Associate Editor: N. Yoshikai Beilstein J. Org. Chem.2025,21, 1422–1453.https://doi.org/10.3762/bjoc.21.106 Received 28 Apr 2025,
Accepted 26 Jun 2025,
Published 11 Jul 2025
Helicenes, a class of non-planar polycyclic aromatic hydrocarbons composed of ortho-fused aromatic rings forming helical architectures, have attracted considerable attention due to their intrinsic chirality and tunable optoelectronic properties. Among them, nitrogen-doped helicenes (azahelicenes) and their heteroatom-co-doped counterparts – such as B/N-, O/N-, S/N-, and Se/N-doped helicenes – have emerged as highly versatile scaffolds for chiral optoelectronic applications. The incorporation of nitrogen enables precise modulation of electronic structures, redox characteristics, and intermolecular interactions, thereby enhancing performance in circularly polarized luminescence (CPL), thermally activated delayed fluorescence (TADF), and chiral sensing. Notably, recent developments have yielded π-extended, structurally robust, and stimuli-responsive azahelicenes exhibiting record-high dissymmetry factors (|gabs| and |glum|), elevated CPL brightness (BCPL), and efficient integration into CPL-OLEDs and redox-switchable emitters. Boron–nitrogen co-doping strategies, in particular, have facilitated the development of materials with ultra-narrowband emissions, near-unity photoluminescence quantum yields, and electroluminescence dissymmetry factors (|gEL|) exceeding 10−3. Likewise, heteroatom co-doping with oxygen, sulfur, or selenium enables spectral tuning across the visible to near-infrared range, improved photostability, and dual-state emissive behavior. In parallel, significant progress in synthetic methodologies – including enantioselective catalysis, electrochemical cyclizations, and multicomponent reaction systems – has granted access to increasingly complex helicene frameworks with well-defined chirality. This review systematically summarizes recent advancements in the synthesis, structural engineering, and chiroptical performance of nitrogen-doped helicenes and their heteroatom-doped derivatives, emphasizing their potential as next-generation chiral optoelectronic materials and outlining future directions toward multifunctional integration and quantum technological applications.
Helicenes, a class of non-planar polycyclic aromatic hydrocarbons characterized by ortho-fused aromatic rings forming a helical framework, have attracted significant attention due to their inherent chirality, unique optoelectronic properties, and wide-ranging applications in asymmetric catalysis [1,2], molecular recognition [3], and organic electronics [4,5]. In recent years, the incorporation of heteroatoms – particularly nitrogen – into the helicene backbone, giving rise to so-called "azahelicenes", has emerged as a powerful strategy to modulate electronic structures, enhance solubility, and expand functional diversity [6]. Substituting carbon atoms with electron-deficient nitrogen atoms introduces new opportunities to fine-tune redox potentials, charge-transport behavior, and intermolecular interactions [7]. These modifications have proven especially valuable in applications such as organic light-emitting diodes (OLEDs) [8], circularly polarized luminescence (CPL) [9], and chiral photocatalysis [10]. In the past decade, heteroatom-containing helicenes have attracted increasing attention due to their tunable optoelectronic properties and potential applications in chiral optoelectronics. Several comprehensive reviews have examined specific classes of these molecules. Crassous and co-workers provided a detailed overview of heterohelicenes up to 2019, focusing on their structural diversity and functional applications [11]. Nowak-Król and colleagues reviewed boron-doped helicenes, emphasizing their roles in chiral materials design [12], while Maeda and Ema explored the circularly polarized luminescence (CPL) properties of azahelicenes [13]. However, despite these valuable contributions, a dedicated and up-to-date overview of nitrogen-doped helicenes – particularly those incorporating additional heteroatoms within the helical π-conjugated framework – remains lacking.
This review addresses this gap by systematically summarizing recent advances (from the past five years) in the synthesis, structural modification, and chiroptical properties of nitrogen-doped helicenes. Particular attention is given to multi-heteroatom systems co-doped with elements such as boron, oxygen, sulfur, and selenium, highlighting their influence on CPL performance and structure–property relationships. We classify the nitrogen-doped helicenes into only N-containing helicenes, B,N-containing helicenes, and X,N-containing helicenes (X = O, S or Se). In each section, structurally similar compounds are categorized into groups to facilitate comparison. Then, the others are discussed in chronological order based on their reported publication dates, with attribution to the respective research groups. Notably, helicenes bearing nitrogen atoms located outside the conjugated system are excluded from this discussion to maintain a consistent focus on electronically integrated heteroatom-doped architectures.
Review
N-Containing helicenes
Among nitrogen-containing helicenes, HBC-fused azahelicenes represent a particularly significant subclass due to their extended π-conjugation and potential for enhanced chiroptical properties. Over the past few years, multiple research groups have investigated their synthesis, structural characteristics, and optoelectronic behavior. Notably, in 2021, Jux and co-workers reported a series of superhelicenes that combine helical and planar π-systems. However, the structural characterization of compound 1 (Table 1) was impeded by its inherent instability, limiting further analysis [14]. In 2024, Liu’s group developed a series of nonalternant nanographenes 2a–c featuring a nitrogen-embedded cyclopenta[ef]heptalene core [15]. These compounds exhibit λabs at 363, 452, and 580 nm, and PLQYs of 0.05, 0.33, and 0.32, respectively. While compounds 2a and 2b display broad emission near 505 nm, 2c shows dual-emission peaks at 588 and 634 nm with an ultranarrow FWHM of 22 nm. Notably, 2b and 2c demonstrate strong chiroptical activity with |gabs| values of 6.7 × 10−3 and 1.0 × 10−2, |glum| of 2.4 × 10−3 and 7.0 × 10−3, and BCPL values of 9.1 and 95.2 M−1 cm−1, respectively. Shortly thereafter, Gong’s group further expanded the π-system by constructing a tris-hexabenzo[7]helicene 3 with a carbazole core, which emits at 595/628 nm (PLQY = 0.40), displays |gabs| = 2.98 × 10−3, and achieves a BCPL of 32.5 M−1 cm−1[16]. In 2025, Babu’s group synthesized two regioisomeric π-extended azahelicenes, 4a and 4b, which differ in the position of attachment to the carbazole core [17]. Compared to 4a, compound 4b exhibits bathochromic shifts of 12 nm in absorption and 45 nm in emission, as well as a higher ΦF (0.75 vs 0.68). Both isomers display TADF at room temperature and phosphorescence at 77 K. Notably, 4a demonstrates a long-lived red afterglow persisting for up to 30 seconds. In contrast, 4b exhibits superior chiroptical properties, with |gabs| and |glum| values of 3.91 × 10−3 and 1.12 × 10−3, respectively, and an impressive BCPL of 45.77 M−1 cm−1 (Table 1).
Table 1:
Structures and optical properties of compounds 1, 2a–c, 3, and 4a,b.a
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
2a
363
508
0.05
–
–
–
2b
452
503
0.33
6.7 × 10−3
2.4 × 10−3
9.1
2c
580
588, 634
0.32
1.0 × 10−2
7.0 × 10−3
95.2
3
525
595, 628
0.40
2.98 × 10−3
4.3 × 10−4
32.5
4a
497
497, 531, 570
0.677
–
–
–
4b
522
542, 581, 630
0.754
3.91 × 10−3
1.12 × 10−3
45.77
aCompound 1 is unstable and characterized only by mass spectrometry.
In 2021, several research groups reported structurally diverse heterohelicene systems exhibiting distinctive chiroptical and photophysical properties, highlighting the expanding potential of these molecules in chiral optoelectronics. Yorimitsu’s group developed a series of dihetero[8]helicenes through a systematic asymmetric synthesis. Among these, diaza[8]helicene 5 exhibited pronounced chiroptical activity, with absorption and emission maxima (λabs = 399 nm, λem = 405 nm), a fluorescence quantum yield (ΦF) of 0.13, and high dissymmetry factors (|gabs| = 1.9 × 10−2, |gabs| = 9.5 × 10−3 at 403 nm) [18] (Table 2). Miura and co-workers employed Pd(II)/Ag(I)-catalyzed cyclizations to construct azahelicenes, with compound 6 exhibiting enhanced chiroptical performance and protonation-induced CPL amplification [19]. Meanwhile, Audisio’s team developed heterohelicenes via regioselective [3 + 2]-cycloadditions, with compound 7 displaying pH-responsive CPL sign inversion (|glum| = +1.1 × 10−3 at 430 nm, −1.2 × 10−3 at 585 nm) attributed to reversible intramolecular charge transfer [20]. In parallel, several groups explored the functional versatility of heterohelicenes in device-oriented and sensing applications. Crassous’s group synthesized bipyridine-embedded helicenes via the Mallory reaction, enabling coordination with Ru(II) to form NIR-emissive complexes that exhibit redox-responsive chiroptical switching, notably with complex 8 showing reversible electronic circular dichroism (ECD) upon oxidation [21]. Liao and co-workers introduced a narrowband CP-TADF emitter 9, characterized by a narrow emission bandwidth (FWHM = 36 nm), |glum| = 1.1 × 10−3, |gEL| = 1.5 × 10−3, and an external quantum efficiency (EQE) of 0.14 – demonstrating promise for CPL-OLED applications [22]. Wanichacheva’s team reported urazole-functionalized aza[5]helicene 10, exhibiting selective Fe(III) sensing, marked solvatochromism, and a large Stokes shift (85 nm) with emission at 530 nm in DMSO [23] (Table 2). Collectively, these studies underscore the structural versatility and functional tunability of heterohelicenes, establishing them as robust platforms for advanced chiral optoelectronic materials. Their diverse response to external stimuli, modular synthetic accessibility, and strong CPL performance render them ideal candidates for applications in molecular sensing, stimuli-responsive switches, and next-generation CPL-active devices.
Table 2:
Structures and optical properties of compounds 5–10.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
5
399
405, 430, 460
0.13
1.9 × 10−2
9.5 × 10−3
6
405
420, 439
0.14
1.1 × 10−2
4.4 × 10−3
7
430
436, 460, 500
0.10
–
1.1 × 10−3
8
(M,Λ,Λ)
522
788
0.10
–
–
(P,Λ,Λ)
512
786
0.25
–
–
9
440
467
0.47a
–
1.1 × 10−3
10
400
485
–
–
–
aAs detected in film.
In 2021, Ema’s group reported the synthesis of carbazole-based azahelicenes 11a–e via intramolecular Scholl reactions [24] (Table 3). All compounds exhibited strong absorption in the UV–vis region (250–450 nm) and fluorescence emission between 400–550 nm. Among these, compound 11c, a saddle-shaped dibenzodiaza[8]circulene, was particularly noteworthy as the first example of its kind synthesized in solution and structurally confirmed via single-crystal X-ray diffraction. It demonstrated the highest CPL performance among the series, with a |glum| value of 3.5 × 10−3 and a photoluminescence quantum yield (PLQY) of 0.31, indicating its potential as a chiral emissive material. Building upon this foundation, the same group in 2024 developed a series of structurally refined aza[7]helicenes (compounds 12a and 12b) under modified Scholl reaction conditions [25]. These products were obtained as optically active diastereomers, which were successfully separated using silica gel chromatography. Additionally, two cyclic dimers, designated as compounds 12c and 12d, were isolated, exhibiting strong absorption bands at 493 and 474 nm, high PLQYs of 0.61 and 0.54, and notable CPL activity (|glum| = 0.74 × 10−3 and 1.3 × 10−3, respectively), with corresponding brightness values (BCPL) reaching 19 and 31 M−1 cm−1 (Table 3). Importantly, both dimers displayed selective fluoride ion recognition through hydrogen bonding, with (M,M)-12c exhibiting a high binding constant (Ka = 2 × 105 M−1). The resulting [12c·F−] and [12d·F−] complexes exhibited red-shifted circular dichroism (CD), fluorescence, and CPL spectra, underscoring the capability of helicene-based frameworks for anion-responsive chiroptical modulation. These findings highlight how precise structural design and supramolecular engineering can facilitate the development of high-performance, stimuli-responsive chiral luminophores.
Table 3:
Structures and optical properties of 11a–e and 12a–d.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
11a
418
432, 454
0.28
4.9 × 10−3
3.2 × 10−3
–
11b
419
432, 455
0.27
5.9 × 10−3
3.4 × 10−3
–
11c
419
432, 456
0.31
5.4 × 10−3
3.5 × 10−3
–
11d
422
458, 480
0.10
3.2 × 10−3
3.9 × 10−4
–
11e
412
456
0.24
4.5 × 10−4
2.9 × 10−4
–
12a
436
447, 474
0.45
4.8 × 10−3
2.6 × 10−3
6.7
12b
423
431, 456
0.32
3.8 × 10−3
2.2 × 10−3
2.8
12c
494
502, 536
0.64
2.4 × 10−3
6.5 × 10−4
19
12d
475
485, 514
0.54
2.7 × 10−3
1.4 × 10−3
31
In 2022, Zhang and co-workers reported a nitrogen-embedded quintuple [7]helicene 13, constructed by hybridizing helicene and azacorannulene π-systems [26] (Table 4). Compound 13 exhibited distinct absorption bands at 408, 611, and 715 nm, with strong near-infrared (NIR) fluorescence centered at 770 nm and a PLQY value of 0.28. Upon coordination with tris(4-bromophenyl)aminium hexachloroantimonate (BAHA), a new absorption band emerged around 900 nm, extending to 1300 nm, indicative of charge-transfer processes. The enantiomers of 13 displayed mirror-image CD signals and showed excellent dispersibility in polar solvents, highlighting their potential for NIR bio-imaging applications. In parallel, Církva’s group synthesized a series of aza[n]helicenes 14a–d via photocyclodehydrochlorination [27]. These compounds exhibited dual fluorescence bands, with emission red-shifting progressively with increasing helical length. Protonation further induced red-shifted emission, with compound 14d-H+ emitting at 542 nm. However, PLQYs decreased significantly from 0.078 to 0.006 with longer helicenes. The CD spectra of 14c and 14d were found to resemble their carbohelicene analogues, underscoring the structural fidelity and chiroptical retention upon nitrogen incorporation. Qian’s group developed a series of azahelicenes 15a–d through Bischler–Napieralski cyclization [28]. Notably, compound 15b displayed a high interconversion barrier of 36.0 kcal mol−1, enabling enantiomeric resolution. All compounds exhibited visible-range fluorescence (400–500 nm) and structured UV–vis absorption spectra. Importantly, 15b showed acid/base-switchable UV and CD spectra, suggesting potential for use in responsive optoelectronic systems. Hu’s group reported an X-shaped double [7]helicene 16 functionalized with four triazole units, which demonstrated absorption at 368 and 516 nm, strong emission at 553 nm, a high PLQY of 0.96, |gabs| of 1.1 × 10−2, |glum| of 9.1 × 10−4, and BCPL of 30.1 M−1 cm−1 – surpassing the performance of its all-carbon and thiadiazole counterparts [29]. In a related study, Hu’s team synthesized double aza[5]helicenes 17a and 17b, among which compound 17b exhibited red-shifted emission (538–632 nm in CHCl3) and the largest Stokes shift (192 nm), attributed to extended conjugation and sulfur incorporation [30] (Table 4). These findings collectively underscore how structural modulation and heteroatom doping can tailor the optical, chiroptical, and stimuli-responsive behavior of azahelicenes, providing strategic design avenues for next-generation chiral optoelectronic materials.
Table 4:
Structures and optical properties of 13, 14a–d, 15a–d, 16, and 17a,b.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
13
715
770
0.28
–
–
–
14a
313
380, 399
0.077
–
–
–
14b
302
410, 431
0.120
–
–
–
14c
311
421, 443
0.067
–
–
–
14d
337
443, 467
0.029
–
–
–
15a
398
408, 430
–
–
–
–
15b
404
408, 434
–
–
–
–
15c
407
413, 437
–
–
–
–
15d
424
434, 456
–
–
–
–
16
516
553
0.96
1.1 × 10−2
9.1 × 10−4
30.1
17a
328
458
0.010
–
–
–
17b
440
632
0.014
–
–
–
In 2023, Langer’s group synthesized a series of double aza[4,6]helicenes 18a–l featuring diverse peripheral substituents through a one-pot, multistep synthetic protocol [31] (Table 5). Selected compounds such as 18b, 18c, 18d and 18l exhibit similar λabs around 410 nm and emit fluorescence centered near 530 nm, demonstrating consistent optical profiles despite structural variation. In a parallel effort, Yang’s group developed an efficient, enantioselective synthetic approach toward azahelicenes via a chiral phosphoric acid-catalyzed multicomponent Povarov reaction or oxidative aromatization [32]. Among the synthesized compounds, compound 19 displayed dual absorption bands at 260 and 325 nm and emission peaks at 420 and 440 nm, which red-shifted to approximately 500 nm upon trifluoroacetic acid treatment. Both the neutral and protonated forms of 19 exhibited mirror-image CD and CPL spectra, with high |glum| values of 1.4 × 10−3 and 1.3 × 10−3, respectively, underscoring their potential for responsive chiral optoelectronic applications. Concurrently, Liu [33] and Ishigaki’s [34] groups independently reported a class of highly twisted nitrogen-doped heptalene derivatives (e.g., compound 20a), which exhibit consistent absorption at 315 nm and blue fluorescence centered near 450 nm, regardless of the substituents. These compounds display redox and electronic behaviors reminiscent of nitrogen-doped azulenes, featuring strong absorption dissymmetry factors (|gabs|) at 345 nm – 1.2 × 10−2 for compound 20a, 1.0 × 10−2 for 20d, and 1.3 × 10−2 for 20e (Table 5). Notably, the radical cation form of compound 20e (20e•+) exhibits pronounced CD signals extending into the near-infrared region, suggesting potential for redox-responsive chiral photonic systems.
Table 5:
Structures and optical properties of 18a–l,19, and 20a–e.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
18b
411
530
0.15
–
–
18c
409
520
0.16
–
–
18d
419
525
0.17
–
–
18i
413
525
0.14
–
–
19
325
420, 440
–
–
1.4 × 10−3
20a
315a, 320b
447
–
1.2 × 10−2
–
20b
315
459
–
–
–
20c
315
446
–
–
–
20d
320
–
–
1.0 × 10−2
–
20e
321
–
–
1.3 × 10−2
–
aBased on reports from Liu's group; bbased on reports from Ishigaki's group.
In 2023, Chen’s group reported three nitrogen–nitrogen (NN)-embedded azahelicenes 21a–c, among which compound 21c, a structurally defined antiaromatic double aza[7]helicene – exhibited distinctive long-wavelength optical and chiroptical properties [35] (Table 6). In the solid state, 21c emitted in the far-red region at 641 nm (ΦF = 0.10) and demonstrated CPL with |glum| = 2.04 × 10−4. In solution, 21c showed a strong absorption band at 560 nm and a high ФF value of 0.86 at 583 nm, yielding a BCPL value of 13.2 M−1 cm−1. Notably, compound 21c undergoes reversible redox interconversion to its radical cation 21c•+ and dicationic 21c2+ states via chemical oxidation, enabling controllable switching between antiaromatic and aromatic configurations. These results provide a compelling strategy for engineering redox-switchable chiral luminophores. In 2024, the same research group expanded on this redox-responsive platform by constructing a polycationic open-shell cyclophane 22, comprising carbazole-embedded aza[7]helicene subunits [36]. Compound 22 displays intense fluorescence (ΦF = 0.99), exceptionally high BCPL as 100.2 M−1 cm−1, and marked chiroptical activity (|gabs| = 2.50 × 10−3 at 435 nm; |glum| = 5.00 × 10−3 at 460 nm) (Table 6). Upon mild oxidation, neutral 22 undergoes stepwise conversion into highly charged, multispin open-shell species 222+2• and 224+4•, preserving strong chiroptical signals. This study presents a novel approach to constructing stable, redox-switchable chiral luminophores based on extended azahelicene architectures, offering broad potential for molecular electronics and spintronic devices.
Table 6:
Structures and optical properties of 21a–c and 22.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
21a
408
423
0.26
9.78 × 10−4
–
–
21a in film
≈410
449
0.15
–
–
–
21b
495
521
0.77
–
–
–
21b in film
≈500
548
0.63
–
–
–
21c
560
583
0.86
4.76 × 10−4
2.22 × 10−4
13.2
21c in film
≈570
641
0.10
–
2.04 × 10−4
–
22
438
480
0.99
2.50 × 10−3
5.00 × 10−3
100.2
In 2024, Qiu’s group synthesized π-extended diaza[7]helicenes 23a–f incorporating dual heptagonal rings [37]. Compound 23a exhibits dynamic chirality, aggregation-induced emission (AIE), and intense CPL (|glum| = 1.7 × 10−2), whereas compound 23f, with lateral π-extension, shows enhanced thermal stability and green emission at 517 nm (Table 7). Kuehne and co-workers reported two radical aza[7]helicenes, 24a and 24b, exhibiting distinct photophysical behaviors [38]. Compound 24b features a higher PLQY (0.43), while 24a demonstrates doublet-state CPL (|glum| = 5.0 × 10−4), highlighting the potential of helicene radicals for spintronic applications. Meng’s group synthesized carbonyl-nitrogen embedded hetero[7]helicenes 25a and 25b bearing axial chirality [39]. Compound 25a displays excellent optical characteristics with ΦF = 0.57, |gabs| = 1.7 × 10−2, |glum| = 1.4 × 10−3, and a BCPL of 8.94 M−1 cm−1. Then, Chen’s group contributed triple aza[6]helicenes 26a and 26b with |glum| values of approximately 3.0 × 10−3, offering new architectures for CPL-active helicenes [40]. Singh’s group developed fluorophore-conjugated aza[7]helicenes 27a–d, with 27b demonstrating pronounced intramolecular charge transfer (ICT), a high ΦF of 0.71 and an extended fluorescence lifetime (τ) of 15.5 ns [41]. Wu’s group synthesized a family of expanded azahelicenes 28a–e, where increasing helical length leads to red-shifted emission, prolonged lifetime, and attenuated PLQY [42]. Nonetheless, these compounds exhibit outstanding chiroptical performance, with |gabs|max reaching 4.8 × 10−2, |glum|max = 2.1 × 10−2, and BCPL values up to 76 M−1 cm−1. Collectively, these investigations underscore the efficacy of heteroatom doping, extended π-conjugation, and radical design in advancing azahelicene-based systems. These approaches significantly enhance optical and chiroptical performance, paving the way for high-efficiency chiral optoelectronic and quantum materials.
Table 7:
Structures and optical properties of 23a–f, 24a,b, 25a,b, 26a,b, 27a–d, and 28a–e.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
23a
360
625
–
–
1.7 × 10−2a
–
23f
462
517
–
–
2.0 × 10−3
–
24a
642
696
0.34
4.4 × 10−4
5 × 10−4
0.25
24b
655
712
0.43
1 × 10−4
–
–
25a
506
525
0.57
1.7 × 10−2
1.4 × 10−3
8.94
25b
513
535
0.55
2.2 × 10−2
8 × 10−4
4.29
26a
388
506, 530
0.055
1.2 × 10−2
3.0 × 10−3
–
26b
393
508, 532
0.058
1.4 × 10−2
3.2 × 10−3
–
27a
483
524
0.38
–
–
–
27b
487
539
0.71
–
–
–
27c
459
590
0.24
–
–
–
27d
470
611
0.53
–
–
–
28a
414
496, 532
0.152
–
–
–
28b
≈475
511, 543
0.116
4.4 × 10−2
3 × 10−3
16
28c
≈475
522, 550
0.089
4.8 × 10−2
1.4 × 10−2
61
28d
≈475
530, 554
0.066
4.3 × 10−2
2.1 × 10−2
76
28e
≈475
530, 555
0.034
–
–
–
aIn the aggregated state.
In 2024, Kivala’s group selectively synthesized highly distorted [6]helicenes 29a and 29b incorporating azocine units via a regioselective Beckmann rearrangement from oxime precursor 29c[43] (Table 8). For comparative evaluation, the corresponding lactams 29d and 29e and amines 29f and 29g were also obtained. Compounds 29a and 29b exhibit λabs centered at 513 nm, while the amines 29f and 29g display high ФF values of 0.48 and 0.56, respectively. Notably, azocine derivative 29b exhibits the highest CPL activity among the series, with a |glum| value of 1.6 × 10−3. In addition, both 29a and 29b demonstrate redox activity, undergoing reversible formation of radical anions, dianions, and radical cations. The radical cation 29b•+, in particular, exhibits a broad near-infrared (NIR) absorption band extending to 3000 nm, highlighting its potential for NIR optoelectronic applications. Building on this work, in 2025 the same group reported the synthesis of a stable N-heterotriangulene dimer (compound 30) bridged by a rigid π-conjugated [5]helicene [44]. This chiral dimer undergoes reversible stepwise oxidation to 30•+ and 302+, accompanied by pronounced NIR Cotton effects extending up to 2000 nm. These results provide critical insights into the rational design of redox-switchable, NIR-active chiral molecular systems, underscoring their promise in advanced optoelectronic and spintronic technologies.
Table 8:
Structures and optical properties of 29a–f and 30.a
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
29a
513
540, 565
0.01
2.5 × 10−3
–
–
29b
513
552, 582
0.12
1.9 × 10−3
1.6 × 10−3
–
29d
510
539, 570
0.52
3.0 × 10−3
6.0 × 10−4
–
29e
510
543, 575
0.51
2.1 × 10−3
2.4 × 10−4
–
29f
510
536, 570
0.48
2.0 × 10−3
9.1 × 10−4
–
29g
547
609, 652
0.56
2.4 × 10−3
6.0 × 10−4
–
30
495
534
0.42
1.25 × 10−3
1.1 × 10−3
7.00
aThe optical properties of compound 29c are not mentioned in the original paper.
In 2024, Tanaka’s group synthesized and characterized a series of length-variable aza[n]helicenes 31a–f via a one-pot intramolecular cyclodehydrogenation [45] (Table 9). Notably, compounds 31e and 31f represent the first examples of triple-layered heterohelicenes with fully conjugated frameworks. All members of the series demonstrate high solubility, attributed to intermolecular hydrogen bonding with solvent molecules. With increasing helical length, both the λabs and λem exhibit progressive bathochromic shifts, while the ФF values systematically decline, without clear saturation within the investigated range. Chiroptical measurements of the N-butylated aza[n]helicenes 31g–j reveal |gabs| and |glum| values on the order of 10−3. These findings address long-standing challenges in the synthesis and stabilization of extended heterohelicenes, paving the way for the development of structurally persistent, π-extended chiral materials. In a parallel effort, Tanaka’s group synthesized benzannulated double aza[9]helicene 32a and its alkylated derivatives 32b and 32c via a one-pot oxidative fusion strategy [46]. Compared to the parent compound 32a (ΦF = 0.07), compounds 32b and 32c exhibit significantly enhanced ΦF (0.35), red-shifted absorption bands, and |gabs| values of 2.4 × 10−3 and 2.3 × 10−3 at 345 nm, respectively. Their corresponding BCPL values reach 16.0 and 19.2 M−1 cm−1. Furthermore, terminus-functionalized aza[9]helicenes 33a, 33b, and 33c were prepared to investigate interlayer interactions [47]. Among them, the pyrene-decorated compound 33c displays red-shifted emission and prolonged fluorescence lifetimes as solvent polarity increases, indicating enhanced excited-state stabilization. Collectively, these studies offer valuable strategies for stabilizing long π-extended helicenes and finely tuning their chiroptical and emissive properties, thereby advancing their application in multifunctional chiral photonic and sensing platforms.
Table 9:
Structures and optical properties of 31a–j, 32a–c, and 33a–c.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
31a
412
437, 466, 500
0.21
–
–
–
31b
425
452, 479, 514
0.17
–
–
–
31c
438
427, 450, 480
0.11
–
–
–
31d
451
466, 491, 530
0.09
–
–
–
31e
388
483, 511
0.18
–
–
–
31f
310
508
0.08
–
–
–
31g
409
465, 495
0.16
5.6 × 10−3
4.5 × 10−3
8.6a
31h
314
482, 508
0.16
4.2 × 10−3
4.2 × 10−3
–
31i
315
508
0.09
4.2 × 10−3
1.7 × 10−3
–
31j
≈385
≈520
0.07
1.7 × 10−3
5.7 × 10−3
–
32a
464
496, 529, 570
0.07b
0.33c
–
–
–
32b
510
521, 555
0.35
2.4 × 10−3
–
16.0
32c
508
522, 556
0.35
2.3 × 10−3
–
19.2
33a
415
441, 466, 500
0.19
–
–
–
33b
414
437, 466, 500
0.21
–
–
–
33c
416
441, 466, 500
0.08
–
–
–
aAccording to reference paper [42]; bin THF; cin DMSO.
In 2025, Gryko’s group synthesized a series of heterohelicenes 34a–c, featuring a 1,4-dihydropyrrolo[3,2-b]pyrrole (DHPP) core [48] (Table 10). The compounds exhibit similar absorption and emission profiles. However, compound 34c stands out due to its pronounced solvatofluorochromism (λem = 546 nm, ΦF = 0.42 in DMSO). Among the series, compound 34b exhibits the highest |glum| of 7.22 × 10−3, while compound 34c shows the greatest BCPL as 29.3 M−1 cm−1. These studies underscore the importance of regioisomerism and molecular core design in optimizing the chiroptical and emissive properties of heteroatom-rich nanographenes, advancing their potential in next-generation optoelectronic and chiral photonic devices.
Table 10:
Structures and optical properties of 33a,b and 34a–c.
compound
λabs(max) [nm]
λem [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
34a
438
460, 481
0.270
–
1.33 × 10−3
2.0
34b
446
463, 488
0.045
–
6.11 × 10−3
4.3
34c
456
483, 505
0.324
–
3.25 × 10−3
29.3
B,N-containing helicenes
Enhancing charge transfer between electron-donating and electron-accepting units, as well as extending π-conjugated frameworks, are widely employed strategies for achieving longer-wavelength emission in optoelectronic materials. Inspired by the electronic configuration of borazine, boron has emerged as a valuable electron-accepting counterpart to electron-donating nitrogen in conjugated systems, enabling the design of donor–acceptor helicenes with tunable photophysical properties.
In 2020, Ema and co-workers developed a series of chiral carbazole-based BODIPY analogues 35a–f, derived from helical carbazole-based BF2 dyes [49] (Table 11). These analogues exhibit red-shifted emission and enhanced CPL compared to their carbazole-based helicene precursors. At λabs (≈500 nm), the compounds display |gabs| values ranging from 1.1 × 10−3 to 3.1 × 10−3, ФF values of 20–36%, and |glum| values between 7.0 × 10−4 and 1.9 × 10−3. In a subsequent study, Ema’s group reported an N-containing hetero[7]helicene 36a containing a boron–nitrogen coordination site [50]. Its chiroptical properties could be modulated through the addition of tetrabutylammonium (TBA) salts, which transformed the boron center from a trigonal planar to a tetrahedral geometry, thereby enhancing the |glum| from 4.7 × 10−4 to 1.5 × 10−3 (OAc−, 36c) and 1.7 × 10−3 (F−/OH−, 36b/36d). Treatment with Ag+ ions reversed this coordination, restoring the neutral trigonal boron center and its initial optical characteristics. These findings underscore the potential of boron–nitrogen-embedded helicene frameworks as tunable chiral luminophores with reversible CPL modulation, offering promising strategies for the development of advanced molecular optoelectronic devices.
Table 11:
Structures and optical properties of 35a–f and 36a–d.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
35a
495
568
0.22
2.7 × 10−3
1.7 × 10−3
35b
508
594
0.20
3.1 × 10−3
1.3 × 10−3
35c
508
566
0.33
1.2 × 10−3
8.7 × 10−4
35d
524
592
0.21
1.1 × 10−3
7.0 × 10−4
35e
(R,P)
508
576
0.30
2.3 × 10−3
1.5 × 10−3
(R,M)
509
571
0.36
1.5 × 10−3
1.2 × 10−3
35f
(R,P)
530
605
0.20
1.8 × 10−3
1.2 × 10−3
(R,M)
532
602
0.26
1.5 × 10−3
8.8 × 10−4
36a
487
493
–
1.6 × 10−3
4.7 × 10−4
36b
502
512
–
3.0 × 10−3
1.7 × 10−3
36c
510
526
–
2.9 × 10−3
1.5 × 10−3
36d
511
520
–
3.2 × 10−3
1.7 × 10−3
In 2021, Hatakeyama and co-workers developed an expanded B,N-containing heterohelicene 37 via a one-step synthesis employing excess BBr3 at 180 °C in an autoclave, achieving a 44% yield [51] (Table 12). In a 1 wt % PMMA-dispersed film, compound 37 exhibited ultra-narrowband emission (FWHM = 16 nm) at 484 nm with an 80% PLQY. OLEDs based on 37 demonstrated excellent external quantum efficiency, current efficiency, and power efficiency. Duan and co-workers reported B,N-containing double hetero[7]helicenes 38a,b, which exhibited deep-red fluorescence emission at 662 and 692 nm, respectively, with narrow emission bandwidths (full width at half maximum, FWHM = 38 nm) and exceptional PLQYs of 100% [52]. Remarkably, they achieved maximum EQEs of 28.1% and 27.6%, representing the highest reported values for thermally activated delayed fluorescence (TADF) emitters operating above 650 nm. Shortly thereafter, Wang’s group reported a related series of B,N-containing compounds 38a–c, which displayed pronounced chiroptical activity in the visible region [53]. These compounds displayed the highest |gabs| values recorded for helicenes to date – 0.033, 0.031, and 0.026 at 502, 518, and 526 nm, respectively. They also showed near-unity ФF values of 100%, 99%, and 90%, with corresponding λem at 660, 684, and 696 nm, and |glum| values of 2 × 10−3. The calculated BCPL reached 28.5, 37.1, and 40.0 M−1 cm−1, positioning these helicenes among the most efficient red CPL emitters reported to date (Table 12).
Table 12:
Structures and optical properties of 37, 38a–c, and 39.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
38a
627
660
1.00
3.3 × 10−2
2.0 × 10−3
38b
650
684
0.99
3.1 × 10−2
2.0 × 10−3
38c
662
696
0.90
2.6 × 10−2
2.0 × 10−3
39
590
617
0.96
1.2 × 10−2
1.4 × 10−3
film
λabs(max) [nm]
λem(max) [nm]
ФF
FWHM [nm]
37 in PMMA
477
484
0.80
16
38a in CBP
–
672
–
48
38b in CBP
–
698
–
49
39 in mCPBC
–
624
0.95
–
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
37
480
–
17
(0.09, 0.21)
22.9a
38a
664
–
48
(0.72, 0.28)
28.1
38b
686
–
49
(0.72, 0.28)
27.6
39
617
1.9 × 10−3
48
(0.67, 0.33)
36.6
aAs detected at 10 cd m−2.
However, such long-wavelength emission poses challenges for achieving optimal color purity in OLED devices. To overcome this limitation, Duan’s group subsequently introduced a covalent B–N bond into the helicene framework in 2023, affording compound 39[54]. This material emits at 617 nm with a FWHM of 38 nm and maintains a near-unity PLQY. Circularly polarized OLEDs (CP-OLEDs) based on 39 exhibit outstanding device performance, achieving a |gEL| of 1.91 × 10−3, a record-high EQE exceeding 36%, and operational stability with an LT95 of approximately 400 h at 10,000 cd m−2. These findings underscore the efficacy of B–N covalent integration in helicene-based frameworks for realizing high-efficiency, spectrally optimized, and robust red CP-OLED emitters.
In 2022, Yang and co-workers reported a W-shaped double hetero[5]helicene 40, incorporating boron, nitrogen, and sulfur atoms within its framework [55] (Table 13). Compound 40 exhibits exceptional photophysical and electroluminescent performance, including a PLQY value of 100% and a |glum| value of 2.1 × 10−3. Circularly polarized organic light-emitting diodes (CP-OLEDs) based on 40 demonstrated a |gEL| of 2.2 × 10−3, a narrow emission bandwidth (FWHM = 49 nm), and a maximum external quantum efficiency (EQE) of 31.5%, placing it among the highest-performing multiple-resonance-induced thermally activated delayed fluorescence (MR-TADF) emitters to date. In 2023, the same group introduced the first deep-blue chiral MR-TADF emitters based on heterohelicene scaffolds 41a–c[56]. These compounds exhibited sharp emissions at 440–444 nm in solution and 445–449 nm in doped films, with emission bandwidths as narrow as 23 nm and PLQYs reaching up to 95%. Notably, racemic 41b and 41c displayed excellent chiroptical properties, with |glum| values ranging from 1.4 to 1.5 × 10−3 and BCPL values exceeding 22 M−1 cm−1. Compound 41c, in particular, achieved a |gEL| of 2.6 × 10−3 and a maximum luminance exceeding 10,000 cd m−2. These findings underscore the significant potential of heteroatom-integrated helicene systems as high-efficiency, CPL-active MR-TADF materials for next-generation OLED technologies, particularly in the development of deep-blue emissive devices with high color purity and device efficiency.
Table 13:
Structure and optical properties of 40 and 41a–c.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
40
483
520
0.98
–
2.1 × 10−3
41a
424
440
0.82
–
–
41b
422
443
0.91
1.4 × 10−3
1.4 × 10−3
41c
427
444
0.95
1.5 × 10−3
1.5 × 10−3
film
λabs(max) [nm]
λem(max) [nm]
ФF
FWHM [nm]
40 in DMIC-TRZ
–
525
–
48
41a in DOBNA-OAr
–
445
0.82
35
41b in DOBNA-OAr
–
448
0.91
28
41c in DOBNA-OAr
–
449
0.95
28
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
40
524
2.2 × 10−3
49
(0.26, 0.66)
31.5
41a
443
–
26
(0.15, 0.05)
23.4
41b
445
2.2 × 10−4
24
(0.15, 0.04)
27.5
41c
447
2.6 × 10−4
24
(0.15, 0.05)
29.3
In 2022, Marder and co-workers introduced various boryl substituents at both termini of a series of nitrogen-doped [5]helicenes, yielding helicenoids 42a–h[57] (Table 14). The Bpin-substituted derivatives 42a–e exhibited broad emission across the 400–800 nm range, whereas their analogues 42f and 42g showed negligible emission, indicating a strong dependence of photophysical behavior on boryl-substituent identity. Compared to their parent azahelicenes, these compounds displayed significantly larger Stokes shifts, highlighting the pronounced electronic effects of boryl incorporation. Notably, when a CF3 group was introduced as a substituent on the azahelicene core, the resulting boryl-functionalized compound 42c exhibited an emission maximum at 563 nm in CH2Cl2, with a quantum yield of 15%, representing the highest emission efficiency observed among the boron-containing quasi-circulenes.
Table 14:
Structure and optical properties of 42a–h.a
compound
λabs(max) [nm]
λem(max) [nm]
ФF
42a
372
520
0.08
42b
373
522
0.08
42c
364
563
0.15
42d
372
530
0.07
42e
407
588
0.05
42f
385
–
–
42g
366
–
–
42h
–
–
–
aNo gabs or glum values were reported.
In 2022, Lu and co-workers developed a series of helical aza-BODIPY analogues 43a–h, featuring a distinctive B–O–B bridge installed within each molecule [58] (Table 15). These compounds display broad chiroptical responses extending from the ultraviolet to the entire visible spectrum – an uncommon characteristic among helicene-type systems. Among them, the phenyl-substituted aza[7]helicene 43f exhibits pronounced chiroptical activity, with |gabs| and |glum| values reaching 3.04 × 10−3 and 1.30 × 10−3, respectively, and a high BCPL of 11.5 M−1 cm−1 in the near-infrared region. In contrast, the corresponding aza[5]helicene analogue shows negligible chiral response, with |gabs| and |glum| values in the 10−5 range. To further enhance chiroptical performance, Lu’s group introduced edge-positioned methyl and ethyl substituents into the helical core, affording 44a and 44b[59]. Compared with 43c, they are with significantly improved |gabs| values of 1.51 × 10−3 and 1.69 × 10−3, respectively. This study underscores the critical importance of molecular design in modulating chiroptical properties and provides valuable insights into the development of helicene-based BODIPY systems for near-infrared CPL applications. In 2024, Shimizu’s group reported azabora[6]helicenes 45a and 45b[60]. However, their enantiomers could not be isolated due to low racemization barriers. The F- and Ph-coordinated derivatives displayed moderate PLQYs in solution (0.26 and 0.18, respectively), which dropped markedly in the solid state (0.02 and 0.04) owing to aggregation-caused quenching (ACQ).
Table 15:
Structure and optical properties of 43a–h, 44a,b, and 45a,b.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
43a
588
625
0.59
4 × 10−5
3 × 10−5
43b
623
649
0.56
–
–
43c
601
640
0.31
–
–
43d
634
668
0.12
–
–
43e
646
682
0.30
2.0 × 10−3
1.3 × 10−3
43f
677
708
0.24
3.0 × 10−3
1.3 × 10−3
43g
660
695
0.16
1.8 × 10−3
1.2 × 10−3
43h
691
719
0.10
–
–
44a
624
665
0.08
1.5 × 10−3
–
44b
625
665
0.07
1.7 × 10−3
–
45a
548
568
0.26
–
–
45b
554
574
0.18
–
–
In 2023, Yang and co-workers reported a pair of (NBN)2-containing double and quadruple helicenes 46a–d[61] (Table 16). The neutral compounds exhibited high PLQYs of 99% and 65% in solution, and 90% and 55% in PMMA-doped films, respectively, with exceptionally narrow full-width (FWHM values as 24 nm and 22 nm). Stepwise titration experiments with fluoride ions induced a change in the coordination number of the boron centers from three to four, forming corresponding anionic species. This coordination triggered red-shifted absorption and CPL responses while maintaining excellent PLQYs – 99% and 90% in solution, and 80% and 77% in PMMA-doped films, respectively.
Table 16:
Structure and optical properties of 46a–d.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
46aa
511
524
0.99
–
–
46ba
507
522
0.65
6.2 × 10−3
1.0 × 10−3
46cb
524
567
0.99
5.0 × 10−3
6.0 × 10−4
46db
518
541
0.90
6.0 × 10−3
7.0 × 10−4
film
λabs(max) [nm]
λem(max) [nm]
ФF
FWHM [nm]
46a in PMMA
–
–
0.95
–
46b in PMMA
–
–
0.55
–
46c in PMMA
–
–
0.80
–
46d in PMMA
–
–
0.77
–
aIn toluene; bin acetone.
In 2024, Wang’s group developed a B,N-embedded hetero[8]helicene 47, exhibiting narrow green emission at 531 nm (FWHM = 36 nm), a high PLQY of 93%, and outstanding CP-OLED performance (EQE = 32.0%; |gEL| = 7.74 × 10−4) [62] (Table 17). Bin’s group introduced orthogonal spiro-structures into hetero[6]helicenes 48a–c, achieving near-unity PLQYs in solution (up to 99%) and OLED external quantum efficiencies (EQEs) exceeding 31% [63]. Chen’s group reported 49, a B,N-containing hetero[9]helicene that emits at 578 nm with a PLQY of 98% and showing excellent chiroptical properties (|glum| = 5.8 × 10−3; BCPL = 220.75 M−1 cm−1) [64]. OLEDs incorporating compound 49 demonstrated an EQE of 35.5% and |gEL| = 6.2 × 10−3. Zhang’s group synthesized 50a–f, with and without installed heptagons [65]. The heptagon-containing derivatives showed red-shifted emission, broader FWHM, lower PLQYs, and diminished BCPL values, indicating a trade-off between extended conjugation and emissive efficiency. Yin’s group introduced 1,4-BN motifs into compounds 51a and 51b, which emitted blue-green light at 474 and 465 nm, respectively, and exhibited moderate CPL activity (|glum| ≈ 5 × 10−4) [66] . OLEDs based on compound 51a emitted at 502 nm and achieved an EQE of 3.18%. Liu’s group positioned B and N atoms on the inner rim of 52a and 52b[67]. While 52b exhibited remarkably high |gabs| and |glum| values (6.1 × 10−2 and 2.4 × 10−2, respectively), its PLQY was relatively low (24%). Further molecular optimization led to the development of compounds 53a–c, which demonstrated ultra-narrow emission bands (FWHM = 16–34 nm), high PLQYs (67–82%), and exceptional CPL brightness (BCPLs of 583, 374, and 349 M−1 cm−1, respectively), with compound 53a setting a new record for BN-containing helicene CPL brightness [68]. These collective findings underscore the critical role of rational BN doping, π-conjugation engineering, and structural rigidity in precisely tuning the photophysical and chiroptical properties of helicene-based materials, thereby advancing the design of next-generation CPL-active optoelectronic systems with superior performance metrics.
Table 17:
Structure and optical properties of 47, 48a–c, 49, 50a–f, 51a,b, 52a,b, and 53a–c.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
47
510
531
0.93
1.4 × 10−3
5.8 × 10−4
48a
482
503
0.91
–
–
48b
495
516
0.99
–
–
48c
493
515
0.94
–
–
49
546
578
0.98
5.6 × 10−3
5.8 × 10−3
50a
548
595
0.68
7.4 × 10−3
2.7 × 10−3
50b
545
585
0.66
8.6 × 10−3
2.5 × 10−3
50c
553
598
0.74
3.1 × 10−3
2.7 × 10−3
50d
622
675
0.11
4.7 × 10−3
2.9 × 10−3
50e
563
623
0.27
–
–
50f
595
641
0.02
6.6 × 10−3
5.0 × 10−3
51a
453
474
0.83
6.2 × 10−3
5.1 × 10−4
51b
447
465
0.54
2.5 × 10−3
4.8 × 10−4
52a
403
409
0.31
3.6 × 10−2
2.4 × 10−2
52b
423
430
0.24
6.1 × 10−2
4.8 × 10−2
53a
506
515
0.82
2.4 × 10−2
1.7 × 10−2
53b
513
529
0.67
1.1 × 10−2
1.2 × 10−2
53c
516
535
0.72
1.1 × 10−2
8.0 × 10−3
film
λabs(max) [nm]
λem(max) [nm]
ФF
FWHM [nm]
46a from CHCl3
–
667
0.02
48
46b from CHCl3
–
632
0.04
35
47 in PhCzBCz
–
≈545
0.92
≈50
51a in DPEPO
–
472
0.32
38
51b in DPEPO
–
467
0.42
29
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
47
536
7.7 × 10−4
38
(0.32, 0.66)
31.1
48a
490
–
30
(0.10, 0.41)
25.2
48b
506
–
37
(0.15, 0.65)
29.2
48c
522
–
37
(0.22, 0.70)
31.0
49
580
6.2 × 10−3
48
(0.53, 0.46)
35.4
51a
502
–
35
(0.14, 0.55)
3.2
However, these findings also suggest that boron may not always be the optimal choice for enhancing charge-transfer properties. The delocalization of electrons between the vacant p-orbital of boron and the electron-rich π-conjugated systems can diminish both the electron-accepting capability of boron and the electron-donating efficiency of the conjugated framework. Additionally, the inherently low electronegativity of boron further limits its effectiveness as an electron acceptor, thereby restricting the achievable red-shift in emission. To overcome these limitations, alternative electron-withdrawing atoms and functional groups have been introduced into nitrogen-doped helicene frameworks to improve their photophysical performance and extend emission into the longer wavelength region.
X,N-containing helicenes (X = O, S or Se)
Imide functional groups are well recognized for their strong electron-accepting character, making them valuable moieties in the design of optoelectronic materials. When incorporated into π-conjugated frameworks, imide groups can significantly modulate electronic structures and enhance properties such as fluorescence efficiency, charge transport, and chiroptical responses. In this section, we begin by summarizing representative imide-functionalized helicenes, highlighting their structural features and photophysical performances. In 2020, Ravat’s group introduced a novel class of helically chiral diimide molecules 54a–c, which integrate the favorable characteristics of arylene diimides within the chiral architecture of [n]helicenes [69]. These compounds exhibit varying PLQYs of 0.22, 0.02, and 0.12 for 54a, 54b, and 54c, respectively, and notably retain fluorescence in the solid state. The |gabs| in the visible region increase systematically with helical length, reaching values as high as ≈10−2 for compounds 54b and 54c – among the highest reported to date – highlighting their strong potential in chiral optoelectronic applications (Table 18). In 2023, the same group reported a stable push–pull [7]helicene diimide (compound 55) that exhibited notable chiroptical performance, with |gabs| and |glum| values of 1.12 × 10−2 and 5.0 × 10−3, respectively, in toluene [70]. Furthermore, compound 55 demonstrated solvent-dependent fluorescence and CPL behavior across the visible spectrum, with both emission intensity and chiroptical properties varying in response to solvent polarity. Concurrently, Würthner’s group developed two naphthalimide-annulated [n]helicenes, compounds 56a and 56b (n = 5, 6), via a concise two-step synthetic route that afforded excellent yields and notable photophysical properties [71]. Both helicenes display high ΦF as 73% for 56a and 69% for 56b. Notably, compound 56b exhibits markedly enhanced |gabs| and |glum| values of 2.1 × 10−3 and 2.3 × 10−3, approximately 4.5-fold greater than that of compound 56a. Its red CPL emission at 615 nm and high BCPL of 66.5 M−1 cm−1 underscore its potential for advanced chiral photonic applications.
Table 18:
Structures and optical properties of 54a–c, 55 and 56a,b.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
54a
417, 442
471, 499
0.22a
0.17b
7 × 10−3
–
–
54b
395
470, 498
0.02a
0.02b
1.75 ×10−2
–
–
54c
452
508
0.12a
0.06b
1.22 × 10−2
–
–
55c
408
532
0.26
8.6 × 10−3
4.2 × 10−3
7.8
56a
629
655
0.73
4.5 × 10−4
5.0 × 10−4
22.0
56b
588
613
0.69
2.1 × 10−3
2.3 × 10−3
66.5
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
56b
618
–
50
–
2.3
aAs detected in solution; bas detected in the solid state; call detected in DCM.
Heteroatom engineering in double helicenes has emerged as a powerful strategy for tuning chiroptical properties and excited-state dynamics. In 2021, Sakamaki’s group synthesized a novel double N,O-hetero[5]helicene (compound 57b) by coupling two 12H-benzo[b]phenoxazine (BPO) units and systematically compared it to its N,N-analogue (compound 57a) derived from 13H-dibenzo[b,i]phenoxazine (DBPO) scaffolds [72] (Table 19). Compound 57b was obtained in significantly higher yield and, like compound 57a, exhibited electron-rich character and compact molecular packing, both favorable for p-type transistor performance. Importantly, both helicenes displayed strong CPL in CH2Cl2, with |glum| values exceeding 10−2. Intriguingly, the CPL signals of the two compounds exhibited opposite signs, underscoring the sensitivity of chiral excited-state properties to heteroatom substitution within the helicene framework. Extending this design principle, the group reported a double N,S-hetero[5]helicene 58 constructed from two benzo[b]phenothiazine units in 2023 [73]. Compared to the N,O-analogue 57b, this new compound showed more intense phosphorescence and an extended emission lifetime in dilute solution. Notably, it demonstrated room-temperature dual-emission CPL originating from both prompt fluorescence and long-lived phosphorescence, a rare feature in helicene systems. In a subsequent study, the same group reported a bis(N,Se)-hetero[4]helicene 59b and systematically compared its structural and dynamic properties with those of its sulfur analogue 59a[74]. Despite their close structural resemblance, the longer C–Se bond in 59b led to a markedly higher racemization barrier (145.7 vs 112.8 kJ/mol), thereby illustrating how subtle atomic substitutions can significantly influence the conformational stability of helical molecules (Table 19). These studies illustrate how precise heteroatom modulation enables fine control over CPL directionality and emission lifetimes, offering promising avenues for the development of multifunctional chiral optoelectronic materials – particularly those capable of simultaneous fluorescence and phosphorescence-based CPL.
Table 19:
Structures and optical properties of 57a,b, 58, and 59a,b.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
57a
≈410
569
0.038
1.7 × 10−2
2.3 × 10−2
57b
≈380
587
0.035
1.3 × 10−2
1.3 × 10−2
58
≈390
547
0.003
0.30a
2.0 × 10−2
1.7 × 10−2 b
59a
380
–
–
–
–
59b
380
–
–
–
–
aPhosphorescence quantum yield ФP; bdoped in β-estradiol matrix.
Recently, thiadiazole-fused helicenes have gradually come into our view. In 2023, Hirose’s group synthesized a series of tetraazadithia[n]helicenes – 60a, 60b, and 60c – featuring 2,1,3-thiadiazole termini [75] (Table 20). Among them, compound 60c exhibited pronounced CPL activity in toluene (|glum| = 0.04, ΦF = 3%), demonstrating the efficacy of terminal heterocycle incorporation for boosting chiroptical performance. In 2024, Babu and co-workers developed two π-extended hetero[6]helicenes – 61a and 61b – incorporating thiadiazole and selenadiazole moieties, respectively [76]. Substitution of sulfur with selenium enhanced intermolecular interactions and led to a notable reduction in the optical bandgap, highlighting the effectiveness of heteroatom modulation in tuning the electronic and photophysical properties of chiral nanographenes. These studies exemplify how strategic structural and electronic design – through π-extension, end-group heteroatom engineering, and atom-specific substitutions – enables precise tuning of chiroptical and photophysical properties in helicene-based materials, advancing their applicability in next-generation optoelectronic devices.
Table 20:
Structures and optical properties of 60a–c and 61a,b.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
BCPL [M−1 cm−1]
60a
391
398
0.005
–
–
–
60b
431
450
0.008
1.5 × 10−2
1.0 × 10−2
2
60c
445
483
0.027
3.7 × 10−2
4.0 × 10−2
15
61a
340
536
0.0735
–
–
–
61b
349
556
0.009
–
–
–
In 2020, Pittelkow’s group developed a unique synthetic strategy that converts a non-planar hetero[7]helicene into a planar hetero[8]circulene featuring an antiaromatic cyclooctatetraene (COT) core (62a–f) [77] (Table 21). Through controlled oxidation of the thiophene units to sulfones, they achieved a systematic red-shift in both absorption and emission spectra. Remarkably, the emission of these derivatives spans nearly the entire visible spectrum. These studies provide innovative molecular design strategies for constructing helically twisted or planarized chiral π-conjugated systems with tunable optical properties, thereby paving the way for the development of multifunctional materials in advanced photonic and electronic technologies.
Table 21:
Structure and optical properties of 62a–f.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
62a
388
429
0.08
–
–
62b
419
484
0.25
–
–
62c
431
518
0.14
–
–
62d
476
574
0.13
–
–
62e
414
436
0.06
–
–
62f
473
485
0.12
–
–
In 2021, Viglianisi’s group synthesized a series of thia-bridged triarylamine[4]helicene-functionalized polynorbornenes 63a–c via ring-opening metathesis polymerization (ROMP), introducing helicene chirality into polymer backbones with tunable electrochromic behavior [78]. These polymers exhibit reversible pH-responsive color changes. For instance, 63a transitions from pale yellow to deep blue in the solid state upon exposure to TFA, while 63b and 63c in CH2Cl2 exhibit new absorption bands at 570 and 575 nm, respectively – reversibly decolorized upon triethylamine treatment (Table 22). This work demonstrates the potential of helicene-containing polymers as stimuli-responsive chiral electrochromic materials. In the same year, You’s group developed a transition-metal-catalyzed C–H/C–H-type regioselective C3-arylation of benzothiophenes using molecular oxygen as the oxidant [79]. This strategy afforded the TADF-active compound 64a, which exhibits efficient blue emission and excellent OLED performance with a maximum EQE of 25.4%. This example highlights the utility of helicene-related heteroaromatic frameworks in the design of high-efficiency emissive materials. Also in 2021, Ema’s group reported a concise Scholl-type cyclodehydrogenation strategy for synthesizing azahelicenes and diaza[8]circulenes 65a–d[24] (Table 22). These molecules exhibited distinct Cotton effects and CPL, with |glum| reaching up to 1.6 × 10−3. This approach offers a generalizable route to structurally diverse chiral polycyclic aromatic hydrocarbons (PAHs) with strong chiroptical responses. Concurrently, Tanaka’s group achieved the enantioselective synthesis of aza[6]- and aza[7]helicene-like molecules via Rh(I)/chiral bisphosphine-catalyzed [2 + 2 + 2] cycloaddition [80]. The resulting S-shaped double aza[6]helicene-like compound 66 displayed high enantiomeric excess (up to 89% ee), pronounced chiroptical activity (|gabs| = 0.0054–0.0056), and substantial ΦF of 0.21–0.32 under both neutral and acidic conditions. This work exemplifies the power of transition-metal catalysis for constructing enantioenriched helicenes with tunable photophysical properties. These contributions from 2021 underscore the synthetic versatility and functional diversity of helicene-based systems, spanning electrochromism, thermally activated delayed fluorescence, and circularly polarized luminescence. Such structural innovations provide valuable frameworks for the development of next-generation chiral optoelectronic materials.
Table 22:
Structures and optical properties of 63a–c, 64a,b, 65a–d, and 66.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
63a
–
–
–
–
–
63b
570
–
–
–
–
63c
575
–
–
–
–
64a
376
456
–
–
–
64b
360
456
–
–
–
65a
401
420, 441
0.30
9.2 × 10−4
7.2 × 10−4
65b
414
432, 457
0.08
1.6 × 10−3
1.1 × 10−3
65c
440
493
0.10
7.3 × 10−4
2.6 × 10−4
65d
420
554
0.02
–
–
66
388, 431
489
0.21
5.59 × 10−3
1.42 × 10−3
66 (+TFA)
290, 389, 439
555
0.32
4.98 × 10−3
1.38 × 10−3
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
64a
474
–
–
(0.15, 0.23)
25.4
In 2022, Furuta’s group developed a one-pot synthetic protocol to access (NH)-phenanthridinone derivatives and chiral amide-functionalized [7]helicene-like molecules 67a,b from biaryl dicarboxylic acids, employing a Curtius rearrangement followed by basic hydrolysis [81] (Table 23). Notably, when chalcogen-containing substrates were used, the process afforded phosphorus ester derivatives of aza[5]helicenes. The chiral nature of the products was confirmed by optical rotation and CD measurements. In parallel, Soni’s group established an efficient three-step synthesis of coumarin-containing hetero[5]- and [6]helicene-like structures 68a–g in high yields [82]. These compounds display diverse photophysical behaviors: compound 68d emits yellow fluorescence in both solution and solid state, exhibiting solvatofluorochromism due to a twisted intramolecular charge transfer (TICT) mechanism, while compound 68e emits blue light (ΦF = 0.37) and demonstrates pronounced AIE in the solid state. Concurrently, Jiang’s group reported 69b, the first hetero[4]helicene-type molecule exhibiting both CPL and TADF [83]. This compound displays a high ΦF of 0.51 and a |glum| of 1.2 × 10−3. OLED devices fabricated using 69b emit sky-blue light with a peak EQE of 10.6% and |gEL| values up to 1.6 × 10−3. Collectively, these studies demonstrate the versatility of helicene-inspired architectures for constructing multifunctional chiral optoelectronic materials, highlighting their growing relevance in next-generation circularly polarized OLED technologies.
Table 23:
Structures and optical properties of 67a,b, 68a–g, and 69a,b.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
67a
–
–
–
–
–
67b
–
–
–
–
–
68a
295
411
0.08
–
–
68b
309
422
0.10
–
–
68c
328
439
0.03
–
–
68d
394
514
0.22
–
–
68e
320
423
0.37
–
–
68f
318
389
0.01
–
–
68g
317
411
0.04
–
–
69a
397
431
–
–
–
69b
400
446
0.51
–
1.2 × 10−3
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
69b
488
1.6 ×10−3
72
(0.17, 0.34)
10.6
Takizawa and co-workers have pioneered electrochemical strategies for synthesizing structurally diverse hetero[7]helicenes with tunable chiroptical properties and excellent configurational stability. In 2022, they introduced two electrochemical routes to construct aza-oxa-dehydro[7]helicenes, yielding helicenes with high racemization barriers and notable chiral stability [84]. The quasicirculenes 70a and 70b demonstrated strong blue CPL activity, with |glum| values of 2.5 × 10−3 at 433 nm and 2.4 × 10−3 at 418 nm, respectively (Table 24). Building on this, the team achieved the enantioselective synthesis of heterodehydrospiroenes on a gram scale using chiral vanadium(V) complexes – marking a significant advancement in asymmetric electrochemical catalysis. In a complementary study that same year, they reported a two-step electrochemical synthesis of a double aza-oxa[7]helicene via oxidative coupling followed by dehydrative cyclization [85]. The resulting meso-isomer (P,M)-71 emerged as the major product, exhibiting dual emission bands at 415 and 440 nm and solvent-independent absorption at 407 nm. Expanding the structural diversity, the group developed a two-pot synthesis of unsymmetrical hetero[7]helicenes 72a–g in 2023 [86], employing p-benzoquinone and N-aryl-2-naphthylamines through acid-promoted cyclization followed by electrochemical domino reactions. This method produced six compounds with yields ranging from 33–45%, all featuring extended π-conjugation and distinct photophysical characteristics. Furthermore, they established a mild electrochemical protocol for synthesizing oxaza[7]helicenes incorporating pyrrole and furan units [87]. This method afforded products in 50–86% yield with Faradaic efficiencies up to 77%. Among them, derivative 73 exhibited CPL activity (|glum| = 3.0 × 10−4), showcasing the ability to modulate chiroptical responses via heteroatom integration. These studies underscore the versatility of electrochemical synthesis in enabling precise structural modulation of heterohelicenes, facilitating access to high-performance chiral optoelectronic materials.
Table 24:
Structures and optical properties of 70a,b, 71, 72a–g, and 73.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
70a
402
433
0.25
–
2.5 × 10−3
70b
–
418
0.16
–
2.4 × 10−3
71
407
415, 440
–
–
–
72a
406
439
–
–
–
72b
403
440
0.065
–
–
72c
402
440
–
–
–
72d
413
450
–
–
–
72e
401
440
–
–
–
72g
405
440
–
–
–
73
–
–
–
–
3.0 × 10−4
In 2023, Zhang’s group introduced a new class of helically chiral double hetero[4]helicenes 74a and 74b exhibiting CP-TADF, constructed on a distinct donor–acceptor core architecture [88] (Table 25). These compounds demonstrate excellent configurational stability and robust CPL signals both in solution and in solid-state films, with a |glum| of 3.1 × 10−3. Corresponding CP-OLEDs based on compound 74a achieved outstanding device performance, reaching a maximum EQE of 20.03% and a |gEL| of 2.9 × 10−3 – underscoring their considerable potential for advanced chiral optoelectronic applications. Building upon this framework, in 2024, the same group developed a novel cove-region bridging strategy to construct double hetero[4]helicenes with enhanced structural rigidity and persistent chirality [89]. By selectively modifying the bay regions of the SPZ (spiro[fluorene-9,9'-xanthene]) scaffold, they successfully converted initially non-emissive helicenes into efficient TADF luminophores with tunable emission wavelengths ranging from sky-blue to deep red. Particularly, the enantiomeric forms of the 75b derivatives emerged as rare examples of red-emissive CPL materials. This innovative design approach offers a versatile and modular platform for engineering chiral multi-helicene systems with customizable optoelectronic properties, paving the way for their deployment in next-generation CPL-active materials and high-performance CP-OLED devices.
Table 25:
Structures and optical properties of 74a,b and 75a–c.
compound
λabs(max) [nm]
λem(max) [nm]
ФF
|gabs|
|glum|
74a
406
493
0.13/0.67a
–
3.1 × 10−3,a
74b
357
450
0.07/0.22a
–
–
75a
612
–
–
–
–
75b
495
656
0.02
–
2.7 × 10−3
75c
436
480
0.09
–
2.5 × 10−2
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
(M,M)-74a
500
2.9 ×10−3
82
(0.24, 0.50)
20.03
rac-74a
500
–
81
(0.24, 0.49)
20.00
aDetected as 20 wt % doped films with the mCBP host.
In 2024, Jančařík and co-workers introduced an intramolecular radical cyclization strategy to synthesize highly luminescent tetraceno[6]helicenone and its aza analogue 76[90] (Table 26). The incorporation of a carbonyl group into the helicene backbone substantially enhanced fluorescence quantum yields and red-shifted the emission into the visible region. The aza analogue demonstrated promising performance in OLEDs, confirming its potential for optoelectronic applications. Concurrently, Shirinian’s group synthesized a series of nitrogen-functionalized quinoline (NFQ)-based aza-oxa[5]helicenes 77a–f exhibiting excellent UV stability and solvent-dependent fluorescence [91]. Protonation significantly enhanced their emission intensity, and the presence of nitrogen facilitated further structural derivatization. In the same year, Alcarazo’s group reported an enantioselective gold-catalyzed synthesis of compound 78, achieving a high enantiomeric excess [92]. They further investigated various post-synthetic modification strategies, demonstrating their potential for application in chiral photonic materials. Collectively, these advances underscore the power of structural tailoring, heteroatom incorporation, and enantioselective strategies in finely tuning the photophysical and chiroptical properties of helicenes, providing a versatile foundation for the development of high-performance chiral optoelectronic materials.
Table 26:
Structures and optical properties of 76, 77a–f, and 78.a
compound
λabs(max) [nm]
λem(max) [nm]
ФF
76
483
561
0.43
77a in CHCl3
352
379, 399
0.39
77b in CHCl3
359
379, 392
0.04
77c in CHCl3
360
397
0.08
77d in CHCl3
362
390, 403
0.09
77a in heptane
347
388
0.21
77e in heptane
348
391
0.20
77f in heptane
348, 358
383
0.19
77a in toluene
352
394, 421
0.56
77e in toluene
353
380, 400
0.44
77f in toluene
353
388
0.28
77a in acetonitrile
348
375
0.48
77e in acetonitrile
348
383
0.48
77f in acetonitrile
349
391
0.42
77a in methanol
351
383
0.48
77e in methanol
349
391
0.47
77f in methanol
352
396
0.27
device
λEL(max) [nm]
|gEL|
FWHM [nm]
CIE coordinate
EQEmax [%]
76
580
–
103
–
0.15
76:MADN 95:5
550
–
93
–
0.7
aNo gabs or glum values were reported, no optical characterization for 78.
Conclusion
Nitrogen-doped helicenes and their heteroatom co-doped analogues constitute a rapidly advancing class of chiral π-conjugated materials, distinguished by exceptional structural tunability, photophysical diversity, and chiroptical functionality. The integration of nitrogen – and its synergistic pairing with heteroatoms such as boron, oxygen, sulfur, and selenium – has significantly expanded the molecular design space, enabling precise control over redox behavior, emission wavelength, CPL, and responsiveness to thermal or redox stimuli. These heteroatom modifications have led to remarkable breakthroughs, including near-unity PLQYs, ultranarrow emission bands, |glum| values exceeding 10−3, and unprecedented BCPL, particularly in the visible to near-infrared (NIR) spectral regions.
Recent advances in synthetic methodology – including electrochemical, Scholl-type, and enantioselective catalytic strategies – have further enabled access to structurally complex helicene topologies with enhanced configurational stability and integrated multifunctionality. These developments have facilitated a growing range of applications in CP-OLEDs, molecular sensing, chiral switches, and photonic devices. Moving forward, key challenges remain, such as mitigating spectral broadening in red/NIR emission, enhancing the chemical and photostability of electron-deficient helicenes, and developing sustainable, scalable synthetic approaches. The integration of computational design with multifunctional molecular engineering is expected to accelerate the deployment of helicene-based materials in next-generation technologies spanning chiral optoelectronics, bioimaging, spintronics, and quantum information science.
Funding
Prof. Dr. H.-Y. Gong acknowledges the financial support provided by the National Natural Science Foundation of China (Grant No. 92156009). Nai-Te Yao appreciates for the financial support from the Interdisciplinary Research Foundation for Doctoral Candidates of Beijing Normal University (Grant No. BNUXKJC2407).
Author Contributions
Meng Qiu: resources; writing – original draft. Jing Du: writing – original draft. Nai-Te Yao: funding acquisition; writing – original draft. Xin-Yue Wang: writing – original draft. Han-Yuan Gong: conceptualization; funding acquisition; writing – review & editing.
Data Availability Statement
Data sharing is not applicable as no new data was generated or analyzed in this study.
References
Chen, J.; Captain, B.; Takenaka, N. Org. Lett.2011,13, 1654–1657. doi:10.1021/ol200102c
Return to citation in text:
[1]
Peng, Z.; Takenaka, N. Chem. Rec.2013,13, 28–42. doi:10.1002/tcr.201200010
Return to citation in text:
[1]
Zhang, G.; Zhang, J.; Tao, Y.; Gan, F.; Lin, G.; Liang, J.; Shen, C.; Zhang, Y.; Qiu, H. Nat. Commun.2024,15, 5469. doi:10.1038/s41467-024-49865-y
Return to citation in text:
[1]
Isla, H.; Saleh, N.; Ou-Yang, J.-K.; Dhbaibi, K.; Jean, M.; Dziurka, M.; Favereau, L.; Vanthuyne, N.; Toupet, L.; Jamoussi, B.; Srebro-Hooper, M.; Crassous, J. J. Org. Chem.2019,84, 5383–5393. doi:10.1021/acs.joc.9b00389
Return to citation in text:
[1]
Meng, D.; Fu, H.; Xiao, C.; Meng, X.; Winands, T.; Ma, W.; Wei, W.; Fan, B.; Huo, L.; Doltsinis, N. L.; Li, Y.; Sun, Y.; Wang, Z. J. Am. Chem. Soc.2016,138, 10184–10190. doi:10.1021/jacs.6b04368
Return to citation in text:
[1]
Jakubec, M.; Storch, J. J. Org. Chem.2020,85, 13415–13428. doi:10.1021/acs.joc.0c01837
Return to citation in text:
[1]
Wang, X.-Y.; Yao, X.; Narita, A.; Müllen, K. Acc. Chem. Res.2019,52, 2491–2505. doi:10.1021/acs.accounts.9b00322
Return to citation in text:
[1]
Feng, Y.; Xu, Y.; Qu, C.; Wang, Q.; Ye, K.; Liu, Y.; Wang, Y. Adv. Mater. (Weinheim, Ger.)2024,36, 2403061. doi:10.1002/adma.202403061
Return to citation in text:
[1]
Zhu, D.; Jiang, W.; Ma, Z.; Feng, J.; Zhan, X.; Lu, C.; Liu, J.; Liu, J.; Hu, Y.; Wang, D.; Zhao, Y. S.; Wang, J.; Wang, Z.; Jiang, L. Nat. Commun.2022,13, 3454. doi:10.1038/s41467-022-31186-7
Return to citation in text:
[1]
Kundu, D.; Rio, N. D.; Crassous, J. Acc. Chem. Res.2024,57, 2941–2952. doi:10.1021/acs.accounts.4c00275
Return to citation in text:
[1]
Dhbaibi, K.; Favereau, L.; Crassous, J. Chem. Rev.2019,119, 8846–8953. doi:10.1021/acs.chemrev.9b00033
Return to citation in text:
[1]
Nowak-Król, A.; Geppert, P. T.; Naveen, K. R. Chem. Sci.2024,15, 7408–7440. doi:10.1039/d4sc01083c
Return to citation in text:
[1]
Maeda, C.; Ema, T. Chem. Commun.2025,61, 4757–4773. doi:10.1039/d4cc06307d
Return to citation in text:
[1]
Reger, D.; Haines, P.; Amsharov, K. Y.; Schmidt, J. A.; Ullrich, T.; Bönisch, S.; Hampel, F.; Görling, A.; Nelson, J.; Jelfs, K. E.; Guldi, D. M.; Jux, N. Angew. Chem., Int. Ed.2021,60, 18073–18081. doi:10.1002/anie.202103253
Return to citation in text:
[1]
Qiu, S.; Valdivia, A. C.; Zhuang, W.; Hung, F.-F.; Che, C.-M.; Casado, J.; Liu, J. J. Am. Chem. Soc.2024,146, 16161–16172. doi:10.1021/jacs.4c03815
Return to citation in text:
[1]
Petdum, A.; Kaewnok, N.; Panchan, W.; Sahasithiwat, S.; Sooksimuang, T.; Sirirak, J.; Chaiyaveij, D.; Wanichacheva, N. J. Mol. Struct.2021,1245, 131250. doi:10.1016/j.molstruc.2021.131250
Return to citation in text:
[1]
Maeda, C.; Nomoto, S.; Akiyama, K.; Tanaka, T.; Ema, T. Chem. – Eur. J.2021,27, 15699–15705. doi:10.1002/chem.202102269
Return to citation in text:
[1]
[2]
Maeda, C.; Yasutomo, I.; Ema, T. Angew. Chem., Int. Ed.2024,63, e202404149. doi:10.1002/anie.202404149
Return to citation in text:
[1]
Wu, Y.-F.; Ying, S.-W.; Su, L.-Y.; Du, J.-J.; Zhang, L.; Chen, B.-W.; Tian, H.-R.; Xu, H.; Zhang, M.-L.; Yan, X.; Zhang, Q.; Xie, S.-Y.; Zheng, L.-S. J. Am. Chem. Soc.2022,144, 10736–10742. doi:10.1021/jacs.2c00794
Return to citation in text:
[1]
Váňa, L.; Jakubec, M.; Sýkora, J.; Císařová, I.; Žádný, J.; Storch, J.; Církva, V. J. Org. Chem.2022,87, 7150–7166. doi:10.1021/acs.joc.2c00375
Return to citation in text:
[1]
Gong, X.; Li, C.; Cai, Z.; Wan, X.; Qian, H.; Yang, G. J. Org. Chem.2022,87, 8406–8412. doi:10.1021/acs.joc.2c00371
Return to citation in text:
[1]
Hong, J.; Xiao, X.; Liu, H.; Dmitrieva, E.; Popov, A. A.; Yu, Z.; Li, M.-D.; Ohto, T.; Liu, J.; Narita, A.; Liu, P.; Tada, H.; Cao, X.-Y.; Wang, X.-Y.; Zou, Y.; Müllen, K.; Hu, Y. Chem. – Eur. J.2022,28, e202202243. doi:10.1002/chem.202202243
Return to citation in text:
[1]
Gan, F.; Zhang, G.; Liang, J.; Shen, C.; Qiu, H. Angew. Chem., Int. Ed.2024,63, e202320076. doi:10.1002/anie.202320076
Return to citation in text:
[1]
Gross, M.; Zhang, F.; Arnold, M. E.; Ravat, P.; Kuehne, A. J. C. Adv. Opt. Mater.2024,12, 2301707. doi:10.1002/adom.202301707
Return to citation in text:
[1]
Borstelmann, J.; Schneider, L.; Rominger, F.; Deschler, F.; Kivala, M. Angew. Chem., Int. Ed.2024,63, e202405570. doi:10.1002/anie.202405570
Return to citation in text:
[1]
Borstelmann, J.; Zank, S.; Krug, M.; Berger, G.; Fröhlich, N.; Glotz, G.; Gnannt, F.; Schneider, L.; Rominger, F.; Deschler, F.; Clark, T.; Gescheidt, G.; Guldi, D. M.; Kivala, M. Angew. Chem., Int. Ed.2025,64, e202423516. doi:10.1002/anie.202423516
Return to citation in text:
[1]
Matsuo, Y.; Gon, M.; Tanaka, K.; Seki, S.; Tanaka, T. J. Am. Chem. Soc.2024,146, 17428–17437. doi:10.1021/jacs.4c05156
Return to citation in text:
[1]
Matsuo, Y.; Gon, M.; Tanaka, K.; Seki, S.; Tanaka, T. Chem. – Asian J.2024,19, e202400134. doi:10.1002/asia.202400134
Return to citation in text:
[1]
Matsuo, Y.; Seki, S.; Tanaka, T. Chem. Lett.2024,53, upae159. doi:10.1093/chemle/upae159
Return to citation in text:
[1]
Kusy, D.; Górski, K.; Bertocchi, F.; Galli, M.; Vanthuyne, N.; Terenziani, F.; Gryko, D. T. Chem. – Eur. J.2025,31, e202404632. doi:10.1002/chem.202404632
Return to citation in text:
[1]
Maeda, C.; Nagahata, K.; Shirakawa, T.; Ema, T. Angew. Chem., Int. Ed.2020,59, 7813–7817. doi:10.1002/anie.202001186
Return to citation in text:
[1]
Maeda, C.; Michishita, S.; Yasutomo, I.; Ema, T. Angew. Chem., Int. Ed.2025,64, e202418546. doi:10.1002/anie.202418546
Return to citation in text:
[1]
Oda, S.; Kawakami, B.; Yamasaki, Y.; Matsumoto, R.; Yoshioka, M.; Fukushima, D.; Nakatsuka, S.; Hatakeyama, T. J. Am. Chem. Soc.2022,144, 106–112. doi:10.1021/jacs.1c11659
Return to citation in text:
[1]
Zhang, Y.; Zhang, D.; Huang, T.; Gillett, A. J.; Liu, Y.; Hu, D.; Cui, L.; Bin, Z.; Li, G.; Wei, J.; Duan, L. Angew. Chem., Int. Ed.2021,60, 20498–20503. doi:10.1002/anie.202107848
Return to citation in text:
[1]
Li, J.-K.; Chen, X.-Y.; Guo, Y.-L.; Wang, X.-C.; Sue, A. C.-H.; Cao, X.-Y.; Wang, X.-Y. J. Am. Chem. Soc.2021,143, 17958–17963. doi:10.1021/jacs.1c09058
Return to citation in text:
[1]
Meng, G.; Zhou, J.; Han, X.-S.; Zhao, W.; Zhang, Y.; Li, M.; Chen, C.-F.; Zhang, D.; Duan, L. Adv. Mater. (Weinheim, Ger.)2024,36, 2307420. doi:10.1002/adma.202307420
Return to citation in text:
[1]
Yang, W.; Li, N.; Miao, J.; Zhan, L.; Gong, S.; Huang, Z.; Yang, C. CCS Chem.2022,4, 3463–3471. doi:10.31635/ccschem.022.202101661
Return to citation in text:
[1]
Ye, Z.; Wu, H.; Xu, Y.; Hua, T.; Chen, G.; Chen, Z.; Yin, X.; Huang, M.; Xu, K.; Song, X.; Huang, Z.; Lv, X.; Miao, J.; Cao, X.; Yang, C. Adv. Mater. (Weinheim, Ger.)2024,36, 2308314. doi:10.1002/adma.202308314
Return to citation in text:
[1]
Zhang, X.; Rauch, F.; Niedens, J.; da Silva, R. B.; Friedrich, A.; Nowak-Król, A.; Garden, S. J.; Marder, T. B. J. Am. Chem. Soc.2022,144, 22316–22324. doi:10.1021/jacs.2c10865
Return to citation in text:
[1]
Huang, T.; Yuan, L.; Lu, X.; Qu, Y.; Qu, C.; Xu, Y.; Zheng, Y.-X.; Wang, Y. Chem. Sci.2024,15, 15170–15177. doi:10.1039/d4sc03854a
Return to citation in text:
[1]
Cheng, H.; Lan, J.; Yang, Y.; Bin, Z. Mater. Horiz.2024,11, 4674–4680. doi:10.1039/d4mh00634h
Return to citation in text:
[1]
Guo, W.-C.; Zhao, W.-L.; Tan, K.-K.; Li, M.; Chen, C.-F. Angew. Chem., Int. Ed.2024,63, e202401835. doi:10.1002/anie.202401835
Return to citation in text:
[1]
Ju, Y.-Y.; Xie, L.-E.; Xing, J.-F.; Deng, Q.-S.; Chen, X.-W.; Huang, L.-X.; Nie, G.-H.; Tan, Y.-Z.; Zhang, B. Angew. Chem., Int. Ed.2025,64, e202414383. doi:10.1002/anie.202414383
Return to citation in text:
[1]
Liu, M.; Li, C.; Liao, G.; Zhao, F.; Yao, C.; Wang, N.; Yin, X. Chem. – Eur. J.2024,30, e202402257. doi:10.1002/chem.202402257
Return to citation in text:
[1]
Yu, Y.; Wang, C.; Hung, F.-F.; Chen, C.; Pan, D.; Che, C.-M.; Liu, J. J. Am. Chem. Soc.2024,146, 22600–22611. doi:10.1021/jacs.4c06997
Return to citation in text:
[1]
Yu, Y.; Wang, C.; Hung, F.-F.; Jiang, L.; Che, C.-M.; Liu, J. Angew. Chem., Int. Ed.2025,64, e202501645. doi:10.1002/anie.202501645
Return to citation in text:
[1]
Saal, F.; Zhang, F.; Holzapfel, M.; Stolte, M.; Michail, E.; Moos, M.; Schmiedel, A.; Krause, A.-M.; Lambert, C.; Würthner, F.; Ravat, P. J. Am. Chem. Soc.2020,142, 21298–21303. doi:10.1021/jacs.0c11053
Return to citation in text:
[1]
Saal, F.; Swain, A.; Schmiedel, A.; Holzapfel, M.; Lambert, C.; Ravat, P. Chem. Commun.2023,59, 14005–14008. doi:10.1039/d3cc04470j
Return to citation in text:
[1]
Tian, X.; Shoyama, K.; Mahlmeister, B.; Brust, F.; Stolte, M.; Würthner, F. J. Am. Chem. Soc.2023,145, 9886–9894. doi:10.1021/jacs.3c03441
Return to citation in text:
[1]
Sakamaki, D.; Tanaka, S.; Tanaka, K.; Takino, M.; Gon, M.; Tanaka, K.; Hirose, T.; Hirobe, D.; Yamamoto, H. M.; Fujiwara, H. J. Phys. Chem. Lett.2021,12, 9283–9292. doi:10.1021/acs.jpclett.1c02896
Return to citation in text:
[1]
Tanaka, S.; Sakamaki, D.; Haruta, N.; Sato, T.; Gon, M.; Tanaka, K.; Fujiwara, H. J. Mater. Chem. C2023,11, 4846–4854. doi:10.1039/d3tc00871a
Return to citation in text:
[1]
Sakamaki, D.; Inoue, Y.; Shimomura, K.; Taura, D.; Yashima, E.; Seki, S. Tetrahedron Lett.2023,114, 154294. doi:10.1016/j.tetlet.2022.154294
Return to citation in text:
[1]
Zhang, Z.; Murata, Y.; Hirose, T. Tetrahedron2023,142, 133514. doi:10.1016/j.tet.2023.133514
Return to citation in text:
[1]
Kumar, V.; Dongre, S. D.; Venugopal, G.; Narayanan, A.; Babu, S. S. Chem. Commun.2024,60, 11944–11947. doi:10.1039/d4cc03707c
Return to citation in text:
[1]
Lousen, B.; Pedersen, S. K.; Bols, P.; Hansen, K. H.; Pedersen, M. R.; Hammerich, O.; Bondarchuk, S.; Minaev, B.; Baryshnikov, G. V.; Ågren, H.; Pittelkow, M. Chem. – Eur. J.2020,26, 4935–4940. doi:10.1002/chem.201905339
Return to citation in text:
[1]
Lupi, M.; Menichetti, S.; Stagnaro, P.; Utzeri, R.; Viglianisi, C. Synthesis2021,53, 2602–2611. doi:10.1055/s-0040-1706743
Return to citation in text:
[1]
Shi, Y.; Yang, G.; Shen, B.; Yang, Y.; Yan, L.; Yang, F.; Liu, J.; Liao, X.; Yu, P.; Bin, Z.; You, J. J. Am. Chem. Soc.2021,143, 21066–21076. doi:10.1021/jacs.1c11277
Return to citation in text:
[1]
Hanada, K.; Nogami, J.; Miyamoto, K.; Hayase, N.; Nagashima, Y.; Tanaka, Y.; Muranaka, A.; Uchiyama, M.; Tanaka, K. Chem. – Eur. J.2021,27, 9313–9319. doi:10.1002/chem.202005479
Return to citation in text:
[1]
Murai, T.; Xing, Y.; Kurokawa, M.; Kuribayashi, T.; Nikaido, M.; Elboray, E. E.; Hamada, S.; Kobayashi, Y.; Sasamori, T.; Kawabata, T.; Furuta, T. J. Org. Chem.2022,87, 5510–5521. doi:10.1021/acs.joc.1c02769
Return to citation in text:
[1]
Soni, R.; Soman, S. S. Asian J. Org. Chem.2022,11, e202100770. doi:10.1002/ajoc.202100770
Return to citation in text:
[1]
Khalid, M. I.; Salem, M. S. H.; Sako, M.; Kondo, M.; Sasai, H.; Takizawa, S. Commun. Chem.2022,5, 166. doi:10.1038/s42004-022-00780-7
Return to citation in text:
[1]
Salem, M. S. H.; Sabri, A.; Khalid, M. I.; Sasai, H.; Takizawa, S. Molecules2022,27, 9068. doi:10.3390/molecules27249068
Return to citation in text:
[1]
Salem, M. S. H.; Khalid, M. I.; Sasai, H.; Takizawa, S. Tetrahedron2023,133, 133266. doi:10.1016/j.tet.2023.133266
Return to citation in text:
[1]
Salem, M. S. H.; Khalid, M. I.; Sako, M.; Higashida, K.; Lacroix, C.; Kondo, M.; Takishima, R.; Taniguchi, T.; Miura, M.; Vo‐Thanh, G.; Sasai, H.; Takizawa, S. Adv. Synth. Catal.2023,365, 373–380. doi:10.1002/adsc.202201262
Return to citation in text:
[1]
Qu, C.; Zhu, Y.; Liang, L.; Ye, K.; Zhang, Y.; Zhang, H.; Zhang, Z.; Duan, L.; Wang, Y. Adv. Opt. Mater.2023,11, 2203030. doi:10.1002/adom.202203030
Return to citation in text:
[1]
Qu, C.; Xu, Y.; Wang, Y.; Nie, Y.; Ye, K.; Zhang, H.; Zhang, Z. Angew. Chem., Int. Ed.2024,63, e202400661. doi:10.1002/anie.202400661
Return to citation in text:
[1]
Sturm, L.; Banasiewicz, M.; Deperasinska, I.; Kozankiewicz, B.; Morawski, O.; Dechambenoit, P.; Bock, H.; Nagata, Y.; Salvagnac, L.; Séguy, I.; Šámal, M.; Jančařík, A. Chem. – Eur. J.2024,30, e202403482. doi:10.1002/chem.202403482
Return to citation in text:
[1]
Balakhonov, R. Y.; Gaeva, E. B.; Mekeda, I. S.; Dolotov, R. A.; Metelitsa, A. V.; Shirinian, V. Z. Dyes Pigm.2024,225, 112032. doi:10.1016/j.dyepig.2024.112032
Return to citation in text:
[1]
Fu, W.; Pelliccioli, V.; Casares-López, R.; Cuerva, J. M.; Simon, M.; Golz, C.; Alcarazo, M. CCS Chem.2024,6, 2439–2451. doi:10.31635/ccschem.024.202404088
Return to citation in text:
[1]
Sakamaki, D.; Tanaka, S.; Tanaka, K.; Takino, M.; Gon, M.; Tanaka, K.; Hirose, T.; Hirobe, D.; Yamamoto, H. M.; Fujiwara, H. J. Phys. Chem. Lett.2021,12, 9283–9292. doi:10.1021/acs.jpclett.1c02896
Salem, M. S. H.; Khalid, M. I.; Sako, M.; Higashida, K.; Lacroix, C.; Kondo, M.; Takishima, R.; Taniguchi, T.; Miura, M.; Vo‐Thanh, G.; Sasai, H.; Takizawa, S. Adv. Synth. Catal.2023,365, 373–380. doi:10.1002/adsc.202201262
Balakhonov, R. Y.; Gaeva, E. B.; Mekeda, I. S.; Dolotov, R. A.; Metelitsa, A. V.; Shirinian, V. Z. Dyes Pigm.2024,225, 112032. doi:10.1016/j.dyepig.2024.112032
Fu, W.; Pelliccioli, V.; Casares-López, R.; Cuerva, J. M.; Simon, M.; Golz, C.; Alcarazo, M. CCS Chem.2024,6, 2439–2451. doi:10.31635/ccschem.024.202404088
Zhang, X.; Rauch, F.; Niedens, J.; da Silva, R. B.; Friedrich, A.; Nowak-Król, A.; Garden, S. J.; Marder, T. B. J. Am. Chem. Soc.2022,144, 22316–22324. doi:10.1021/jacs.2c10865