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

Robust π-conjugated radical cations

Shilong Sua and Qian Miao*ab
aDepartment of Chemistry, The Chinese University of Hong Kong, Shatin, New Territories, Hong Kong, China. E-mail: miaoqian@cuhk.edu.hk
bState Key Laboratory of Synthetic Chemistry, The Chinese University of Hong Kong, Shatin, New Territories, Hong Kong, China

Received 1st September 2025 , Accepted 18th November 2025

First published on 19th November 2025


Abstract

π-Conjugated radical cations, open-shell species carrying a positive charge, serve as charge carriers in p-type organic semiconductors, underpinning the operation of various organic electronic devices. Although they are traditionally perceived as highly reactive and challenging to isolate under ambient conditions, recent advances in electronic and steric stabilization have enabled the isolation and full characterization of a number of robust π-conjugated radical cations. This review provides a comprehensive overview of fully characterized π-conjugated radical cations, with particular emphasis on species identified via single-crystal X-ray crystallography within the last two decades. We highlight structural features and stabilization strategies that enable ambient stability, and explore structure–property relationships critical to their application in organic electronic materials, indicating the potential to enhance material stability and improve device performance. Based on this analysis, we assess the current state of the field and outline promising future directions.


1. Introduction

π-Conjugated radical cations are open-shell species that carry a positive charge, typically generated by oxidation of a neutral, closed-shell π-conjugated system.1 They play a central role in organic electronic materials and devices because the doped states of p-type organic semiconductors are commonly associated with radical cations.2 As a result, most organic conductors rely on π-conjugated radical cation salts, and the operation of various organic electronic devices depends on the controlled generation of π-conjugated radical cations. This is well illustrated by early examples of conducting radical cation salts, as shown in Fig. 1. Oxidation of tetrathiafulvalene (TTF) with Cl2 results in the radical cation salt TTF˙+Cl,3 which exhibits a room-temperature conductivity of 0.27 ± 0.1 S cm−1 in microcrystalline form.4 Similarly, electrochemical oxidation of bis(ethylenedithiolo)tetrathiafulvalene (BEDT-TTF) in the presence of CuSCN produces the crystal of (BEDT-TTF)2˙+ Cu(SCN)2, in which two BEDT-TTF molecules dimerize through short S⋯S contacts and share a single unpaired electron. The resulting single crystal displays a room-temperature conductivity of 14 S cm−1 and becomes a superconductor at 10.4 K under ambient pressure.5 The third example is the perchlorate salt of the tetrakis(methylthio)pyrene (TMTP) radical cation, (TMTP)2˙+ClO4, which exhibits an exceptionally high room-temperature conductivity in the range of 156 to 667 S cm−1.6,7 Consistent with this behavior, the neutral TMTP molecule functions as an organic semiconductor, recently reported to exhibit a remarkably high hole mobility of 32 cm2 V−1 s−1 in single-crystal organic field-effect transistors (OFETs).8
image file: d5sc06726j-f1.tif
Fig. 1 Early examples of organic conductors and superconductors based on radical cation salts.

The unique electronic structures of π-conjugated radical cations—combining both radical and ionic characteristics—give rise not only to electrical conductivity9 but also to magnetic10 and optical properties11,12 due to their unpaired spin densities. These features open avenues for unconventional applications,13–16 such as doublet emission in organic light-emitting diodes (OLEDs) with 100% internal quantum efficiency.17 Despite their importance, π-conjugated radical cations typically exhibit high reactivity, which often results in difficulties in isolating them in pure form and limits their detailed structural and spectroscopic characterization. Furthermore, their paramagnetic nature renders NMR signals undetectable. Consequently, full characterization of these species typically requires single-crystal X-ray crystallography. Over the past two decades, advances in electronic and steric stabilization have enabled the isolation of a number of ambient-stable π-conjugated radical cations. This review highlights these advances and shows how the full characterization of these stable systems has revealed key structural motifs governing stability. These insights are crucial for the rational design of more robust, high-performance organic electronic materials.

While recent reviews have separately covered organic radicals and organic ions, they have not comprehensively addressed π-conjugated radical cations. Reviews on π-conjugated radicals have predominantly focused on neutral species,13,14,18–21 while those on hydrocarbon-based cations have overlooked their radical character.22 This review fills this gap by providing a comprehensive overview of fully characterized π-conjugated radical cations, with a particular emphasis on structures resolved by single-crystal X-ray crystallography. We examine key structural motifs that impart stability and highlight their applications in organic electronics. Finally, we assess the current state of the field and suggest future research directions. To systematically organize this body of work, we classify the discussed radical cations into three categories. Mono-radical cations refer to species containing a single unpaired electron and a single positive charge. Multiple radical cations are those that possess more than one unpaired electron and/or carry more than a single positive charge. Fractional radical cations describe systems in which the unpaired electrons and/or positive charges are shared by multiple π-conjugated molecules, resulting in a non-integer (fractional) distribution of spin or charge per molecule. In each category, π-conjugated radical cations are grouped according to the stabilization strategies employed.

2. Mono-radical cations

This section classifies mono-radical cations into three groups based on their structural features. When a mono-radical cation exhibits more than one such feature—for example, containing both a nitrogen and a sulfur atom—the structural element that contributes more significantly to the stabilization of the radical cation takes precedence in classification. Only in rare cases are radical cations with different structural features discussed together, as their conjugated backbones are closely related.

2.1 Conjugated hydrocarbons

Radical cations of conjugated hydrocarbons are usually stabilized by two strategies: steric protection and delocalization of spin density and charge. Steric protection involves introducing bulky substituents to kinetically protect the radical cation from dimerization or nucleophilic attacks, albeit at the cost of hindering or even blocking π–π interactions. Bulky groups employed in the reported radical cations of polycyclic arenes include tert-butyl and mesityl groups as well as bicyclo[2.2.2]octene and norbornene units, as demonstrated with Fig. 2. Hexa-peri-hexabenzocoronene (HBC) and quaterphenyl represent fragments of graphene and poly-p-phenylene, respectively, and compounds 1 and 2 are their tert-butylated derivatives, respectively. Chemical oxidation of 1 with Ag+(Al(OC(CF3)3)4) affords its radical cation, which forms a π-stacked dimer exhibiting short C⋯C distances ranging from 3.3 to 3.5 Å in the crystal of 1˙+(Al(OC(CF3)3)4).23 In comparison, electrochemical oxidation of 1 in the presence of tetrabutylammonium hexafluoroarsenate at −30 °C results in the crystal of (1)2˙+AsF6, where two molecules of 1 form a π-stacked dimer.24 Averagely, each molecule of 1 possesses half a radical and half a positive charge although bond length analysis of the crystal structure shows the two molecules in the dimer are different, suggesting the radical and positive charge are not evenly distributed among them. Oxidation of compound 2 with nitrosonium hexachloroantimonate (NO+SbCl6) under an argon atmosphere at −10 °C results in 2˙+. It crystallizes from solution at −30 °C as 2˙+SbCl6, where two molecules of 2 also form a π-stacked dimer.25 Notably, in the above radical cation salts, the π-stacked dimers of 1 and 2 are segregated by the counter anions or solvent molecules without further π–π interactions possibly due to steric hindrance of the bulky substituents.
image file: d5sc06726j-f2.tif
Fig. 2 Sterically protected conjugated hydrocarbons whose radical cations were fully characterized.

The radical cations of compounds 3[thin space (1/6-em)]26 and 4[thin space (1/6-em)]27 are sterically shielded by bulky substituents: mesityl and 2,6-dimethyl-4-tert-butylphenyl groups, respectively. Notably, these substituents are strategically positioned on carbon atoms with the highest spin density, ensuring efficient protection of the radical centre. For dibenzoperopyrene 3, the radical cation is further stabilized through delocalization of both positive charge and unpaired spin across the polycyclic framework. This stabilization is enhanced by global aromaticity, as evidenced by diatropic ring currents observed along the edges of the polycyclic system in the anisotropy of the induced current density (ACID) plot. As a result, the radical cation salt 3˙+SbCl6 exhibits high air stability in the solid state, remaining intact for at least two weeks without significant degradation. Compound 4 similarly benefits from delocalization of its unpaired spin and positive charge. Structural analysis comparing the neutral and radical cation forms of 4 in the crystal reveals that the external naphthalene units undergo notable changes in bond lengths due to charge redistribution.

In 1993, Komatsu et al. demonstrated that the radical cation of cyclooctatetraene could be kinetically stabilized in its sterically congested derivative 5a (Fig. 2), which features four fused bicyclo[2.2.2]octene units.28 The single crystal of 5a˙+SbCl6 remains stable for several hours under ambient conditions and can be stored under refrigeration for months without decomposition. Employing the same steric protection strategy, Komatsu et al. later synthesized bicyclo[2.2.2]octene-fused derivatives of naphthalene (5b in Fig. 2), biphenylene (5c in Fig. 2) and anthracene. These radical cations were crystallized as hexachloroantimonate salts, which are persistent at ambient temperature even in air. Notably, crystal structural analysis of 5c reveals that oxidation of biphenylene to its radical cation form leads to a significant shortening (by 0.042 Å) of the bonds connecting the two benzene rings, consistent with the reduction of destabilizing 4π antiaromaticity in the neutral state. Similarly, Kochi et al. reported compound 6, a naphthalene derivative sterically shielded by fusing with tetramethylcyclohexane moieties.29 The corresponding radical cation, 6˙+, was isolated as 6˙+SbCl6 crystals. Interestingly, cocrystallization of 6˙+SbCl6 with unsubstituted naphthalene (naph) in CH2Cl2 solution afforded dark-green crystals of (6˙+)2(naph)(SbCl6)2, which are stable in air at room temperature for several days. In this crystal, two 6˙+ cations sandwich a neutral naphthalene molecule, forming a diradical dicationic (hetero)trimer in which the central naphthalene is rotated by 90° relative to the radical cations. Bond length analysis suggests that negative charge is transferred from naphthalene to 6˙+, which in turn is not a true mono radical cation. Likewise, Ishihara et al. used norbornene units to sterically protect the radical cation of p-dimethoxybenzene in compound 7.30 Oxidation of 7 by FeCl3 produces a persistent radical cation, which crystallizes as 7˙+FeCl4. This result confirms that single-electron oxidation of arenes to yield radical cations is achievable under typical Scholl reaction conditions, providing evidence to support the radical cation mechanism for Scholl reactions.31

The unpaired spin and positive charge in radical cations of conjugated hydrocarbons can be effectively delocalized by π-systems such as phenalenyl, Thiele's hydrocarbon, and azulene. As shown in Fig. 3a, phenalenyl delocalizes a radical or a positive charge across seven carbon atoms via resonance. Leveraging this capability, Kubo et al. designed and synthesized acetylene-linked bisphenalenyl 8,32 which resonates with a closed-shell cumulene structure (Fig. 3b). Compound 8 exhibits biradical character, confirmed experimentally and theoretically. It reacts with 2,3,5,6-tetrafluoro-tetracyanoquinodimethane (F4-TCNQ) to form a charge–transfer complex, 8˙+(F4-TCNQ), with complete charge transfer, as determined by Raman spectroscopy. This crystal remains stable in ambient conditions for weeks and shows a conductivity of 1.43 S cm−1 at 280 K owning to efficient π–π overlap in a π-stacked one-dimensional chain with a staggered packing motif. Notably, cooling 8˙+(F4-TCNQ) to 90 K induces formation of a C–C σ bond between a F4-TCNQ's C(CN)2 group and a phenalenyl carbon. Unlike 8, O- and N-substituted bisphenalenyls (9 and 10, respectively)12,33 feature quinoid-linked phenalenyl units shown in Fig. 3b and were synthesized in radical cation form via reduction of their dications. Their triflate salts, 9˙+(SO3CF3) and 10˙+(SO3CF3), exhibit π-stacking between phenalenyl units with intermolecular C⋯C contacts as short as 3.07 Å for 9. DFT calculations link these intermolecular closer-than-van-de-Waals contacts to intermolecular spin–spin coupling. Single crystals of 9˙+(SO3CF3) and 10˙+(SO3CF3) exhibit room-temperature conductivities of 1.31 × 10−2 S cm−1 and 1.16 × 10−2 S cm−1, respectively. In addition, thin films of 10˙+(SO3CF3) function as ambipolar semiconductors with hole and electron mobilities on the order of 10−5 cm2 V−1 s−1.


image file: d5sc06726j-f3.tif
Fig. 3 (a) Resonance structures of phenalenyl radical or cation; (b) bisphenalenyls 8–10.

As shown in Fig. 4a, the radical cation of Thiele's hydrocarbon (TH) is stabilized by the formation of a central benzene ring. Such aromaticity-enhanced stabilization has been used to develop stable conjugated radical cations based on TH (shown in blue), as demonstrated by compounds 11–13 in Fig. 4b. Bunz et al. found that the radical cations of N,N′-diarylated dihydrodiazapentacene 11a/b remained stable in CH2Cl2 for 24 h under ambient condition.34 In the crystal structure of 11a˙+SbF6, the centre ring and adjacent nitrogen-embedded six-membered ring have less bond length alternation than the neutral form, indicating increased aromaticity in these rings, which is also supported by nucleus-independent chemical shift (NICS) calculation. Such an increase in aromaticity can explain the observed high stability of the radical cation of 11a/b. For the same reason, the radical cation of compound 12 exhibits similar stability as its N-substituted analogue (11a/b). Its salt 12˙+SbF6 in CH2Cl2 solution showed negligible spectral change after one day under aerated conditions at room temperature.35 Compound 13 is similar to 12 by containing oxygen-incorporated TH in spite of extra carbonyl groups in the π-backbone.36 The radical cation of 13 is thermodynamically stabilized by delocalizing its unpaired spin across the whole molecule including aniline nitrogen and benzoyl carbon atoms. As a result, 13˙+SbF6 is very stable in solid state with no degradation after 10 months and in CH2Cl2 solution with no degradation after 7 days under ambient conditions and room light. Spin-coated films of neutral 13 function as a p-type semiconductor with field-effect mobility of 1.3 × 10−5 cm2 V−1 s−1, while spin-coated films of 13˙+DDQ exhibits ohmic behavior with a room-temperature conductivity of 7.7 × 10−3 S cm−1. Having two C atoms replaced by B and N atoms, respectively, compounds 14 is a B–N analogue of TH.37 Oxidation of 14 with Ag+SbF6 results in 14˙+SbF6, which represents a rare example of an isolable boron-containing radical cation.


image file: d5sc06726j-f4.tif
Fig. 4 (a) Thiele's hydrocarbon (TH) and its radical cation; (b) compounds derived from TH (11–13) and a BN analogue of TH (14).

A seven-membered carbocycle capable of forming an aromatic tropylium cation has also been employed as a structural motif to stabilize radical cations, as exemplified by compounds 15 and 16 (Fig. 5).38 DFT calculations revealed that upon one-electron oxidation, the seven-membered ring in 15˙+ becomes more positively charged compared to neutral 15, while the unpaired spin is delocalized across the framework, with higher density localized on the seven-membered ring and the inner rim of the five-membered rings, the latter being shielded by the helical conformation of 15˙+. Stepwise chemical oxidation of 16 with NO+SbF6 generated the radical cation (16˙+) and the dication (162+), which were isolated as their hexafluoroantimonate salts.39 DFT calculations indicate that in 16˙+, the spin is delocalized across the π-backbone, with higher density on the seven-membered rings, while in 162+, the positive charges are predominantly located on the seven-membered carbocycles. In the single crystal of 16˙+SbF6, 16˙+ adopts a cisoid conformation, and packs into a cofacial π-dimer. Partial oxidation with NO+SbF6 can yield fractional radical cation salt. Interestingly, partial oxidation of 16 with NO+SbF6 resulted in crystals of a mixed-valence species 16· (16˙+)3 consisting of one-quarter equivalent of the neutral molecule and three-quarters equivalent of the radical cation. The crystal structure of this mixed-valence species features a huge trigonal unit cell that contains 72 molecules of 16, 54 SbF6 counter anions and 101 disordered hexane molecules.


image file: d5sc06726j-f5.tif
Fig. 5 π-Systems containing seven-membered carbocycles.

Double helicenes were recently been demonstrated by Miao and coworkers to be an effective structural motif for stabilizing radical cations. Chemical oxidation of double [5]helicene 17a/b (Fig. 6) yielded a robust radical cation, showing negligible spectral changes after its CH2Cl2 solution was exposed to ambient air in the dark for 60 days.40 Such a high stability is attributed to spin delocalization in the π-backbone and the twisted structure impeding dimerization. Notably, crystal structures of neutral 17a and radical cation salt 17a˙+PF6 exhibit nearly identical π–π stacking, with PF6 occupying the site of CH2Cl2 in the neutral crystal. Dip-coated films of 17a˙+PF6 showed conductivity up to 1.32 ± 0.04 S cm−1, while 17a behaved as a p-type semiconductor with mobility of 0.064 ± 0.003 cm2 V−1 s−1. Moreover, photochemical oxidation of 17b in the solid state with O2 generated its radical cation, enabling OFETs to function as nonvolatile optoelectronic memory with switching contrast above 103 and long-term stability. Similarly, S-containing double helicenes 18a/b (Fig. 6) formed robust radical cations. 18a˙+ remained stable in CH2Cl2 under ambient air in the absence of light for four weeks and retained ∼85% absorption in water-saturated CH2Cl2 after 8 days.41 The salt 18b˙+SbF6 achieved a conductivity of 0.16 S cm−1 at room temperature, due to strong π–π interactions in its crystal structure.


image file: d5sc06726j-f6.tif
Fig. 6 (a) Double helicenes 17a/b and 18a/b; (b) π–π stacking in the crystal of 17a˙+PF6, where the (M, M) enantiomer is shown in blue, and the (P, P) enantiomer is colored according to the elements. Reproduced with permission from ref. 40. Copyright 2022 American Chemical Society.

2.2 Nitrogen-containing systems

Stable radical cations often feature heteroatoms that act as spin-bearing sites,42 and tertiary arylamines have long been used to develop radical cation species. For example, N,N,N′,N′-tetramethyl-p-phenylenediamine (Fig. 7), can be readily converted by one-electron oxidation into a stable semiquinone radical, known as Wurster's blue.43 Another example is tris(4-bromophenyl)ammoniumyl hexachloroantimonate, which is commonly known as magic blue (Fig. 7) and used as a popular oxidizing agent in organic and organometallic chemistry.
image file: d5sc06726j-f7.tif
Fig. 7 Resonance structure of Wurster's blue and structure of magic blue.

Triphenylamine radical cations lacking para substituents are inherently unstable due to high spin densities at para positions.44 Stabilization can be achieved by ortho-bridging aryl groups, as seen in triarylamines 19–23 (Fig. 8), which enhance planarity of π-frameworks and delocalize spin density. Crystal structures reveal that neutral trioxytriphenylamine 19 is shaped like a shallow bowl with a sp3-hybridized central N atom, whereas its radical cation is flat with shorter C–N bonds.45 The delocalization of unpaired spin across the π-framework of 19˙+, including oxygen atoms, prevents dimerization or oxygenation under ambient conditions. Similarly, the radical cations of S,C,C-bridged triarylamine 20[thin space (1/6-em)]46 and O,C,C-bridged triarylamines 21a/b[thin space (1/6-em)]47 all exhibited appreciable stability, with the absorption spectra of their CH2Cl2 solutions remaining virtually unchanged after one day and the crystals of their hexachloroantimonate salts being stable under ambient conditions. In the crystals, the π-frameworks of 20˙+ and 21a˙+ are essentially flat while that of 21b˙+ is slightly twisted. DFT calculations indicate spin delocalization across all three radicals, with the highest density on the central nitrogen. Notably, sulfur in 20˙+ shares more spin density than oxygen in 21a˙+ and 21b˙+. Spin density on the central N atom increases in the order of 20˙+ (+0.294) < 21a˙+ (+0.328) < 21b˙+ (+0.332), while Mulliken charges on it decrease in the order of 20˙+ (+0.345) > 21a˙+ (+0.290) > 21b˙+ (+0.194).


image file: d5sc06726j-f8.tif
Fig. 8 Structures of triarylamines 19–23.

Very recently, Kivala et al. synthesized trithiatriphenylamine 22a (Fig. 8) and compared it to 22b/c (Fig. 8) with fewer thia bridges.48 The radical cation of 22a, in the form of 22a˙+SbCl6, remained stable for months under ambient conditions, showing no degradation in CHCl2CHCl2 solution. In contrast, 22b˙+ and 22c˙+ exhibited significantly lower stabilities, with half-lives of 1.06 and 6.04 days, respectively, in tetrachloroethane at 80 °C under inert atmosphere. Spin delocalization differences explain this stability: the spin density of 22a˙+ is delocalized across the whole molecule while that of 22b˙+ is mostly distributed over the methoxylated phenothiazine subunit and that of 22c˙+ is localized on the phenothiazine ring. The crystal structure of 22a˙+SbCl6 exhibits a planarized π-backbone relative to the saddle-shaped neutral molecule and one-dimensional stacking of 22a˙+. Additionally, 22a forms charge-transfer complexes with 7,7,8,8-tetracyanoquinodimethane (TCNQ) or F4-TCNQ, transferring 0.29 e and 0.42 e to TCNQ in the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 2[thin space (1/6-em)]:[thin space (1/6-em)]1 complexes, respectively, or 0.90 e to F4-TCNQ in the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complex, as determined from anion bond lengths. In these crystals, molecules of 22a and TCNQ or F4-TCNQ stack in a column with an alternative arrangement. However, the electrical conductivities of these charge transfer complexes as well as 22a˙+SbCl6 were not reported.

Compound 23a, a sp3 hybridized carbon-bridged triphenylamine, is oxidized by Ag+ to it radical cation, which is isolated in the form of 23a˙+(Al(OC(CF3)3)4) and characterized by X-ray crystallography.49 Similar to 19˙+, the polycyclic framework of 23a˙+ adopts a planar conformation, in contrast to the shallow bowl-shaped geometry of neutral 23a. DFT calculations indicate that the central N atom in 23a˙+ carries a higher spin density (+0.366) compared to 19˙+ (+0.328), indicating greater localization of the unpaired electron. When oxidized with 0.5 equivalent of Ag+(Al(OC(CF3)3)4), 23a forms a dimerized product, suggesting the reactivity of 23a˙+. Compound 23b, a sterically protected triphenylamine derivative of 23a, incorporates spiro-fluorene moieties perpendicular to the planar triphenylamine core and tert-butyl groups at para positions.50 Its radical cation remains stable in CH2Cl2 solution for months at 7 °C under N2 and in solid form under ambient conditions for weeks.

Incorporating ortho-bridged triphenylamine into a double helicene structure provides additional kinetic protection for its radical cation through helical curvature around the spin centre, as demonstrated in compounds 24–26 (Fig. 9). The radical cations of 24 and 25 were crystallized as 24˙+SbCl6 and 25˙+NTf2, respectively.51,52 Crystal structures reveal that oxidation induces planarization in both molecules, accompanied by shortened helical pitches in their radical cation states. Notably, the absorption spectrum of 24˙+SbCl6 in CH2Cl2 solution remained unchanged for 2 weeks under ambient air and room light, whereas 25˙+NTf2 retained its spectral profile for 1 day under an inert atmosphere. Compound 26, a double hetero[5]helicene synthesized via oxidative dimerization of benzo[b]phenoxazine,53 forms stable radical cations in both racemic and enantiopure forms with different counteranions, which influence the molecular packing in the resulting radical cation salts.54 Racemic and enantiopure 26˙+NTf2 exhibit one-dimensional π–π stacking with counteranions intercalated between stacks, though enantiomers in the racemic crystal are separated by a larger π–π distance. In contrast, in the racemic and enantiopure 26˙+SbCl6 crystals, the interactions between 26˙+ and SbCl6 prevent continuous π–π stacking of 26˙+ likely due to the matching between the curvature of 26˙+ and the spherical shape of the SbCl6 anion. Despite the close π–π stacking, racemic and enantiopure crystals 26˙+NTf2 exhibited low conductivity of 1.22 × 10−8 S cm−1 and 2.86 × 10−8 S cm−1, respectively. For comparison, vacuum-deposited films of neutral 26 behaved as a p-type semiconductor with a field effect mobility of 2 × 10−5 cm2 V−1 s−1.53 Incorporating triphenylamine into a larger curved π-scaffold yields phenylenediamine-linked nanographene 27 (Fig. 9) whose radical cation remains stable in CH2Cl2 solution for 8 days under ambient conditions without degradation.55 Interestingly, 27 adopts a Z-shaped anti conformation in its neutral state, whereas its radical cation assumes a C-shaped syn conformation in the 27˙+SbF6 crystal. The stability of 27˙+ can be attributed to contribution from quinoidal resonance in 27˙+ and kinetic protection from the curved π-scaffold.


image file: d5sc06726j-f9.tif
Fig. 9 Triarylamine-incorporated double helicenes 24–26 and phenylenediamine-linked nanographene 27.

N,N′-Disubstituted dihydrodiazaacenes, such as compounds 28a–c[thin space (1/6-em)]56 (Fig. 10), serve as general π-scaffolds to stabilize radical cations because one-electron oxidation transforms their central dihydropyrazine ring from antiaromatic to non-aromatic or weakly aromatic, as evidenced by DFT-calculated NICS values. This stabilization mechanism also applies to the radical cation of compound 29 (Fig. 10).57 Consequently, the radical cations of 28a–c and 29 exhibit comparable stability with their absorption spectra essentially unchanged under ambient conditions for at least 24 h. Crystal structures reveal that the diazaacene backbones of 28a–c are flat in both neutral and radical cations forms, while the dibenzo[a,c]phenazine backbone of 29 changes from a V-shaped conformation in the neutral state to a twisted conformation in the radical cation, with the dihydropyrazine ring flattened. Notably, the formation of stable radical cation of 28a is related to the doping of p-type semiconductors based on N,N′-disubstituted dihydrodiazapentacenes.58,59


image file: d5sc06726j-f10.tif
Fig. 10 N,N′-Disubstituted-dihydrodiazaacenes 28a–c and related structures.

Based on the radical cation of N,N′-disubstituted dihydrophenazine, Wang et al. developed a water-soluble bis-sulfonate salt 30 (Fig. 10).60 The radical cation in 30 resists dimerization and disproportionation, attributable to intermolecular coulombic repulsion and steric hindrance from sulfonate groups. It also has an appropriate electron density to avoid reactions with oxygen or water. As a result, 30 exhibits remarkable stability, with no decomposition observed in aqueous solution under ambient conditions for more than 70 days (monitored by UV-Vis spectroscopy). Aqueous organic redox flow batteries using 30 as the posolyte and ZnCl2 as the negolyte demonstrated extremely stable performance over 2500 cycles (∼27 days). Radical cation 31˙+ (Fig. 10), also known as viridium,61 features a structure related to N,N′-disubstituted dihydrophenazine. It exhibits exceptional thermodynamic stability in solution under ambient conditions, as confirmed by Electron Paramagnetic Resonance (EPR) and UV-Vis spectroscopy over two months, due to full spin delocalization across its π-backbone. In the solid state, 31˙+ arranges into one-dimensional columns with a π–π distance of 3.29 Å. In aqueous or perfluorohydrocarbon solutions, it forms π-stacked dimers.

As an electron-rich aromatic heterocycle prone to oxidation, pyrrole serves as a building block for π-conjugated radical cations. While N-annulated rylenes formally incorporate fused pyrrole units, their radical cations are not further discussed here because their spin density and singly occupied molecular orbitals are delocalized across the rylene backbone but located minimally on the nitrogen atoms.62 Pyrrole-containing multiple cations are discussed in the next session. N,N′-Diarylated tetrabenzotetraaza[8]circulene 32a/b (Fig. 11) yields highly stable radical cations upon chemical oxidation.63 This stability is highlighted by the 83% yield of 32a˙+ obtained after aqueous work-up, CH2Cl2 extraction, and silica gel column chromatography under ambient conditions. The stability arises from delocalization of spin density across the tetrabenzotetraaza[8]circulene backbone, with the highest spin density localized on the α-carbons of the N-octylated pyrrole units. In their hexachloroantimonate salt crystals, 32a˙+ and 32b˙+ both exhibit planar structures but adopt distinct stacking arrangements: 32a˙+ forms eclipsed stacking with an intermolecular π–π distance of 3.32 Å, whereas 32b˙+ adopts a slip-stacked arrangement. Pyrrole-embedded buckybowl 33a and its planar analogue 33b generate radical cations of different reactivities in solution.64 Despite similar spin delocalization across the π-backbones with higher spin density localized on pyrrole carbons, the reactivity divergence arises from the strain induced by the positive curvature of 33a. In both single crystals and low temperature solution states, 33a˙+ undergoes reversible σ-dimerization between the internal pyrrole α-carbons, whereas 33b˙+ selectively forms a π-dimer with a short π–π distance of 3.14 Å.


image file: d5sc06726j-f11.tif
Fig. 11 Pyrrole-embedded molecules 32 and 33.

2.3 Sulfur-containing systems

Sulfur-containing π-conjugated radical cations often feature sulfur atoms in thiophene or tetrathiafulvalene (TTF) units, as exemplified by the molecules shown in Fig. 12a and b, respectively. The radical cations of bithiophene and terthiophene are sterically protected in their bicyclo[2.2.2]octene-fused derivatives 34 and 35a (Fig. 12a),65 similar to the sterically congested hydrocarbons 5 and 6 (Fig. 2). Such kinetical stabilization allows the crystals of 34˙+SbF6 and 35a˙+SbF6 to be stable without decomposition after the crystals were left to stand at room temperature under air for one month. The structure of 35˙+ involves a significant contribution from the quinoidal resonance as indicated by crystal structure and ESR measurement as well as theoretical calculations. Notably, oxidation of sterically protected bithiophene and terthiophene (34 and 35) with NO+SbF6 yields radical cations, whereas longer oligothiophenes under the same conditions form dications. Without a bicyclo[2.2.2]octene moiety at the centre thiophene ring, 35b˙+ forms π-dimer with intermolecular distance as short as 2.976 Å between the beta carbon atoms of the central thiophene units.66 Compound 36 is a macrocycle composed of four ethynylene-thienylene and two vinylene-thienylene units, whose radical cation remains stable in solid state but degrades in CH2Cl2 solution under ambient condition.67 In the crystal of 36˙+SbF6, 36˙+ forms a π-dimer, (36˙+)2, with intermolecular S⋯S distances of 3.7–3.8 Å comparable to twice the standard van der Waals radius of sulfur atom. Spin-coated films of 36 from its CS2 solution exhibited a room temperature conductivity of 5.7 × 10−8 S cm−1, but became a p-type semiconductor after annealing at 100 °C, with a lower conductivity of 9.2 × 10−11 S cm−1 and a field effect mobility of 1.4 × 10−4 cm2 V−1 s−1, suggesting formation of 36˙+ in the thin film by self-oxidation of 36 in air.68
image file: d5sc06726j-f12.tif
Fig. 12 (a) Thiophene-containing molecules; (b) TTF-containing zwitterions.

Tetrathiafulvalene (TTF) is a well-established π-scaffold for stable radical cations, as its five-membered ring becomes aromatic with six π-electrons upon one-electron oxidation. A recent notable example is zwitterion 37 (Fig. 12b), which has a carboxylate anion attached to the π-conjugated radical cation core. It exhibits exceptionally high conductivity: 530 S cm−1 at 300 K and 1000 S cm−1 at 50 K.69 DFT calculations reveal that the spin density in 37 is localized primarily on the carboxylate-connected TTF unit, balancing the negative charge of the carboxylate with the positive charge of the TTF moiety. Notably, 37 is unique among the radical cations discussed here, as its single-crystal structure has not been reported. Due to lack of single crystals, the high conductivity of 37 was measured from disordered bulk solids rather than crystalline materials. DFT calculations suggest stable dimeric structures in the solid state, featuring anti-parallel arrangements in both face-to-face and side-by-side stacking. Zwitterion 38, with a structure closely related to 37, form hydrogen-bonded organic frameworks (HOFs) in the solid state.70 Tuning crystallization conditions yields two polymorphs of 38, which exhibit room-temperature conductivities of 6.07 × 10−7 and 1.35 × 10−6 S cm−1, respectively, as measured in pressed pellets of crystalline powder.

3. Multiple radical cations

Conjugated multiple radical cations, which possess more than one unpaired electron and/or carry more than a single positive charge, can in principle be created by incorporating several redox active π-units into a single framework. The structural factors that stabilize monoradical cations, discussed previously, are not reiterated here to avoid redundancy. This approach is exemplified by the diradical dication 392(˙+), which results from reduction of cyclobis(paraquat-p-phenylene) tetracation (394+),71,72 as shown in Fig. 13. Specifically, 392(˙+) forms an inclusion complex with methyl viologen radical cation (40˙+) by three-electron reduction of an equimolar mixture of 394+ and 402+. This spontaneous encapsulation of 40˙+ inside the cavity of 392(˙+) arises from favorable radical–radical interactions occurring between the three 4,4′-bipyridinium radical cations. The binding constant of this inclusion complex was determined using isothermal titration calorimetry and UV-Vis spectroscopy as (5.0 ± 0.6) × 104 M−1 and (7.9 ± 5.5) × 104 M−1, respectively. Crystal structure shows that the bipyridinium radical cation subunits in 40˙+ and 392(˙+) are separated by 3.22 Å as measured from their centroids, and quantum chemistry calculations indicate partial charge transfer from the guest (40˙+) to the host (392(˙+)), with calculated charges of +1.18 and +1.71, respectively. In the crystal of the complex, 392(˙+) arranges into one-dimensional columns with π–π distance of 3.28 Å. These π–π interactions enable the salt of 40˙+392(˙+) 3PF6 to perform as a p-type semiconductor in a single-crystal field effect transistor with a hole mobility up to 0.05 cm2 V−1 s−1.73 Notably, 392(˙+) was also crystallized alone as its hexafluorophosphate salt, where 392(˙+) stacks into a one-dimensional column with an even shorter π–π distance (3.12 Å), highlighting the dominance of strong radical–radical interactions in driving solid-state packing. In addition, the radical–radical interactions of 392(˙+) have been applied in developing rotaxanes and catenanes.74,75
image file: d5sc06726j-f13.tif
Fig. 13 Macrocyclic multiple radical cations based on 4,4′-bipyridinium.

On the basis of tetracation 394+, Stoddart et al. further developed a hexacationic organic cage, 416+, by incorporating three redox-active building blocks of 4,4′-bipyridinium dications into a triangular prism-shaped symmetry.76 Reduction of 416+ with cobaltocene results in its triradical trication 413(˙+), which has a doublet ground state. Its quartet state is only 0.08 kcal mol−1 higher in energy than the doublet, suggesting a strong intramolecular exchange coupling. 413(˙+) was crystallized in the form of CH3CN ⊂ 413(˙+) 3PF6, where a molecule of CH3CN occupies the cavity of 413(˙+) (Fig. 13). In the crystal structure, 413(˙+) arranges into a hexagonal porous superstructure with interconnected one-dimensional channels. Each channel is formed by taking advantage of the strong radical-pairing interactions with a stacking distance of 3.15 Å between the bipyridinium radical cations and the π–π stacking between the phenylene units in the adjacent cages.

Incorporating several bridged triarylamines units in one molecule allows formation of multiple radical cations through oxidation. The overall radical nature of the resulting cationic species depends on the interactions between the individual radical sites. Macrocycles 42a/b (Fig. 14) contain S-bridged triphenylamine units connected by methylene groups. 42a3(˙+) and 42b4(˙+) are formed by oxidation of each S-bridged triphenylamine unit to the radical cation, with spin density delocalized on each planarized phenothiazine unit.77 They were both isolated in the form of hexafluoroantimonate salts, and 42a3(˙+) appeared more stable than 42b4(˙+), with its characteristic absorption barely changed after 48 h at room temperature. The capability of forming stable radical cations allow 42a and 42b to function as cathode materials in lithium-ion batteries with initial discharge capacities of 19.3 mA h g−1 and 28.7 mA h g−1, respectively.


image file: d5sc06726j-f14.tif
Fig. 14 Multiple triarylamine-based molecules that afford multiple radical cations.

Compound 43a comprises two sulfur and carbon-bridged triphenylamine (SCBT) units linked by a 1,8-naphthalene moiety, while 43b contains two oxygen and carbon-bridged triphenylamine (OCBT) units. Stepwise chemical oxidation of 43a results in the radical cation and diradical dication, which were isolated in the form of 43a˙+SbCl6 and 43a2(˙+)2NTf2, and both species are stable under ambient conditions.78 Single-crystal X-ray diffraction analyses revealed that both 43a˙+ and 43a2(˙+) adopt intramolecular face-to-face close π-stacking of the SCBT units, which arises from covalent-like bonding interactions between the sulfur atoms. This is evidenced by significantly shortened S⋯S distances of 3.37 Å in 43a˙+ and 3.03 Å in 43a2(˙+), both well below twice the van der Waals radius of sulfur (3.60 Å). In contrast, the crystal structures of neutral 43b, 43b˙+SbCl6 and 43b2(˙+)2SbCl6 reveal that the two OCBT units progressively separate upon oxidation, with no intramolecular π–π stacking observed in the dicationic state 43b2(˙+). In the crystal of 43a˙+SbCl6, the SCBT units stack into a one-dimensional column, while in of 43a2(˙+)2NTf2, the naphthalene linkers stack into one-dimensional columns, with no intermolecular overlap between SCBT units. Owing to the π–π interactions in the solid state, compressed pellets of 43a˙+SbCl6 and 43a2(˙+)2NTf2 exhibited room-temperature conductivity of 1.9 × 10−6 S cm−1 and 1.4 × 10−7 S cm−1, respectively. Dication 442+ has two carbon-bridged triphenylamine units linked through a C–C single bond and was obtained by oxidative dimerization of C-bridged triphenylamine 23a (Fig. 8).49 It has an open-shell singlet ground state, with a diradical character of 0.77. The radicals in 442+ interact through resonance as shown in Fig. 14, unlike the radicals in 42a3(˙+) and 43a2(˙+). Compound 45 contains two C-bridged triphenylamine units bridged with a π-extended [5]helicene.79 The dication 452+ contains small contribution of open-shell singlet, with a diradical character of 0.20, which is in line with the shorter C–N bond upon oxidation, suggesting the small contribution of diradical dication character. In both 45˙+ and 452+, π–π stacking are absent.

Conjugated macrocycles containing multiple triphenylamine units, such as compounds 46–48 (Fig. 15), enable stable multiple radical cations. Chemical oxidation of compounds 46a and 46b yields stable diradical dication that can be stored under ambient conditions.80–82 Having 1,4-phenylene linkers, 46a2(˙+) exhibits a triplet ground state with spins and charges delocalized over two 1,4-phenylenediamine moieties. In contrast, having 9,10-anthrylene linkers, 46b2(˙+), exhibits a singlet ground state with spins and charges localized on two 1,3-phenylenediamine moieties, supported by shortening of the corresponding C–N bonds. In the solid state, 46a2(˙+) arranges into a one-dimensional chain in the crystal structures of 46a2(˙+)2SbF6 and 46a2(˙+)2(Al(OC(CF3)3)4), but not in the crystal structure of 46a2(˙+)2(B(C6F5)4). Macrocycles 47a/b and 48 (Fig. 15) exhibit rich redox properties, with several cationic species isolated and characterized.83,84 47a2+ and 47a4+ were both isolated as hexachloroantimonate salts. The open-shell singlet ground states of 47a2+ and 47a4+ exhibit diradical characters of 0.862 and 0.379, respectively, with 47a2+ being globally aromatic and 47a4+ globally antiaromatic. In contrast, the oxidized species of 47b was isolated as a radical trication. Similarly, chemical oxidation of bismacrocycle 48 generates 482(˙+), which possesses an open-shell singlet ground state with a high diradical character of 0.96. Similar redox behavior is observed for their hydrocarbon analogue 49 (Fig. 15).85 492+, isolated in the form of 492+2SbF6, exhibits a globally aromatic, open-shell singlet ground state, with multiple diradical characters (y0 = 0.46, y1 = 0.37, y2 = 0.16) calculated from the crystal structure.


image file: d5sc06726j-f15.tif
Fig. 15 Conjugated macrocycles that afford multiple radical cations.

Sessler et al. synthesized cyclo[8]pyrrole in its dicationic form, 502+SO42− (Fig. 15) and subsequently oxidized it to the radical trication 50˙3+ using I2.86,87 The side chains in 50˙3+ play a crucial role in determining the packing motifs of the radical trications in the resulting crystals. Diffusion of hexane vapor into a dichloromethane solution yielded (50a˙3+SO42−)2·I252−·I2, in which 50a˙3+SO42− units form slipped π-dimers without interdimer π–π interactions. Using ethyl acetate instead of hexane resulted in (50a˙3+SO42−)2·I7·I24, where two 50a˙3+SO42− units sandwich an I7 ion, and these sandwiches are intercalated by layers of 1D polyiodide I24 (Fig. 16a). In the crystal structure of 50b˙3+SO42−·I12, the 50b˙3+SO42− units are intercalated by 1D polyiodide chains, forming a linear 1D stacked donor–acceptor structure. For 50c˙3+SO42−·I16, the 50c˙3+SO42− units assemble into 2D highly ordered layers, which are intercalated with 2D polyiodide layers of I8 and I24, leading to a linear 1D donor–acceptor cocrystal structure (Fig. 16b). Single-crystal of (50a˙3+SO42−)2·I7·I24, 50b˙3+SO42−·I12 and 50c˙3+SO42−·I16 exhibited conductivity of 3.6 ± 0.4 × 10−4 S cm−1, 7.2 ± 0.7 × 10−3 S cm−1 and 6.1 ± 0.6 × 10−1 S cm−1, respectively, under ambient temperature. In addition, 50a2+SO42− can transfer electrons to macrocycle 51 via proton-coupled electron transfer (PCET) or through an artificial electron transport chain (ETC) using I2 as mediators and trifluoroacetic acid as proton source. This process generates the radical trication 50a˙3+ in the crystal 50a˙3+SO42−(CO2CF3) and the radical dication H351˙2+ in the crystals H351˙2+2(CO2CF3)·CHCl3 or (H351˙2+)2SO42−·2I3·2.5I2·6.25H2O.


image file: d5sc06726j-f16.tif
Fig. 16 Crystal structures of (50a˙3+SO42−)2·I7·I24 (a) and 50c˙3+SO42−·I16 (b) with cartoon representation. Reproduced with permission from ref. 86. Copyright 2025 American Chemical Society.

Other notable examples of multiple radical cations include 52(2˙)+, 532(˙+) and 542(˙+), as shown in Fig. 17. Azatriangulene diradical monocation 52(2˙)+ features a polycyclic framework isoelectronic with the triangulene diradical, where a cationic sp2-hybridized nitrogen atom replaces the central carbon atom of triangulene. Two derivatives, 52(2˙)+, 532(˙+), were isolated as 52a(2˙)+SbCl6 and 52b(2˙)+(SO3CF3), respectively.88,89 They both have a triplet ground state, analogous to triangulene, with spin density delocalized along the periphery of the aza-triangulene backbone. This spin is kinetically stabilized by trichlorophenyl and mesityl substituents. As a result, 52b(2˙)+(SO3CF3) exhibits excellent stability, showing no changes in its absorption spectrum after 10 hours or in fluorescent intensity after 5000 seconds. Oxidation of diindinodihydrophenazine 53 with I2 generates its diradical dication as the solid 532(˙+)2I3, which is benchtop stable for several months under ambient conditions.90 A large excess of reducing reagents and base is required to reduce 532(˙+) back to 53. Di(benzothiino)pyrene 54 can be stepwise oxidized to 54˙+ and 542(˙+) using 1 or 2 equivalents of NO+(Al(OC(CF3)3)4).91 Upon one-electron oxidation, one 1,4-dithiane ring becomes planar, with the spin density primarily localized on this ring and a concomitant contraction of the S–C bond lengths. As a result, 542(˙+) adopts a fully planar backbone and possesses a triplet ground state.


image file: d5sc06726j-f17.tif
Fig. 17 Structures of cation 52a/b(2˙)+ and compounds 53 and 54.

4. Fractional radical cations

Fractional radical cations here refer to charged π-systems that, on average, bear non-integer charges and/or unpaired spins. These fractional charges or unpaired spins typically arise from the association of radical cations with neutral molecules in specific stoichiometric ratios, resulting in mixed-valence species. An early example is the mixed-valence salt of naphthalene (naph), reported by Fritz et al. in 1978.92 In this crystal, the naphthalene radical cation (naph)˙+ forms a mixed-valence dimer, (naph)2˙+, where the two crystallographically indistinguishable components are arranged in a face-to-face π-stack with an interplanar separation (3.2 Å) apparently shorter than the sum of their van der Waals radii. Polycrystalline pellets of (naph)2˙+PF6 exhibit a room-temperature conductivity of 0.12 ± 0.046 S cm−1. Mixed-valence salts, exemplified by (naph)2˙+PF6, played a pivotal role in the development of organic conductors and superconductors.93 Besides naphthalene, other polycyclic aromatic hydrocarbons, such as pyrene (pyr), perylene (per), triphenylene (TP) and fluoranthene (flu), form mixed-valence salts through electrocrystallization.94–97 Single crystals of (pyr)12(AsF6)7 exhibits metallic conducting behaviour in 200–300 K, and those of (per)2(AsF6)1.1·(CH2Cl2)0.7 and (per)2(AsF6)0.75(PF6)0.35·(CH2Cl2)0.85 exhibit very high conductivities of 1200 S cm−1 at 300 K and 70[thin space (1/6-em)]000 S cm−1 around 285 K, respectively, owning to the 1D stacks of perylene molecules in the mixed-valence salts. In contrast, polycrystalline samples of (TP)2PF6, (flu)2PF6 exhibit low conductivity of 7 × 10−3 S cm−1 and 0.05 S cm−1 at room temperature, respectively.98

Tetrathiafulvalene (TTF) not only forms radical cation salts as shown in Fig. 1, but also participates in mixed-valence systems, exemplified by the TTF-TCNQ complex, the first purely organic material reported to exhibit electrical conductivity comparable to that of metals.99 The high conductivity of TTF-TCNQ arises from its mixed-valence state, in which TTF and TCNQ carry partial charges of +0.59 and −0.59, respectively, and form segregated stacks in the solid state.100,101 Similarly, TTF analogues and derivatives also support mixed-valence states. A prominent example is tetramethyltetraselenafulvalene (TMTSF, Fig. 17), which forms superconducting mixed-valence salts. (TMTSF)2˙+PF6 exhibits a conductivity of 105 S cm−1 around 20 K at ambient pressure and was the first organic superconductor discovered, with a superconducting transition temperature of 0.9 K under a pressure of 12 kbar,102 while single crystals of (TMTSF)2˙+ClO4 transit to a superconductor below 1.4 K in the absence of applied pressure.103

Compound 55 (Fig. 18) can be regarded as an indenofluorene-extended tetrathiafulvalene (TTF). Oxidation of 55 via electrocrystallization yields either the mixed-valence salt 55·(BF4)1.5 or the radical cation salts 55˙+PF6 and 55˙+TaF6, depending on the electrolyte employed.104 The mixed-valence salt contains 55˙+ and 552+ in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio. In the crystal structures of all three salts, the 55 cations form one-dimensional (1D) π-stacks. However, 55·(BF4)1.5 exhibits a larger π–π distance (3.69 Å) than the other two salts, indicating stronger electrostatic repulsion between 55˙+ and 552+ compared to that between 55˙+ ions. As measured from the compressed powdered samples, at room temperature, 55·(BF4)1.5, 55˙+PF6 and 55˙+TaF6 exhibit room-temperature conductivities of 1.3 × 10−3, 3.07 × 10−2, and 5.32 × 10−2 S cm−1, respectively. All three salts behave as semiconductors, with conductivity decreasing by four to five orders of magnitude upon cooling to 77 K. Consistent with the intrinsic conductivity observed in the radical cation salts of 55, a butylated derivative of 55 functions as a p-type organic semiconductor in single crystals transistors with a field effect mobility of 1.44 cm2 V−1 s−1.105 In contrast to the 1,3-dithiole rings in TTF connected through a C–C double bond, anion 56˙ (Fig. 18) features two 1,3-thiazole rings connected through conjugated nitrogen atoms and has a radical cation in its π-backbone. Adding tetramethylbenzidine (57 in Fig. 18) to a solution of 56˙ led to the formation of the 56·572 complex as blue crystals.106 Mulliken population analysis indicates an overall charge of −1.6 on 56 and +0.8 on 57, confirming significant electron transfer from 57 to 56. The charge–transfer complex of 56·572 effectively absorbs across the full solar spectrum, extending beyond 2500 nm, and was therefore employed as a photothermal material, achieving a photothermal conversion efficiency of 49.6%.


image file: d5sc06726j-f18.tif
Fig. 18 Structures of TTF analogues and derivatives.

Compound 58 (Fig. 19), a derivative of bispyrrole-fused bisanthene, presents a rare example for radical cations exhibiting symmetry-broken intermolecular charge separation in the solid state.107 Stepwise oxidation of 58 with NO+SbF6 generates 58˙+ and 582+, distinguishable by absorption spectroscopy. DFT-calculated spin density indicates that spin is largely distributed along the periphery of 58˙+. Interestingly, X-ray crystallography combined with DFT calculation-based Mulliken population analysis reveals that the crystal of 58˙+SbF6 is in fact a mixed valence complex comprising two different forms of 58: a shallow bowl-shaped form bearing more positive charge (58m+) and an essentially flat form bearing less positive charge (58n+). In the crystal, the two forms of 58 arrange into a π-stack with the repeating sequence of 58m+58n+58n+58m+. Upon increasing the temperature from 173 K to 298 K, the degree of charge transfer decreases, with m decreasing from 1.4 to 1.3 and n increasing from 0.6 to 0.7. This reduced charge transfer is accompanied with a reduced bowl depth and an increased π–π distance between 58m+ and 58n+.


image file: d5sc06726j-f19.tif
Fig. 19 Polycyclic heteroarenes that form fractional radical cations.

Compounds 59a and 59b (Fig. 19) are O and S-containing anthanthrenes, which form mixed valence salts, 59a2ClO4 and 59b3(ClO4)2, via electrocrystallization.108 In the crystals of these mixed valence salts, 59a and 59b molecules form 1D column, with shorter π–π distances (<3.40 Å) and greater π–π overlaps than the corresponding neutral crystals. The single crystals of 59a2ClO4 and 59b3(ClO4)2 exhibit room temperature conductivities up to 0.11 S cm−1 and 0.031 S cm−1, respectively. In connection with electrical conductivities of these mixed valence salts, compounds 59a and 59b and their derivatives in neutral forms have been applied as p-type organic semiconductors in OFETs. For example, the 3,9-diphenyl derivative of 59a in vacuum-deposited films achieved a field effect mobility of 0.43–0.46 cm2 V−1 s−1, and the devices were demonstrated to be stable for five months under ambient conditions.109

Compounds 60a and 60b (Fig. 19) are a pair of regioisomeric polycyclic O-heteroarenes, which form mixed valence salts, 60a10(PF6)6·(THF)16 and 60b3(ClO4)2·THF·(H2O)0.5, via electrocrystallization.110 In the crystals, the 60a units assemble into a π-stacked column with an antiparallel arrangement and an average interplanar spacing of 3.33(7) Å, in contrast to the herringbone packing of neutral 60a. Similarly, the 60b units arrange into a π-stacked column with both pincer-like and head-to-tail packing modes and π–π distances varying from 3.21 to 3.42 Å (Fig. 20a). The crystallographically independent molecules of 60a and 60b within each stack display nearly identical bond lengths, indicating effective charge delocalization across the entire π-column. Two-terminal devices fabricated from a single crystal of 60b3(ClO4)2·THF·(H2O)0.5 exhibited semiconducting behavior, demonstrated by its sigmoidal IV curves and an increase in conductivity with rising temperature. The room-temperature conductivity reached values of up to 3.7 × 10−3 S cm−1 (ohmic) and 5.1 × 10−3 S cm−1 (non-ohmic).


image file: d5sc06726j-f20.tif
Fig. 20 (a) π–π stacking in the crystal structure of 60b3(ClO4)2·THF·(H2O)0.5; (b) π-stacked trimer in the crystal structure of 61b32(˙+)(BArF4)2; (c) a unit cell in the crystal structure of 62a43+(PMo12O40)3−; (d) π–π stacking in the crystal structure of 62a43+(PMo12O40)3−. Reproduced with permission from ref. 110–112. Copyright 2020, 2023 and 2024 Wiley.

Benzannulated B,N-doped corannulenes 61a and 61b (Fig. 19) undergo one-electron oxidation with NO+SbF6 in the presence of NaBArF4 (ArF = 3,5-bis(trifluoromethyl)phenyl) to form mixed-valence salts 61a32(˙+)(BArF4)2 and 61b32(˙+)(BArF4)2. These salts exhibit high thermal stabilities under an inert atmosphere and could be stored for several weeks at room temperature.111 In these crystals, both 61a and 61b exist as π-stacked trimers, 61a32(˙+) and 61b32(˙+) (Fig. 20b), with intra-trimer π–π distances of 3.28–3.29 Å and 3.34–3.35 Å, respectively. These π-stacked trimers further stack to form a column with intertrimer distances of 3.50 Å for 61a and 3.50 Å for 61b. EPR spectroscopy and DFT calculations indicate that 61a32(˙+) and 61b32(˙+) possess an open-shell singlet ground state, with the majority of the unpaired electrons located on the B–N bonds. As measured from the compressed pellets under a pressure of 3 MPa, 61a32(˙+) and 61b32(˙+) exhibit low conductivity of 2.48 × 10−7 S cm−1 and 3.44 × 10−7 S cm−1, respectively.

Trithiasumanene 62a and triselenasumanene 62b (Fig. 19) form four mixed-valence salts via electrocrystallization: 62a53+(BArF4)3, 62b53+(BArF4)3, 62a43+(PMo12O40)3− and 62b43+(PMo12O40)3− (ArF = 3,5-bis(trifluoromethyl)phenyl).112 In the BArF4 salts, the BArF4 anions organize into hexagonal channels that accommodate 1D π-stacked columns of 62a53+ or 62b53+. The crystal structure of 62a53+(BArF4)3 contains 10 crystallographically independent 62a molecules, each adopting a shallow bowl-shaped conformation with bowl depths ranging from 0.51 to 0.73 Å and Mulliken charges varying from +0.24 to +1.10. These units pack in a convex–concave arrangement with an average π–π distance of 3.52 Å. The crystal structure of 62b53+(BArF4)3 contains 5 crystallographically independent molecules of 62b, which are all nearly flat and have Mulliken charges varying from +0.41 to +1.0. In the crystal structure of 62a43+(PMo12O40)3− (Fig. 20c), three crystallographically independent 62a molecules (labeled A, B, and C in Fig. 20d) adopt a flat π-framework with Mulliken charges of +0.99, +0.65, and +0.71, respectively. They assemble into π-stacked columns with a repeating ⋯A–B–C–B⋯ pattern and an average interplanar distance of 3.42 Å. Temperature-dependent electrical conductivity, measured on single crystals using a four-probe method, indicates that all four salts behave as semiconductors with thermally activated conductivity. 62b43+(PMo12O40)3− exhibit room temperature conductivity of 3.5 × 10−3 S cm−1, while other three mixed-valence salts exhibit conductivity around 1.5 × 10−4 S cm−1. The conductivity observed in the mixed-valence salt of 62b is closely related to the p-type semiconductor behavior exhibited by a butylated derivative of 62b in single-crystal transistors, which achieves a field-effect mobility of 0.37 cm2 V−1 s−1.113

5. Conclusions and outlook

The exploration of robust, fully characterized π-conjugated radical cations reveals a clear evolutionary path in their stabilization and application. While early strategies relied heavily on steric protection to isolate radical cations of simple polycyclic arenes, this approach often comes at the cost of electronic communication in the solid state, limiting its utility for functional materials. A more powerful paradigm has emerged through the use of π-units, such as phenalenyl, flattened triarylamines, and TTF, which confer thermodynamic stability by efficiently delocalizing spin and charge. This fundamental insight has been pivotal, enabling the development of organic conductors and semiconductors with high conductivity114 without sacrificing intermolecular interactions. The recent identification of promising new motifs, such as double (hetero)helicenes and conjugated macrocycles, further enriches this toolkit for intrinsic stabilization.

In p-type organic semiconductors, the neutral molecule represents the undoped state, enabling high on–off ratios in OFETs, while the radical cations (including mixed-valence states) constitute the doped, higher-conductivity state. However, studies on the neutral and radical cation forms of the same π-system are often conducted in isolation. Only a handful of systems, such as TMTP (Fig. 1), 17a (Fig. 6), and 59a (Fig. 19), have been investigated in both states, typically using field-effect transistors for the neutral form and two-terminal devices for the radical cation. Bridging these largely separate research domains is a promising direction for future research. π-Systems proven in conductive radical cation salts could inspire design of new high-mobility semiconductors for OFETs. Simultaneously, robust radical cations from high-mobility neutral semiconductors offer a direct pathway to novel, highly conductive organic materials. A concerted effort to explore both states of the same molecular core will provide deeper insights into organic semiconductor design and accelerate the development of new materials.

Author contributions

S. Su and Q. Miao wrote the manuscript, and Q. Miao revised the manuscript according to the reviewers' comments.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results have been included, and no new data were generated or analyzed as a part of this review article.

Acknowledgements

This research was supported by the Research Grants Council of Hong Kong (CRF C4001-23G) and the State Key Laboratory of Synthetic Chemistry.

Notes and references

  1. N. L. Bauld, in Radicals, Ion Radicals, and Triplets, Wiley-VCH, New York, 1997, p. 141 Search PubMed.
  2. Y. Shirota and H. Kageyama, Chem. Rev., 2007, 107, 953–1010 CrossRef CAS PubMed.
  3. F. Wudl, G. M. Smith and E. J. Hufnagel, J. Chem. Soc. D, 1970, 1453–1454 RSC.
  4. F. Wudl, D. Wobschall and E. J. Hufnagel, J. Am. Chem. Soc., 1972, 94, 670–672 CrossRef CAS.
  5. H. Urayama, H. Yamochi, G. Saito, K. Nozawa, T. Sugano, M. Kinoshita, S. Sato, K. Oshima, A. Kawamoto and J. Tanaka, Chem. Lett., 1988, 17, 55–58 CrossRef.
  6. B. Dhara, M. Nakamura, K. Bulgarevich and K. Takimiya, Cryst. Growth Des., 2024, 24, 5826–5833 CrossRef CAS.
  7. G. Heywang and S. Roth, Angew Chem. Int. Ed. Engl., 1991, 30, 176–177 CrossRef.
  8. K. Takimiya, K. Bulgarevich, M. Abbas, S. Horiuchi, T. Ogaki, K. Kawabata and A. Ablat, Adv. Mater., 2021, 33, e2102914 CrossRef PubMed.
  9. U. Geiser and J. A. Schlueter, Chem. Rev., 2004, 104, 5203–5242 CrossRef CAS PubMed.
  10. T. Sugawara, H. Komatsu and K. Suzuki, Chem. Soc. Rev., 2011, 40, 3105–3118 RSC.
  11. Z. Mi, P. Yang, R. Wang, J. Unruangsri, W. Yang, C. Wang and J. Guo, J. Am. Chem. Soc., 2019, 141, 14433–14442 CrossRef CAS PubMed.
  12. M. Imran, C. M. Wehrmann and M. S. Chen, J. Am. Chem. Soc., 2020, 142, 38–43 CrossRef CAS PubMed.
  13. Z. X. Chen, Y. Li and F. Huang, Chem, 2021, 7, 288–332 CAS.
  14. D. Yuan, W. Liu and X. Zhu, Chem, 2021, 7, 333–357 CAS.
  15. S. Dong and Z. Li, J. Mater. Chem. C, 2022, 10, 2431–2449 RSC.
  16. L. Ji, J. Shi, J. Wei, T. Yu and W. Huang, Adv. Mater., 2020, 32, e1908015 CrossRef PubMed.
  17. Q. Peng, A. Obolda, M. Zhang and F. Li, Angew. Chem., Int. Ed., 2015, 54, 7091–7095 CrossRef CAS PubMed.
  18. W. Zeng and J. Wu, Chem, 2021, 7, 358–386 CAS.
  19. B. Tang, J. Zhao, J. F. Xu and X. Zhang, Chem. Sci., 2020, 11, 1192–1204 RSC.
  20. Z. Cui, A. Abdurahman, X. Ai and F. Li, CCS Chem., 2020, 2, 1129–1145 CrossRef CAS.
  21. F. Tani, M. Narita and T. Murafuji, ChemPlusChem, 2020, 85, 2093–2104 CrossRef CAS PubMed.
  22. T. Harimoto and Y. Ishigaki, ChemPlusChem, 2022, 87, e202200013 CrossRef CAS PubMed.
  23. W. Wang, P. Sun, X. Liu, X. Zhang, L. Zhang, Y. Z. Tan and X. Wang, Org. Lett., 2024, 26, 1017–1021 CrossRef CAS PubMed.
  24. P. T. Herwig, V. Enkelmann, O. Schmelz and K. Müllen, Chem. – Eur. J., 2000, 6, 1834–1839 CrossRef CAS PubMed.
  25. M. Banerjee, S. V. Lindeman and R. Rathore, J. Am. Chem. Soc., 2007, 129, 8070–8071 CrossRef CAS PubMed.
  26. J. Guo, C. Zhou, S. Xie, S. Luo, T. Y. Gopalakrishna, Z. Sun, J. Jouha, J. Wu and Z. Zeng, Chem. Mater., 2020, 32, 5927–5936 CrossRef CAS.
  27. H. Hayashi, J. E. Barker, A. Cardenas Valdivia, R. Kishi, S. N. MacMillan, C. J. Gomez-Garcia, H. Miyauchi, Y. Nakamura, M. Nakano, S. I. Kato, M. M. Haley and J. Casado, J. Am. Chem. Soc., 2020, 142, 20444–20455 CrossRef CAS PubMed.
  28. T. Nishinaga, K. Komatsu, N. Sugita, H. J. Lindner and J. Richter, J. Am. Chem. Soc., 1991, 115, 11642–11643 CrossRef.
  29. P. L. Magueres, S. V. Lindeman and J. K. Kochi, Org. Lett., 2000, 2, 3567–3570 CrossRef CAS PubMed.
  30. T. Horibe, S. Ohmura and K. Ishihara, J. Am. Chem. Soc., 2019, 141, 1877–1881 CrossRef CAS PubMed.
  31. Y. Zhang, S. H. Pun and Q. Miao, Chem. Rev., 2022, 122, 14554–14593 CrossRef CAS PubMed.
  32. T. Kubo, Y. Goto, M. Uruichi, K. Yakushi, M. Nakano, A. Fuyuhiro, Y. Morita and K. Nakasuji, Chem.–Asian J., 2007, 2, 1370–1379 CrossRef CAS PubMed.
  33. C. M. Wehrmann, R. T. Charlton and M. S. Chen, J. Am. Chem. Soc., 2019, 141, 3240–3248 CrossRef CAS PubMed.
  34. G. Xie, V. Brosius, J. Han, F. Rominger, A. Dreuw, J. Freudenberg and U. H. F. Bunz, Chem. – Eur. J., 2020, 26, 160–164 CrossRef CAS PubMed.
  35. C. Sato, S. Suzuki, K. Okada and M. Kozaki, Chem.–Asian J., 2018, 13, 3729–3736 CrossRef CAS PubMed.
  36. M. Harada, M. Tanioka, A. Muranaka, T. Aoyama, S. Kamino and M. Uchiyama, Chem. Commun., 2020, 56, 9565–9568 RSC.
  37. Y. K. Loh, P. Vasko, C. McManus, A. Heilmann, W. K. Myers and S. Aldridge, Nat. Commun., 2021, 12, 7052 CrossRef CAS PubMed.
  38. M. Narita, T. Teraoka, T. Murafuji, Y. Shiota, K. Yoshizawa, S. Mori, H. Uno, S. Kanegawa, O. Sato, K. Goto and F. Tani, Bull. Chem. Soc. Jpn., 2019, 92, 1867–1873 CrossRef CAS.
  39. C. Zhu, K. Shoyama and F. Wurthner, Angew. Chem., Int. Ed., 2020, 59, 21505–21509 CrossRef CAS PubMed.
  40. Y. Wang, Q. Gong, S. H. Pun, H. K. Lee, Y. Zhou, J. Xu and Q. Miao, J. Am. Chem. Soc., 2022, 144, 16612–16619 CrossRef CAS PubMed.
  41. L. Zhang, M. Gao, S. Su, Z. Zhou, H. K. Lee, X. Chen, Z. Huang and Q. Miao, Chem. – Eur. J., 2025, 31, e202501062 CrossRef CAS PubMed.
  42. A. Ito, Y. Ono and K. Tanaka, Angew. Chem., Int. Ed., 2000, 39, 1072–1075 CrossRef CAS PubMed.
  43. C. Wurster and E. Schobig, Ber. Dtsch. Chem. Ges., 1879, 12, 1807–1813 CrossRef.
  44. E. T. Seo, R. F. Nelson, J. M. Fritsch, L. S. Marcoux, D. W. Leedy and R. N. Adams, J. Am. Chem. Soc., 1966, 88, 3498–3503 CrossRef CAS.
  45. M. Kuratsu, M. Kozaki and K. Okada, Angew. Chem., Int. Ed., 2005, 44, 4056–4058 CrossRef CAS PubMed.
  46. S. I. Kato, T. Matsuoka, S. Suzuki, M. S. Asano, T. Yoshihara, S. Tobita, T. Matsumoto and C. Kitamura, Org. Lett., 2020, 22, 734–738 CrossRef CAS PubMed.
  47. S. Kataoka, S. Suzuki, Y. Shiota, K. Yoshizawa, T. Matsumoto, M. S. Asano, T. Yoshihara, C. Kitamura and S. I. Kato, J. Org. Chem., 2021, 86, 12559–12568 CrossRef CAS PubMed.
  48. J. Borstelmann, V. Gensch, D. Fehn, M. E. Miehlich, F. Hampel, F. Rominger, K. Meyer and M. Kivala, Angew. Chem., Int. Ed., 2025, 64, e202423802 CrossRef CAS PubMed.
  49. X. Zheng, X. Wang, Y. Qiu, Y. Li, C. Zhou, Y. Sui, Y. Li, J. Ma and X. Wang, J. Am. Chem. Soc., 2013, 135, 14912–14915 CrossRef CAS PubMed.
  50. T. A. Schaub, T. Mekelburg, P. O. Dral, M. Miehlich, F. Hampel, K. Meyer and M. Kivala, Chem. – Eur. J., 2020, 26, 3264–3269 CrossRef CAS PubMed.
  51. D. Sakamaki, D. Kumano, E. Yashima and S. Seki, Chem. Commun., 2015, 51, 17237–17240 RSC.
  52. K. Harada, C. Hasegawa, T. Matsumoto, H. Sugishita, C. Kitamura, S. Higashibayashi, M. Hasegawa, S. Suzuki and S. I. Kato, Chem. Commun., 2023, 59, 1301–1304 RSC.
  53. D. Sakamaki, S. Tanaka, K. Tanaka, M. Takino, M. Gon, K. Tanaka, T. Hirose, D. Hirobe, H. M. Yamamoto and H. Fujiwara, J. Phys. Chem. Lett., 2021, 12, 9283–9292 CrossRef CAS PubMed.
  54. D. Sakamaki, H. Sekiguchi, S. Suzuki and H. Fujiwara, Chem. – Eur. J., 2025, 31, e202500942 CrossRef CAS PubMed.
  55. L. Ruan, R. Li, M. Li, Y. Huang and P. An, J. Org. Chem., 2025, 90, 4365–4373 CrossRef CAS PubMed.
  56. G. Xie, N. M. Bojanowski, V. Brosius, T. Wiesner, F. Rominger, J. Freudenberg and U. H. F. Bunz, Chem. – Eur. J., 2021, 27, 1976–1980 CrossRef CAS PubMed.
  57. B. Huang, H. Kang, X.-L. Zhao, H.-B. Yang and X. Shi, Cryst. Growth Des., 2022, 22, 3587–3593 CrossRef CAS.
  58. X. Gu, B. Shan, Z. He and Q. Miao, ChemPlusChem, 2017, 82, 1034–1038 CrossRef CAS PubMed.
  59. L. Zhang, Y. Zhao, J. Li, Y. Fu, B. Peng, J. Yang, X. Lu and Q. Miao, J. Am. Chem. Soc., 2025, 147, 3459–3467 CrossRef CAS PubMed.
  60. L. Li, Y. Su, Y. Ji and P. Wang, J. Am. Chem. Soc., 2023, 145, 5778–5785 CrossRef CAS PubMed.
  61. H.-P. J. de Rouville, C. Gourlaouen, D. Bardelang, N. Le Breton, J. S. Ward, L. Ruhlmann, J. M. Vincent, D. Jardel, K. Rissanen, J. L. Clement, S. Choua and V. Heitz, J. Am. Chem. Soc., 2025, 147, 1823–1830 CrossRef PubMed.
  62. Q. Qi, P. M. Burrezo, H. Phan, T. S. Herng, T. Y. Gopalakrishna, W. Zeng, J. Ding, J. Casado and J. Wu, Chem. – Eur. J., 2017, 23, 7595–7606 CrossRef CAS PubMed.
  63. Y. Matsuo, T. Tanaka and A. Osuka, Chem. – Eur. J., 2020, 26, 8144–8152 CrossRef CAS PubMed.
  64. H. Yokoi, S. Hiroto and H. Shinokubo, J. Am. Chem. Soc., 2018, 140, 4649–4655 CrossRef CAS PubMed.
  65. T. Nishinaga, A. Wakamiya, D. Yamazaki and K. Komatsu, J. Am. Chem. Soc., 2004, 126, 3163–3174 CrossRef CAS PubMed.
  66. D. Yamazaki, T. Nishinaga, N. Tanino and K. Komatsu, J. Am. Chem. Soc., 2006, 128, 14470–14471 CrossRef CAS PubMed.
  67. T. Fujiwara, A. Muranaka, T. Nishinaga, S. Aoyagi, N. Kobayashi, M. Uchiyama, H. Otani and M. Iyoda, J. Am. Chem. Soc., 2020, 142, 5933–5937 CrossRef CAS PubMed.
  68. T. Fujiwara, M. Takashika, M. Hasegawa, Y. Ie, Y. Aso, S. Aoyagi, H. Otani and M. Iyoda, ChemPlusChem, 2019, 84, 694–703 CrossRef CAS PubMed.
  69. Y. Kobayashi, T. Terauchi, S. Sumi and Y. Matsushita, Nat. Mater., 2017, 16, 109–114 CrossRef CAS PubMed.
  70. M. Vicent-Morales, M. Esteve-Rochina, J. Calbo, E. Orti, I. J. Vitorica-Yrezabal and G. Minguez Espallargas, J. Am. Chem. Soc., 2022, 144, 9074–9082 CrossRef CAS PubMed.
  71. A. Trabolsi, N. Khashab, A. C. Fahrenbach, D. C. Friedman, M. T. Colvin, K. K. Coti, D. Benitez, E. Tkatchouk, J. C. Olsen, M. E. Belowich, R. Carmielli, H. A. Khatib, W. A. Goddard, M. R. Wasielewski and J. F. Stoddart, Nat. Chem., 2010, 2, 42–49 CrossRef CAS PubMed.
  72. A. C. Fahrenbach, J. C. Barnes, D. A. Lanfranchi, H. Li, A. Coskun, J. J. Gassensmith, Z. Liu, D. Benitez, A. Trabolsi, W. A. Goddard, M. Elhabiri and J. F. Stoddart, J. Am. Chem. Soc., 2012, 134, 3061–3072 CrossRef CAS PubMed.
  73. A. C. Fahrenbach, S. Sampath, D. J. Late, J. C. Barnes, S. L. Kleinman, N. Valley, K. J. Hartlieb, Z. Liu, V. P. Dravid, G. C. Schatz, R. P. Van Duyne and J. F. Stoddart, ACS Nano, 2012, 6, 9964–9971 CrossRef CAS PubMed.
  74. Y. Jiao, Y. Qiu, L. Zhang, W. G. Liu, H. Mao, H. Chen, Y. Feng, K. Cai, D. Shen, B. Song, X. Y. Chen, X. Li, X. Zhao, R. M. Young, C. L. Stern, M. R. Wasielewski and R. D. Astumian, Nature, 2022, 603, 265–270 CrossRef CAS PubMed.
  75. Y. Jiao, H. Mao, Y. Qiu, G. Wu, H. Chen, L. Zhang, H. Han, X. Li, X. Zhao, C. Tang, X. Y. Chen, Y. Feng, C. L. Stern, M. R. Wasielewski and J. F. Stoddart, J. Am. Chem. Soc., 2022, 144, 23168–23178 CrossRef CAS PubMed.
  76. H. Han, Y. Huang, C. Tang, Y. Liu, M. D. Krzyaniak, B. Song, X. Li, G. Wu, Y. Wu, R. Zhang, Y. Jiao, X. Zhao, X. Y. Chen, H. Wu, C. L. Stern, Y. Ma, Y. Qiu, M. R. Wasielewski and J. F. Stoddart, J. Am. Chem. Soc., 2023, 145, 18402–18413 CrossRef CAS PubMed.
  77. L. Mao, M. Zhou, T. Wu, D. Ma, G. Dai and X. Shi, Org. Lett., 2024, 26, 7244–7248 CrossRef CAS PubMed.
  78. H. Morishita, K. Sambe, S. Dekura, H. Sugishita, C. Kitamura, T. Matsumoto, T. Takeda, R. Uesugi, T. Ishida, T. Akutagawa, S. Suzuki and S. I. Kato, Chem. – Eur. J., 2025, 31, e202500576 CrossRef CAS PubMed.
  79. J. Borstelmann, S. Zank, M. Krug, G. Berger, N. Frohlich, G. Glotz, F. Gnannt, L. Schneider, F. Rominger, F. Deschler, T. Clark, G. Gescheidt, D. M. Guldi and M. Kivala, Angew. Chem., Int. Ed., 2025, 64, e202423516 CrossRef CAS PubMed.
  80. R. Kurata, D. Sakamaki and A. Ito, Org. Lett., 2017, 19, 3115–3118 CrossRef CAS PubMed.
  81. W. Wang, L. Wang, S. Chen, W. Yang, Z. Zhang and X. Wang, Sci. China Chem., 2017, 61, 300–305 CrossRef.
  82. W. Wang, C. Chen, C. Shu, S. Rajca, X. Wang and A. Rajca, J. Am. Chem. Soc., 2018, 140, 7820–7826 CrossRef CAS PubMed.
  83. S. Dong, T. Y. Gopalakrishna, Y. Han and C. Chi, Angew. Chem., Int. Ed., 2019, 58, 11742–11746 CrossRef CAS PubMed.
  84. Y. Ma, Y. Han, X. Hou, S. Wu and C. Chi, Angew. Chem., Int. Ed., 2024, 63, e202407990 CrossRef CAS PubMed.
  85. Z. Li, X. Hou, Y. Han, W. Fan, Y. Ni, Q. Zhou, J. Zhu, S. Wu, K. W. Huang and J. Wu, Angew. Chem., Int. Ed., 2022, 61, e202210697 CrossRef CAS PubMed.
  86. Y. D. Yang, M. Leng, Q. Zhang, X. Jin, C. V. Chau, J. Yang, S. Vasylevskyi, G. Henkelman, H. Y. Gong, L. Fang and J. L. Sessler, J. Am. Chem. Soc., 2025, 147, 19364–19371 CrossRef CAS PubMed.
  87. Y. D. Yang, Q. Zhang, L. Khrouz, C. V. Chau, J. Yang, Y. Wang, C. Bucher, G. Henkelman, H. Y. Gong and J. L. Sessler, ACS Cent. Sci., 2024, 10, 1148–1155 CrossRef CAS PubMed.
  88. H. Wei, X. Hou, T. Xu, Y. Zou, G. Li, S. Wu, Y. Geng and J. Wu, Angew. Chem., Int. Ed., 2022, 61, e202210386 CrossRef CAS PubMed.
  89. S. Arikawa, A. Shimizu, D. Shiomi, K. Sato, T. Takui, H. Sotome, H. Miyasaka, M. Murai, S. Yamaguchi and R. Shintani, Angew. Chem., Int. Ed., 2023, 62, e202302714 CrossRef CAS PubMed.
  90. M. Ahmed, Y. Wu, M. R. Schiavone, K. Lang, L. You, M. Zeller and J. Mei, Org. Lett., 2023, 25, 6363–6367 CrossRef CAS PubMed.
  91. S. Tang, L. Zhang, H. Ruan, Y. Zhao and X. Wang, J. Am. Chem. Soc., 2020, 142, 7340–7344 CrossRef CAS PubMed.
  92. H. P. Fritz, H. Gebauer, P. Friedrich, P. Ecker, R. Artes and U. Schubert, Z. Naturforsch., B: Chem. Sci., 1978, 33, 498–506 CrossRef.
  93. G. Saito and T. Murata, Philos. Trans. R. Soc., A, 2008, 366, 139–150 CrossRef CAS PubMed.
  94. A. Schätzle, J. U. Von SchÜTz, H. C. Wolf, H. Schäfer and H. W. Helberg, Mol. Cryst. Liq. Cryst., 1985, 120, 229–232 CrossRef.
  95. H. J. Keller, D. Nöthe, H. Pritzkow, D. Wehe, M. Werner, P. Koch and D. Schweitzer, Mol. Cryst. Liq. Cryst., 1980, 62, 181–199 CrossRef CAS.
  96. P. Koch, D. Schweitzer, R. H. Harms, H. J. Keller, H. Schäfer, H. W. Helberg, R. Wilckens, H. P. Geserich and W. Ruppel, Mol. Cryst. Liq. Cryst., 1982, 86, 87–101 CrossRef.
  97. H. Endres, H. J. Keller, B. Müller and D. Schweitzer, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1985, 41, 607–613 CrossRef.
  98. C. Kröhnke, V. Enkelmann and G. Wegner, Angew Chem. Int. Ed. Engl., 1980, 19, 912–919 CrossRef.
  99. J. Ferraris, D. O. Cowan, V. Walatka and J. H. Perlstein, J. Am. Chem. Soc., 1973, 95, 948–949 CrossRef CAS.
  100. R. Comès, S. M. Shapiro, G. Shirane, A. F. Garito and A. J. Heeger, Phys. Rev. Lett., 1975, 35, 1518–1521 CrossRef.
  101. T. E. Phillips, T. J. Kistenmacher, J. P. Ferraris and D. O. Cowan, J. Chem. Soc., Chem. Commun., 1973, 471–472 RSC.
  102. D. Jérome, A. Mazaud, M. Ribault and K. Bechgaard, J. Phys., Lett., 1980, 41, 95–98 CrossRef.
  103. K. Bechgaard, K. Carneiro, M. Olsen, F. B. Rasmussen and C. S. Jacobsen, Phys. Rev. Lett., 1981, 46, 852–855 CrossRef CAS.
  104. M. A. Christensen, C. R. Parker, T. J. Sørensen, S. de Graaf, T. J. Morsing, T. Brock-Nannestad, J. Bendix, M. M. Haley, P. Rapta, A. Danilov, S. Kubatkin, O. Hammerich and M. B. Nielsen, J. Mater. Chem. C, 2014, 2, 10428–10438 RSC.
  105. L. Feng, H. Dong, Q. Li, W. Zhu, G. Qiu, S. Ding, Y. Li, M. A. Christensen, C. R. Parker, Z. Wei, M. B. Nielsen and W. Hu, Sci. China Mater., 2016, 60, 75–82 CrossRef.
  106. J. Xu, Q. Chen, S. Li, J. Shen, P. Keoingthong, L. Zhang, Z. Yin, X. Cai, Z. Chen and W. Tan, Angew. Chem., Int. Ed., 2022, 61, e202202571 CrossRef CAS PubMed.
  107. G. Liu, L. Gao, Y. Han, Y. Xiao, B. Du, J. Gong, J. Hu, F. Zhang, H. Meng, X. Li, X. Shi, Z. Sun, J. Wang, G. Dai, C. Chi and Q. Wang, Angew. Chem., Int. Ed., 2023, 62, e202301348 CrossRef CAS PubMed.
  108. O. Matuszewska, T. Battisti, R. R. Ferreira, N. Biot, N. Demitri, C. Meziere, M. Allain, M. Salle, S. Manas-Valero, E. Coronado, E. Fresta, R. D. Costa and D. Bonifazi, Chem. – Eur. J., 2023, 29, e202203115 CrossRef CAS PubMed.
  109. N. Kobayashi, M. Sasaki and K. Nomoto, Chem. Mater., 2009, 21, 552–556 CrossRef CAS.
  110. L. Ethordevic, C. Valentini, N. Demitri, C. Meziere, M. Allain, M. Salle, A. Folli, D. Murphy, S. Manas-Valero, E. Coronado and D. Bonifazi, Angew. Chem., Int. Ed., 2020, 59, 4106–4114 CrossRef PubMed.
  111. Y. Gao, Z. Liu, T. Li and W. Zhao, Angew. Chem., Int. Ed., 2023, 62, e202314006 CrossRef CAS PubMed.
  112. L. Wu, Y. Li, X. Hua, L. Ye, C. Yuan, Z. Liu, H. L. Zhang and X. Shao, Angew. Chem., Int. Ed., 2024, 63, e202319587 CrossRef CAS PubMed.
  113. B. Fu, X. Hou, C. Wang, Y. Wang, X. Zhang, R. Li, X. Shao and W. Hu, Chem. Commun., 2017, 53, 11407–11409 RSC.
  114. M. Ikeda, Y. Sasaki, Y. Fujikawa, S. Mori, K. Konishi, K. Ohara, H. Dekura, H. Toyota, M. Takase, A. Mi Shirai, Y. Murotani, R. Matsunaga and T. Naito, J. Mater. Chem. C, 2025, 13, 12650–12656 RSC.

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