energy levels measured in electrochemical experiments (Fig. S70). The photoluminescence spectra of all nanorings and their enantiomers are shown in Fig. 5b and S71, respectively. All samples were measured in toluene solution. The excitation wavelengths were set at 360 nm for D3-(P/M)-NR1, 370 nm for D4-(P/M)-NR1, and 380 nm for both D3-(M/P)-NR2 and D4-(M/P)-NR2. The fluorescence peaks of D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2 are located at 561, 520, 519, and 492 nm, respectively. The measured fluorescence quantum yields are 42%, 80%, 25%, and 43%, respectively. Notably, compared to the larger D4 nanorings, the smaller D3 nanorings exhibit red-shifted emission peaks and reduced quantum yields, similar to the size-dependent luminescence properties observed in CPP and methylene-bridged cycloparaphenylene (MCPP) nanorings.67–69


image file: d5sc06445g-f5.tif
Fig. 5 The UV-Vis absorption spectra (a) and photoluminescence spectra (b) for D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2. The CD (left) and CPL (right) spectra for D3-(P)-NR1 and D3-(M)-NR1 (c and d), D4-(P)-NR1 and D4-(M)-NR1 (e and f), D3-(P)-NR2 and D3-(M)-NR2 (g and h), D4-(P)-NR2 and D4-(M)-NR2 (i and j) in toluene. Note: the dotted lines in the CD spectra are DFT-predicted ones.

Since the nanorings prepared via the chiral tether-guided approach possess uniquely defined configurations, chiral resolution is unnecessary, allowing direct testing of their chiroptical properties. The circular dichroism (CD) and circularly polarized luminescence (CPL) spectra of all nanorings and their enantiomers are shown in Fig. 5c–j. All measurements were conducted in dilute solutions (∼1 × 10−5 M) to ensure the signals originated from single molecules rather than aggregates. Each pair of enantiomers exhibited mirror-image CD and CPL spectra. For the CD spectra, all P-helical nanorings displayed a negative Cotton effect near the bandgap, while all M-helical nanorings showed a positive Cotton effect. The theoretically calculated CD spectra matched well with the experimental results. We further compared the CD spectra of the fused-lactam nanorings D3-(P)-NR1 and D4-(P)-NR1 before and after heating at 100 °C for 24 hours (Fig. S72a), revealing nearly overlapping spectra. Since the fused-carbazole nanorings D3-(M)-NR2 and D4-(M)-NR2 decompose upon heating, we assessed their chiral stability at room temperature. After one week of storage, no significant changes were observed in their CD spectra (Fig. S72b). These experiments confirm the earlier theoretical predictions that these molecules generally possess high atropisomerization barriers (Fig. S65), ensuring excellent chiral stability. All nanorings exhibited strong CPL signals. Consistent with the CD signals near the bandgap, all P-helical nanorings displayed a negative Cotton effect in CPL, while all M-helical nanorings showed a positive Cotton effect. Additionally, all nanorings demonstrated large luminescence dissymmetry factors (glum), exceeding 10−2 in magnitude. Notably, D3-(P)-NR2 exhibited a remarkably high glum of −0.076, making it one of the rare purely organic molecules with a large intrinsic glum. To understand why these nanorings generally exhibit large dissymmetry factors, we analyzed their S1 → S0 transitions by DFT. Fig. 6 illustrates their transition characteristics, revealing that all molecules possess a substantial magnetic transition dipole moment (m) along the helical axis. The |m| values for D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2 are 8.32, 12.55, 9.14, and 14.18 Bohr magnetons (μB), respectively—among the largest reported for organic chiral molecules.70 Moreover, the electric (μ) and magnetic (m) transition dipole moments are consistently aligned either parallel or antiparallel. In P-helical nanorings, the angles between μ and m are 179.5° and 180°, whereas in M-helical nanorings, they are 1.3° and 0°. These factors collectively contribute to the high glum. Furthermore, the carbazole-type nanorings exhibited larger m and smaller μ compared to the lactam-type nanorings, explaining their higher glum values. In both carbazole- and lactam-type nanorings, m increased with ring size, consistent with the reported “molecular solenoid inner area rule”.70 From D3 to D4 nanorings, the increase in m was comparable to that in μ, resulting in minor changes (or even slight decreases) in glum. However, based on theoretical calculations by Long et al., synthesizing higher-order Dn-symmetric nanorings could further enhance m relative to μ, potentially leading to breakthroughs in g-values.50


image file: d5sc06445g-f6.tif
Fig. 6 The optimized geometry of S1 state and the S1 → S0 transition characteristics for D3-(P)-NR1, D4-(P)-NR1, D3-(M)-NR2 and D4-(M)-NR2. Note: “a.u.” denotes the atomic unit; the vector arrow for μ is scaled up by a factor of 60.

Conclusion

In summary, we have developed a chiral tether-guided strategy to selectively prepare Dn-symmetric chiral conjugated nanorings. The high selectivity stems from the chiral tether induced strain, which amplifies the energy differences among the various atropisomers of the nanorings, thereby directing one Dn-symmetric isomer to become the thermodynamically favored product. All synthesized D3 and D4 nanorings exhibit exceptional circularly polarized luminescence properties, with a maximum luminescence dissymmetry factor reaching 0.076 and photoluminescence quantum yields up to 80%, demonstrating their great potential as high-performance CPL emitters. In the future, we will extend this strategy to synthesize higher-order Dn-symmetric nanorings, which may lead to breakthroughs in the development of organic molecules with intrinsically giant glum.

Author contributions

Z. X. conceptualized the work. T. A., Q. Y., Y. S., Z. D., X. Z. synthesized the compounds. T. A. and Y. W. performed the theoretical calculations. J. Y., Z. X., Y. G., Z. H., C. Z. and L. D. directed the experiments and discussed the results. Z. X., Z. H., M. L., C. Z. and L. D. directed the project. Z. X. wrote the manuscript.

Conflicts of interest

The authors declare no conflict of interest.

Data availability

CCDC 2440960, 2440962 and 2440966 contain the supplementary crystallographic data for this paper.71a–c

All other data supporting the findings of this study are available within the article and its supplementary information (SI), as well as available from the corresponding authors upon reasonable request. Supporting information available: materials synthesis and characterization; theoretical calculation; Scheme S1–5, Fig. S1–72, Table S1. See DOI: https://doi.org/10.1039/d5sc06445g.

Acknowledgements

We thank the National Key Research and Development Program of China (2022YFB3803300, 2023YFE0116800), National Natural Science Foundation of China (22571057), and Beijing Natural Science Foundation (IS23037) for financial support. This work is supported by High Performance Computing Center of NCNST, China.

References

  1. T. Zhang, Y. Xiao, H. Wang, S. Kong, R. Huang, V. Ka-Man Au, T. Yu and W. Huang, Highly Twisted Thermally Activated Delayed Fluorescence (Tadf) Molecules and Their Applications in Organic Light-Emitting Diodes (Oleds), Angew. Chem., Int. Ed., 2023, 62, e202301896 CrossRef CAS PubMed.
  2. A. Concellón, J. Castro-Esteban and T. M. Swager, Ultratrace Pfas Detection Using Amplifying Fluorescent Polymers, J. Am. Chem. Soc., 2023, 145, 11420–11430 CrossRef PubMed.
  3. Y. Huang, W. Chen, J. Chung, J. Yin and J. Yoon, Recent Progress in Fluorescent Probes for Bacteria, Chem. Soc. Rev., 2021, 50, 7725–7744 RSC.
  4. W. Xu, Q. He and J. Cheng, Organic Semiconductor Fluorescent Sensing Materials and Hazardous Chemicals Detection Applications, Sci. Sin Chem., 2020, 50, 92–107 CrossRef.
  5. G. Yin, J. Zhou, W. Lu, L. Li, D. Liu, M. Qi, B. Z. Tang, P. Théato and T. Chen, Targeting Compact and Ordered Emitters by Supramolecular Dynamic Interactions for High-Performance Organic Ambient Phosphorescence, Adv. Mater., 2024, 36, 2311347 CrossRef CAS PubMed.
  6. H. Zhu, J. Fan, J. Du and X. Peng, Fluorescent Probes for Sensing and Imaging within Specific Cellular Organelles, Acc. Chem. Res., 2016, 49, 2115–2126 CAS.
  7. Y. Zhang, J. Guan, L. Luo, X. Han, J. Wang, Y. Zheng and J. Xu, Chiral Twisted Molecular Carbons: Synthesis, Properties, and Applications, Interdiscip. Mater., 2024, 3, 453–479 CAS.
  8. M. Zhang, Q. Guo, Z. Li, Y. Zhou, S. Zhao, Z. Tong, Y. Wang, G. Li, S. Jin and M. Zhu, et al., Processable Circularly Polarized Luminescence Material Enables Flexible Stereoscopic 3D Imaging, Sci. Adv., 2023, 9, eadi9944 CAS.
  9. B. L. Feringa, R. A. van Delden, N. Koumura and E. M. Geertsema, Chiroptical Molecular Switches, Chem. Rev., 2000, 100, 1789–1816 CrossRef CAS PubMed.
  10. P. Stachelek, S. Serrano-Buitrago, B. L. Maroto, R. Pal and S. de la Moya, Circularly Polarized Luminescence Bioimaging Using Chiral Bodipys: A Model Scaffold for Advancing Unprecedented CPL Microscopy Using Small Full-Organic Probes, ACS Appl. Mater. Interfaces, 2024, 16, 67246–67254 CAS.
  11. I. Song, J. Ahn, H. Ahn, S. H. Lee, J. Mei, N. A. Kotov and J. H. Oh, Helical Polymers for Dissymmetric Circularly Polarized Light Imaging, Nature, 2023, 617, 92–99 CAS.
  12. X. Liang, W. Liang, P. Jin, H. Wang, W. Wu and C. Yang, Advances in Chirality Sensing with Macrocyclic Molecules, Chemosensors, 2021, 9, 279 CAS.
  13. T. Kawasaki, M. Sato, S. Ishiguro, T. Saito, Y. Morishita, I. Sato, H. Nishino, Y. Inoue and K. Soai, Enantioselective Synthesis of near Enantiopure Compound by Asymmetric Autocatalysis Triggered by Asymmetric Photolysis with Circularly Polarized Light, J. Am. Chem. Soc., 2005, 127, 3274–3275 CAS.
  14. Z. Shang, T. Liu, Q. Yang, S. Cui, K. Xu, Y. Zhang, J. Deng, T. Zhai and X. Wang, Chiral-Molecule-Based Spintronic Devices, Small, 2022, 18, 2203015 CAS.
  15. W. Chen, B. Li, G. Gao and T. Sun, Chiral Supramolecular Nanomaterials: From Chirality Transfer and Amplification to Regulation and Applications, Interdiscip. Mater., 2023, 2, 689–713 Search PubMed.
  16. L. Liu, Y. Yang and Z. Wei, Chiral Organic Optoelectronic Materials and Circularly Polarized Light Luminescence and Detection, Acta Chim. Sinica, 2022, 80, 970–992 CAS.
  17. F. S. Richardson and J. P. Riehl, Circularly Polarized Luminescence Spectroscopy, Chem. Rev., 1977, 77, 773–792 CrossRef CAS.
  18. E. M. Sánchez-Carnerero, A. R. Agarrabeitia, F. Moreno, B. L. Maroto, G. Muller, M. J. Ortiz and S. de la Moya, Circularly Polarized Luminescence from Simple Organic Molecules, Chem.–Eur. J., 2015, 21, 13488–13500 CrossRef PubMed.
  19. M. Kasha, Characterization of Electronic Transitions in Complex Molecules, Discuss. Faraday Soc., 1950, 9, 14–19 RSC.
  20. T. Zhao, J. Han, P. Duan and M. Liu, New Perspectives to Trigger and Modulate Circularly Polarized Luminescence of Complex and Aggregated Systems: Energy Transfer, Photon Upconversion, Charge Transfer, and Organic Radical, Acc. Chem. Res., 2020, 53, 1279–1292 CrossRef CAS PubMed.
  21. Z.-L. Gong, X. Zhu, Z. Zhou, S.-W. Zhang, D. Yang, B. Zhao, Y.-P. Zhang, J. Deng, Y. Cheng and Y.-X. Zheng, et al., Frontiers in Circularly Polarized Luminescence: Molecular Design, Self-Assembly, Nanomaterials, and Applications, Sci. China Chem., 2021, 64, 2060–2104 CrossRef CAS.
  22. S. Guo, L. Liu, X. Li, G. Liu, Y. Fan, J. He, Z. Lian, H. Yang, X. Chen and H. Jiang, Highly Luminescent Chiral Carbon Nanohoops Via Symmetry Breaking with a Triptycene Unit: Bright Circularly Polarized Luminescence and Size-Dependent Properties, Small, 2024, 20, 2308429 CrossRef CAS PubMed.
  23. Y. Liu, Z. Li, M.-W. Wang, J. Chan, G. Liu, Z. Wang and W. Jiang, Highly Luminescent Chiral Double π-Helical Nanoribbons, J. Am. Chem. Soc., 2024, 146, 5295–5304 CrossRef CAS PubMed.
  24. Z. Sun, W. Xu, S. Qiu, Z. Ma, C. Li, S. Zhang and H. Wang, Thia[n]helicenes with Long Persistent Phosphorescence, Chem. Sci., 2024, 15, 1077–1087 RSC.
  25. L. Bian, M. Tang, J. Liu, Y. Liang, L. Wu and Z. Liu, Luminescent Chiral Triangular Prisms Capable of Forming Double Helices for Detecting Traces of Acids and Anion Recognition, J. Mater. Chem. C, 2022, 10, 15394–15399 RSC.
  26. X. Kong, X. Zhang, B. Yuan, W. Zhang, D. Lu and P. Du, Synthesis and Photophysical Properties of a Chiral Carbon Nanoring Containing Rubicene, J. Org. Chem., 2024, 89, 8255–8261 CrossRef CAS PubMed.
  27. G. Li, L.-L. Mao, J.-N. Gao, X. Shi, Z.-Y. Huo, J. Yang, W. Zhou, H. Li, H.-B. Yang and C.-H. Tung, et al., A Helical Tubular Dyad of [9]Cycloparaphenylene: Synthesis, Chiroptical Properties and Post-Functionalization, Angew. Chem., Int. Ed., 2025, 64, e202419435 CrossRef CAS PubMed.
  28. Y.-F. Wu, S.-W. Ying, L.-Y. Su, J.-J. Du, L. Zhang, B.-W. Chen, H.-R. Tian, H. Xu, M.-L. Zhang and X. Yan, et al., Nitrogen-Embedded Quintuple [7]Helicene: A Helicene–Azacorannulene Hybrid with Strong near-Infrared Fluorescence, J. Am. Chem. Soc., 2022, 144, 10736–10742 CrossRef CAS PubMed.
  29. F. Zhao, J. Zhao, H. Liu, Y. Wang, J. Duan, C. Li, J. Di, N. Zhang, X. Zheng and P. Chen, Synthesis of π-Conjugated Chiral Organoborane Macrocycles with Blue to near-Infrared Emissions and the Diradical Character of Cations, J. Am. Chem. Soc., 2023, 145, 10092–10103 CrossRef CAS PubMed.
  30. X.-Y. Wang, J. Bai, Y.-J. Shen, Z.-A. Li and H.-Y. Gong, A Carbazole-Centered Expanded Helicene Stabilized with Hexabenzocoronene (HBC) Units, Angew. Chem., Int. Ed., 2025, 64(1), e202417745 CAS.
  31. Y. Chen, C. Lin, Z. Luo, Z. Yin, H. Shi, Y. Zhu and J. Wang, Double π-Extended Undecabenzo[7]helicene, Angew. Chem., Int. Ed., 2021, 60, 7796–7801 CrossRef CAS PubMed.
  32. J.-K. Li, X.-Y. Chen, W.-L. Zhao, Y.-L. Guo, Y. Zhang, X.-C. Wang, A. C.-H. Sue, X.-Y. Cao, M. Li and C.-F. Chen, et al., Synthesis of Highly Luminescent Chiral Nanographene, Angew. Chem., Int. Ed., 2023, 62, e202215367 CrossRef CAS PubMed.
  33. Y. Zhu, G. Chen, Y. Deng, H. Yang, C. Wang, J. Gao, C. Zhang and J. Xiao, Synthesis and Characterization of Asymmetric Azatwistarenes with Chiroptical Property, Org. Lett., 2024, 26, 9486–9491 CrossRef CAS PubMed.
  34. J. Hu, Q. Xiang, X. Tian, L. Ye, Y. Wang, Y. Ni, X. Chen, Y. Liu, G. Chen and Z. Sun, S-Shaped Helical Singlet Diradicaloid and Its Transformation to Circumchrysene via a Two-Stage Cyclization, J. Am. Chem. Soc., 2024, 146, 10321–10330 CrossRef CAS PubMed.
  35. G.-F. Huo, W.-T. Xu, J. Hu, Y. Han, W. Fan, W. Wang, Z. Sun, H.-B. Yang and J. Wu, Perylene-Embedded Helical Nanographenes with Emission up to 1010 nm: Synthesis, Structures, and Chiroptical Properties, Angew. Chem., Int. Ed., 2025, 64, e202416707 CrossRef CAS PubMed.
  36. D. Yang, K. M. Cheung, Q. Gong, L. Zhang, L. Qiao, X. Chen, Z. Huang and Q. Miao, Synthesis, Structures and Properties of Trioxa[9]circulene and Diepoxycyclononatrinaphthalene, Angew. Chem., Int. Ed., 2024, 63, e202402756 CrossRef CAS PubMed.
  37. W. Niu, Y. Fu, Z.-L. Qiu, C. J. Schürmann, S. Obermann, F. Liu, A. A. Popov, H. Komber, J. Ma and X. Feng, π-Extended Helical Multilayer Nanographenes with Layer-Dependent Chiroptical Properties, J. Am. Chem. Soc., 2023, 145, 26824–26832 CrossRef CAS PubMed.
  38. G.-F. Huo, T. M. Fukunaga, X. Hou, Y. Han, W. Fan, S. Wu, H. Isobe and J. Wu, Facile Synthesis and Chiral Resolution of Expanded Helicenes with up to 35 cata-Fused Benzene Rings, Angew. Chem., Int. Ed., 2023, 62, e202218090 CrossRef CAS PubMed.
  39. M. Krzeszewski, H. Ito and K. Itami, Infinitene: A Helically Twisted Figure-Eight [12]Circulene Topoisomer, J. Am. Chem. Soc., 2022, 144, 862–871 CrossRef CAS PubMed.
  40. Y. Matsuo, M. Gon, K. Tanaka, S. Seki and T. Tanaka, Synthesis of Aza[n]helicenes up to n = 19: Hydrogen-Bond-Assisted Solubility and Benzannulation Strategy, J. Am. Chem. Soc., 2024, 146, 17428–17437 CrossRef CAS PubMed.
  41. X. Xiao, Q. Cheng, S. T. Bao, Z. Jin, S. Sun, H. Jiang, M. L. Steigerwald and C. Nuckolls, Single-Handed Helicene Nanoribbons Via Transfer of Chiral Information, J. Am. Chem. Soc., 2022, 144, 20214–20220 CrossRef CAS PubMed.
  42. R. Ammenhäuser, J. M. Lupton and U. Scherf, Chain-Length Dependence of the Optical Activity of Helical Triptycene-Based π-Conjugated Ladder Polymers, Adv. Opt. Mater., 2024, 12, 2301857 CrossRef.
  43. S. Jena, A. Thayyil Muhammed Munthasir, S. Pradhan, M. Kitahara, S. Seika, Y. Imai and P. Thilagar, Single Molecular Persistent Room-Temperature Phosphorescence and Circularly Polarized Luminescence from Binaphthol-Decorated Optically Innocent Cyclotriphosphazenes, Chem.–Eur. J., 2023, 29, e202301924 CrossRef CAS PubMed.
  44. H. Kubo, D. Shimizu, T. Hirose and K. Matsuda, Circularly Polarized Luminescence Designed from Molecular Orbitals: A Figure-Eight-Shaped [5]Helicene Dimer with D2 Symmetry, Org. Lett., 2020, 22, 9276–9281 CrossRef CAS PubMed.
  45. S. Sato, A. Yoshii, S. Takahashi, S. Furumi, M. Takeuchi and H. Isobe, Chiral Intertwined Spirals and Magnetic Transition Dipole Moments Dictated by Cylinder Helicity, Proc. Natl. Acad. Sci. U. S. A, 2017, 114, 13097–13101 CrossRef CAS PubMed.
  46. J. Wang, G. Zhuang, M. Chen, D. Lu, Z. Li, Q. Huang, H. Jia, S. Cui, X. Shao and S. Yang, et al., Selective Synthesis of Conjugated Chiral Macrocycles: Sidewall Segments of (−)/(+)-(12,4) Carbon Nanotubes with Strong Circularly Polarized Luminescence, Angew. Chem., Int. Ed., 2020, 59, 1619–1626 CrossRef CAS PubMed.
  47. F. Xu, H. Su, J. J. B. van der Tol, S. A. H. Jansen, Y. Fu, G. Lavarda, G. Vantomme, S. Meskers and E. W. Meijer, Supramolecular Polymerization as a Tool to Reveal the Magnetic Transition Dipole Moment of Heptazines, J. Am. Chem. Soc., 2024, 146, 15843–15849 CrossRef CAS PubMed.
  48. K. Kogashi, T. Matsuno, S. Sato and H. Isobe, Narrowing Segments of Helical Carbon Nanotubes with Curved Aromatic Panels, Angew. Chem., Int. Ed., 2019, 58, 7385–7389 CrossRef CAS PubMed.
  49. Q. Zhou, X. Hou, J. Wang, Y. Ni, W. Fan, Z. Li, X. Wei, K. Li, W. Yuan and Z. Xu, et al., A Fused [5]Helicene Dimer with a Figure-Eight Topology: Synthesis, Chiral Resolution, and Electronic Properties, Angew. Chem., Int. Ed., 2023, 62, e202302266 CrossRef CAS PubMed.
  50. T. He, M. Lin, H. Wang, Y. Zhang, H. Chen, C.-L. Sun, Z. Sun, X.-Y. Wang, H.-L. Zhang and Y. Chen, et al., Chiral Cylindrical Molecule with Absorption Dissymmetry Factor Towards Theoretical Limit of 2, Adv. Theory Simul., 2024, 7, 2300573 CAS.
  51. J. Nogami, Y. Nagashima, K. Miyamoto, A. Muranaka, M. Uchiyama and K. Tanaka, Asymmetric Synthesis, Structures, and Chiroptical Properties of Helical Cycloparaphenylenes, Chem. Sci., 2021, 12, 7858–7865 CAS.
  52. K. Jin, Z. Xiao, H. Xie, X. Shen, J. Wang, X. Chen, Z. Wang, Z. Zhao, K. Yan and Y. Ding, et al., Tether-Entangled Conjugated Helices, Chem. Sci., 2024, 15, 17128–17149 CAS.
  53. G. Povie, Y. Segawa, T. Nishihara, Y. Miyauchi and K. Itami, Synthesis of a Carbon Nanobelt, Science, 2017, 356, 172–175 CAS.
  54. Y. Li, Y. Segawa, A. Yagi and K. Itami, A Nonalternant Aromatic Belt: Methylene-Bridged [6]Cycloparaphenylene Synthesized from Pillar[6]Arene, J. Am. Chem. Soc., 2020, 142, 12850–12856 CAS.
  55. G. Povie, Y. Segawa, T. Nishihara, Y. Miyauchi and K. Itami, Synthesis and Size-Dependent Properties of [12], [16], and [24]Carbon Nanobelts, J. Am. Chem. Soc., 2018, 140, 10054–10059 CrossRef CAS PubMed.
  56. H. Wang, H. Zhao, S. Chen, L. Bai, Z. Su and Y. Wu, Effective Synthesis of Ladder-type Oligo(p-aniline)s and Poly(p-aniline)s via Intramolecular Snar Reaction, Org. Lett., 2021, 23, 2217–2221 CAS.
  57. T. Fukushima, H. Sakamoto, K. Tanaka, Y. Hijikata, S. Irle and K. Itami, Polymorphism of [6]Cycloparaphenylene for Packing Structure-Dependent Host–Guest Interaction, Chem. Lett., 2017, 46, 855–857 CAS.
  58. F. Sibbel, K. Matsui, Y. Segawa, A. Studer and K. Itami, Selective Synthesis of [7]- and [8]Cycloparaphenylenes, Chem. Commun., 2014, 50, 954–956 CAS.
  59. S. Hitosugi, W. Nakanishi and H. Isobe, Atropisomerism in a Belt-Persistent Nanohoop Molecule: Rotational Restriction Forced by Macrocyclic Ring Strain, Chem.–Asian J., 2012, 7, 1550–1552 CAS.
  60. Y. Han, S. Dong, J. Shao, W. Fan and C. Chi, Synthesis of a Sidewall Fragment of a (12,0) Carbon Nanotube, Angew. Chem., Int. Ed., 2021, 60, 2658–2662 CAS.
  61. M. Toya, T. Omine, F. Ishiwari, A. Saeki, H. Ito and K. Itami, Expanded [2,1][n]Carbohelicenes with 15- and 17-Benzene Rings, J. Am. Chem. Soc., 2023, 145, 11553–11565 CAS.
  62. P. V. R. Schleyer, C. Maerker, A. Dransfeld, H. Jiao and N. J. R. van Eikema Hommes, Nucleus-Independent Chemical Shifts: A Simple and Efficient Aromaticity Probe, J. Am. Chem. Soc., 1996, 118, 6317–6318 CAS.
  63. Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta and P. V. R. Schleyer, Nucleus-Independent Chemical Shifts (NICS) as an Aromaticity Criterion, Chem. Rev., 2005, 105, 3842–3888 CAS.
  64. H. Xie, Z. Xiao, Y. Song, K. Jin, H. Liu, E. Zhou, J. Cao, J. Chen, J. Ding and C. Yi, et al., Tethered Helical Ladder-Type Aromatic Lactams, J. Am. Chem. Soc., 2024, 146, 11978–11990 CAS.
  65. L. B. Casabianca, Effect of Curvature on Carbon Chemical Shielding in Extended Carbon Systems, J. Phys. Chem. A, 2016, 120, 7011–7019 CrossRef CAS PubMed.
  66. T. Iwamoto, Y. Watanabe, Y. Sakamoto, T. Suzuki and S. Yamago, Selective and Random Syntheses of [n]Cycloparaphenylenes (n = 8–13) and Size Dependence of Their Electronic Properties, J. Am. Chem. Soc., 2011, 133, 8354–8361 CrossRef CAS PubMed.
  67. H. Kono, Y. Li, R. Zanasi, G. Monaco, F. F. Summa, L. T. Scott, A. Yagi and K. Itami, Methylene-Bridged [6]-, [8]-, and [10]Cycloparaphenylenes: Size-Dependent Properties and Paratropic Belt Currents, J. Am. Chem. Soc., 2023, 145, 8939–8946 CrossRef CAS PubMed.
  68. E. R. Darzi, T. J. Sisto and R. Jasti, Selective Syntheses of [7]–[12]Cycloparaphenylenes Using Orthogonal Suzuki–Miyaura Cross-Coupling Reactions, J. Org. Chem., 2012, 77, 6624–6628 CrossRef CAS PubMed.
  69. S. Guo, L. Liu, L. Liu, Y. Fan, H. Yang, J. He, Y. Wang, Z. Bo, X. Xu and X. Chen, et al., Naphthalene Diimide-Embedded Donor–Acceptor Carbon Nanohoops: Photophysical, Photoconductive, and Charge Transport Properties, ACS Appl. Mater. Interfaces, 2025, 17, 5202–5212 CrossRef PubMed.
  70. R. G. Uceda, C. M. Cruz, S. Míguez-Lago, L. Á. de Cienfuegos, G. Longhi, D. A. Pelta, P. Novoa, A. J. Mota, J. M. Cuerva and D. Miguel, Can Magnetic Dipole Transition Moment Be Engineered?, Angew. Chem., Int. Ed., 2024, 63, e202316696 CrossRef CAS PubMed.
  71. (a) CCDC 2440960: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2my0n8; (b) CCDC 2440962: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2my0qb; (c) CCDC 2440966: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2my0vg.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.