DOI:
10.1039/D5QO00746A
(Review Article)
Org. Chem. Front., 2025, Advance Article
Multistage organic redox systems. Aromaticity as a guiding principle for materials chemistry
Received
11th May 2025
, Accepted 7th July 2025
First published on 10th July 2025
Abstract
Organic π-conjugated molecules capable of undergoing multistage redox processes have emerged as versatile platforms for applications in energy storage, molecular electronics, and optical devices. This review highlights the critical role of aromaticity in shaping the redox behavior of such systems. Structural factors including topology of the ring system, heteroatom incorporation, and conjugation pathways are examined with respect to their influence on redox ranges, stability, and properties of oxidation levels. The discussion encompasses monomeric aromatic systems and their linked and fused covalent assemblies with linear, branched, and cyclic topologies. The survey provides a unifying perspective on multiredox design principles, offering insights for future development of functional organic redox materials.
 Marcin Stępień (left), Ishfaq A. Bhat (top), Ankit Kumar Gaur (right), Natasza Sprutta (center) | Marcin Stępień (PhD 2003, habilitation 2010, titular professorship 2017) works at the Faculty of Chemistry, University of Wrocław (UWr), Poland. His team studies the synthesis, properties, and theory of π-conjugated systems, including nanographene analogues, curved aromatics, and open-shell organics, with a focus on aromaticity-driven functions. In 2023, his contributions to the field were recognized with the Foundation for Polish Science Prize. Ankit Kumar Gaur (PhD 2023 IISER Mohali) and Ishfaq A. Bhat (PhD 2023, Hyderabad) have joined Marcin's research group at UWr as NCN postdoctoral fellows to develop next-generation multiredox macrocycles. Natasza Sprutta (PhD 2001) teaches organic chemistry at UWr; conducting research on porphyrin analogues and azulene-containing aromatics. |
1. Introduction
The concept of the oxidation state can be traced back to Friedrich Wöhler's 1835 textbook Unorganische Chemie,1,2 in which he used the term Oxydationsstufe (“oxidation grade”) to denote the stoichiometric ratios found in different oxides of the same element. While the existence of multiple oxidation states in inorganic compounds was recognized early on, the development of organic redox chemistry progressed more slowly. This delay was likely due to the structural complexity of organic compounds and the broader range of redox processes they can undergo. Over the course of the XIX century, it was appreciated that oxidations and reductions were associated with changes in the amount of either oxygen or hydrogen in an organic compound. However, although the first redox processes in which electron transfer was not tied to atom transfer were discovered around the same time, their full understanding was only achieved in the 20th century.3 For instance, Wurster's blue, containing the radical cation [S1.1]˙+ (Scheme 1), was discovered in 1879,4 but its identity as a discrete oxidation state of S1.1 was only confirmed in 1926.5–7
 |
| Scheme 1 Relationship between oxidation level and aromaticity in Weitz- and Wurster-type multistage redox systems according to Deuchert and Hünig's classification.7 Clar sextets are indicated in light red. The semiquinone [S1.1]˙+ is Wurster's blue. In the schemes and figures of this review, fully localized structures are typically shown for charged and radical states, to accurately represent the electron counts in the π systems. | |
In organic chemistry, systems comprising π bonds are the most effective electron donors and acceptors, and even relatively small π-conjugated molecules can exist in multiple redox states. However, the thermodynamic accessibility and properties of these states depend on the size, topology, and geometrical distortion of the π system, as well as on the inclusion of heteroatom dopants and peripheral substitution. All these structural features directly affect the key characteristics that define multiredox behavior, i.e., the HOMO–LUMO gap, redox potentials, and spectral properties of individual redox levels. These relationships are relevant to diverse classes of compounds and various fields of application, which are seldom discussed from a unified perspective. The objective of this review is to explore strategies developed across these diverse research areas to create multiredox π-conjugated molecules. We aim to provide a broader look at the relationship between multiredox behavior and π-conjugation with a particular focus on aromaticity, which can change dramatically upon either oxidation or reduction.
Our title is a reference to Deuchert and Hünig's classic review,7 which summarized the state of the field in the late 1970s, and formulated general structural principles underlying many of the then-known multiredox designs. Since then, the area has expanded considerably, driven by advances in synthetic organic chemistry and by growing interest in materials science. This progress is evident from some of the recently published reviews covering selected aspects of multiredox organics,8–11 and their application in such fields as energy storage and conversion,12–17 supramolecular chemistry,9,18 near-infrared absorbers,11 superconductivity,19 and electrochromism.20 The aim of this review is not to compile an exhaustive bibliography, but rather to provide a comprehensive overview of recent developments, emphasizing the diversity of molecular structures, properties, and applications. To that end, we focus primarily on literature from the past two decades, with occasional references to earlier key contributions for context.
We will begin by discussing general principles underlying the design of multistage redox systems. After that, we will take a detailed look at the most prominent classes of multiredox organic compounds. We will start with monocyclic aromatics, which not only provide a simple illustration of principles underlying multiredox behavior but also serve as structural components of larger assemblies. Different types of connectivity between redox-active building blocks, i.e., linear, two-dimensional, and cyclic, will be discussed in subsequent chapters. Because of our focus on aromaticity, the scope of the review will be limited to systems in which the multiredox behavior emerges in a contiguously conjugated π system. Covalent and non-covalent assemblies of multiple π-conjugated units, have been reviewed elsewhere,9,17,18 and will generally be excluded from this review, except for some recent examples of hyperconjugatively coupled systems.
In the context of this review, a multistage redox (multiredox) system is a species that has at least three accessible oxidation levels. We will avoid the ambiguous term “multivalent”, which is occasionally used to describe multiredox organics21,22 but can also refer to e.g. Mn+ cations (n > 1) in metal-ion batteries,16 or to supramolecular interactions.23 The terms redox state, oxidation state, and oxidation level will be used interchangeably, and will be identified with the charge of the species, provided that the composition of the system does not change upon reduction or oxidation. The latter rule may not hold when proton transfer or metal coordination occurs to (partially) compensate the change of total charge caused by the redox process, as often seen e.g. in porphyrin analogues (see section 5).
The redox range of an organic system encompasses all its accessible oxidation states. In the graphics illustrating this review, redox ranges are highlighted in rounded boxes accompanying molecular structures. The accurate determination of the redox range of an organic system is complicated by the different ways in which the existence and stability of an oxidation state can be determined. Ideally, all oxidation levels should be isolable as pure compounds; however, such a requirement would be overly restrictive given the experimental challenges associated with the isolation of highly charged organic ions. The redox ranges given in this review attempt to include all states whose stability is sufficiently documented by the reported chemical and electrochemical experiments performed in solution. It should be noted, however, that these two experimental approaches often produce somewhat different redox ranges, and in some cases, the different states may not be stable under a single set of conditions. Larger ranges are often observed electrochemically, however, the reversibility of the outer oxidations and reductions is not always clearly determined in the original papers. It can also happen that oxidation states that do not seem to be accessible via reversible electrochemical events, have been obtained using chemical means. If not indicated otherwise, all potentials are referenced to the ferrocene/ferrocenium couple (Fc/Fc+).
The multistage redox behavior of organic molecules most often relies on the ability of a π-electron system to act as a reservoir for electrons or holes. While the resulting redox ranges are occasionally spectacular, π-conjugated molecules usually show a strong preference for a single oxidation level, which is typically characterized by the neutral charge, closed-shell configuration, and aromaticity. The inability of a system to simultaneously satisfy all three requirements often leads to redox amphoterism, e.g. when the system can switch between a diradicaloid neutral state and an aromatic yet charged state. For larger systems, usually containing multiple rings, heteroatoms, labile hydrogens, multiple oxidation levels can be observed, which derive their stability from extensive charge and spin delocalization, favorable substituent effects, and changes in their aromatic character. The oxidation levels may differ only in the number of electrons, but in some cases, especially in porphyrinoids, the changes of total charge may be partly or completely compensated by protonation or deprotonation. Usually, the geometry and topology of the π system is insignificantly affected by the redox state, although exceptions are known. Planarization of the cyclooctatetraene dianion is a classic example; in more complex cases, recently termed dynamic redox (DYREX) systems,24 a change of the oxidation state can induce large torsional changes,24–26 or rearrangements of the ring system.27–29
Given their great diversity, it is hardly possible to provide a general an unambiguous structural classification of multiredox organic molecules. A simple, yet useful typology differentiates between Wurster- and Weitz-type multiredox organics.7 Wurster systems, named after Wurster's blue [S1.1]˙+, feature charge-carrying sites (“end groups”) outside the ring system, and are aromatic in the neutral state. Inverse Wurster systems (e.g. S1.2) also have exocyclic redox-active groups, but they gain aromaticity in the oxidized state. In contrast, Weitz systems contain endocyclic end groups and are quinoidal in the reduced state, as exemplified by viologen S1.3. S1.4 is an example of an inverse Weitz system which becomes quinoidal upon oxidation.
The classification of multiredox organics used in this review focuses on the subunit-level connectivity of the π system. Indeed, the majority of systems discussed herein can be interpreted as covalent assemblies of smaller π-conjugated units, often monocyclic, each with at least two accessible redox states. The subunits can be linked into linear, two-dimensional (branched, radial), or macrocyclic assemblies (Fig. 1). The coupling between subunits depends on the linker type (bold red lines), which may be a direct bond, consist of one or more atoms (e.g. methine or vinylene bridges). A similar classification can be used for some fused ring systems, for which, however, delineation of subunits may be somewhat arbitrary.
 |
| Fig. 1 Topology-based classification of multiredox π-conjugated molecules used in this review. Shaded rings represent π-conjugated subunits with arbitrary internal topology (number and size of fused rings), and heteroatom content. Linking bonds and fused ring edges are indicated in red and green, respectively. | |
The multiredox behavior critically depends on the interaction strength between subunits: in very weakly coupled systems, all subunits will be reduced or oxidized in unison. Conversely, in the limit of very strong coupling, achieved, e.g., by cross-conjugated linkers such as meso bridges in porphyrins, the number of accessible oxidation states is usually limited, because of the greater separation of MO levels. In such a case, the π system starts behaving as a single “larger subunit”. The multiredox behavior is however difficult to predict in general, because the interaction strength depends not only on the connectivity between the subunits but also on the oxidation state of the system. Furthermore, while the neutral state (and some of the limiting oxidized or reduced states) often feature closed-shell configurations and large energy gaps, the electronic structure of intermediate redox levels is often additionally complicated by their mixed-valence character.30,31
2. π-Conjugated monomers
Even though small aromatic molecules typically show a limited number of oxidation states, their redox chemistry is of considerable fundamental and practical importance. Small monocyclic aromatics serve as models for the behavior of more complex systems. Additionally, given their low molecular weights and ease of preparation, small molecules are suitable for charge-storage13 and electrochromic applications.20 Redox chemistry of monocyclic π-conjugated systems typically evolves from the 6-electron aromaticity found in benzene, cyclopentadienyl anion, and tropylium cation, with the stability of oxidized and reduced states being affected by their open-shell character or antiaromaticity.32,33 In this section we highlight some recent cases of multiredox behavior observed in simple, often monocyclic π-conjugated structures.
Benzene and its simple derivatives are inherently difficult to reduce or oxidize because of their aromaticity and large energy gaps. In sandwich complexes of d-electron metals, ηn-coordinated benzene rings are typically considered to be neutral ligands, while s-block metals such as magnesium can produce fairly localized bonding that disrupts π-conjugation.34 Highly charged benzene anions can, however, be stabilized by the more electropositive lanthanide cations. The dinuclear lanthanum(II) complex [2.1]−, reported by Lappert et al., contained a formally monoanionic benzene ring (Fig. 2), providing an early example of such stabilization.35 Benzene dianion complexes can be stabilized by both mono- and dinuclear complexes of various LnII/III ions, such as [2.2]− (ref. 36 and 37) and [2.3]2−,38 respectively. By judicious choice of ligands and metal ions, it was possible to obtain further inverse-sandwich complexes containing benzene tetraanion [C6H6]4− and related arenes as bridging ligands.39–43 Dithorium(IV) species such as 2.4 (Fig. 2) are chargeless and act as four-electron reductants toward reactive arenes such as anthracene. These species provide unique experimental insight into the properties of the benzene tetraanion, which as a 10π-electron system, is predicted to be Hückel-aromatic. In contrast, The 8 π-electron benzene dianion [C6H6]2− is expected to be Hückel-antiaromatic in its singlet state and Baird-aromatic in the triplet state. The two multiplicities are inherently close in energy and the ground state of the dianion is apparently controlled by the nature of the interacting cations and ancillary ligands. For instance, the YIII-containing dianion [S2.2-Y]2− (Scheme 2), reported by Long et al.,44 was found to be a ground-state singlet with a very small ΔEST = 0.023(4) kcal mol−1. In comparison, the gadolinium analogue [S2.2-Gd]2− exhibited an S = 6 ground state, which was interpreted as originating from antiferromagnetic coupling between the two Gd3+ ions (S = 7/2) and 3[C6H6]2− (S = 1). In contrast, the neutral complex S2.4-Y (Scheme 2), reported recently by Delano IV and Demir,45 contains a well defined singlet 1[C6H6]2−, displaying quinone-like distortion and paratropicity.
 |
| Fig. 2 Sandwich complexes of benzene anions. Countercations are not shown for clarity. | |
 |
| Scheme 2 Synthesis of dinuclear lanthanide complexes stabilizing a benzene dianion ring. Reagents and conditions: (a)44 1. M(CH2SiMe3)3(THF)2; 2. n KC8, n 18-crown-6; (b)45 20 equiv. KC8, C6H6. | |
Oxidation states of benzene and other small rings are occasionally “virtual”, i.e. they appear in specifically substituted or coordinated species, but have not been generated via a sequence of redox processes for a single derivative. Indeed, the chemistry of π-conjugated seven-membered carbocycles is largely dominated by the most stable oxidation level, the 6π-aromatic tropylium cation [3.1]+ (Fig. 3).46,47 The 7π-electron tropyl radical is unstable,48 but its fused oligocyclic analogues, such as 3.2a, can be isolated, some of them being redox-active.49–53 Similarly, the cycloheptatrienyl anion is rarely observed in solution chemistry: it can be stabilized by benzannulation, as in [3.3]−,54 or by substitution with electron withdrawing groups as in [3.4]−.55 The former system showed magnetic characteristics consistent with paratropicity of the seven-membered ring. In the crystal of the potassium salt K[3.4], the anion was non-planar, showing significant localization of one endocyclic double bond. The first evidence for the formation of the cycloheptatrienyl trianion [3.1]3− was obtained in 1977 by Bates et al.56 As a 10π-electron species, [3.1]3− is expected to be Hückel-aromatic, which however has been considered too unstable to exist without the stabilization provided by lithium cations.57 C7H7 rings bound to early transition metals have been described as 7-electron donors,58 however, the trianionic state is believed to exist in the inverse-sandwich lanthanide complexes with the general structure of 3.5.59–61 The elusive [C6H6]4− and [C7H7]3− ions may be seen as lower homologues of the more easily accessible cyclooctatetraene (COT) dianion [3.6]2−, obtainable directly via two-electron reduction of COT.62 COT dianions have enjoyed continued interest as ligands in lanthanide-based single-molecule magnets.63–65 The recently reported the diaza-COT derivative 3.7 could similarly be reduced to the corresponding dianion using potassium metal.29 However, the presence of two nitrogens made 3.7 susceptible to a reductive transannular contraction reaction, which was induced by a Co(I) complex used as an alternative reductant. Interestingly, similar bicyclo rearrangements can be induced in tetrabenzo-COTs by either two-electron oxidation27 or reduction28 (cf. 14.2, Fig. 14).
 |
| Fig. 3 Examples of charged states of 7- and 8-membered carbocycles. | |
Redox chemistry of higher annulenes was extensively investigated in the second half of the 20th century, greatly contributing to our current understanding of aromaticity and π-electron delocalization.66 This field still offers unexpected research opportunities, as illustrated by the recent reinvestigation of the reduction chemistry of [18]annulene (4.1, Fig. 4), carried out by the groups of Petrukhina and Anderson.67 The mono-, di- and tetraanions of 4.1 were obtained by reduction with lithium metal; 1H NMR spectroscopy showed that [4.1]2− and [4.1]4− were respectively para- and diatropic, in line with their 20π and 22π electron counts. The crystal structure of the Li4[4.1] salt revealed a quasi-dimeric sandwich structure, with five of the eight Li cations intercalated between the two annulene ligands. The new analysis further revealed, in contrast to the original report,68 that the (−2) and (−4) states of 4.1 adopt a C2v symmetric conformation.
 |
| Fig. 4 Top: Reduction chemistry of [18]annulene 4.1. Arbitrary localized valence structures are shown for [4.1]2− and [4.1]4−. Bottom: Single-crystal X-ray structure of Li8[4.1]2. The asymmetric unit (left, interstitial solvent, hydrogen atoms and minor disorder components omitted for clarity). Two orthogonal views on the sandwich part (right). Color code: C, grey; O, red and Li, blue. Adapted from ref. 67, licensed under a Creative Commons Attribution 4.0 International License (CC BY 4.0, https://creativecommons.org/licenses/by/4.0/). | |
Reduced states of benzenes can be stabilized by introduction of substituents that can delocalize the negative charge, thus compensating for the loss of aromatic stabilization. Aromatic carboxylic esters such as diethyl terephthalate 5.1a and tetraethyl pyromellitate 5.1b can be sequentially reduced to the corresponding radical anions and dianions, providing a working principle for multi-color electrochromic devices (Fig. 5 and 6).69 The radical anions [5.1a]˙− and [5.1b]˙− are respectively pink- and magenta-colored, whereas the dianion [5.1b]2− is yellow. A similar reduction sequence is thought to occur in pyromellitic diimide salts, which have been employed as electrode materials in Li-ion and Na-ion batteries.70,71 The initial diimide dianion [5.2]2− (Fig. 5) is aromatic and undergoes a two-electron reduction to the quinoidal [5.2]4− state. Interestingly, the latter structure may be considered to have some s-indacene conjugation, a feature that may affect the relative stability of the tetraanion. The performance of these imides was found to be better in Li-ion setups, while in their sodium-based counterparts, prolonged cycling resulted in faster decomposition, which was attributed to overreduction of the imides beyond the [5.2]4− state. The range of accessible oxidation levels can be extended by simultaneously substituting the system with electron-donating and electron-withdrawing groups. For instance, the tetraanion of 2,5-dihydroxyterephthalic acid, [5.3]4−, can be both reduced to the quinodimethane-like hexaanion [5.3]6− and oxidized to the benzoquinone dianion [5.3]2− (Fig. 5). These complementary capabilities were employed in the construction of a symmetrical rocking-chair sodium-ion cell, capable of delivering an average voltage of +1.8 V and energy density of ca. 65 Wh kg−1.72
 |
| Fig. 5 The use of electron-withdrawing and electron-donating groups for stabilization of charged oxidation states in simple benzene derivatives. | |
 |
| Fig. 6 Color changes of electrochromic devices containing 5.1a, 5.1b, 8.15, and 15.2 as a function of applied potential. Adapted with permission of the Royal Society of Chemistry, from ref. 69; permission conveyed through Copyright Clearance Center, Inc. | |
In contrast to the above carboxylic derivatives, which gain aromaticity upon reduction, the systems 5.473 and 5.5,74 reported by the group of Jana, are aromatic in the oxidized state, and can thus be classified as inverse Wurster species. The benzenoid state [5.4]2+ is additionally stabilized by delocalization of positive charges into the iminium groups away from the central ring, but it is easily reduced to the corresponding radical cation and neutral state at relatively high potentials. The dipolar 5.5 is a radical in the neutral state, with the spin delocalized across the aromatic backbone. The radical is easily oxidized to the cation [5.5]+, which has a conventional aminidium-like structure. In contrast, the monoanion [5.5]−, which was accessed by further reduction of 5.5 with KC8, is a ground-state singlet diradical with a ΔEST = –0.73 kcal mol−1.
Oxocarbons75 CnOn (S3.1-n, n = 3 to 6, Scheme 3) may be viewed as fully oxidized analogues of monocyclic hydrocarbons and as formal precursors of radialenes (section 4). Neutral oxocarbons tend to exist only in covalently hydrated forms, although syntheses of pristine S3.1–4, S3.1–5, and S3.1–6 were recently claimed.76,77 These results need to be taken with caution given the known thermodynamic instability of the lowest oxocarbons and the triplet ground state of S3.1–4.78,79 All known oxocarbons are most stable as dianions [S3.1-n]2− which are conjugate bases of the corresponding acids (deltic, squaric, croconic, and rhodizonic). The dianions are fully conjugated and symmetric, but aromatic stabilization is present only in the deltate dianion [S3.1–3]2−.80 The redox activity of the oxocarbons is complex and dependent on the ring size, but it has received considerable interest because of the possible application of these materials in energy storage.81,82 Oxocarbon-based batteries are attractive because they can potentially be made from renewable sources, but their capacity retentions are at present relatively low. While neutral states of some oxocarbons, e.g., S3.1–5,83 can be electrochemically generated in solution, charge storage is typically realized by reduction of the dianions to more highly charged states, i.e., [S3.1–5]4− and the presumably aromatic [S3.1–6]6−.77,84
 |
| Scheme 3 Structure and selected redox transformations of tripak (S3.2). Reagents and conditions:22 (a) Br2, DCM; (b) [FeIII(phen)3][PF6]3; (c) 1 equiv. KC8, THF; (d) 2.2 equiv. KC8, THF. | |
A remarkable way of converting a single benzenoid ring into a multiredox system was devised by Pakulski, Pinkowicz et al.22 Their design, called tripak (S3.2, Scheme 3) consists of three thiadiazole dioxide rings fused to a central C6 core, and was found to stabilize the (−2) oxidation level. Tripak can be electrochemically switched between oxidation levels from 0 to −6, in a potential window of ca. +0.7 V through −2.7 V (in THF, vs. Fc/Fc+). Oxidation levels from 0 through −4 were also generated chemically and isolated (Scheme 3). Computational investigations showed that S3.2 has three low-lying virtual orbitals, the upper two showing twofold degeneracy. As a consequence, while [S3.2]0 and [S3.2]2− were closed shell singlets, [S3.2]4− was predicted to be a ground-state triplet with a relatively small singlet–triplet gap (+2.3 kcal mol−1), in agreement with the experimentally observed paramagnetism of the tetraanion. On the basis of computational data, the 1[S3.2]0, 1[S3.2]2−, and 3[S3.2]4− states were classified as non-aromatic, weakly Hückel-aromatic, and Baird-aromatic, respectively. Bond length variations occurring during stepwise reduction suggested that the tripak π system is gradually transformed from the fully quinoidal conjugation in the neutral state toward benzenoid conjugation in the central ring in the negatively charged states. While the −6 oxidation state was not investigated, it is possible that it contains a fully aromatic benzene ring, which is however destabilized by the significant accumulation of negative charge at the periphery.
3. Linear systems
Linearly linked systems
One of the oldest-known redox-active aromatic molecules,85 the parent viologen system ([S4.1]2+, Scheme 4, various R groups), undergoes sequential one-electron reductions to the highly colored radical cation [S4.1]˙+ and the weakly colored neutral S4.1. The viologens were recognized early as useful redox indicators,86 and they have remained of high continuing interest because of their electrochromic and energy-storage applications.20,87,88 Viologen subunits have also been extensively explored as components of mechanically interlocked molecules,89 which are outside the scope of this review. The properties of the viologen can be modified by inserting conjugated linkers between the two pyridinium moieties, as in, e.g., [S4.2]2+ through [S4.11]2+ (Scheme 4).90–92 The reduced states of extended viologens show highly variable properties, dependent on the length and nature of the linker. Upon reduction to the neutral state, the linker switches from an aromatic (A) to a cross-conjugated or quinoidal form (Q, Scheme 4). For weakly coupled systems, the quinoidal neutral state may have a considerable open-shell character, and equilibrate between singlet and triplet configurations, as experimentally observed for S4.8.93
 |
| Scheme 4 Extended viologens and other linearly linked systems. The charges correspond only to the viologen units (charges located on R groups, if present, are not included). Clar sextets and their heterocyclic equivalents are indicated in light red. | |
An illustration of such dependences is provided by the series of bispyridinium electrolytes [S4.1]2+–[S4.10]2+ (R = CH2CH2CH2NMe3+, Cl− counteranions) developed by Grey, Scherman et al.92 These systems were employed in redox flow batteries, whose remarkable air tolerance was explained by the stabilizing effect of π-dimerization of viologen radical cations. In Scheme 4, these viologens are ordered according to the increasing singlet–triplet gap of the neutral state, which provides a measure of the interaction strength between the pyridine subunits. In electrochemical measurements, [S4.1]2+ and [S4.2]2+ showed two reversible one-electron reductions, characteristic of a strongly conjugated π system. A single two-electron reduction was observed for [S4.3]2+ through [S4.6]2+, in line with the greater separation between the pyridinium ends in these viologens. The remaining four systems showed non-reversible electrochemical reductions, suggesting relatively low stability of the neutral states, which was attributed to their open-shell character, which may be expressed both in the singlet and triplet state. In particular, S4.8 bears a structural similarity to the Chichibabin hydrocarbon, whereas the meta-linked S4.9 and S4.10 contain no closed-shell quinoidal forms and may be seen as analogues of the Schlenk hydrocarbon.
Because their redox properties are coupled to color changes, viologens can double as both electrolytes and electrochromic materials. For instance, when paired with a soluble TEMPO derivative, thiazolothiazole [S4.11]2+ (R = CH2CH2CH2NMe3+, Scheme 4) was used as a two-electron storage anolyte in all-organic aqueous redox flow batteries (Fig. 7).94 When substituted with chargeless groups, [S4.11]2+ was employed in electrochromic devices, wherein it was switchable from the yellow, and strongly fluorescent dicationic state to the deep-blue and non-emissive neutral form.95 A related case of electrofluorochromism was demonstrated for liquid-crystalline derivatives of [S4.5]2+, whose strong fluorescence in the bulk state, could be switched by application of reductive voltage or electric field.96
 |
| Fig. 7 Aqueous organic redox flow battery based on the extended viologen motif [S4.11]2+ (cf. Scheme 4). (A) Cyclic voltammograms of 4.0 mM [(NPr)2TTz]Cl4 and 4.0 mM NMe-TEMPO in 0.5 m NaCl solution. The gray dash curve is the cyclic voltammogram of only the 0.5 m NaCl electrolyte, with labels for the onset potentials for the hydrogen evolution reaction (HER, −1.00 V) and oxygen evolution reaction (OER, +1.50 V). The red and green dash curves are the fitted redox waves for the 1st and 2nd electron reductions, respectively. (B) Schematic representation of the [(NPr)2TTz]4+/NMe-TEMPO battery and its anodic and cathodic half-cell reactions. Reprinted with permission from ref. 94. Copyright © 2018 Wiley-VCH GmbH. | |
Ring fusion provides modified viologens that do not contain an extended conjugation pathway between the pyridinium units, but can benefit from the greater conformational rigidity and heteroatom effects imparted by the additional rings. Phosphaviologens [S4.12]2+ (R = alkyl, benzyl, aryl), developed by Baumgartner et al.,97,98 are reduced to the corresponding [S4.12]˙+ and S4.12 states at relatively high potentials (−0.60 V and −1.0 V, respectively, for R = Me), benefiting from the hyperconjugative electron-withdrawing effect of the phosphole oxide ring. The large potential separation of the two reductions enabled two-stage electrochromism: for instance, the benzyl-substituted [S4.12]2+ switches from colorless through blue-purple, to orange upon sequential electrochemical or chemical reduction. The similarly structured chalcogenoviologens [S4.13-E]2+ (E = S, Se, Te), showed a remarkable dependence of the optical energy gap on the chalcogen center (3.00, 2.78, and 2.33 eV, respectively, for R = Bn, TfO− counteranions).99 The latter systems could be used not only in electrochromic devices, but also for visible light-driven hydrogen evolution. In those experiments, the chalcogenoviologens played a dual role of a photosensitizer and an electron mediator.
Typically, in extended viologens, the accessible oxidation levels remain limited to 0, +1, and +2, although anionic states are occasionally observed, as in S4.8 (R = n-octyl), which was electrochemically reduced to [S4.8]˙− and [S4.8]2−.91 Limited redox ranges are typical in linear systems, as illustrated by Reynolds’ electrochromic oligomers, which consist of up to six thiophene or benzene subunits but were oxidized only to +1 and +2 states.100 Extending the redox range is possible by connecting a greater number of electroactive subunits, e.g. using radial designs described below. Linear multiredox systems can be obtained by judiciously linking multiple charged rings, as illustrated by [S4.14]4+, reported by Hansmann et al. (Scheme 4).101 The nearly colorless tetracation could be electrochemically reduced to the +3, +2, and 0 states, which were respectively green blue and purple in solution, thus providing a particularly rich electrochromic response. The radical trication [S4.14]˙3+ additionally absorbed in the near-infrared region, with the strongest maximum at 1513 nm). A phenanthroline analogue of [S4.14]4+ produced a complete range of discrete oxidation levels from 0 to +4, and showed high stability when tested as a potential anolyte material.
Given the redox amphotericity of many organic radicals, building conjugated oligoradical oligomers is an attractive route toward multiredox organics. The principal challenge of achieving sufficient chemical stability of such assemblies, was successfully addressed by Wu et al. in their work on fluorenyl oligomers S4.15–n (n = 1 through 6).102 In these systems, the inter-subunit coupling is ensured by quinoidal contributions, such as S4.15–2′ (Scheme 4). In electrochemical measurements, each of these systems showed n one-electron oxidations and n one-electron reductions, although in the longer oligomers some of these events considerably overlapped. In the limit of complete reduction to the [S4.15–n]n− level, each subunit bears a unit negative charge, and can be treated as an aromatic fluorenyl monoanion, decoupled from the adjacent subunits.
In a somewhat similar fashion, oligo(biindenylidene)s S4.16–n, developed by Fukazawa et al. as linearized substructures of C60,103 are capable of stabilizing up to one negative charge per indenyl subunit. The largest member of the series with n = 3, could be reduced up to the pentaanion [S4.16–3]5−. Further reduction was likely limited by the accessible potential range. The difference between the first two reductions decreased with the increasing length of the oligomer (+0.31 V, +0.15 V, and 0.00 V, for n = 1, 2, and 3, respectively), reflecting the diminishing interaction between the charges. The stability of the oligoanions can be attributed to locally aromatic 6π/10π-electron cyclopentadienyl/indenyl anion resonance contributions emerging upon reduction.
Linearly fused systems
Acenes are the simplest class of aromatics with one-dimensional fusion. Their redox chemistry typically covers oxidation levels from −2 to +2, and underlies potential applications of acene-based materials as semiconductors,104,105 superconductors,106 and quantum spin liquids.107 Linearly fused acenes (e.g., 8.1–4 through 8.1–7, Fig. 8) contain a single “migrating” Clar sextet regardless of the fusion length, a feature that contributes to the decreasing stability of the higher members of the series (n > 4). Since acenes are formally (4n + 2)-electron systems (where n coincides with number of fused rings), their dications could in principle be expected to show antiaromatic behavior associated with the 4n-electron count. However, two-electron oxidation of 8.1–4–8.1–7, which occurs spontaneously in fuming sulfuric acid, produces persistent and strongly diatropic dications [8.1–n]2+ (Fig. 8).108 This counterintuitive enhancement of aromaticity was rationalized by the presence of two Clar sextets in the dications. Dianions of higher acenes, e.g., [8.1–5]2− (ref. 109 and 110) and [8.1–6]2−,111 can be generated under strongly reducing conditions. In contrast to the corresponding dications, the dianions show global aromaticity, apparently originating from the peripheral (4n + 4)-electron circuit.
 |
| Fig. 8 Aromaticity changes in acenes and their derivatives. Red arrows indicate possible relocations of Clar sextet in the acene ring system. Experimental discharge voltages and experimental (theoretical) discharge capacities are given for 8.12–8.14. | |
A similar switching between acene and oligoquinone conjugation can be found in some linear heteroacenes, such as 5,6,11,12-tetraazanaphthacene 8.2. The latter species can be reduced to the dihydro derivative, fluoflavine,112 whose dianion [8.2]2− can act as a bridging ligand.113 The range of oxidation levels attainable in metal complexes of 8.2 extends from −1 to −3, as recently demonstrated by Benner and Demir.114 Dithiaacenes 8.3, 8.6, and 8.7, developed by Chi et al.,115 show a reversed charge–aromaticity relationship relative to the corresponding hydrocarbons. The former species are oligoquinoidal in their neutral states, with a progressively increasing diradicaloid character. Their dications, in contrast, are closed-shell species, isoelectronic with neutral acenes. Analogs of 8.3 containing selenium, nitrogen, and boron sites were reported respectively by the groups of Nagahora,116 Bunz (8.4),117 and Yamaguchi (8.8).118 8.4 has three oxidation states, isoelectronic with those of 8.3, with significant diatropicity of the radical cation [8.4]˙+ and dication [8.4]2+ determined in NICS calculations. In the B2O2 system 8.8, notable for its sharp NIR emission, low-lying oxidations (−0.07 V, +0.27 V) are observed in addition to reductions (−1.44 V, −1.83 V). In the latter system the stabilization of cationic and anionic states was associated with the oxygen and boron sites, respectively.
In the linear acene-like and oligoquinone conjugation patterns discussed above, the number of Clar sextets remains 1 and 2, respectively, regardless of the acene length. A larger number of Clar sextets can be produced by introduction of multiple benzoquinone units into the acene structure, leading to greater aromatic stabilization. This well-established strategy also works with benzoquinodimethane units, which were implemented by Ishigaki, Suzuki et al. in pentacene derivatives such as 8.5.25,26 The latter system shows an unusual hysteretic redox behavior: the neutral 8.5 is oxidized to the tetracation [8.5]4+ in a single four-electron step (+1.12 V vs. SCE, R = p-anisyl), whereas the reduction of the tetracation takes place in two two-electron steps (+0.55 and +0.27 V). This remarkable characteristic may originate from the significant conformational difference between the quinodimethane and dicarbocationic sections and the associated isomerization barriers.
The properties of acenes can be significantly altered by heteroatom doping or by substitution. A simple substitution with ester groups is sufficient to produce a useful shift of reduction potentials to higher voltages, as exemplified by the tetraester 8.15 which was found to be a useful electrochromic material.69 The nitrogen-doped azaacenes,119 are typically more electron deficient than the parent hydrocarbons, and are consequently of interest as n-type semiconductors. For instance, 5,7,12,14-tetraazapentacene 8.9120 shows two one-electron reductions at relatively high potentials (Ered1 = −0.79 V and Ered2 = −1.23 V, vs. Ered1 = −1.87 V for 8.1–5121) to produce a NIR-absorbing radical anion and a photostable and highly fluorescent dianion (λem = 598 nm, QY = 95%).122 Interestingly, while 8.9 turned out to be an excellent n-type organic semiconductor, with an electron mobility of up to 3.3 cm2 V−1 s−1, the related pentacene derivatives 8.10 and 8.11 showed respectively ambipolar and p-type behavior.123 8.11 is a doubly hydrogenated derivative of 8.10 and is thus not isoelectronic with pentacene. In fact, 8.11 is quite electron-rich, with HOMO and LUMO levels at −5.06 eV and −2.66 eV, respectively.
Acene-derived quinones have been considered as potential cathode materials for Li-ion batteries, as they can potentially provide higher discharge capacities than conventional inorganic cathodes. Their practical application is currently limited by relatively low discharge voltages and rapid capacity fade. For instance, anthraquinone 8.12, a simple and inexpensive material, produced a high discharge capacity in the first cycle (up to 251 mAh g−1), but the discharge voltage was only ca. +2.2 V.124 In 5,7,12,14-pentacenetetraone 8.13, which can be reduced to the tetraanionic state [8.13]4−, the number of quinone groups per mass unit is increased, leading to a higher discharge capacity; however, the attainable voltage (+2.1 V) was even lower than for anthraquinone.125 In both, 8.12 and 8.13, the neutral form of the quinone (corresponding to the charged state of the battery) is more aromatic than the fully reduced form, as can be inferred from the number of Clar sextets in both states. Chen et al. proposed that a higher discharge potential can be achieved if the relationship is reversed, i.e. if the aromaticity increases in the reduced (discharged) state.126 This assumption was confirmed experimentally by using 1,4,5,8-phenanthrenediquinone 8.14 as the cathode material, which yielded a discharge voltage of +2.77 V. As a consequence of the angular fusion present in 8.14, the number of Clar sextets increases from one to two upon reduction of the tetraanion [8.14]4−.
Reduced states of acenes can display unusual magnetic and electronic properties in the solid state.19 In 2010, Kubozono et al. reported the first hydrocarbon-based superconductor, Kx[8.17], obtained by annealing picene with potassium.106 The properties of the material strongly depended on the potassium content, varying from Pauli-like for x < 2, through superconducting, with a maximum Tc = 18 K obtained for x = 3.3, to Curie-like, for x ≥ 4. Superconductors based on doped phenanthrene, dibenzopentacene, and other hydrocarbons have also been reported.19 While their behavior is still incompletely understood, the doping factor x dramatically affects the solid-state interactions between the metal and the hydrocarbon.107,109 Significant changes can occur also at low doping levels, for instance phenanthrene salts Cs2[8.17] and Cs[8.17] are respectively diamagnetic and antiferromagnetic, the latter material displaying features of a quantum spin liquid.107
Reduction of the energy gap of an acene molecule may be seen as a way of simultaneously bringing the oxidation and reduction events into a chemically accessible potential regime. Such redox amphoterism is most readily achievable in open-shell derivatives, such as diindenoarenes.127,128 In such systems, each indene unit can be considered as a carrier of either positive or negative charge, potentially stabilizing oxidation states from (−2) to (+2). For smaller diindenoarenes, such as 8.18 (Eox1 = +0.13 V, Ered1 = −1.13 V),129 the stability of the (+2) and (−2) states is usually limited. A full range of expected oxidation states can however be observed in systems containing larger carbocyclic130–134 (e.g. 8.19130) or heterocyclic135,136 motifs. Similar behavior is observed in oligoradicaloids based on other structural motifs, e.g. phenalenyls.137
4. Two-dimensional systems
[6]Radialenes
A multiredox organic system can be built by radially connecting multiple redox-active units around a common “hub” moiety. The electronic structure such as assembly depends on the properties of the subcomponents and nature of the possible interactions between these redox units and the hub. Benzene is a particularly useful hub motif, offering the possibility of connecting six redox units in a symmetrical and chemically robust fashion. Vaid's [6]radialene S5.1138 provides an example of a strongly coupled system, which adopts radialene-like conjugation in its neutral state. S5.1, which revealed a chair-like conformation in the crystal structure, is air-sensitive. It was electrochemically oxidized to [S5.1]4+ and then to [S5.1]6+ in 4- and 2-electron processes, respectively (−1.33 V and −1.14). This behavior, notably the simultaneous transfer of four electrons, is unique, and is apparently related to the interplay between π-conjugation and conformational changes caused by oxidation. It is worth noting that in hexaferrocenylbenzene, which is expected to be weakly π-coupled across the hub, the electrochemical oxidation to the corresponding hexacation occurs in a sequence of 1-, 2- and 3-electron events.139
Hexaaniline S5.2, investigated by Kochi et al.,140,141 was sequentially oxidized to [S5.2]˙+, [S5.2]2+, [S5.2]3+, and [S5.2]6+ at potentials in the 0.5–0.9 V range (vs. SCE). The radical monocation of this species, was of particular interest as a potential example of toroidal delocalization of the hole among the six aniline arms (green arrows, Scheme 5).142 All cationic states [S5.2]n+ (n = 1 to 6) could apparently be observed in chemical oxidation experiments. A radialene-like conjugation can be envisaged for the hexacationic state [S5.2]6+, in an interesting opposition to S5.1, the latter having a radialene structure in its neutral state. Kochi's group also investigated the redox properties of hexacarbazolylbenzene S5.3,141,143 from which only the first four electrons could be removed reversibly. Because of the large torsional twist of the carbazoles relative to the central benzene, S5.3 is likely to exhibit limited π conjugation across the hub unit.
 |
| Scheme 5 [6]Radialenes and related structures based on benzene hubs. | |
Some features of the radialene design can be extended to other systems. Hexakis(guanidino)benzene S5.4, developed by Himmel et al.,144 contains no peripheral aromatic units and is a very strong electron donor. It is quadruply oxidized via two-electron low-potential events (−0.94, −0.41 V) and can form a pentacation at ca. +0.85 V. The latter oxidation level corresponds to a high theoretical charge capacity of 180 Ah kg−1, comparable with capacities achievable by Li-ion batteries (ca. 150–170 Ah kg−1).
The extent of coupling between substituents in radialene-like systems depends on their intrinsic redox properties, their number, and sterics. This dependence is illustrated by Diederich's charge-transfer systems S5.5 and S5.6:21 the former system showed significant differentiation of the first six reductions (−0.46 to −1.07 V), but further electrochemical reduction occurred as a single six-electron event (−1.57 V, Scheme 5). Similarly, a six-electron oxidation, likely involving the amino groups, was observed for S5.5 at +0.89 V. In the less densely substituted S5.6, there were three two-electron events, corresponding to the oxidation of the 1,2-di(1,3-dithiol-2-ylidene)ethane moieties (+0.41, +0.65, and +0.81 V) followed by a three-electron event at +0.96 V, likely involving the amine termini. Similarly to S5.5, compound S5.6 appeared to undergo sequential reduction, but these electrochemical events were somewhat less resolved in the latter case. Since S5.6 contains meta-connected arms and is relatively crowded, the differentiation of oxidation and reduction events may be largely caused by through-space electrostatic interactions, especially for the oxidations which occur closer to the benzene hub. When the charge is delocalized further away from the ring as in the oxidation of hexa- and tripyrrolylbenzenes S5.7 and S5.8, all TTF-pyrrole arms are oxidized almost simultaneously.145 In particular, S5.7 was oxidized to the NIR-absorbing hexa(radical cation) [S5.7]6+ and hexa(dication) [S5.7]12+. For S5.8, additional mixed-valence states were observed, which were attributed to the formation of aggregates, likely facilitated by the more planarized geometry of this species.
Non-benzenoid radialenes and related systems
The 2π-electron cyclopropenyl cation is the smallest entity exhibiting aromatic stabilization. Its diminutive π-conjugated system needs to be appropriately functionalized to exhibit multiredox behavior. Because of their small size and stability of their cationic state, such substituted cyclopropenyl cations have been researched as organic catholytes in redox-flow batteries, a field recently reviewed by Sigman, Sanford et al.146 The triaminocyclopropenyl cations typically used for this purpose ([9.1]+, Fig. 9) undergo just a single oxidation to the corresponding radical dication [9.1]˙2+, however, systems with a more extended redox behavior, such as [9.2]+,147 can provide higher energy densities.
 |
| Fig. 9 Non-benzenoid radialenes and related systems with a multiple redox states. Hückel-aromatic rings are shaded in color. | |
[3]Radialenes, the smallest members of the radialene family,148 provide another example of using cyclopropenyl cation contributions for stabilization of charged states. In particular, the electron-deficient hexacyano[3]radialene 9.3, reported by Fukunaga in 1976, can stabilize oxidation levels from 0 to −2 (Fig. 9).149,150 The dianion derives its unique stability from the aromaticity of the cyclopropenyl cation. 9.3 and related derivatives have been recently explored as soluble p-dopants in conductive polymer films,151 organic catholytes in redox flow batteries,152 and multimodal information switches.153 The electron-rich [3]radialenes 9.4a and 9.4b, reported by Matsumoto et al.,154 display four reversible one-electron oxidations and one reversible one-electron reduction in cyclic voltammograms. While detailed investigations of the tetracations [9.4a]4+ and [9.4b]4+ were not reported, their stability may likewise originate from the local aromaticity of the three-membered ring (Fig. 9). The deep-red [4]radialene 9.5, described by Horner and Hünig in 1977,155,156 can be oxidized up to the (+4) level. The initial oxidation is effectively a two-electron process (∼0.1 V vs. AgCl/Ag), leading to a stable dication [9.5]2+ characterized by a sharp red-shifted absorption. Formation of [9.5]4+ required a relatively high potential of +1.39 V (vs. AgCl/Ag), implying that the tetracation is strongly destabilized by the antiaromaticity of the central cyclobutadiene ring.
The radialene design principle can be applied to more complex core motifs and peripheral units, containing, e.g., multiple fused rings, triple bonds or heteroatoms. By judicious combination of two redox-active motifs, i.e. tetrathiafulvalene (TTF) and radiannulene, the group of Nielsen produced the hybrid system 9.6 with a significant redox range (Fig. 9).157 The system was oxidizable up to the tetracation level, relying on the ability of the TTF units to easily release two electrons each. The dianion [9.6]2− achieved its stability by producing an aromatic [14]annulene pathway in the macrocycle and delocalizing the negative charges in the outer branches of the oligoyne π system. The radialene-like 9.7, reported by the Yamaguchi group,158 contains an azatriangulene core and three dibenzo[c,g]fluorenylidene arms. The system has a relatively small energy gap of ca. 1.3 eV, and shows two reversible electrochemical one electron oxidations at potentials lower than 1 V. 9.7 undergoes a two-electron reduction at −1.13 V, which is followed by two one electron events.
Multiredox alkenes
The C
C double bond can be seen as the smallest conjugated entity enabling quasi-radial linking of multiple redox-active groups. Tetraarylethylenes bearing electron rich groups such as 10.1 (Fig. 10) can be reversibly oxidized to the corresponding radical cations and dications.159 The structurally related 9,9′-bifluorenylidene 10.2 forms an antiaromatic dication,160 as well as the anionic states, [10.2]˙− and [10.2]2−,161 the latter stabilized by fluorenyl anion-like aromaticity. The redox range of 10.2 can be extended by fusion of additional rings, as in the recently reported tetrafluorenofulvalene 10.3, for which all oxidation levels from (−4) to (+3) could be generated electrochemically.162 The cross-conjugated structure of 10.3 leads to a complex relationship between the oxidation state, aromaticity and the strength of the central carbon–carbon bond. For instance, in the [10.3]4− tetraanion, the central bond becomes the strongest, and the system behaves as a union of four aromatic fluorenyl anions. Conversely, in [10.3]2+, the two monocationic halves of the molecule become almost completely decoupled, each showing considerable antiaromaticity, reminiscent of the fluorenyl cation.
 |
| Fig. 10 Multiredox alkenes. | |
Compound 10.4, reported by Hein, Feringa et al.,24 provides an example of a dynamic redox (DYREX) system in which changes of the oxidation state are coupled to conformational dynamics of the overcrowded alkene bond. The neutral 10.4 adopts an anti-folded conformation. When electrochemically oxidized, it undergoes a three-step process consisting of (1) the formation of the transient anti-folded radical cation [10.4]˙+, (2) its rapid isomerization to the most stable twisted form, and (3) immediate oxidation to the dication [10.4]2+, resulting in an apparent 2-electron electrochemical event. The dication has an orthogonal conformation similar to the twisted conformation of [10.4]˙+, and redox switching between the two species occurs with a minimal steric constraint. The reduction of [10.4]˙+ to the neutral 10.4 involves a final conformational change, which makes the process electrochemically non-reversible.
Through-space conjugated systems
In [6]radialene systems, such as hexacarbazolyl S5.3 (Scheme 5), the redox behavior is affected by a fairly significant through space component, but the inner core of the radialene is, at least in principle, π-conjugated. Multiredox systems relying solely on through-space interactions have been constructed using triptycene and trinaphtho[3.3.3]propellane (TNP) cores (Fig. 11). The principle is illustrated by triptycene triquinone 11.1, explored by Kwon et al. as a cathode material for Li-ion batteries.163 This species undergoes three redox events at −0.58, −0.77, and −0.94 V vs. Ag/Ag+, corresponding consecutive reductions of the three quinone arms. Interestingly, the first reduction occurs at a higher potential than in benzoquinone (+0.79 V), suggesting that the radical anion [11.1]˙− is stabilized by electron sharing, while the relative lowering of the third reduction potential may originate from electrostatic repulsion. Further reduction of 11.1 (−1.44 and −1.59 V vs. Ag/Ag+) was limited to two one-electron events, suggesting that complete reduction of all three quinone units was thermodynamically unfeasible. In Li-ion half-cell tests, the overall transfer of up to five electrons was confirmed by the measured initial discharge capacity of 387 mAh g−1, very close to the theoretical value of 389 mAh g−1.
 |
| Fig. 11 Multiredox systems with radial through-space conjugation. | |
Triradical 11.2 consisting of three Blatter-radical units attached to a triptycene core (Fig. 11) was obtained by Matsuda, Shimizu, et al.,164 as two separable isomers, syn-11.2 and anti-11.2, with nearly identical properties. In electrochemical measurements, the C3v-symmetric syn-11.2 showed three-closely spaced oxidations (−0.08, −0.15, −0.27 V) and three reductions (−1.23, −1.31, −1.38 V), all corresponding to reversible events. For larger systems, such as the family of 3D nanostructures 11.3-n, described by Sisto et al.,165 the interaction between the three π-conjugated arms is hardly discernible. Thus, 11.3–3 underwent six consecutive electrochemical reductions in a small voltage window from approximately −1 to −2 V. Each of these events corresponded to simultaneous transfer of three electrons. The number of observed reduction events was 2n, for each member of the 11.3-n series, suggesting that each PDI units behaves as a 2-electron acceptor that is strongly coupled to other subunits in the ribbon.
The TNP triradical 11.4, reported by Kubo et al.,166 featured three reversible one-electron oxidations (−0.47, −0.02, +0.22 V) and three reductions (−1.46, −1.71, −2.00 V). The observed differentiation of potentials implies a relatively strong communication between the three phenalenyl subunits. The effectiveness of TNP as a hyperconjugative coupler unit has also been demonstrated in the triimide 11.5, developed by Zhang et al.167 This electron-deficient system consisting of three naphthalenemonoimide (NMI) arms showed three one-electron reductions, the first one occurring at −1.36 V. i.e. at a potential higher by +0.29 V than the reduction of the NMI parent. This difference was interpreted as a sign of additional stabilization of the monoanion by electron sharing between the three NMI subunits.
2D-fused hydrocarbons with peripheral conjugation
Two-dimensionally fused aromatics, often containing peri-condensed rings, provide more opportunities for multiredox behavior because of the typically large size of their π systems, and a greater number of potential functionalization sites. Furthermore, complex ring fusion patterns can lead to unusual conjugation effects, involving a combination of local and global contributions, which may variously affect the stability of neutral and charged states.
Corannulene 12.1, the first curved fullerene fragment synthesized, apparently has no stable oxidized states,168 but it can accept up to four electrons via both electrochemical and chemical reduction. Both the dianion [12.1]2− (ref. 169) and tetraanion [12.1]4− (ref. 170–172) can be described using the annulene-within-the-annulene (AWA) model, with the 6π-electron cyclopentadienide anion as the inner annulene circuit. In [12.1]2−, the outer circuit may be interpreted as the 16π-electron [15]annulene anion, in line with the paratropic character of the dianion. By analogy, the tetraanion [12.1]4− contains a 18π-electron [15]annulene-trianion peripheral pathway, which explains its diatropicity and unusual stability. [12.1]4− is capable binding multiple lithium cations in solution and in the solid state, indicating a potential utility of corannulene and related systems in lithium-ion batteries. Interestingly, the bicorannulenyl 12.2 forms a dianion [12.2]2− with two local six-electron Cp rings and an oligoene periphery. The tetraanion [12.2]4− behaves as a dimer of two corannulene dianions, as inferred from its paratropic character.
The bis-pentannulated bisanthene diradicaloid 12.3, reported by Chi et al.,174 showed stable oxidation states in the (−2,+2) range, forming at clearly separated redox potentials. The neutral state was globally paratropic and can formally be characterized as a 10π + 20π AWA system, although according to ACID calculations, the inner naphthalene substructure was negligibly aromatic. The inner 10π-electron circulation was similarly weak in the moderately aromatic closed-shell dianion [12.3]2−, which contains a 22π-electron periphery. In contrast, the diradicaloid dication [12.3]2+ featured strong diatropicity in 1H NMR, while the ACID analysis revealed the presence of both 10π and 18π circulations confirming an AWA structure. Similar relationships between the oxidation level and aromaticity were established for Wu's 12.4,175 a smaller homologue of 12.3. The magnetism of the neutral state appeared to combine features of a globally antiaromatic AWA contribution and a diradicaloid contribution with local benzenoid aromaticity. The dianion [12.4]2− showed moderate peripheral aromaticity with no significant inner 6π contribution. The [12.4]2+ dication, which was not experimentally accessible, showed AWA aromaticity analogous to that of [12.3]2+. Tetracyclopenta[def,jkl,pqr,vwx]tetraphenylene 12.5, described by Tobe et al., has an open-shell quinoid structure of D2h symmetry, and does not exhibit AWA conjugation.176 In the doubly charged states, the AWA conjugation is predicted by theoretical methods, leading to significant decoupling of the periphery from the inner 8π circuit.177 The open-shell dianion [12.5]2− and closed-shell dication [12.5]2+ are both globally aromatic, with their peripheral circuits containing respectively, 22 and 18 π electrons.
The fullerenes, whose spherical shapes provided inspiration for much of the research on curved aromatics, are fundamentally important as multistage electron acceptors and as examples of three-dimensional electron delocalization. The two most studied fullerenes, C60 and C70, can be oxidized to the (+1) state with considerable difficulty,178,179 but each of them can reversibly accept up to 6 electrons in electrochemical experiments.180 While the initial reductions occur at relatively high potentials (Ered1 = −0.98 V and −0.97 V for C60 and C70, respectively), the last reductions require potentials lower than −3 V vs. Fc/Fc+. In C60, the six reductions correspond to sequentially populating the low-lying triply degenerate LUMO level (t1u, Fig. 12, inset). The (−3) state, containing a half-filled LUMO, is responsible for the metallic character of the AxA′3−x[C60] solids (A, A′ = Li, Na, K, Rb, Cs, 3 ≥ x ≥ 0), many of which show superconducting properties at low temperatures.181–183 When generated electrochemically, the lower oxidation levels of fullerenes show only limited stability in solution;178 however, bulk amounts of [C60]6− and [C70]6− were reported to form upon reduction with lithium metal.184 Reduction of C60 with magnesium metal was recently found to produce polymeric polyfullerides, which open a promising path to new carbon allotropes.185,186
 |
| Fig. 12 Multiredox aromatics with two- and three-dimensional fusion. Substituents are omitted for clarity. Relevant conjugated circuits are indicated with bold bonds: aromatic in red, and antiaromatic in blue. Open-shell contributors are not shown. Relative contributions of various conjugation types are discussed in the text. For clarity, π-conjugation is not indicated for the fullerenes. The molecular orbital diagram for neutral C60 (bottom right) is based on published Hückel energy levels.173 | |
With a completely filled t1u level, the hexaanion [C60]6− is a closed-shell species, yielding insulating solids, such as K6[C60].182 The relatively low energy of the L + 1 level in C60, suggests that further reduction, down to the (−12) state, could in principle be possible. The [C60]12− anion can be envisaged as being aromatic both in the local sense (12 aromatic sextets associated with the five-membered rings) and in terms of the spherical aromaticity model (2·(n + 1)2 π electrons).173,187 Clusters with a composition Li12[C60], potentially containing [C60]12−, were indeed observed using mass spectrometry.188,189 However, the solid Li12[C60] was found to contain Li4 clusters, with the [C60]6− oxidation level indicated by NMR and Raman spectroscopic data.190,191 Overall, from the practical viewpoint, the stability of fulleride oligoanions is relatively low, limiting the use of pristine fullerenes in rechargeable batteries.192
Azacoronenes
Hexapyrrolohexaazacoronenes (HPHACs, e.g. 13.1a–b, Fig. 13) may be viewed as analogues of hexa-peri-hexabenzocoronene (p-HBC), in which the peri-fused outer benzene rings are replaced with pyrroles. The majority of known derivatives posses 12 peripheral aryl193,194 or alkyl195 substituents, which affect the redox potentials of these species. Except for NMI-fused analogues51,196,197 (see below), these pyrrole-based nanographenoids are highly electron-rich and stabilize a variety of oxidized states, while being relatively unsusceptible to reduction. Both 13.1a
194 and 13.1b
195 showed three one-electron oxidations in electrochemical experiments. In each case, the oxidation states [13.1]˙+ and [13.1]2+ are stable and can be generated by chemical oxidation. The stability of these dications is derived from the presence of a 22-electron Hückel-aromatic peripheral conjugation pathway (Fig. 13). Likely because of this stabilization, the third oxidation requires a much higher potential than the second one (ΔE ≈ +0.6 V and +1.0 V for 13.1b and 13.1a, respectively).
 |
| Fig. 13 Azacoronenes and their selected oxidized states. Peripheral conjugation paths are shown in red. Aromatic sextets are shaded in light red. | |
The accessibility of higher oxidation states in azacoronenes can be influenced by various modifications of the fused ring system. The indole-containing 13.2
198 shows a fairly similar redox behavior, even though the accessible conjugation pathway in the dication encompasses 26 π-electrons. Other expanded HPHACs were however found to produce monoradical states and switch between global aromaticity and antiaromaticity.51,199 The juxtaposition of pyrrole and benzene rings 13.3 and related systems200 leads to partial decoupling of the fused subunits, which interestingly made the higher oxidation states more accessible than in unmodified HPHACs. Closed-shell configurations, such as that shown in Fig. 13, can be constructed the [13.3]2+ dication, however, its topology precludes the existence of peripheral aromatic pathways. Experimental data suggested that the latter dication could exist in open-shell singlet and triplet states. An even stronger case of decoupling was achieved in the doubly bridged azacoronene analogue 13.4,193 for which four reversible oxidation events could be observed. In fact, the tetracation [13.4]4+ could be obtained by chemical oxidation and was proven to be diamagnetic. The core-expanded azacoronene analogues 13.5 and 13.6 reveal an effective strategy for increasing the redox range of the π system. In the naphthalene-based 13.5, four oxidations were observed,201 the first two located at very low potentials (−0.52 V and −0.32 V respectively). The highly contorted anthracene analogue 13.6202 showed four reversible and two pseudoreversible oxidations. The dications [13.5]2+ and [13.6]2+ may be seen as globally aromatic, with respectively 30- and 38-electron peripheral pathways analogous to that found in [13.1]2+.
Eight-membered rings
Peripheral fusion of π-conjugated subunits can be used to extend the intrinsic range of redox states of cyclooctatetraene (vide supra). The outcome may depend on the extent of planarization induced by fusion203 and the chemical stability of the charged states. The saddle-shaped thiazole cyclotetramer 14.1 (Fig. 14), reported by the group of Yamaguchi,204 showed a one-electron oxidation, apparently producing the corresponding radical cation, and could be reduced in two steps to the nearly planar aromatic dianion [14.1]2−. The heavily substituted tetraphenylene 14.2 was explored by the teams of Petrukhina and Kivala,205 who showed that it could be reduced with lithium to yield [14.2]4−. In the solid state, the latter tetraanion was highly contorted, with two internally coordinated Li+ ions. The Kivala group subsequently reported the indenoannulated tridecacyclene 14.3,206 which is notable for two reversible one-electron oxidations occurring at relatively low potentials (+0.33 V and +0.56 V). The system undergoes four consecutive reductions above −2.2 V, with an additional, probably also reversible event at −2.44 V.
 |
| Fig. 14 Redox-active cyclooctatetraene and [8]circulene derivatives. | |
The rich redox chemistry of corannulene ([5]circulene) and derivatives of coronene ([6]circulene), such as HPHACs, suggests that higher [n]circulenes might also act as suitable multiredox platforms. With the notable exception of Tobe's 12.5 (see above), carbocyclic [8]circulenes typically show just one electrochemically reversible oxidation,207,208 and their reduction chemistry has not been explored. The electron-rich tetraoxa[8]circulenes appear to be more redox active, and a low-symmetry derivative 14.4 (Fig. 14) was reported by Pittelkow et al. to undergo three electrochemical oxidations and three reductions in a potential range between −2.75 and +1.20 V, the first oxidation occurring at +0.76 V.209 Tetraaza[8]circulenes developed by Tanaka et al.210 are apparently more electron rich than their tetraoxa congeners. For instance, in the desymmetrized derivative 14.5, three reversible one-electron oxidations were observed at +0.22, +0.63, and +1.00 V, and the radical cation [14.5]˙+ was isolated and structurally characterized.211 As with other systems, the redox activity of circulenes can be considerably extended by fusion of π-conjugated units. The PDI-fused system 14.6 featured two reversible reduction events (−1.21, −1.50 V), which were interpreted as corresponding to four simultaneous electron transfers.212 Such behavior is consistent with a weak interaction between the PDI fragments.
Electron deficient 2D-fused systems
In comparison with linearly fused aromatics, 2D-fused systems offer, perhaps non-intuitively, a smaller number of functionalization sites per π-electron center. Consequently, introduction of electron-withdrawing functional groups can produce high theoretical charge capacities only for relatively small π-conjugated cores. Pyrene-4,5,9,10-tetraone 15.1 (Fig. 15) can be reduced to the tetraanion [15.1]4−, which has a greater flexibility in the placement of Clar sextets than the neutral state. The resulting increase of aromaticity in the tetraanion may account for a relatively high discharge potential achieved in Li-ion batteries based on 15.1 (+2.59 V).213 Quadruple reduction of the relatively small 15.1 corresponds to a high theoretical specific capacity of these batteries (409 mAh g−1), with experimentally observed values of up to 360 mAh g−1. The trioxotriangulene motif (TOT, [15.5]˙, Fig. 15), explored by the groups of Morita and Takui as a material for organic Li-ion batteries,214 provides an instructive example of engineering a multielectron system using keto groups. The neutral [15.5]˙ is an ambient-stable monoradical, which may be viewed as an electron-poor, flattened version of the trityl radical. [15.5]˙ has a doubly degenerate LUMO level that creates a relatively small energy gap above the SOMO (0.82 eV in [15.5a]˙ and 0.56 eV in [15.5b]˙). These features could in principle make [15.5]˙ a 5-electron acceptor, however, a maximum of four reductions were observed in actual experiments. The first electron is transferred with particular ease, leading to the monoanion [15.5]−, which is stabilized by a combination of carbanionic (I, NCS = 3) and phenolate resonance contributors (II, NCS = 2). Further electrochemical reduction in solution leads to three one-electron events at closely spaced potentials. In the tetraanion [15.5]4−, the three carbonyls are likely completely reduced and mostly decoupled from the aromatic core, which attains triangulene-like conjugation. The use of [15.5a]˙ in Li-ion batteries led to high discharge capacities (>300 Ah kg−1). In comparison, batteries based on the more electron deficient [15.5b]˙ produced a higher output voltage and improved cycle performance.
 |
| Fig. 15 Electron-deficient 2D-fused aromatics. Aromatic sextets are shaded in light red. | |
A different type of stabilization of reduced states is achieved in rylene tetraesters (e.g. 15.2) and diimides (e.g. 15.3a–d). These systems form easily accessible radical anions and dianions. The particular stability of the dianions originates from cross-conjugation of the aromatic core with the electron efficient substituents, which can occur without decreasing the Clar sextet count (Fig. 15). Because the ester groups are twisted away from the plane of the rylene, the conjugation is weaker, but the reductions nevertheless occur at relatively high potentials, making e.g. 15.2 useful in electrochromic applications (Fig. 6).69 The smallest members of the series, naphthalene tetraesters and diimides (NDIs),215 are of particular interest in battery applications, as they can provide the highest theoretical specific capacities. Their redox potentials can be tuned over a substantial range via substitution of the naphthalene core (e.g., 15.3a–d).216 Thanks to their chemical robustness, and good crystallinity, NDIs found use as electrode materials in Li-ion batteries,216 and as anolytes in aqueous organic redox flow batteries.217 According to NICS calculations, the NDI dianion [15.3b]2− shows enhanced diatropicity in the imide rings,217 suggesting that the stability of the dianion is derived from the charge-separated bis-pyridinium contributions (II), which have a higher Clar-sextet count than the quinoidal contributors such as I (Fig. 15). Such structures resemble those proposed for reduced pentacosacyclene tetraimide (see below).
By simultaneous application of two or more design principles, it is possible to dramatically extend the multiredox character of 2D-fused aromatics. Simultaneous use of imide fusion, nitrogen doping and oligomerization is illustrated by the tetraimide 15.4, reported by Shinokubo et al.,218 which contains two diazazethrene diimide subunits. Even though linked via a single bond and twisted out of planarity, the two units are apparently quite strongly conjugated, yielding four well-resolved one-electron reductions in the −0.36 V to −0.91 V range, and several non-reversible oxidation events. Radial fusion of redox-active units is a potentially more efficient strategy, as it can capitalize on the higher symmetry of the resulting structures, and provide stronger inter-subunit communication than the complementary radial linking discussed above. These advantages are evident even in relatively small molecules such as diquinoxalino[2,3-a:2′,3′-c]phenazine 15.6, which is susceptible to six consecutive one-electron reductions,219 a feature that has recently been exploited in a variety of battery setups.219–222
Multiredox nanographenoids have been obtained by fusion of naphthalimide (NMI) units. The Würthner group developed efficient benzannulation methods suitable for making both planar and curved aromatic molecules, such as the pyrene-based C64 system 16.1
223 or the C80N5 nanocone 16.2 containing a corannulene core224 (Fig. 16). Pentannulation of NMI units, explored by our group, produced a variety of extended systems, such azacoronenes (16.3),196,197 metalloporphyrins (16.4),225 and cyclooctatetraenes (16.5
226 and 16.6
227). In these systems, the oxidation behavior usually reflects the properties of the fusion-free core, for instance, 16.3 and its derivatives, inherited the +1 and +2 oxidation states from the parent HPHAC motif. The accessibility of reduced states was found to correlate with the number of low-lying virtual molecular orbitals (MOs). Thus, 16.2, which has three low-energy empty MO levels, underwent six consecutive one-electron reductions, and as many as 10 reductions were observed for 16.3, matching its five low-energy virtual MOs. As shown for 16.4, in the initial reduction products, i.e. [16.4]˙− and [16.4]2−, most of the negative charge is located in the porphyrin core, and the contribution of NMI units to charge delocalization gradually increases upon progressive reduction.225 Theoretical data obtained for 16.5 and 16.6 showed that the stability of anionic states in the pentannulated systems can be linked to an increase of local aromaticity caused by reduction. For instance, in the octaanionic [16.6]8−, the conjugation in each of the four subunits can be described using local 6e, 10e, and 14e contributions (I–V, Fig. 16). The role of the five membered ring is particularly important, as it can accept one of the two negative charges, yielding an aromatic cyclopentadienyl anion. Anionic states can also be stabilized by multiple cyanation, instead of imide fusion, as shown recently by the Kivala group for an octacyanotridecacyclene derivative.228
 |
| Fig. 16 Imide-fused nanographenoids. dipp = 2,6-di(isopropyl)phenyl. | |
5. Macrocycles
Porphyrinoids
Oxidation and reduction of free-base porphyrinoids are often associated with a change in the protonation status of the pyrrolic nitrogens. Protonation-coupled redox reactions are of significant interest in their own right,229 but are outside the scope of this review. Proton transfer can be suppressed, e.g., by complete protonation of the macrocycle,230 by metal coordination, or by structural modifications at the pyrrolic nitrogens (see below). In porphyrins and their analogues, a two electron oxidation of reduction results in switching of macrocyclic aromaticity (between Hückel-aromatic and either Hückel-antiaromatic or Möbius-aromatic), although analogous changes can occur also in intrinsically non-aromatic macrocycles.231 Metal complexation typically prevents redox-coupled proton transfer, leading to well-defined oxidation and reduction behavior. Typical metalloporphyrins have stable oxidation states in the (−2) to (+2) range,232 although electrochemical reduction to the hexaanion was achieved for zinc(II) 5,10,15,20-tetra(p-tolyl)porphyrin (17.1a, Fig. 17).233 The neutral state of porphyrins is globally aromatic, with an 18π-electron conjugation pathway. The dicationic and dianionic states feature respectively 16π- and 20π-electron pathways, both of which correspond to Hückel antiaromaticity. Both of these redox regimes can be employed in battery applications, as shown by Fichtner et al. in their work on a diethynyl-substituted copper(II) porphyrin 17.1b.234 The latter compound was employed in two different configurations: as a cathode in a Li–metal cell, operating in the (+2) to (−2) redox range; and as an anode coupled with a graphite cathode, with a working redox range from (0) to (−2). Both of these configurations yielded supercapacitor-like power densities and long cycle life.
 |
| Fig. 17 Tuning of multiredox behavior in porphyrins and their analogues. Aromatic, antiaromatic, and odd-electron conjugation pathways are shown respectively in red, blue, and green. The conjugation pathway shown for 17.8 corresponds to the proposed [16]annulene dianion-like conjugation. In the structure of 17.10a and 17.10b, π-bonds are not shown for clarity. | |
Redox ranges of porphyrins can be controlled by diverse structural modifications, including macrocycle expansion and contraction, peripheral substitution and fusion, heteroatom doping, subunit replacement, and metal coordination. Heteroatom doping at the meso positions has been recently employed by the group of Matano in their metallo-5,15-diazaporphyrins 17.3 (M = Ni, Zn, Cu).235,236 The neutral state of the latter species is antiaromatic and is very easily oxidized to the corresponding radical cation and dication. The dication, isoelectronic with an undoped neutral metalloporphyrin, is globally aromatic. The related oxygen-doped system 17.4 (M = Ni), reported by Shimizu et al.,237 features somewhat diminished paratropicity in the neutral state but largely reproduces the redox behavior of 17.3.
The use of non-pyrrolic subunits for tailoring the redox properties of porphyrinoids is illustrated by the chalcogen-containing diazuliporphyrins 17.4 and 17.5, synthesized by the group of Latos-Grażyński.238,239 In their neutral states, these systems retain the local aromaticity of individual azulene units. When treated with relatively mild oxidants, such as elemental bromine, they undergo sequential oxidation to the respective radical cations and dications, all of which absorbed intensely in the visible and NIR regions. The dications are stabilized by the combination of porphyrin-like 18π-electron macrocyclic aromaticity with tropylium-like aromaticity of the outer seven-membered rings. The color changes observed on going from 17.4 (dark red), through [17.4]˙+ (purple) to [17.4]2+ (navy blue) were proposed to be useful in electrochromic applications.
From the viewpoint of porphyrinoid chemistry, nitridomanganese(V) phthalocyanine 17.8, reported by Ménard et al.,240 may be interpreted as combining peripheral fusion, heteroatom doping and metal coordination. Similarly to typical metalloporphyrins, the redox range of 17.8 covers states from (−2) to (+2), which are separated by electrochemical events at −1.75, −1.38, +0.02, and +0.45 V. All five oxidation levels were characterized structurally in the solid state and the even-electron states were shown to be diamagnetic by 1H NMR. The coordinated MnVN was found to insignificantly participate in the redox chemistry of 17.8. The aromaticity of the neutral 17.8 was interpreted as corresponding to either neutral [18]annulene or [16]annulene dianion, whereas the dication [17.8]2+ was found to be mostly non-aromatic in spite of its formally [16]annulene structure. In contrast, the dianion [17.8]2− was decisively antiaromatic, showing a NICS pattern consistent with an [18]annulene-dianion conjugation.
Radial fusion has been used to produce porphyrin analogues both electron-deficient, such as 16.4 (Fig. 16), or electron-rich. For the latter purpose, fusion of tetrathiafulvalene (TTF) and related units is particularly attractive, because of their easy oxidation to the corresponding radical cations and dications.241 Such porphyrins, containing from one to four TTF subunits fused at the β positions as in 17.11a–b, were originally developed by the groups of Jeppsen and Sessler.242,243 Each of these systems showed two one-electron reductions, likely associated with the metalloporphyrin core, and an increasing number of low-potential oxidation events, occurring in the 0.0–0.7 V range. The tetra-TTF macrocycle 17.11a, underwent at least 5 overlapping oxidations, but the total number of transferred electrons may have been higher. Expanded porphyrins containing benzo-TTF units showed a similar redox behavior.244
Resizing of the macrocycle affects the redox range in the expected way, however, no specific correlation seems to exist between the size of the porphyrinoid and the number of accessible stable oxidation levels. Metal complexes of norcorroles, the smallest contracted porphyrin analogues, are chemically stable in spite of their pronounced 16π-electron antiaromaticity.245,246 Just as typical metalloporphyrins, nickel(II) norcorrole 17.6 features two reversible reductions and two reversible oxidations. However, all these events occur over a fairly narrow potential range of ca. +2.5 V. The relative ease of oxidation and reduction is rationalized by the global aromaticity of both the dication and dianion (respectively 14π and 18π-electron). 17.6 was utilized in rechargeable batteries both as a cathode active material with a Li metal anode and in lithium-free batteries exploiting the bipolar character of the norcorrole (Fig. 18).247 In the former setup, a maximum discharge capacity of about 207 mAh g−1 was maintained after 100 charge/discharge cycles.
 |
| Fig. 18 Charge/discharge reactions for (a) Li-NiNC batteries and (b) NiNC–NiNC batteries (NiNC = 17.6). Reprinted with permission from ref. 247. Copyright © 2014 Wiley-VCH GmbH. | |
Porphyrinoids containing solely furan or thiophene subunits, cannot undergo deprotonation when oxidized, which makes them useful for controllable redox switching. In such electron-rich systems, two or three neutral and cationic states can be chemically accessed,248,249 but the range may be extended in larger macrocycles. For instance, electrochemical data obtained for a [38]octaphyrin diradicaloid 17.9, reported by Wu et al.,250 featured two one-electron reductions and four one-electron oxidations at potentials lower than +0.9 V. In contrast, pyrrole-based expanded porphyrins are usually quite electron-deficient, a feature that may be partly caused by the usual substitution with electron-withdrawing pentafluorophenyl groups. Achieving well-defined redox behavior in these systems is however, challenging, because of their conformational flexibility and potentially limited chemical stability. For instance, voltammetric data reported for a mononuclear ZnII complex of [62]tetradecaphyrin by Yoneda, Osuka et al.251 were somewhat poorly resolved, but they nevertheless showed up to three oxidations and up to 8 reductions in the +0.64 V to −2.13 V potential range. Fusing expanded porphyrins into aromatic tapes such as 17.7
252 apparently yields more robust redox systems, capable of accepting up to 5 electrons. Less obvious fusion patterns, such as the recently reported COT fusion of orangarins,253 can also lead to oligomacrocyclic systems showing simultaneously multiple oxidation and reduction events.
In the examples of metal complexes highlighted above, the coordinated ions played mainly a structural role, stabilizing the geometry of the macrocycle and preventing proton transfer reactions. In the triple-decker complexes 17.10a and 17.10b, investigated by Ogawa et al.,254 terbium(III) and yttrium(III), respectively, act as linkers between the outer porphyrin decks and the inner fused bisporphyrin. The neutral states of these complexes featured remarkably small energy gaps (ca. 0.33 eV) and were characterized as diradicaloids with singlet ground states and thermally accessible triplets. Voltammograms recorded for 17.10a revealed seven reversible one-electron events: two reductions and five oxidations in the range of −0.73 to +1.44 V. At least 10 electron transfers were observed for a DBU salt of the dianionic [17.10a]2−, implying that the redox range may extend to much lower potentials. While the role of the metal ions in the redox processes was not clarified, it seems likely that the reductions and oxidations are primarily associated with the π-systems, with a good electronic communication ensured by the triple-decker structure of the complexes.
Nanohoops and nanorings
[n]Cycloparaphenylenes (CPPs, 19.1–n, n ≥ 5, Fig. 19), first synthesized by Jasti, Bertozzi et al. (n = 9, 12, 18),255 may be viewed as highly strained paracyclophanes containing no bridges between the para-phenylene rings. The smallest of these systems, [5]CPP (19.1–5),256,257 has a remarkably narrow electrochemical gap and shows two pseudoreversible one-electron oxidations (at +0.25 and +0.46 V) and two reductions (at −2.27 and −2.55 V), but none of the corresponding charged states has been isolated. Radical cations [19.1–n]˙+ and dications [19.1–n]2+ can be chemically generated258–260 by oxidation with, e.g., [NO][SbF6], and they were subsequently characterized using Raman spectroscopy (5 ≤ n ≤ 12),261 magnetic circular dichroism (n = 8),262 and NIR fluorescence spectroscopy (5 ≤ n ≤ 9).263 Unlike the neutral states, which have benzenoid conjugation, the dications show global in-plane aromaticity, corresponding to the (4n − 2)π pathways emerging upon oxidation. For n > 9, the dications undergo symmetry breaking, yielding diradicaloid configurations.261 It can be expected that most cycloparaphenylenes should stabilize negative states down to at least (−2). CPP anions characterized to date include [19.1–6]˙−, [19.1–6]2−,264 and [19.1–8]4−.265 The latter species shows a significant elliptical distortion in the solid state, which is partly attributable to interactions with the internally bound potassium cations.
 |
| Fig. 19 Multiredox nanohoops and nanorings. For clarity, π-conjugation is not shown in porphyrin units of 19.2. | |
Remarkable instances of multiredox systems are found among Anderson's porphyrin nanorings, exemplified here by 19.2.266 In this system, the macrocycle consisting of 12 zinc(II) 5,15-di(ethynyl)porphyrin subunits is rigidified by coordination of two stacked star-shaped hexapyridine ligands. 19.2 could be oxidized chemically up to the dodecacation [19.2]12+. Such oxidations tend to occur sequentially, in one-electron steps, and using 19F NMR spectroscopy of the CF3 groups installed on the inner ligands, it is possible to directly observe the majority of intermediate states [19.2]n+ (0 < n < 12), including the odd-electron ones.266,267 Given the extraordinary dimensions of 19.2, the charge density achieved upon oxidation is relatively low; however, the electronic properties of the oligocations strongly depend on the oxidation level, providing evidence for global aromaticity at a previously unexpected scale. While local aromaticity of individual porphyrins prevails in the neutral 19.2 (168π) and the fully charged [19.2]12+ (156π) intermediate states, i.e. [19.2]6+ (162π), [19.2]8+ (160π), and [19.2]10+ (158π) are, respectively, globally aromatic, antiaromatic, and again aromatic.
Other bridgeless cyclophanes
In spite of their structural simplicity, CPPs are rather unusual cyclophanes, because of their high internal strain and in-plane conjugation. Direct linking of larger, angular subunits leads to less strained and potentially more robust systems. Phenanthrenequinone cyclotrimer S6.1 (Scheme 6), was implemented in rechargeable aluminum–organic batteries by the groups of Stoddart and Choi.268 In this particular application, S6.1 underwent three-electron reduction to the complex S6.2, containing the Al-coordinated triradical trianion [S6.1]3(˙−). S6.1 formed layered superstructures which facilitated reversible insertion and extraction of aluminum during, respectively, discharging and charging of the battery. This feature led to outstanding electrochemical performance, with a reversible capacity of 110 mAh g−1 and cyclability of up to 5000 cycles.
 |
| Scheme 6 Multiredox cyclophanes containing no conjugated meso bridges. R = 10-(3,5-di-tert-butylphenyl)anthracen-9-yl in S6.4–n. | |
The accessible redox range typically increases with the size of the macrocycle, but in order to observe multiple redox events, the interaction between subunits must be sufficiently strong. This requirement is met by the series of oligoradicaloid fluorenyl macrocycles S6.4–n (n = 4 through 6), synthesized by the Wu group.269 Linking the fluorenyls at positions 3,6 enables quinoidal conjugation reminiscent of the Chichibabin hydrocarbon. Electrochemical oxidation and reduction of S6.4–n occurred sequentially, up to, respectively, [S6.4–n]n+ and [S6.4–n]n− oxidation levels, and the n-cations could in all cases be observed using spectroelectrochemistry. A particularly well resolved cyclic voltammogram was obtained for S6.4–6, revealing a number of redox events at +0.03 V, +0.12 V, +0.25 V, +0.44 V (3e) and at −1.07 V (2e), −1.23 V, 1.30 V, and −1.51 V (2e).
To ensure well-defined redox behavior, it is advantageous to stabilize the conformation of very large macrocycles, e.g., by coordinating templates, as in porphyrin nanorings, or by covalent linking. The latter approach is illustrated by the hexacarbazole bis-macrocycle S6.5, reported recently by Jin, Chen et al.270 In this species, the π system is rigidified by inclusion of a 2,2′,7,7′-spirobifluorenyl subunit, which interconnects two tercarbazole sections. In voltammetric experiments, up to six electrons can be removed from S6.5 (2 e at +0.42 V, 1 e at +0.60 V, 1 e at +0.71 V, 2 e at +1.10 V), and the presence of one-electron events indicated an electronic interaction between the two halves of the molecule. The even-electron cationic states [S6.5]2+, [S6.5]4+, and [S6.5]6+ were all predicted to have singlet ground states with open-shell oligoradicaloid configurations.
Through-space coupling in cyclophanes
In appropriately rigidified cyclophanes, through-space interactions between π-conjugated subunits provide a mechanism of electronic communication in the charged states, resulting in a multistage redox behavior. This concept has recently been reviewed by Stoddart et al.271 and is only highlighted here, using the molecular triangle S6.3
272,273 (Scheme 6) as a representative example. S6.3 undergoes six one-electron electrochemical reductions (−0.61, −0.70, −0.77, −1.33, −1.55, −1.72 V). Electron sharing among subunits in such reduced states can be explained in terms of spatial overlap of the lowest virtual orbitals, occurring between adjacent imide moieties at the corners of the triangle.274 When used as a cathode in Li-ion batteries, S6.3 showed a galvanostatic voltage profile with a single plateau at +2.33 V vs. Li+/Li, and a specific capacity of 163 mAh g−1, indicating acceptance of 5.8 Li+ per molecule.272 It was proposed that the unique performance of S6.3 originated from its strain-free structure, minimization of electrostatic repulsion in the anions, and from the ability to chelate lithium cations by pairs of adjacent imide oxygens.
Phenylene cyclophanes with meso bridges
[24]Paracyclophanetetraene 20.1 is a classic example of the interplay between the oxidation level and aromaticity.275–279 The neutral state of 20.1 features locally aromatic benzene rings, while its dianion [20.1]2− and tetraanion [20.1]4− contain macrocyclic conjugation pathways corresponding respectively to [26]annulenoid aromaticity and [28]annulenoid antiaromaticity. The compound was recently revisited by the groups of Petrukhina and Anderson, who provided crystallographic evidence for aromaticity changes in the dianionic and tetraanionic states of 20.1.280 The switching between local and global aromaticity of 20.1 and [20.1]2− was exploited by Choi, Glöcklhofer et al. in organic sodium-ion battery anodes.281 The resulting batteries combined stable cycling behavior and showed excellent performance during fast charging and discharging. Squarephaneic tetraimide 20.2, obtained by the Glöcklhofer group,282 features significantly elevated redox potentials relative to those of 20.1, and it could be implemented in Li-ion batteries, albeit, with a moderate cycling performance. Interestingly, the tetraanion [20.2]4− was computationally predicted to have a Baird-aromatic 28π-electron ground state.
Tetraaza[1.1.1.1]m,p,m,p-cyclophane, reported by the groups of Hartwig283 and Tanaka (20.3, Fig. 20),284 may be viewed as a macrocyclic analog of Wurster's blue. Consequently, its most stable oxidation level is the diradical dication [20.3]2+: when aryl-substituted, it is air-stable and has a triplet ground state with ΔEST ≈ 0.5 kcal mol−1.285 In electrochemical experiments, 20.3 was found to reversibly undergo four one-electron oxidations at −0.03, +0.16, +0.47, and +0.62 V. 20.3 is a versatile building block for the construction of multiredox aromatics, and covalent assemblies containing two copies of 20.3 were subsequently shown to be oxidizable up to a +12 oxidation level.286 20.4, synthesized by Ito, Tanaka et al.,287 is another example of such a structure: it contains two tetraazacyclophane moieties bound into a macrocycle, and can be oxidized up to a hexacation. In general, cage-like structures containing multiple phenylene rings, aza bridges, and electron rich substituents, often achieve multiple redox states, as exemplified by Ito's double-decker hexaamine 20.5.288 The latter system showed a range of oxidized states in electrochemical experiments: the lowest three species, i.e., [20.5]+, [20.5]2+, [20.5]4+ (the latter formed via a two-electron oxidation), were studied in detail and found to have a delocalized intervalence character. The cage design found in 20.4 and 20.5 apparently offers a general approach to multiredox systems, as illustrated by a bis-macrocyclic oligothiophene cage, recently reported by the Wu group, which yielded oxidation states up to +6.289
 |
| Fig. 20 Multiredox phenylene macrocycles containing meso bridges. | |
While nitrogen meso bridges, present e.g. in 20.3–20.5, are relatively weakly coupled with phenylene rings in neutral macrocycles, the single-carbon meso bridges promote strong cross-conjugation, leading to enhanced interactions between subunits in the cyclophane. Such phenylene rings can sustain global aromaticity in neutral macrocycles290,291 but can also increase the open-shell characteristics. Both features are evident in [6]cyclo-para-phenylmethine 20.6–3, reported by the groups of Casanova and Wu,292 which has a singlet ground state combining 30π-electron global aromaticity with a diradicaloid character. The species stabilizes oxidation levels from −2 through +3, as evidenced by electrochemical and chemical experiments. In analogy to 20.6–3, which could be viewed as an expanded benzene, its larger congeners, 20.6–4
293 and 20.6–5,294 may be likened to cyclooctatetraene and [10]annulene, respectively. Both macrocycles display localized quinoidal structures in their neutral states and can both be oxidized up to a tetracation level, with concomitant changes of their global aromatic character.
Replacing some of the methine carbons with nitrogens increases the number of Clar sextets in the neutral state of the cyclophane, as illustrated by the nitrogen-doped superbenzene 20.7
295 and supernaphthalene 20.8,296 reported by the Chi group. The first of these systems could be oxidized up to the tetracation level, and [20.7]2+, [20.7]3+ (30π-aromatic) and [20.7]4+ (28π-antiaromatic) were characterized crystallographically in the solid state. For the bis-macrocyclic 20.8, only the even-electron states could be isolated, i.e., [20.8]2+ (diradicaloid), [20.8]4+ (52π-antiaromatic) and [20.8]6+ (50π-aromatic, diradicaloid).
[n]Cyclo-para-biphenylmethines 20.9-n, synthesized by the Wu group,297 can be viewed as higher benzologues of the 20.6–n series, or as macrocyclic analogues of the Chichibabin hydrocarbon. The latter relationship is reflected in their oligoradicaloid configurations containing only Clar sextets in the macrocycle, and methine-localized radical sites. For even n values, closed-shell configurations 20.9–n′ can also be constructed; indeed, 20.9–4 is actually quinoidally localized, undergoing rapid valence tautomerization via a tetraradicaloid transition state. Because of their open-shell nature, the 20.6–n macrocycles show redox amphoterism, and higher members of the series (n > 4) have shown up to four oxidations and up to four reductions in electrochemical measurements.
Coronoids and related systems
Coronoids (cycloarenes) consist of contiguously fused rings fully enclosing a macrocyclic opening.298,299 Investigations of non-functionalized coronoids, such as kekulene (21.1a, Fig. 21),300 have often been hampered by their very low solubility; however, electrochemical data reported for substituted kekulenes and octulenes (21.1b–c, 21.2b–c) show several reductions and oxidations at relatively high absolute potentials.301,302 To date, no charged states have been isolated for such benzenoid cycloarenes, perhaps because the available substitution patterns are insufficient to protect the resulting coronoid ions from decomposition. Cycloarene analogues featuring five-membered rings, heteroatoms, or incomplete fusion (e.g., 21.3–21.8, Fig. 21) are often much easier to oxidize and reduce than their benzenoid congeners. These features can often be linked to the open-shell character of the neutral state, which results in the narrowing of the energy gap, and to aromatic stabilization achieved by certain charged levels. Because of their sizes and limited stability, cycloarene oligoradicaloids remain challenging to analyze, especially using electrochemical methods.
 |
| Fig. 21 Redox-active coronoids. Fully radicaloid configurations are shown for the neutral states of 21.3–21.7. | |
[4]Chrysaorene 21.7, reported independently by the group of Wu303 and ourselves,304 is possibly the best-studied case of multiredox behavior in coronoids. The neutral state of 21.7 has a singlet ground state with significant tetraradicaloid character. A theoretical analysis showed that the experimentally observed diatropicity of the neutral 21.7 stems from the global 36π-electron contribution 21.7′, rather than from AWA conjugation. In electrochemical experiments, 21.7 showed four one-electron oxidations and four one-electron oxidations. The radical cation [21.7]˙+ and dication [21.7]2+ could be generated chemically using elemental iodine, which acts not only as a mild oxidant but also as a source of iodides which were bound in the cavity of the cationic states.304 Reduction with sodium anthracenide in turn produced the anionic states [21.7]2−, [21.7]˙3−, and [21.7]4−.303 Both the dication and the dianion were diamagnetic and globally diatropic, consistent with the 38π-electron aromatic pathway shown in Fig. 21. Unlike all the other attainable redox states, the tetraanion [21.7]4− showed a 1H NMR spectrum indicative of local aromaticity, and a large optical energy gap. Both of these features were consistent with the benzenoid character of the tetraanion, isoelectronic with octulene.
The above results suggest that for the open-shell coronoids, the lower reduction limit can be expected to equal the number of formal radical sites (i.e., 4, 8, 10, and 6 for 21.3,305 21.4, 21.5,306 and 21.6,307 respectively), since the corresponding oligoanions are strongly stabilized by local aromaticity. An analogous argument will not, however, work for the corresponding cations, which will be destabilized by local antiaromaticity, similar to that of the fluorenyl cation. In many cases, however, the first two oxidations occur at very low potentials, and the corresponding cation, such as [21.7]2+ or [21.6]2+, is ambient-stable and globally aromatic. The actual range of accessible oxidation levels will depend on the global aromatic/antiaromatic contributions, charge density (which should be higher in systems such as 21.4 and 21.5), and the stability of the cations and anions against decomposition.
The nitrogen-doped coronoid 21.8, obtained by Lu, Zhao et al.,308 is a higher homologue of the bowl-shaped [3]chrysaorole.309 Unlike the tetraradicaloid 21.7, compound 21.8 is a closed-shell species, featuring up to 6 low-potential electrochemical oxidations (+0.08, +0.30, +0.38, +0.60, +0.75 and +0.93 V), yielding hole mobilities of up to 2.06 cm2 V−1 s−1 in single-crystal transistors. In this regard, it significantly outperformed the related carbazole cyclooligomer, suggesting potential applications of appropriately designed cycloarenes as organic semiconductors.
7. Conclusions and outlook
Multistage redox behavior represents a compelling expression of π-electron delocalization in organic molecules. As this review illustrates, aromaticity serves as a fundamental organizing principle for understanding and engineering such systems. The emergence of distinct redox states is closely tied to topological, electronic, and conformational properties of the π framework. Linear, cyclic, and radial arrays of redox-active units provide access to particularly diverse and effective multiredox systems. Recent structural developments, which range from new monomeric motifs to complex multiredox arrays based on, e.g. nanographenes and oligoradicaloids such as, rely on the judicious use of global and local aromatic stabilization at specific oxidation levels and on proper tuning of inter-subunit coupling. Continued exploration of these themes promises new materials with tailored redox characteristics for energy, optoelectronic, and molecular switching applications.
Author contributions
M.S. writing – original draft. I. A. B, A. K. G. N. S. writing – review & editing, visualization.
Conflicts of interest
There are no conflicts to declare.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Acknowledgements
Financial support was kindly provided by the National Science Center of Poland (UMO-2022/47/B/ST4/00798 to M. S.).
References
- F. Wöhler, Grundriss der Chemie, Duncker und Humblot, 1854 Search PubMed.
- P. Karen, Oxidation State, A Long-Standing Issue!, Angew. Chem., Int. Ed., 2015, 54, 4716–4726 CrossRef CAS PubMed.
- H. D. Roth, Early contributions to the chemistry of organic radical ions, Tetrahedron, 1986, 42, 6097–6100 CrossRef CAS.
- C. Wurster and E. Schobig, Ueber die Einwirkung oxydirender Agentien auf Tetramethylparaphenylendiamin, Ber. Dtsch. Chem. Ges., 1879, 12, 1807–1813 CrossRef.
- E. Weitz and K. Fischer, Radikale und merichinoide Verbindungen, I.: Die Dipyridinium-subhalogenide, Ber. Dtsch. Chem. Ges. B, 1926, 59, 432–445 CrossRef.
- E. Weitz, Radikale, Quasi-Radikale, merichinoide Verbindungen und Chinhydrone Ein Beitrag zur Farbentheorie, Angew. Chem., 1954, 66, 658–677 CrossRef CAS.
- K. Deuchert and S. Hünig, Multistage Organic Redox Systems—A General Structural Principle, Angew. Chem., Int. Ed. Engl., 1978, 17, 875–886 CrossRef.
- T. Nishinaga, Organic Redox Systems: Synthesis, Properties, and Applications, John Wiley & Sons, 2015 Search PubMed.
- J. Shukla, V. P. Singh and P. Mukhopadhyay, Molecular and Supramolecular Multiredox Systems, ChemistryOpen, 2020, 9, 304–324 CrossRef CAS PubMed.
- H. Ueda and S. Yoshimoto, Multi-Redox Active Carbons and Hydrocarbons: Control of their Redox Properties and Potential Applications, Chem. Rec., 2021, 21, 2411–2429 CrossRef CAS PubMed.
- T. Harimoto and Y. Ishigaki, Recent Advances in NIR-Switchable Multi-Redox Systems Based on Organic Molecules, Chem. – Eur. J., 2025, 31, e202403273 CrossRef CAS PubMed.
- D. Xu, M. Liang, S. Qi, W. Sun, L.-P. Lv, F.-H. Du, B. Wang, S. Chen, Y. Wang and Y. Yu, The Progress and Prospect of Tunable Organic Molecules for Organic Lithium-Ion Batteries, ACS Nano, 2021, 15, 47–80 CrossRef CAS PubMed.
- T. B. Schon, B. T. McAllister, P.-F. Li and D. S. Seferos, The rise of organic electrode materials for energy storage, Chem. Soc. Rev., 2016, 45, 6345–6404 RSC.
- Q. Zhao, Z. Zhu and J. Chen, Molecular Engineering with Organic Carbonyl Electrode Materials for Advanced Stationary and Redox Flow Rechargeable Batteries, Adv. Mater., 2017, 29, 1607007 CrossRef PubMed.
- P. Poizot, J. Gaubicher, S. Renault, L. Dubois, Y. Liang and Y. Yao, Opportunities and Challenges for Organic Electrodes in Electrochemical Energy Storage, Chem. Rev., 2020, 120, 6490–6557 CrossRef CAS PubMed.
- Y. Chen, K. Fan, Y. Gao and C. Wang, Challenges and Perspectives of Organic Multivalent Metal-Ion Batteries, Adv. Mater., 2022, 34, 2200662 CrossRef CAS PubMed.
- Q. Zhu, D. Fu, Q. Ji and Z. Yang, A Review of Macrocycles Applied in Electrochemical Energy Storge and Conversion, Molecules, 2024, 29, 2522 CrossRef CAS PubMed.
- K. Cai, L. Zhang, R. D. Astumian and J. F. Stoddart, Radical-pairing-induced molecular assembly and motion, Nat. Rev. Chem., 2021, 5, 447–465 CrossRef CAS PubMed.
- Y. Kubozono, R. Eguchi, H. Goto, S. Hamao, T. Kambe, T. Terao, S. Nishiyama, L. Zheng, X. Miao and H. Okamoto, Recent progress on carbon-based superconductors, J. Phys.: Condens. Matter, 2016, 28, 334001 CrossRef PubMed.
- M. Stolar, Organic electrochromic molecules: synthesis, properties, applications and impact, Pure Appl. Chem., 2020, 92, 717–731 CrossRef CAS.
- M. Kivala, C. Boudon, J.-P. Gisselbrecht, P. Seiler, M. Gross and F. Diederich, Charge-Transfer Chromophores by Cycloaddition–Retro-electrocyclization: Multivalent Systems and Cascade Reactions, Angew. Chem., Int. Ed., 2007, 46, 6357–6360 CrossRef CAS PubMed.
- P. Pakulski, M. Magott, S. Chorazy, M. Sarewicz, M. Srebro-Hooper, D. Tabor, Ł. Łapok, D. Szczepanik, S. Demir and D. Pinkowicz, A multifunctional pseudo-[6]oxocarbon molecule innate to six accessible oxidation states, Chem, 2024, 10, 971–997 CAS.
- R. Haag, Multivalency as a chemical organization and action principle, Beilstein J. Org. Chem., 2015, 11, 848–849 CrossRef CAS PubMed.
- R. Hein, C. N. Stindt and B. L. Feringa, Mix and Match Tuning of the Conformational and Multistate Redox Switching Properties of an Overcrowded Alkene, J. Am. Chem. Soc., 2024, 146, 26275–26285 CrossRef CAS PubMed.
- T. Harimoto, T. Tadokoro, S. Sugiyama, T. Suzuki and Y. Ishigaki, Domino-Redox Reaction Induced by An Electrochemically Triggered Conformational Change, Angew. Chem., Int. Ed., 2024, 63, e202316753 CrossRef CAS PubMed.
- Y. Ishigaki, T. Harimoto, K. Sugawara and T. Suzuki, Hysteretic Three-State Redox Interconversion among Zigzag Bisquinodimethanes with Non-fused Benzene Rings and Twisted Tetra-/Dications with [5]/[3]Acenes Exhibiting Near-Infrared Absorptions, J. Am. Chem. Soc., 2021, 143, 3306–3311 CrossRef CAS PubMed.
- R. Rathore, P. Le Magueres, S. V. Lindeman and J. K. Kochi, A Redox-Controlled Molecular Switch Based on the Reversible C−C Bond Formation in Octamethoxytetraphenylene, Angew. Chem., Int. Ed., 2000, 39, 809–812 CrossRef CAS PubMed.
- Z. Zhou, Y. Zhu, Z. Wei, J. Bergner, C. Neiß, S. Doloczki, A. Görling, M. Kivala and M. A. Petrukhina, Reversible structural rearrangement of π-expanded cyclooctatetraene upon two-fold reduction with alkali metals, Chem. Commun., 2022, 58, 3206–3209 RSC.
- Z. Huang, G. Tan, C. Chen, W.-X. Zhang, X. Wang and Z. Xi, Selective reduction of 1,5-diazacyclooctatetraenes: synthesis and structures of aromatic diazacyclooctatetraenyl dianions and a 2,6-bipyrrolinyl dianionic Co(II) complex, Chem. Commun., 2019, 55, 2648–2651 RSC.
- J. Hankache and O. S. Wenger, Organic Mixed Valence, Chem. Rev., 2011, 111, 5138–5178 CrossRef CAS PubMed.
- A. Heckmann and C. Lambert, Organic Mixed-Valence Compounds: A Playground for Electrons and Holes, Angew. Chem., Int. Ed., 2012, 51, 326–392 CrossRef CAS PubMed.
- A. D. Allen and T. T. Tidwell, Antiaromaticity in Open-Shell Cyclopropenyl to Cycloheptatrienyl Cations, Anions, Free Radicals, and Radical Ions, Chem. Rev., 2001, 101, 1333–1348 CrossRef CAS PubMed.
- K. B. Wiberg, Antiaromaticity in Monocyclic Conjugated Carbon Rings, Chem. Rev., 2001, 101, 1317–1332 CrossRef CAS PubMed.
- M. J. Evans, J. Mullins, R. Mondal and C. Jones, Reductions of Arenes using a Magnesium-Dinitrogen Complex, Chem. – Eur. J., 2024, 30, e202401005 CrossRef CAS PubMed.
- M. C. Cassani, D. J. Duncalf and M. F. Lappert, The First Example of a Crystalline Subvalent Organolanthanum Complex: [K([18]crown-6)- (η2-C6H6)2][(LaCptt2)2(μ-η6:η6-C6H6)]·2C6H6 (Cptt = η5-C5H3But2–1,3), J. Am. Chem. Soc., 1998, 120, 12958–12959 CrossRef CAS.
- M. C. Cassani, Y. K. Gun'ko, P. B. Hitchcock and M. F. Lappert, The first metal complexes containing the 1,4-cyclohexa-2,5-dienyl ligand (benzene 1,4-dianion); synthesis and structures of [K(18-crown-6)][Ln{η5-C5H3(SiMe3)2-1,3}2(C6H6)](Ln = La, Ce), Chem. Commun., 1996, 1987–1988 RSC.
- M. C. Cassani, Y. K. Gun'ko, P. B. Hitchcock, M. F. Lappert and F. Laschi, Synthesis and Characterization of Organolanthanidocene(III) (Ln = La, Ce, Pr, Nd) Complexes Containing the 1,4-Cyclohexa-2,5-dienyl Ligand (Benzene 1,4-Dianion): Structures of [K([18]-crown-6)][Ln{η5-C5H3(SiMe3)2-1,3}2(C6H6)] [Cp′′ = η5-C5H3(SiMe3)2-1,3; Ln = La, Ce, Nd], Organometallics, 1999, 18, 5539–5547 CrossRef CAS.
- C. M. Kotyk, M. E. Fieser, C. T. Palumbo, J. W. Ziller, L. E. Darago, J. R. Long, F. Furche and W. J. Evans, Isolation of +2 rare earth metal ions with three anionic carbocyclic rings: bimetallic bis(cyclopentadienyl) reduced arene complexes of La2+ and Ce2+ are four electron reductants, Chem. Sci., 2015, 6, 7267–7273 RSC.
- D. Patel, F. Moro, J. McMaster, W. Lewis, A. J. Blake and S. T. Liddle, A Formal High Oxidation State Inverse-Sandwich Diuranium Complex: A New Route to f-Block-Metal Bonds, Angew. Chem., Int. Ed., 2011, 50, 10388–10392 CrossRef CAS PubMed.
- W. Huang, F. Dulong, T. Wu, S. I. Khan, J. T. Miller, T. Cantat and P. L. Diaconescu, A six-carbon 10π-electron aromatic system supported by group 3 metals, Nat. Commun., 2013, 4, ncomms2473 CrossRef PubMed.
- C. Yu, J. Liang, C. Deng, G. Lefèvre, T. Cantat, P. L. Diaconescu and W. Huang, Arene-Bridged Dithorium Complexes: Inverse Sandwiches Supported by a δ Bonding Interaction, J. Am. Chem. Soc., 2020, 142, 21292–21297 CrossRef CAS PubMed.
- Y. Wang, Y. Zhang, J. Liang, B. Tan, C. Deng and W. Huang, Neutral inverse-sandwich rare-earth metal complexes of the benzene tetraanion, Chem. Sci., 2024, 15, 8740–8749 RSC.
- S. K. Thakur, N. Roig, R. Monreal-Corona, J. Langer, M. Alonso and S. Harder, Similarities and Differences in Benzene Reduction with Ca, Sr, Yb and Sm: Strong Evidence for Tetra-Anionic Benzene, Angew. Chem., Int. Ed., 2024, 63, e202405229 CrossRef CAS PubMed.
- C. A. Gould, J. Marbey, V. Vieru, D. A. Marchiori, R. David Britt, L. F. Chibotaru, S. Hill and J. R. Long, Isolation of a triplet benzene dianion, Nat. Chem., 2021, 13, 1001–1005 CrossRef CAS PubMed.
- F. Delano IV and S. Demir, Stabilization of the Compressed Planar Benzene Dianion in Inverse-Sandwich Rare Earth Metal Complexes, Angew. Chem., Int. Ed., 2025, 64, e202417217 CrossRef PubMed.
- W. Von, E. Doering and L. H. Knox, The Cycloheptatrienylium (Tropylium) Ion, J. Am. Chem. Soc., 1954, 76, 3203–3206 CrossRef.
- F. T. Zahra, A. Saeed, K. Mumtaz and F. Albericio, Tropylium Ion, an Intriguing Moiety in Organic Chemistry, Molecules, 2023, 28, 4095 CrossRef CAS PubMed.
- A. G. Harrison, L. R. Honnen, H. J. Dauben and F. P. Lossing, Free Radicals by Mass Spectrometry. XX. Ionization Potentials of Cyclopentadienyl and Cycloheptatrienyl Radicals, J. Am. Chem. Soc., 1960, 82, 5593–5598 CrossRef.
- K. Asai, A. Fukazawa and S. Yamaguchi, A Near–Infrared Dye That Undergoes Multiple Interconversions through Acid–Base Equilibrium and Reversible Redox Processes, Angew. Chem., Int. Ed., 2017, 56, 6848–6852 CrossRef CAS PubMed.
- T. Nishiuchi, R. Ito, A. Takada, Y. Yasuda, T. Nagata, E. Stratmann and T. Kubo, Anthracene-Attached Persistent Tricyclic Aromatic Hydrocarbon Radicals, Chem. – Asian J., 2019, 14, 1830–1836 CrossRef CAS PubMed.
- L. Moshniaha, M. Żyła-Karwowska, P. J. Chmielewski, T. Lis, J. Cybińska, E. Gońka, J. Oschwald, T. Drewello, S. M. Rivero, J. Casado and M. Stępień, Aromatic Nanosandwich Obtained by σ-Dimerization of a Nanographenoid π-Radical, J. Am. Chem. Soc., 2020, 142, 3626–3635 CrossRef CAS PubMed.
- J. Bloch, S. Kradolfer, T. L. Gianetti, D. Ostendorf, S. Dey, V. Mougel and H. Grützmacher, Synthesis and Characterization of Ion Pairs between Alkaline Metal Ions and Anionic Anti-Aromatic and Aromatic Hydrocarbons with π-Conjugated Central Seven- and Eight-Membered Rings, Molecules, 2020, 25, 4742 CrossRef CAS PubMed.
- Y. Hamamoto, K. Ochiai, Y. Li, E. Tapavicza and S. Ito, Synthesis and Properties of Azahomocorannulenyl Cations and Radicals, Angew. Chem., Int. Ed., 2024, 63, e202319022 CrossRef CAS PubMed.
- L. M. Tolbert and M. Z. Ali, 1-Phenyl-3,4:5,6-dibenzocycloheptatrienyl anion. A stable antiaromatic carbanion, J. Org. Chem., 1982, 47, 4793–4795 CrossRef CAS.
- Y. V. Tomilov, D. N. Platonov, R. F. Salikov and G. P. Okonnishnikova, Synthesis and properties of stable 1,2,3,4,5,6,7-heptamethoxycarbonylcyclohepta-2,4,6-trien-1-yl potassium and its reactions with electrophilic reagents, Tetrahedron, 2008, 64, 10201–10206 CrossRef CAS.
- J. J. Bahl, R. B. Bates, W. A. Beavers and C. R. Launer, Cycloheptatrienyl and heptatrienyl trianions, J. Am. Chem. Soc., 1977, 99, 6126–6127 CrossRef CAS.
- P. Von Ragué Schleyer, D. Wilhelm and T. Clark, On the nature of the “cycloheptatrienyl trianion”: The decisive importance of counterion stabilization in polymetallated “carbanion” systems, J. Organomet. Chem., 1985, 281, c17–c20 CrossRef.
- M. L. H. Green and D. K. P. Ng, Cycloheptatriene and -enyl Complexes of the Early Transition Metals, Chem. Rev., 1995, 95, 439–473 CrossRef CAS.
- A. Hauser, L. Münzfeld, S. Schlittenhardt, C. Uhlmann, L. Leyen, E. Moreno-Pineda, M. Ruben and P. W. Roesky, Cycloheptatrienyl-Bridged Triple-Decker Complexes, J. Am. Chem. Soc., 2024, 146, 13760–13769 CrossRef CAS PubMed.
- T. Arliguie, M. Lance, M. Nierlich and M. Ephritikhine, Inverse cycloheptatrienyl sandwich complexes of uranium and neodymium, J. Chem. Soc., Dalton Trans., 1997, 2501–2504 RSC.
- K. L. M. Harriman, J. J. L. Roy, L. Ungur, R. J. Holmberg, I. Korobkov and M. Murugesu, Cycloheptatrienyl trianion: an elusive bridge in the search of exchange coupled dinuclear organolanthanide single-molecule magnets, Chem. Sci., 2016, 8, 231–240 RSC.
- T. J. Katz, The Cycloöctatetraenyl Dianion, J. Am. Chem. Soc., 1960, 82, 3784–3785 CrossRef CAS.
- J. J. Le Roy, L. Ungur, I. Korobkov, L. F. Chibotaru and M. Murugesu, Coupling Strategies to Enhance Single-Molecule Magnet Properties of Erbium–Cyclooctatetraenyl Complexes, J. Am. Chem. Soc., 2014, 136, 8003–8010 CrossRef CAS PubMed.
- S.-D. Jiang, B.-W. Wang, H.-L. Sun, Z.-M. Wang and S. Gao, An Organometallic Single-Ion Magnet, J. Am. Chem. Soc., 2011, 133, 4730–4733 CrossRef CAS PubMed.
- K. R. Meihaus and J. R. Long, Magnetic Blocking at 10 K and a Dipolar-Mediated Avalanche in Salts of the Bis(η8-cyclooctatetraenide) Complex [Er(COT)2]−, J. Am. Chem. Soc., 2013, 135, 17952–17957 CrossRef CAS PubMed.
- K. Müllen, Reduction and oxidation of annulenes, Chem. Rev., 1984, 84, 603–646 CrossRef.
- W. Stawski, Y. Zhu, I. Rončević, Z. Wei, M. A. Petrukhina and H. L. Anderson, The anti-aromatic dianion and aromatic tetraanion of [18]annulene, Nat. Chem., 2024, 16, 998–1002 CrossRef CAS PubMed.
- J. F. M. Oth, E. P. Woo and F. Sondheimer, Unsaturated macrocyclic compounds. LXXXIX. Dianion of [18]annulene, J. Am. Chem. Soc., 1973, 95, 7337–7345 CrossRef CAS.
- P. Zhang, X. Xing, Y. Wang, I. Murtaza, Y. He, J. Cameron, S. Ahmed, P. J. Skabara and H. Meng, Multi-colour electrochromic materials based on polyaromatic esters with low driving voltage, J. Mater. Chem. C, 2019, 7, 9467–9473 RSC.
- S. Renault, J. Geng, F. Dolhem and P. Poizot, Evaluation of polyketones with N-cyclic structure as electrode material for electrochemical energy storage: case of pyromellitic diimide dilithium salt, Chem. Commun., 2011, 47, 2414–2416 RSC.
- S. Renault, V. A. Mihali, K. Edström and D. Brandell, Stability of organic Na-ion battery electrode materials: The case of disodium pyromellitic diimidate, Electrochem. Commun., 2014, 45, 52–55 CrossRef CAS.
- S. Wang, L. Wang, Z. Zhu, Z. Hu, Q. Zhao and J. Chen, All Organic Sodium-Ion Batteries with Na4C8H2O6, Angew. Chem., Int. Ed., 2014, 53, 5892–5896 CrossRef CAS PubMed.
- A. Mahata, S. Chandra, A. Maiti, D. K. Rao, C. B. Yildiz, B. Sarkar and A. Jana, α,α′-Diamino-p-quinodimethanes with Three Stable Oxidation States, Org. Lett., 2020, 22, 8332–8336 CrossRef CAS PubMed.
- A. Maiti, F. Zhang, I. Krummenacher, M. Bhattacharyya, S. Mehta, M. Moos, C. Lambert, B. Engels, A. Mondal, H. Braunschweig, P. Ravat and A. Jana, Anionic Boron- and Carbon-Based Hetero-Diradicaloids Spanned by a p -Phenylene Bridge, J. Am. Chem. Soc., 2021, 143, 3687–3692 CrossRef CAS PubMed.
- G. Seitz and P. Imming, Oxocarbons and pseudooxocarbons, Chem. Rev., 1992, 92, 1227–1260 CrossRef CAS.
- X. Hou, Y. Lu, Y. Ni, D. Zhang, Q. Zhao and J. Chen, Synthesis of a class of oxocarbons (C4O4, C5O5) and the application as high-capacity cathode materials for lithium-ion batteries, Sci. China: Chem., 2023, 66, 2780–2784 CrossRef CAS.
- Y. Lu, X. Hou, L. Miao, L. Li, R. Shi, L. Liu and J. Chen, Cyclohexanehexone with Ultrahigh Capacity as Cathode Materials for Lithium-Ion Batteries, Angew. Chem., Int. Ed., 2019, 58, 7020–7024 CrossRef CAS PubMed.
- J.-C. Guo, G.-L. Hou, S.-D. Li and X.-B. Wang, Probing the Low-Lying Electronic States of Cyclobutanetetraone (C4O4) and Its Radical Anion: A Low-Temperature Anion Photoelectron Spectroscopic Approach, J. Phys. Chem. Lett., 2012, 3, 304–308 CrossRef CAS PubMed.
- B. Chen, D. A. Hrovat, R. West, S. H. M. Deng, X.-B. Wang and W. T. Borden, The Negative Ion Photoelectron Spectrum of Cyclopropane-1,2,3-Trione Radical Anion, (CO)3·–—A Joint Experimental and Computational Study, J. Am. Chem. Soc., 2014, 136, 12345–12354 CrossRef CAS PubMed.
- D. Quiñonero, C. Garau, A. Frontera, P. Ballester, A. Costa and P. M. Deyà, Quantification of Aromaticity in Oxocarbons: The Problem of the Fictitious “Nonaromatic” Reference System, Chem. – Eur. J., 2002, 8, 433–438 CrossRef.
- M. Armand and J.-M. Tarascon, Building better batteries, Nature, 2008, 451, 652–657 CrossRef CAS PubMed.
- H. Chen, M. Armand, G. Demailly, F. Dolhem, P. Poizot and J.-M. Tarascon, From Biomass to a Renewable LiC6O6 Organic Electrode for Sustainable Li-Ion Batteries, ChemSusChem, 2008, 1, 348–355 CrossRef CAS.
- L. M. Doane and A. J. Fatiadi, Electrochemical oxidation of several oxocarbon salts in N,N-dimethylformamide, J. Electroanal. Chem. Interfacial Electrochem., 1982, 135, 193–209 CrossRef CAS.
- Q. Zhao, J. Wang, Y. Lu, Y. Li, G. Liang and J. Chen, Oxocarbon Salts for Fast Rechargeable Batteries, Angew. Chem., Int. Ed., 2016, 55, 12528–12532 CrossRef CAS PubMed.
- Th. Anderson, Ueber die Producte der trockenen Destillation thierischer Materien, Justus Liebigs Ann. Chem., 1870, 154, 270–286 CrossRef.
- L. Michaelis and E. S. Hill, The Viologen Indicators, J. Gen. Physiol., 1933, 16, 859–873 CrossRef CAS PubMed.
- K. Madasamy, D. Velayutham, V. Suryanarayanan, M. Kathiresan and K.-C. Ho, Viologen-based electrochromic materials and devices, J. Mater. Chem. C, 2019, 7, 4622–4637 RSC.
- K. W. Shah, S.-X. Wang, D. X. Y. Soo and J. Xu, Viologen-Based Electrochromic Materials: From Small Molecules, Polymers and Composites to Their Applications, Polymers, 2019, 11, 1839 CrossRef CAS PubMed.
- J. F. Stoddart, Mechanically Interlocked Molecules (MIMs)—Molecular Shuttles, Switches, and Machines (Nobel Lecture), Angew. Chem., Int. Ed., 2017, 56, 11094–11125 CrossRef CAS.
- K. Takahashi, T. Nihira, K. Akiyama, Y. Ikegami and E. Fukuyo, Synthesis and characterization of new conjugation-extended viologens involving a central aromatic linking group, J. Chem. Soc., Chem. Commun., 1992, 620–622 RSC.
- W. W. Porter, T. P. Vaid and A. L. Rheingold, Synthesis and Characterization of a Highly Reducing Neutral “Extended Viologen” and the Isostructural Hydrocarbon 4,4′′′′-Di-n-octyl-p-quaterphenyl, J. Am. Chem. Soc., 2005, 127, 16559–16566 CrossRef CAS PubMed.
- M. E. Carrington, K. Sokołowski, E. Jónsson, E. W. Zhao, A. M. Graf, I. Temprano, J. A. McCune, C. P. Grey and O. A. Scherman, Associative pyridinium electrolytes for air-tolerant redox flow batteries, Nature, 2023, 623, 949–955 CrossRef CAS.
- J. Casado, S. Patchkovskii, M. Z. Zgierski, L. Hermosilla, C. Sieiro, M. Moreno Oliva and J. T. López Navarrete, Raman Detection of “Ambiguous” Conjugated Biradicals: Rapid Thermal Singlet-to-Triplet Intersystem Crossing in an Extended Viologen, Angew. Chem., Int. Ed., 2008, 47, 1443–1446 CrossRef CAS PubMed.
- J. Luo, B. Hu, C. Debruler and T. L. Liu, A π-Conjugation Extended Viologen as a Two-Electron Storage Anolyte for Total Organic Aqueous Redox Flow Batteries, Angew. Chem., Int. Ed., 2018, 57, 231–235 CrossRef CAS.
- A. N. Woodward, J. M. Kolesar, S. R. Hall, N.-A. Saleh, D. S. Jones and M. G. Walter, Thiazolothiazole Fluorophores Exhibiting Strong Fluorescence and Viologen-Like Reversible Electrochromism, J. Am. Chem. Soc., 2017, 139, 8467–8473 CrossRef CAS PubMed.
- A. Beneduci, S. Cospito, M. La Deda, L. Veltri and G. Chidichimo, Electrofluorochromism in π-conjugated ionic liquid crystals, Nat. Commun., 2014, 5, 3105 CrossRef PubMed.
- M. Stolar, J. Borau-Garcia, M. Toonen and T. Baumgartner, Synthesis and Tunability of Highly Electron-Accepting, N-Benzylated “Phosphaviologens”, J. Am. Chem. Soc., 2015, 137, 3366–3371 CrossRef CAS PubMed.
- C. Reus, M. Stolar, J. Vanderkley, J. Nebauer and T. Baumgartner, A Convenient N-Arylation Route for Electron-Deficient Pyridines: The Case of π-Extended Electrochromic Phosphaviologens, J. Am. Chem. Soc., 2015, 137, 11710–11717 CrossRef CAS.
- G. Li, L. Xu, W. Zhang, K. Zhou, Y. Ding, F. Liu, X. He and G. He, Narrow-Bandgap Chalcogenoviologens for Electrochromism and Visible-Light-Driven Hydrogen Evolution, Angew. Chem., Int. Ed., 2018, 57, 4897–4901 CrossRef CAS PubMed.
- C. B. Nielsen, A. Angerhofer, K. A. Abboud and J. R. Reynolds, Discrete Photopatternable π-Conjugated Oligomers for Electrochromic Devices, J. Am. Chem. Soc., 2008, 130, 9734–9746 CrossRef CAS PubMed.
- P. W. Antoni, C. Golz and M. M. Hansmann, Organic Four-Electron Redox Systems Based on Bipyridine and Phenanthroline Carbene Architectures, Angew. Chem., Int. Ed., 2022, 61, e202203064 CrossRef CAS.
- X. Lu, S. Lee, J. O. Kim, T. Y. Gopalakrishna, H. Phan, T. S. Herng, Z. Lim, Z. Zeng, J. Ding, D. Kim and J. Wu, Stable 3,6-Linked Fluorenyl Radical Oligomers with Intramolecular Antiferromagnetic Coupling and Polyradical Characters, J. Am. Chem. Soc., 2016, 138, 13048–13058 CrossRef CAS PubMed.
- M. Hayakawa, N. Sunayama, S. I. Takagi, Y. Matsuo, A. Tamaki, S. Yamaguchi, S. Seki and A. Fukazawa, Flattened 1D fragments of fullerene C60 that exhibit robustness toward multi-electron reduction, Nat. Commun., 2023, 14, 2741 CrossRef CAS.
- J. E. Anthony, The Larger Acenes: Versatile Organic Semiconductors, Angew. Chem., Int. Ed., 2008, 47, 452–483 CrossRef CAS PubMed.
- H. Jiang, S. Zhu, Z. Cui, Z. Li, Y. Liang, J. Zhu, P. Hu, H.-L. Zhang and W. Hu, High-performance five-ring-fused organic semiconductors for field-effect transistors, Chem. Soc. Rev., 2022, 51, 3071–3122 RSC.
- R. Mitsuhashi, Y. Suzuki, Y. Yamanari, H. Mitamura, T. Kambe, N. Ikeda, H. Okamoto, A. Fujiwara, M. Yamaji, N. Kawasaki, Y. Maniwa and Y. Kubozono, Superconductivity in alkali-metal-doped picene, Nature, 2010, 464, 76–79 CrossRef CAS.
- Y. Takabayashi, M. Menelaou, H. Tamura, N. Takemori, T. Koretsune, A. Štefančič, G. Klupp, A. J. C. Buurma, Y. Nomura, R. Arita, D. Arčon, M. J. Rosseinsky and K. Prassides, π-electron S = ½ quantum spin-liquid state in an ionic polyaromatic hydrocarbon, Nat. Chem., 2017, 9, 635–643 CrossRef CAS PubMed.
- R. Einholz and H. F. Bettinger, Heptacene: Increased Persistence of a 4n+2 π-Electron Polycyclic Aromatic Hydrocarbon by Oxidation to the 4n π-Electron Dication, Angew. Chem., Int. Ed., 2013, 52, 9818–9820 CrossRef CAS PubMed.
- F. D. Romero, M. J. Pitcher, C. I. Hiley, G. F. S. Whitehead, S. Kar, A. Y. Ganin, D. Antypov, C. Collins, M. S. Dyer, G. Klupp, R. H. Colman, K. Prassides and M. J. Rosseinsky, Redox-controlled potassium intercalation into two polyaromatic hydrocarbon solids, Nat. Chem., 2017, 9, 644–652 CrossRef CAS PubMed.
- M. Pennachio, Z. Zhou, Z. Wei, S. Liu, A. Yu. Rogachev and M. A. Petrukhina, Doubly-Reduced Pentacene in Different Coordination Environments: X-ray Crystallographic and Theoretical Insights into Structural and Electronic Changes, Chem. – Eur. J., 2022, 28, e202104194 CrossRef CAS PubMed.
- H. Angliker, E. Rommel and J. Wirz, Electronic spectra of hexacene in solution (ground state. Triplet state. Dication and dianion), Chem. Phys. Lett., 1982, 87, 208–212 CrossRef CAS.
- J. Armand, L. Boulares, C. Bellec and J. Pinson, Electrochemical reduction of quinoxalino[2,3-b]quinoxaline, Can. J. Chem., 1982, 60, 2797–2803 CrossRef CAS.
- F. A. Cotton, Z. Li, C. Y. Liu, C. A. Murillo and D. Villagrán, Strong Electronic Interaction between Two Dimolybdenum Units Linked by a Tetraazatetracene, Inorg. Chem., 2006, 45, 767–778 CrossRef CAS PubMed.
- F. Benner and S. Demir, Isolation of Elusive Fluoflavine Radicals in Two Differing Oxidation States, J. Am. Chem. Soc., 2024, 146, 26008–26023 CrossRef CAS PubMed.
- S. Dong, T. S. Herng, T. Y. Gopalakrishna, H. Phan, Z. L. Lim, P. Hu, R. D. Webster, J. Ding and C. Chi, Extended Bis(benzothia)quinodimethanes and Their Dications: From Singlet Diradicaloids to Isoelectronic Structures of Long Acenes, Angew. Chem., Int. Ed., 2016, 55, 9316–9320 CrossRef CAS PubMed.
- N. Nagahora, T. Kushida, K. Shioji and K. Okuma, Dicationic Heteroacenes Containing Thio- or Selenopyrylium Moieties, Organometallics, 2019, 38, 1800–1808 CrossRef CAS.
- G. Xie, V. Brosius, J. Han, F. Rominger, A. Dreuw, J. Freudenberg and U. H. F. Bunz, Stable Radical Cations of N,N′-Diarylated Dihydrodiazapentacenes, Chem. – Eur. J., 2020, 26, 160–164 CrossRef CAS PubMed.
- M. Ito, M. Sakai, N. Ando and S. Yamaguchi, Electron-Deficient Heteroacenes that Contain Two Boron Atoms: Near-Infrared Fluorescence Based on a Push–Pull Effect, Angew. Chem., Int. Ed., 2021, 60, 21853–21859 CrossRef CAS PubMed.
- K. Kotwica, I. Wielgus and A. Proń, Azaacenes Based Electroactive Materials: Preparation, Structure, Electrochemistry, Spectroscopy and Applications—A Critical Review, Materials, 2021, 14, 5155 CrossRef CAS.
- S. Miao, A. L. Appleton, N. Berger, S. Barlow, S. R. Marder, K. I. Hardcastle and U. H. F. Bunz, 6,13-Diethynyl-5,7,12,14-tetraazapentacene, Chem. – Eur. J., 2009, 15, 4990–4993 CrossRef CAS PubMed.
- Y. Sakamoto, T. Suzuki, M. Kobayashi, Y. Gao, Y. Fukai, Y. Inoue, F. Sato and S. Tokito, Perfluoropentacene: High-Performance p−n Junctions and Complementary Circuits with Pentacene, J. Am. Chem. Soc., 2004, 126, 8138–8140 CrossRef CAS PubMed.
- L. Ji, M. Haehnel, I. Krummenacher, P. Biegger, F. L. Geyer, O. Tverskoy, M. Schaffroth, J. Han, A. Dreuw, T. B. Marder and U. H. F. Bunz, The Radical Anion and Dianion of Tetraazapentacene, Angew. Chem., Int. Ed., 2016, 55, 10498–10501 CrossRef CAS PubMed.
- Z. Liang, Q. Tang, J. Xu and Q. Miao, Soluble and Stable N-Heteropentacenes with High Field-Effect Mobility, Adv. Mater., 2011, 23, 1535–1539 CrossRef CAS PubMed.
- Z. Song, H. Zhan and Y. Zhou, Anthraquinone based polymer as high performance cathode material for rechargeable lithium batteries, Chem. Commun., 2009, 448–450 RSC.
- M. Yao, H. Senoh, T. Sakai and T. Kiyobayashi, 5,7,12,14-Pentacenetetrone as a High-Capacity Organic Positive-Electrode Material for Use in Rechargeable Lithium Batteries, Int. J. Electrochem. Sci., 2011, 6, 2905–2911 CrossRef CAS.
- D. Wu, Z. Xie, Z. Zhou, P. Shen and Z. Chen, Designing high-voltage carbonyl-containing polycyclic aromatic hydrocarbon cathode materials for Li-ion batteries guided by Clar's theory, J. Mater. Chem. A, 2015, 3, 19137–19143 RSC.
- C. K. Frederickson, B. D. Rose and M. M. Haley, Explorations of the Indenofluorenes and Expanded Quinoidal Analogues, Acc. Chem. Res., 2017, 50, 977–987 CrossRef CAS PubMed.
- Y. Tobe, Quinodimethanes Incorporated in Non-Benzenoid Aromatic or Antiaromatic Frameworks, Top. Curr. Chem., 2018, 376, 12 CrossRef PubMed.
- A. Shimizu, R. Kishi, M. Nakano, D. Shiomi, K. Sato, T. Takui, I. Hisaki, M. Miyata and Y. Tobe, Indeno[2,1-b]fluorene: A 20-π-Electron Hydrocarbon with Very Low-Energy Light Absorption, Angew. Chem., Int. Ed., 2013, 52, 6076–6079 CrossRef CAS PubMed.
- G. E. Rudebusch, J. L. Zafra, K. Jorner, K. Fukuda, J. L. Marshall, I. Arrechea-Marcos, G. L. Espejo, R. Ponce Ortiz, C. J. Gómez-García, L. N. Zakharov, M. Nakano, H. Ottosson, J. Casado and M. M. Haley, Diindeno-fusion of an anthracene as a design strategy for stable organic biradicals, Nat. Chem., 2016, 8, 753–759 CrossRef CAS PubMed.
- M. A. Majewski, P. J. Chmielewski, A. Chien, Y. Hong, T. Lis, M. Witwicki, D. Kim, P. M. Zimmerman and M. Stępień, 5,10-Dimesityldiindeno[1,2-a:2′,1′-i]phenanthrene: a stable biradicaloid derived from Chichibabin's hydrocarbon, Chem. Sci., 2019, 10, 3413–3420 RSC.
- H. Hayashi, J. E. Barker, A. Cárdenas Valdivia, R. Kishi, S. N. MacMillan, C. J. Gómez-García, H. Miyauchi, Y. Nakamura, M. Nakano, S. Kato, M. M. Haley and J. Casado, Monoradicals and Diradicals of Dibenzofluoreno[3,2-b]fluorene Isomers: Mechanisms of Electronic Delocalization, J. Am. Chem. Soc., 2020, 142, 20444–20455 CrossRef CAS PubMed.
- A. Borissov, P. J. Chmielewski, C. J. Gómez García, T. Lis and M. Stępień, Dinor[7]helicene and Beyond: Divergent Synthesis of Chiral Diradicaloids with Variable Open-Shell Character, Angew. Chem., Int. Ed., 2023, e202309238 CAS.
- A. Borissov, P. J. Chmielewski, A. C. Valdivia, C. J. G. García, J. Casado and M. Stępień, Triindenotriphenylenes: Spin-Frustrated Triskelion Triradicals with Excellent Ambient Stability, Angew. Chem., Int. Ed., 2024, 63, e202408510 CrossRef CAS PubMed.
- S. Mori, M. Akita, S. Suzuki, M. S. Asano, M. Murata, T. Akiyama, T. Matsumoto, C. Kitamura and S. Kato, Open-shell singlet diradicaloid difluoreno[4,3-b:3′,4′-d]furan and its radical cation and dianion, Chem. Commun., 2020, 56, 5881–5884 RSC.
- B. Prajapati, T. Kwenda, T. Lis, P. J. Chmielewski, C. J. Gómez-García, M. A. Majewski and M. Stępień, Difluorenoheteroles: topological control of π conjugation in diradicaloids and mixed-valence radical ions, Chem. Sci., 2024, 15, 10101–10109 RSC.
- T. Kubo, M. Sakamoto, M. Akabane, Y. Fujiwara, K. Yamamoto, M. Akita, K. Inoue, T. Takui and K. Nakasuji, Four-Stage Amphoteric Redox Properties and Biradicaloid Character of Tetra-tert-butyldicyclopenta[b;d]thieno[1,2,3-cd;5,6,7-c′d′]diphenalene, Angew. Chem., Int. Ed., 2004, 43, 6474–6479 CrossRef CAS PubMed.
- Z. Han, T. P. Vaid and A. L. Rheingold, Hexakis(4-(N-butylpyridylium))benzene: A Six-Electron Organic Redox System, J. Org. Chem., 2008, 73, 445–450 CrossRef CAS PubMed.
- Y. Yu, A. D. Bond, P. W. Leonard, U. J. Lorenz, T. V. Timofeeva, K. P. C. Vollhardt, G. D. Whitener and A. A. Yakovenko, Hexaferrocenylbenzene, Chem. Commun., 2006, 2572–2574 RSC.
- D. Sun, S. V. Rosokha and J. K. Kochi, Through-Space (Cofacial) π-Delocalization among Multiple Aromatic Centers: Toroidal Conjugation in Hexaphenylbenzene-like Radical Cations, Angew. Chem., Int. Ed., 2005, 44, 5133–5136 CrossRef CAS PubMed.
- S. V. Rosokha, I. S. Neretin, D. Sun and J. K. Kochi, Very Fast Electron Migrations within p-Doped Aromatic Cofacial Arrays Leading to Three-Dimensional (Toroidal) π-Delocalization, J. Am. Chem. Soc., 2006, 128, 9394–9407 CrossRef CAS PubMed.
- C. Lambert, Hexaarylbenzenes—Prospects for Toroidal Delocalization of Charge and Energy, Angew. Chem., Int. Ed., 2005, 44, 7337–7339 CrossRef CAS PubMed.
- J. J. Piet, H. A. M. Biemans, J. M. Warman and E. W. Meijer, Rapid rotation of energy in the excited state of a circular hexa-carbazole array, Chem. Phys. Lett., 1998, 289, 13–18 CrossRef CAS.
- B. Eberle, E. Kaifer and H.-J. Himmel, A Stable Hexakis(guanidino)benzene: Realization of the Strongest Neutral Organic Four-Electron Donor, Angew. Chem., Int. Ed., 2017, 56, 3360–3363 CrossRef CAS PubMed.
- M. Takase, N. Yoshida, T. Nishinaga and M. Iyoda, Star-Shaped Pyrrole-Fused Tetrathiafulvalene Oligomers: Synthesis and Redox, Self-Assembling, and Conductive Properties, Org. Lett., 2011, 13, 3896–3899 CrossRef CAS PubMed.
- R. Walser-Kuntz, Y. Yan, M. S. Sigman and M. S. Sanford, A Physical Organic Chemistry Approach to Developing Cyclopropenium-Based Energy Storage Materials for Redox Flow Batteries, Acc. Chem. Res., 2023, 56, 1239–1250 CrossRef CAS PubMed.
- Y. Yan, D. B. Vogt, T. P. Vaid, M. S. Sigman and M. S. Sanford, Development of High Energy Density Diaminocyclopropenium-Phenothiazine Hybrid Catholytes for Non-Aqueous Redox Flow Batteries, Angew. Chem., Int. Ed., 2021, 60, 27039–27045 CrossRef CAS PubMed.
- G. Maas, in Cross Conjugation, John Wiley & Sons, Ltd, 2016, pp. 79–116 Search PubMed.
- T. Fukunaga, Negatively substituted trimethylenecyclopropane dianions, J. Am. Chem. Soc., 1976, 98, 610–611 CrossRef CAS.
- T. Fukunaga, M. D. Gordon and P. J. Krusic, Negatively substituted trimethylenecyclopropanes and their radical anions, J. Am. Chem. Soc., 1976, 98, 611–613 CrossRef CAS.
- Y. Karpov, N. Kiriy, M. Al-Hussein, M. Hambsch, T. Beryozkina, V. Bakulev, S. C. B. Mannsfeld, B. Voit and A. Kiriy, Hexacyano-[3]-radialene anion-radical salts: a promising family of highly soluble p-dopants, Chem. Commun., 2018, 54, 307–310 RSC.
- N. A. Turner, M. B. Freeman, H. D. Pratt, A. E. Crockett, D. S. Jones, M. R. Anstey, T. M. Anderson and C. M. Bejger, Desymmetrized hexasubstituted [3]radialene anions as aqueous organic catholytes for redox flow batteries, Chem. Commun., 2020, 56, 2739–2742 RSC.
- G. Xu, H. Liu, Z. Zhou, W. Lai, B. Li, Y. Zhou, R. Hu, W. Yao, K. Yang, S. Xie and Z. Zeng, α-Cyano Triaryl[3]radialene: Unsymmetrical Stereo-configuration, Clustering-enhanced Excimer Emission, and Radical-involved Multimodal Information Switching, Angew. Chem., Int. Ed., 2023, 62, e202305011 CrossRef CAS PubMed.
- K. Matsumoto, N. Yamada, T. Enomoto, H. Kurata, T. Kawase and M. Oda, Triarylamines on [3]Radialene Scaffold: Novel [3]Radialene-based, Multistep, Wide-range Redox Systems with Remarkably Low Oxidation Potentials, Chem. Lett., 2011, 40, 1033–1035 CrossRef CAS.
- M. Horner and S. Hünig, Reversible Transition between a [4]Radialene and a Cyclobutadiene by Electron Transfer, Angew. Chem., Int. Ed. Engl., 1977, 16, 410–411 CrossRef.
- M. Horner, S. Hünig and H.-U. Reißig, Mehrstufige, reversible Redoxsysteme, XXXIII. Reversible Umwandlung eines [4]Radialens in ein Cyclobutadien durch vierstufige Elektronenübertragung, Liebigs Ann. Chem., 1983, 1983, 658–667 CrossRef.
- K. Lincke, A. Floor Frellsen, C. R. Parker, A. D. Bond, O. Hammerich and M. Brøndsted Nielsen, A Tetrathiafulvalene-Functionalized Radiaannulene with Multiple Redox States, Angew. Chem., Int. Ed., 2012, 51, 6099–6102 CrossRef CAS PubMed.
- C.-M. Chou, S. Saito and S. Yamaguchi, Heterotriangulenes π-Expanded at Bridging Positions, Org. Lett., 2014, 16, 2868–2871 CrossRef CAS PubMed.
- R. Rathore, S. V. Lindeman, A. S. Kumar and J. K. Kochi, Disproportionation and Structural Changes of Tetraarylethylene Donors upon Successive Oxidation to Cation Radicals and to Dications, J. Am. Chem. Soc., 1998, 120, 6931–6939 CrossRef CAS.
- N. S. Mills, J. L. Malandra, E. E. Burns, A. Green, K. E. Unruh, D. E. Kadlecek and J. A. Lowery, Dications of Fluorenylidenes: Conformational and Electronic Effects on the Paratropicity/Antiaromaticity of Fluorenyl Cations with Cyclic Substituents, J. Org. Chem., 1997, 62, 9318–9322 CrossRef CAS.
- Y. Cohen, J. Klein and M. Rabinovitz, Stable polycyclic anions: dianions from overcrowded ethylenes, J. Chem. Soc., Chem. Commun., 1986, 1071 RSC.
- B. Prajapati, M. D. Ambhore, D.-K. Dang, P. J. Chmielewski, T. Lis, C. J. Gómez-García, P. M. Zimmerman and M. Stępień, Tetrafluorenofulvalene as a sterically frustrated open-shell alkene, Nat. Chem., 2023, 15, 1541–1548 CrossRef CAS PubMed.
- J. E. Kwon, C.-S. Hyun, Y. J. Ryu, J. Lee, D. J. Min, M. J. Park, B.-K. An and S. Y. Park, Triptycene-based quinone molecules showing multi-electron redox reactions for large capacity and high energy organic cathode materials in Li-ion batteries, J. Mater. Chem. A, 2018, 6, 3134–3140 RSC.
- T. Aoki, H. Sotome, D. Shimizu, H. Miyasaka and K. Matsuda, Propeller-Shaped Blatter-Based Triradicals: Distortion-Free Triangular Spin System and Spin-State-Dependent Photophysical Properties, Angew. Chem., Int. Ed., 2025, 64, e202418655 CrossRef CAS PubMed.
- S. R. Peurifoy, E. Castro, F. Liu, X.-Y. Zhu, F. Ng, S. Jockusch, M. L. Steigerwald, L. Echegoyen, C. Nuckolls and T. J. Sisto, Three-Dimensional Graphene Nanostructures, J. Am. Chem. Soc., 2018, 140, 9341–9345 CrossRef CAS PubMed.
- T. Kodama, M. Aoba, Y. Hirao, S. M. Rivero, J. Casado and T. Kubo, Molecular and Spin Structures of a Through-Space Conjugated Triradical System, Angew. Chem., Int. Ed., 2022, 61, e202200688 CrossRef CAS PubMed.
- Y. Chen, Y. Zhao, Y. Zhao, X. Chen, X. Liu, L. Li, D. Cao, S. Wang and L. Zhang, A Novel Homoconjugated Propellane Triimide: Synthesis, Structural Analyses, and Gas Separation, Angew. Chem., Int. Ed., 2024, 63, e202401706 CrossRef CAS PubMed.
- C. Bruno, E. Ussano, G. Barucca, D. Vanossi, G. Valenti, E. A. Jackson, A. Goldoni, L. Litti, S. Fermani, L. Pasquali, M. Meneghetti, C. Fontanesi, L. T. Scott, F. Paolucci and M. Marcaccio, Wavy graphene sheets from electrochemical sewing of corannulene, Chem. Sci., 2021, 12, 8048–8057 RSC.
- M. Baumgarten, L. Gherghel, M. Wagner, A. Weitz, M. Rabinovitz, P.-C. Cheng and L. T. Scott, Corannulene Reduction: Spectroscopic Detection of All Anionic Oxidation States, J. Am. Chem. Soc., 1995, 117, 6254–6257 CrossRef CAS.
- A. Ayalon, M. Rabinovitz, P.-C. Cheng and L. T. Scott, Corannulene Tetraanion: A Novel Species with Concentric Anionic Rings, Angew. Chem., Int. Ed. Engl., 1992, 31, 1636–1637 CrossRef.
- A. Ayalon, A. Sygula, P.-C. Cheng, M. Rabinovitz, P. W. Rabideau and L. T. Scott, Stable High-Order Molecular Sandwiches: Hydrocarbon Polyanion Pairs with Multiple Lithium Ions Inside and Out, Science, 1994, 265, 1065–1067 CrossRef CAS PubMed.
- A. V. Zabula, A. S. Filatov, S. N. Spisak, A. Yu. Rogachev and M. A. Petrukhina, A Main Group Metal Sandwich: Five Lithium Cations Jammed Between Two Corannulene Tetraanion Decks, Science, 2011, 333, 1008–1011 CrossRef CAS PubMed.
- M. Bühl and A. Hirsch, Spherical Aromaticity of Fullerenes, Chem. Rev., 2001, 101, 1153–1184 CrossRef PubMed.
- Q. Wang, T. Y. Gopalakrishna, H. Phan, T. S. Herng, S. Dong, J. Ding and C. Chi, Cyclopenta Ring Fused Bisanthene and Its Charged Species with Open-Shell Singlet Diradical Character and Global Aromaticity/Anti-Aromaticity, Angew. Chem., Int. Ed., 2017, 56, 11415–11419 CrossRef CAS PubMed.
- Y. Zou, W. Zeng, T. Y. Gopalakrishna, Y. Han, Q. Jiang and J. Wu, Dicyclopenta[4,3,2,1-ghi:4′,3′,2′,1′-pqr]perylene: A Bowl-Shaped Fragment of Fullerene C70 with Global Antiaromaticity, J. Am. Chem. Soc., 2019, 141, 7266–7270 CrossRef CAS PubMed.
- S. Nobusue, H. Miyoshi, A. Shimizu, I. Hisaki, K. Fukuda, M. Nakano and Y. Tobe, Tetracyclopenta[def,jkl,pqr,vwx]tetraphenylene: A Potential Tetraradicaloid Hydrocarbon, Angew. Chem., Int. Ed., 2015, 54, 2090–2094 CrossRef CAS PubMed.
- H. Miyoshi, R. Sugiura, R. Kishi, S. N. Spisak, Z. Wei, A. Muranaka, M. Uchiyama, N. Kobayashi, S. Chatterjee, Y. Ie, I. Hisaki, M. A. Petrukhina, T. Nishinaga, M. Nakano and Y. Tobe, Dianion and Dication of Tetracyclopentatetraphenylene as Decoupled Annulene-within-an-Annulene Models, Angew. Chem., Int. Ed., 2022, 61, e202115316 CrossRef CAS PubMed.
- L. Echegoyen and L. E. Echegoyen, Electrochemistry of Fullerenes and Their Derivatives, Acc. Chem. Res., 1998, 31, 593–601 CrossRef CAS.
- C. A. Reed and R. D. Bolskar, Discrete Fulleride Anions and Fullerenium Cations, Chem. Rev., 2000, 100, 1075–1120 CrossRef CAS PubMed.
- Q. Xie, E. Perez-Cordero and L. Echegoyen, Electrochemical detection of C606- and C706-: Enhanced stability of fullerides in solution, J. Am. Chem. Soc., 1992, 114, 3978–3980 CrossRef CAS.
- A. F. Hebard, M. J. Rosseinsky, R. C. Haddon, D. W. Murphy, S. H. Glarum, T. T. M. Palstra, A. P. Ramirez and A. R. Kortan, Superconductivity at 18 K in potassium-doped C60, Nature, 1991, 350, 600–601 CrossRef CAS.
- M. J. Rosseinsky, Recent Developments in the Chemistry and Physics of Metal Fullerides, Chem. Mater., 1998, 10, 2665–2685 CrossRef CAS.
- Y. Takabayashi and K. Prassides, in 50 Years of Structure and Bonding – The Anniversary Volume, ed. D. M. P. Mingos, Springer International Publishing, Cham, 2017, pp. 119–138 Search PubMed.
- E. Shabtai, A. Weitz, R. C. Haddon, R. E. Hoffman, M. Rabinovitz, A. Khong, R. J. Cross, M. Saunders, P.-C. Cheng and L. T. Scott, 3He NMR of He@C606- and He@C706-. New Records for the Most Shielded and the Most Deshielded 3He Inside a Fullerene1, J. Am. Chem. Soc., 1998, 120, 6389–6393 CrossRef CAS.
- L. Hou, X. Cui, B. Guan, S. Wang, R. Li, Y. Liu, D. Zhu and J. Zheng, Synthesis of a monolayer fullerene network, Nature, 2022, 606, 507–510 CrossRef CAS PubMed.
- E. Meirzadeh, A. M. Evans, M. Rezaee, M. Milich, C. J. Dionne, T. P. Darlington, S. T. Bao, A. K. Bartholomew, T. Handa, D. J. Rizzo, R. A. Wiscons, M. Reza, A. Zangiabadi, N. Fardian-Melamed, A. C. Crowther, P. J. Schuck, D. N. Basov, X. Zhu, A. Giri, P. E. Hopkins, P. Kim, M. L. Steigerwald, J. Yang, C. Nuckolls and X. Roy, A few-layer covalent network of fullerenes, Nature, 2023, 613, 71–76 CrossRef CAS PubMed.
- A. Hirsch, Z. Chen and H. Jiao, Spherical Aromaticity in Ih Symmetrical Fullerenes: The 2(N+1)2 Rule, Angew. Chem., Int. Ed., 2000, 39, 3915–3917 CrossRef CAS PubMed.
- U. Zimmermann, N. Malinowski, A. Burkhardt and T. P. Martin, Metal-coated fullerenes, Carbon, 1995, 33, 995–1006 CrossRef CAS.
- J. Kohanoff, W. Andreoni and M. Parrinello, A possible new highly stable fulleride cluster: Li12C60, Chem. Phys. Lett., 1992, 198, 472–477 CrossRef CAS.
- L. Cristofolini, M. Riccò and R. De Renzi, NMR and high-resolution X-ray diffraction evidence for an alkali-metal fulleride with large interstitial clusters: ${\mathrm{Li}}_{12}{\mathrm{C}}_{60}$, Phys. Rev. B:Condens. Matter Mater. Phys., 1999, 59, 8343–8346 CrossRef CAS.
- F. Giglio, D. Pontiroli, M. Gaboardi, M. Aramini, C. Cavallari, M. Brunelli, P. Galinetto, C. Milanese and M. Riccò, Li12C60: A lithium clusters intercalated fulleride, Chem. Phys. Lett., 2014, 609, 155–160 CrossRef CAS.
- Z. Jiang, Y. Zhao, X. Lu and J. Xie, Fullerenes for rechargeable battery applications: Recent developments and future perspectives, J. Energy Chem., 2021, 55, 70–79 CrossRef CAS.
- E. Gońka, P. J. Chmielewski, T. Lis and M. Stępień, Expanded Hexapyrrolohexaazacoronenes. Near-Infrared Absorbing Chromophores with Interrupted Peripheral Conjugation, J. Am. Chem. Soc., 2014, 136, 16399–16410 CrossRef PubMed.
- M. Takase, V. Enkelmann, D. Sebastiani, M. Baumgarten and K. Müllen, Annularly Fused Hexapyrrolohexaazacoronenes: An Extended π
System with Multiple Interior Nitrogen Atoms Displays Stable Oxidation States, Angew.
Chem., Int. Ed., 2007, 46, 5524–5527 CrossRef CAS PubMed. - K. Oki, M. Takase, N. Kobayashi and H. Uno, Synthesis and Characterization of Peralkylated Pyrrole-Fused Azacoronene, J. Org. Chem., 2021, 86, 5102–5109 CrossRef CAS PubMed.
- M. Żyła-Karwowska, H. Zhylitskaya, J. Cybińska, T. Lis, P. J. Chmielewski and M. Stępień, An Electron–Deficient Azacoronene Obtained by Radial π Extension, Angew. Chem., Int. Ed., 2016, 55, 14658–14662 CrossRef PubMed.
- M. Navakouski, H. Zhylitskaya, P. J. Chmielewski, T. Lis, J. Cybińska and M. Stępień, Stereocontrolled Synthesis of Chiral Heteroaromatic Propellers with Small Optical Bandgaps, Angew. Chem., Int. Ed., 2019, 58, 4929–4933 CrossRef CAS PubMed.
- M. Żyła, E. Gońka, P. J. Chmielewski, J. Cybińska and M. Stępień, Synthesis of a peripherally conjugated 5–6–7 nanographene, Chem. Sci., 2016, 7, 286–294 RSC.
- K. Oki, M. Takase, S. Mori and H. Uno, Synthesis and Isolation of Antiaromatic Expanded Azacoronene via Intramolecular Vilsmeier-Type Reaction, J. Am. Chem. Soc., 2019, 141, 16255–16259 CrossRef CAS PubMed.
- M. Takase, T. Narita, W. Fujita, M. S. Asano, T. Nishinaga, H. Benten, K. Yoza and K. Müllen, Pyrrole-Fused Azacoronene Family: The Influence of Replacement with Dialkoxybenzenes on the Optical and Electronic Properties in Neutral and Oxidized States, J. Am. Chem. Soc., 2013, 135, 8031–8040 CrossRef CAS PubMed.
- K. Oki, M. Takase, S. Mori, A. Shiotari, Y. Sugimoto, K. Ohara, T. Okujima and H. Uno, Synthesis, Structures, and Properties of Core-Expanded Azacoronene Analogue: A Twisted π-System with Two N-Doped Heptagons, J. Am. Chem. Soc., 2018, 140, 10430–10434 CrossRef CAS PubMed.
- A. A. Suleymanov, Q. He, P. Müller and T. M. Swager, Highly Contorted Rigid Nitrogen-Rich Nanographene with Four Heptagons, Org. Lett., 2024, 26, 5227–5231 CrossRef PubMed.
- T. Ohmae, T. Nishinaga, M. Wu and M. Iyoda, Cyclic Tetrathiophenes Planarized by Silicon and Sulfur Bridges Bearing Antiaromatic Cyclooctatetraene Core: Syntheses, Structures, and Properties, J. Am. Chem. Soc., 2010, 132, 1066–1074 CrossRef CAS PubMed.
- K. Mouri, S. Saito and S. Yamaguchi, Highly Flexible π-Expanded Cyclooctatetraenes: Cyclic Thiazole Tetramers with Head-to-Tail Connection, Angew. Chem., Int. Ed., 2012, 51, 5971–5975 CrossRef CAS PubMed.
- Z. Zhou, Y. Zhu, Z. Wei, J. Bergner, C. Neiß, S. Doloczki, A. Görling, M. Kivala and M. A. Petrukhina, Reduction of π-Expanded Cyclooctatetraene with Lithium: Stabilization of the Tetra-Anion through Internal Li+ Coordination, Angew. Chem., Int. Ed., 2021, 60, 3510–3514 CrossRef CAS PubMed.
- E. Misselwitz, J. Spengler, F. Rominger and M. Kivala, Indenoannulated Tridecacyclene: An All-Carbon Seven-Stage Redox-Amphoter, Chem. – Eur. J., 2024, 30, e202400696 CrossRef CAS PubMed.
- Y. Sakamoto and T. Suzuki, Tetrabenzo[8]circulene: Aromatic Saddles from Negatively Curved Graphene, J. Am. Chem. Soc., 2013, 135, 14074–14077 CrossRef CAS PubMed.
- S. H. Pun, Y. Wang, M. Chu, C. K. Chan, Y. Li, Z. Liu and Q. Miao, Synthesis, Structures, and Properties of Heptabenzo[7]circulene and Octabenzo[8]circulene, J. Am. Chem. Soc., 2019, 141, 9680–9686 CrossRef CAS PubMed.
- C. B. Nielsen, T. Brock-Nannestad, T. K. Reenberg, P. Hammershøj, J. B. Christensen, J. W. Stouwdam and M. Pittelkow, Organic Light-Emitting Diodes from Symmetrical and Unsymmetrical π-Extended Tetraoxa[8]circulenes, Chem. – Eur. J., 2010, 16, 13030–13034 CrossRef CAS PubMed.
- F. Chen, Y. S. Hong, S. Shimizu, D. Kim, T. Tanaka and A. Osuka, Synthesis of a Tetrabenzotetraaza[8]circulene by a “Fold-In” Oxidative Fusion Reaction, Angew. Chem., Int. Ed., 2015, 54, 10639–10642 CrossRef CAS PubMed.
- Y. Matsuo, T. Tanaka and A. Osuka, Highly Stable Radical Cations of N,N′-Diarylated Tetrabenzotetraaza[8]circulene, Chem. – Eur. J., 2020, 26, 8144–8152 CrossRef CAS PubMed.
- V. B. R. Pedersen, S. K. Pedersen, Z. Jin, N. Kofod, B. W. Laursen, G. V. Baryshnikov, C. Nuckolls and M. Pittelkow, Electronic Materials: An Antiaromatic Propeller Made from the Four–Fold Fusion of Tetraoxa[8]circulene and Perylene Diimides, Angew. Chem., Int. Ed., 2022, 61, e202212293 CrossRef CAS PubMed.
- Y. Liang, P. Zhang and J. Chen, Function-oriented design of conjugated carbonyl compound electrodes for high energy lithium batteries, Chem. Sci., 2013, 4, 1330–1337 RSC.
- Y. Morita, S. Nishida, T. Murata, M. Moriguchi, A. Ueda, M. Satoh, K. Arifuku, K. Sato and T. Takui, Organic tailored batteries materials using stable open-shell molecules with degenerate frontier orbitals, Nat. Mater., 2011, 10, 947–951 CrossRef CAS PubMed.
- S.-L. Suraru and F. Würthner, Strategies for the Synthesis of Functional Naphthalene Diimides, Angew. Chem., Int. Ed., 2014, 53, 7428–7448 CrossRef CAS PubMed.
- G. S. Vadehra, R. P. Maloney, M. A. Garcia-Garibay and B. Dunn, Naphthalene Diimide Based Materials with Adjustable Redox Potentials: Evaluation for Organic Lithium-Ion Batteries, Chem. Mater., 2014, 26, 7151–7157 CrossRef CAS.
- X. Liu, H. Zhang, C. Liu, Z. Wang, X. Zhang, H. Yu, Y. Zhao, M.-J. Li, Y. Li, Y.-L. He and G. He, Commercializable Naphthalene Diimide Anolytes for Neutral Aqueous Organic Redox Flow Batteries, Angew. Chem., Int. Ed., 2024, 63, e202405427 CrossRef CAS PubMed.
- K. Tajima, K. Matsuo, H. Yamada, N. Fukui and H. Shinokubo, Diazazethrene bisimide: a strongly electron-accepting π-system synthesized via the incorporation of both imide substituents and imine-type nitrogen atoms into zethrene, Chem. Sci., 2023, 14, 635–642 RSC.
- C. Peng, G.-H. Ning, J. Su, G. Zhong, W. Tang, B. Tian, C. Su, D. Yu, L. Zu, J. Yang, M.-F. Ng, Y.-S. Hu, Y. Yang, M. Armand and K. P. Loh, Reversible multi-electron redox chemistry of π-conjugated N-containing heteroaromatic molecule-based organic cathodes, Nat. Energy, 2017, 2, 1–9 Search PubMed.
- B. Wang, J. Li, M. Ye, Y. Zhang, Y. Tang, X. Hu, J. He and C. C. Li, Dual-Redox Sites Guarantee High-Capacity Sodium Storage in Two-Dimension Conjugated Metal–Organic Frameworks, Adv. Funct. Mater., 2022, 32, 2112072 CrossRef CAS.
- S. Li, S. Hu, H. Li and C. Han, Initiating a High-Rate and Stable Aqueous Air Battery by Using Organic N-Heterocycle Anode, Angew. Chem., Int. Ed., 2024, 63, e202318885 CrossRef CAS PubMed.
- M. Yang, X. Zeng, M. Xie, Y. Wang, J.-M. Xiao, R.-H. Chen, Z.-J. Yi, Y.-F. Huang, D.-S. Bin and D. Li, Conductive Metal–Organic Framework with Superior Redox Activity as a Stable High-Capacity Anode for High-Temperature K-Ion Batteries, J. Am. Chem. Soc., 2024, 146, 6753–6762 CrossRef CAS PubMed.
- S. Seifert, K. Shoyama, D. Schmidt and F. Würthner, An Electron-Poor C64 Nanographene by Palladium-Catalyzed Cascade C−C Bond Formation: One-Pot Synthesis and Single-Crystal Structure Analysis, Angew. Chem., Int. Ed., 2016, 55, 6390–6395 CrossRef CAS PubMed.
- K. Shoyama and F. Würthner, Synthesis of a Carbon Nanocone by Cascade Annulation, J. Am. Chem. Soc., 2019, 141, 13008–13012 CrossRef CAS PubMed.
- H. Zhylitskaya, J. Cybińska, P. Chmielewski, T. Lis and M. Stępień, Bandgap Engineering in π-Extended Pyrroles. A Modular Approach to Electron-Deficient Chromophores with Multi-Redox Activity, J. Am. Chem. Soc., 2016, 138, 11390–11398 CrossRef CAS PubMed.
- R. Kumar, P. J. Chmielewski, T. Lis, D. Volkmer and M. Stępień, Tridecacyclene Tetraimide: An Easily Reduced Cyclooctatetraene Derivative, Angew. Chem., Int. Ed., 2022, 61, e202207486 CrossRef CAS PubMed.
- R. Kumar, P. J. Chmielewski, T. Lis, M. Czarnecki and M. Stępień, Pentacosacyclenes: cruciform molecular nanocarbons based on cyclooctatetraene, Chem. Sci., 2024, 15, 18640–18649 RSC.
- E. Misselwitz, J. Spengler, F. Rominger and M. Kivala, Octacyano-substituted tridecacyclene: A non-benzenoid cyanocarbon with low-lying LUMO and multistage redox properties, Chem. Commun., 2025, 61, 5754–5757 RSC.
- M. Ishida, S.-J. Kim, C. Preihs, K. Ohkubo, J. M. Lim, B. S. Lee, J. S. Park, V. M. Lynch, V. V. Roznyatovskiy, T. Sarma, P. K. Panda, C.-H. Lee, S. Fukuzumi, D. Kim and J. L. Sessler, Protonation-coupled redox reactions in planar antiaromatic meso-pentafluorophenyl-substituted o-phenylene-bridged annulated rosarins, Nat. Chem., 2012, 5, 15–20 CrossRef PubMed.
- T. Sarma, G. Kim, S. Sen, W.-Y. Cha, Z. Duan, M. D. Moore, V. M. Lynch, Z. Zhang, D. Kim and J. L. Sessler, Proton-Coupled Redox Switching in an Annulated π-Extended Core-Modified Octaphyrin, J. Am. Chem. Soc., 2018, 140, 12111–12119 CrossRef CAS PubMed.
- R. Li, C.-H. Mak, K. Luo, F. Wang, G. Liu, B. Tan, G. Liu, Y. Xu, Y.-D. Yang, H. Wei, Z. Huang, L. Wu, X. Zhao, H.-Y. Hsu, Q. He, J. L. Sessler and Z. Zhang, Cyclo[2]pyridine[8]pyrrole: An Expanded Porphyrinoid with Three Closed-Shell Oxidation States, J. Am. Chem. Soc., 2024, 146, 3585–3590 CrossRef CAS PubMed.
- K. M. Kadish, W. R. Osterloh and E. V. Caemelbecke, Electrochemistry Of Metalloporphyrins, World Scientific, 2023 Search PubMed.
- R. Cosmo, C. Kautz, K. Meerholz, J. Heinze and K. Müllen, Highly Reduced Porphyrins, Angew. Chem., Int. Ed. Engl., 1989, 28, 604–607 CrossRef.
- P. Gao, Z. Chen, Z. Zhao-Karger, J. E. Mueller, C. Jung, S. Klyatskaya, T. Diemant, O. Fuhr, T. Jacob, R. J. Behm, M. Ruben and M. Fichtner, A Porphyrin Complex as a Self-Conditioned Electrode Material for High-Performance Energy Storage, Angew. Chem., Int. Ed., 2017, 56, 10341–10346 CrossRef CAS PubMed.
- T. Satoh, M. Minoura, H. Nakano, K. Furukawa and Y. Matano, Redox-Switchable 20π-, 19π-, and 18π-Electron 5,10,15,20-Tetraaryl-5,15-diazaporphyrinoid Nickel(II) Complexes, Angew. Chem., Int. Ed., 2016, 55, 2235–2238 CrossRef CAS PubMed.
- K. Sudoh, T. Satoh, T. Amaya, K. Furukawa, M. Minoura, H. Nakano and Y. Matano, Syntheses, Properties, and Catalytic Activities of Metal(II) Complexes and Free Bases of Redox-Switchable 20π, 19π, and 18π 5,10,15,20-Tetraaryl-5,15-diazaporphyrinoids, Chem. – Eur. J., 2017, 23, 16364–16373 CrossRef CAS PubMed.
- J. Hao, A. Nishiyama, S. Mori, K. Furukawa and S. Shimizu, Oxidation of 5,15-Dioxaporphyrin: Its Generality and Novelty as an Oxaporphyrin Analogue, Angew. Chem., Int. Ed., 2023, 62, e202307862 CrossRef CAS PubMed.
- N. Sprutta, M. Świderska and L. Latos-Grażyński, Dithiadiazuliporphyrin: Facile Generation of Carbaporphyrinoid Cation Radical and Dication, J. Am. Chem. Soc., 2005, 127, 13108–13109 CrossRef CAS PubMed.
- N. Sprutta, M. Siczek, L. Latos-Grażyński, M. Pawlicki, L. Szterenberg and T. Lis, Dioxadiazuliporphyrin: A Near-IR Redox Switchable Chromophore, J. Org. Chem., 2007, 72, 9501–9509 CrossRef CAS PubMed.
- C. Hunt, M. Peterson, C. Anderson, T. Chang, G. Wu, S. Scheiner and G. Ménard, Switchable Aromaticity in an Isostructural Mn Phthalocyanine Series Isolated in Five Separate Redox States, J. Am. Chem. Soc., 2019, 141, 2604–2613 CrossRef CAS PubMed.
- A. Jana, M. Ishida, J. S. Park, S. Bähring, J. O. Jeppesen and J. L. Sessler, Tetrathiafulvalene- (TTF-) Derived Oligopyrrolic Macrocycles, Chem. Rev., 2017, 117, 2641–2710 CrossRef CAS PubMed.
- A. Jana, M. Ishida, K. Kwak, Y. M. Sung, D. S. Kim, V. M. Lynch, D. Lee, D. Kim and J. L. Sessler, Comparative Electrochemical and Photophysical Studies of Tetrathiafulvalene-Annulated Porphyrins and Their ZnII Complexes: The Effect of Metalation and Structural Variation, Chem. – Eur. J., 2013, 19, 338–349 CrossRef CAS PubMed.
- K. A. Nielsen, E. Levillain, V. M. Lynch, J. L. Sessler and J. O. Jeppesen, Tetrathiafulvalene Porphyrins, Chem. – Eur. J., 2009, 15, 506–516 CrossRef CAS PubMed.
- A. Jana, M. Ishida and H. Furuta, Benzo-Tetrathiafulvalene- (BTTF-) Annulated Expanded Porphyrins: Potential Next-Generation Multielectron Reservoirs, Chem. – Eur. J., 2021, 27, 4466–4472 CrossRef CAS PubMed.
- M. Bröring, S. Köhler and C. Kleeberg, Norcorrole: Observation of the Smallest Porphyrin Variant with a N4 Core, Angew. Chem., Int. Ed., 2008, 47, 5658–5660 CrossRef PubMed.
- T. Ito, Y. Hayashi, S. Shimizu, J.-Y. Shin, N. Kobayashi and H. Shinokubo, Gram-Scale Synthesis of Nickel(II) Norcorrole: The Smallest Antiaromatic Porphyrinoid, Angew. Chem., Int. Ed., 2012, 51, 8542–8545 CrossRef CAS PubMed.
- J.-Y. Shin, T. Yamada, H. Yoshikawa, K. Awaga and H. Shinokubo, An Antiaromatic Electrode-Active Material Enabling High Capacity and Stable Performance of Rechargeable Batteries, Angew. Chem., Int. Ed., 2014, 53, 3096–3101 CrossRef CAS PubMed.
- T. Y. Gopalakrishna and V. G. Anand, Reversible Redox Reaction Between Antiaromatic and Aromatic States of 32π-Expanded Isophlorins, Angew. Chem., Int. Ed., 2014, 53, 6678–6682 CrossRef CAS PubMed.
- T. Y. Gopalakrishna, J. S. Reddy and V. G. Anand, An Amphoteric Switch to Aromatic and Antiaromatic States of a Neutral Air-Stable 25π Radical, Angew. Chem., Int. Ed., 2014, 53, 10984–10987 CrossRef CAS PubMed.
- Y. Ni, T. Y. Gopalakrishna, S. Wu and J. Wu, A Stable All-Thiophene-Based Core-Modified [38]Octaphyrin Diradicaloid: Conformation and Aromaticity Switch at Different Oxidation States, Angew. Chem., Int. Ed., 2020, 59, 7414–7418 CrossRef CAS PubMed.
- T. Yoneda, T. Soya, S. Neya and A. Osuka, [62]Tetradecaphyrin and Its Mono- and Bis-ZnII Complexes, Chem. – Eur. J., 2016, 22, 14518–14522 CrossRef PubMed.
- H. Mori, T. Tanaka, S. Lee, J. M. Lim, D. Kim and A. Osuka, meso–meso Linked Porphyrin–[26]Hexaphyrin–Porphyrin Hybrid Arrays and Their Triply Linked Tapes Exhibiting Strong Absorption Bands in the NIR Region, J. Am. Chem. Soc., 2015, 137, 2097–2106 CrossRef CAS PubMed.
- Y. Rao, J. Lee, J. Chen, L. Xu, M. Zhou, B. Yin, J. Kim, A. Osuka and J. Song, 5,18-Dimesitylorangarin: a Stable Antiaromatic [20]Pentaphyrin(1.0.1.0.0) Displaying Remarkable Oxidative Self-Coupling Reactions, Angew. Chem., Int. Ed., 2024, 63, e202409655 CrossRef CAS PubMed.
- S. Lee, K. Yamashita, N. Sakata, Y. Hirao, K. Ogawa and T. Ogawa, Stable Singlet Biradicals of Rare–Earth–Fused Diporphyrin–Triple–Decker Complexes with Low Energy Gaps and Multi–Redox States, Chem. – Eur. J., 2019, 25, 3240–3243 CrossRef CAS PubMed.
- R. Jasti, J. Bhattacharjee, J. B. Neaton and C. R. Bertozzi, Synthesis, Characterization, and Theory of [9]-, [12]-, and [18]Cycloparaphenylene: Carbon Nanohoop Structures, J. Am. Chem. Soc., 2008, 130, 17646–17647 CrossRef CAS PubMed.
- P. J. Evans, E. R. Darzi and R. Jasti, Efficient room-temperature synthesis of a highly strained carbon nanohoop fragment of buckminsterfullerene, Nat. Chem., 2014, 6, 404–408 CrossRef CAS PubMed.
- E. Kayahara, V. K. Patel and S. Yamago, Synthesis and Characterization of [5]Cycloparaphenylene, J. Am. Chem. Soc., 2014, 136, 2284–2287 CrossRef CAS PubMed.
- E. Kayahara, T. Kouyama, T. Kato, H. Takaya, N. Yasuda and S. Yamago, Isolation and Characterization of the Cycloparaphenylene Radical Cation and Dication, Angew. Chem., Int. Ed., 2013, 52, 13722–13726 CrossRef CAS PubMed.
- M. R. Golder, B. M. Wong and R. Jasti, Photophysical and theoretical investigations of the [8]cycloparaphenylene radical cation and its charge-resonance dimer, Chem. Sci., 2013, 4, 4285–4291 RSC.
- E. Kayahara, T. Kouyama, T. Kato and S. Yamago, Synthesis and Characterization of [n]CPP (n = 5, 6, 8, 10, and 12) Radical Cation and Dications: Size-Dependent Absorption, Spin, and Charge Delocalization, J. Am. Chem. Soc., 2016, 138, 338–344 CrossRef CAS PubMed.
- M. P. Alvarez, M. C. R. Delgado, M. Taravillo, V. G. Baonza, J. T. L. Navarrete, P. Evans, R. Jasti, S. Yamago, M. Kertesz and J. Casado, The Raman fingerprint of cyclic conjugation: the case of the stabilization of cations and dications in cycloparaphenylenes, Chem. Sci., 2016, 7, 3494–3499 RSC.
- N. Toriumi, A. Muranaka, E. Kayahara, S. Yamago and M. Uchiyama, In-Plane Aromaticity in Cycloparaphenylene Dications: A Magnetic Circular Dichroism and Theoretical Study, J. Am. Chem. Soc., 2015, 137, 82–85 CrossRef CAS PubMed.
- Y. Masumoto, N. Toriumi, A. Muranaka, E. Kayahara, S. Yamago and M. Uchiyama, Near-Infrared Fluorescence from In-Plane-Aromatic Cycloparaphenylene Dications, J. Phys. Chem. A, 2018, 122, 5162–5167 CrossRef CAS PubMed.
- S. N. Spisak, Z. Wei, E. Darzi, R. Jasti and M. A. Petrukhina, Highly strained [6]cycloparaphenylene: crystallization of an unsolvated polymorph and the first mono- and dianions, Chem. Commun., 2018, 54, 7818–7821 RSC.
- A. V. Zabula, A. S. Filatov, J. Xia, R. Jasti and M. A. Petrukhina, Tightening of the Nanobelt upon Multielectron Reduction, Angew. Chem., Int. Ed., 2013, 52, 5033–5036 CrossRef CAS PubMed.
- M. Rickhaus, M. Jirasek, L. Tejerina, H. Gotfredsen, M. D. Peeks, R. Haver, H.-W. Jiang, T. D. W. Claridge and H. L. Anderson, Global aromaticity at the nanoscale, Nat. Chem., 2020, 12, 236–241 CrossRef CAS PubMed.
- S. M. Kopp, H. Gotfredsen, J. Hergenhahn, A. Rodríguez-Rubio, J.-R. Deng, H. Zhu, W. Stawski and H. L. Anderson, Charge delocalization and global aromaticity in a partially fused 12-porphyrin nanoring, Chem, 2024, 10, 3410–3427 CAS.
- D. J. Kim, D.-J. Yoo, M. T. Otley, A. Prokofjevs, C. Pezzato, M. Owczarek, S. J. Lee, J. W. Choi and J. F. Stoddart, Rechargeable aluminium organic batteries, Nat. Energy, 2019, 4, 51–59 CrossRef CAS.
- X. Lu, S. Lee, Y. Hong, H. Phan, T. Y. Gopalakrishna, T. S. Herng, T. Tanaka, M. E. Sandoval-Salinas, W. Zeng, J. Ding, D. Casanova, A. Osuka, D. Kim and J. Wu, Fluorenyl Based Macrocyclic Polyradicaloids, J. Am. Chem. Soc., 2017, 139, 13173–13183 CrossRef CAS PubMed.
- Y. Shi, C. Li, J. Di, Y. Xue, Y. Jia, J. Duan, X. Hu, Y. Tian, Y. Li, C. Sun, N. Zhang, Y. Xiong, T. Jin and P. Chen, Polycationic Open-Shell Cyclophanes: Synthesis of Electron-Rich Chiral Macrocycles, and Redox-Dependent Electronic States, Angew. Chem., Int. Ed., 2024, 63, e202402800 CrossRef CAS PubMed.
- Y. Wang, H. Wu and J. F. Stoddart, Molecular Triangles: A New Class of Macrocycles, Acc. Chem. Res., 2021, 54, 2027–2039 CrossRef CAS PubMed.
- D. J. Kim, K. R. Hermann, A. Prokofjevs, M. T. Otley, C. Pezzato, M. Owczarek and J. F. Stoddart, Redox-Active Macrocycles for Organic Rechargeable Batteries, J. Am. Chem. Soc., 2017, 139, 6635–6643 CrossRef CAS PubMed.
- J. Gawroński, M. Brzostowska, K. Gawrońska, J. Koput, U. Rychlewska, P. Skowronek and B. Nordén, Novel Chiral Pyromellitdiimide (1,2,4,5-Benzenetetracarboxydiimide) Dimers and Trimers: Exploring Their Structure, Electronic Transitions, and Exciton Coupling, Chem. – Eur. J., 2002, 8, 2484–2494 CrossRef.
- S. T. Schneebeli, M. Frasconi, Z. Liu, Y. Wu, D. M. Gardner, N. L. Strutt, C. Cheng, R. Carmieli, M. R. Wasielewski and J. F. Stoddart, Electron Sharing and Anion–π Recognition in Molecular Triangular Prisms, Angew. Chem., Int. Ed., 2013, 52, 13100–13104 CrossRef CAS PubMed.
- B. Thulin, O. Wennerström, H.-E. Högberg, O. Smidsrød, O. Dahl, O. Buchardt and G. Schroll, A Simple Synthesis of [2.2.2.2]-Paracyclophane-1,9,17,25-tetraene by a Wittig Reaction, Acta Chem. Scand., 1975, 29b, 138–139 CrossRef.
- K. Ankner, B. Lamm, B. Thulin, O. Wennerström, S. Sprake, J.-E. Berg and A.-M. Pilotti, Electrochemical Reduction of [2(4)]paracyclophanetetraene. A Reversible Two-electron Process., Acta Chem. Scand., 1978, 32b, 155–156 CrossRef.
- E. Ljungström, O. Lindqvist and O. Wennerström, The crystal and molecular structure of [24]paracyclophanetetraene, Acta Crystallogr., Sect. B, 1978, 34, 1889–1893 CrossRef.
- W. Huber, K. Müllen and O. Wennerström, Dianion and Tetraanion of [24] Paracyclophanetetraene—New Unusual Perimeter Systems, Angew. Chem., Int. Ed. Engl., 1980, 19, 624–625 CrossRef.
- K. Muellen, H. Unterberg, W. Huber, O. Wennerstroem, U. Norinder and D. Tanner, Diatropic cyclophane dianions, J. Am. Chem. Soc., 1984, 106, 7514–7522 CrossRef CAS.
- W. Stawski, Y. Zhu, Z. Wei, M. A. Petrukhina and H. L. Anderson, Crystallographic evidence for global aromaticity in the di-anion and tetra-anion of a cyclophane hydrocarbon, Chem. Sci., 2023, 14, 14109–14114 RSC.
- S. Eder, D.-J. Yoo, W. Nogala, M. Pletzer, A. Santana Bonilla, A. J. P. White, K. E. Jelfs, M. Heeney, J. W. Choi and F. Glöcklhofer, Switching between Local and Global Aromaticity in a Conjugated Macrocycle for High-Performance Organic Sodium-Ion Battery Anodes, Angew. Chem., Int. Ed., 2020, 59, 12958–12964 CrossRef CAS PubMed.
- S. Eder, B. Ding, D. B. Thornton, D. Sammut, A. J. P. White, F. Plasser, I. E. L. Stephens, M. Heeney, S. Mezzavilla and F. Glöcklhofer, Squarephaneic Tetraanhydride: A Conjugated Square-Shaped Cyclophane for the Synthesis of Porous Organic Materials, Angew. Chem., Int. Ed., 2022, 61, e202212623 CrossRef CAS PubMed.
- S. I. Hauck, K. V. Lakshmi and J. F. Hartwig, Tetraazacyclophanes by Palladium-Catalyzed Aromatic Amination. Geometrically Defined, Stable, High-Spin Diradicals, Org. Lett., 1999, 1, 2057–2060 CrossRef CAS.
- A. Ito, Y. Ono and K. Tanaka, The Tetraaza[1.1.1.1]m,p,m,p-cyclophane Dication: A Triplet Diradical Having Two m-Phenylenediamine Radical Cations Linked by Twisted Benzenes, Angew. Chem., Int. Ed., 2000, 39, 1072–1075 CrossRef CAS PubMed.
- W. Wang, C. Chen, C. Shu, S. Rajca, X. Wang and A. Rajca, S = 1 Tetraazacyclophane Diradical Dication with Robust Stability: A Case of Low-Temperature One-Dimensional Antiferromagnetic Chain, J. Am. Chem. Soc., 2018, 140, 7820–7826 CrossRef CAS PubMed.
- D. Sakamaki, A. Ito, K. Furukawa, T. Kato and K. Tanaka, High-spin polycationic states of an alternate meta-para-linked oligoarylamine incorporating two macrocycles, Chem. Commun., 2009, 4524–4526 RSC.
- D. Sakamaki, A. Ito, K. Furukawa, T. Kato, M. Shiro and K. Tanaka, A Polymacrocyclic Oligoarylamine with a Pseudobeltane Motif: Towards a Cylindrical Multispin System, Angew. Chem., Int. Ed., 2012, 51, 12776–12781 CrossRef CAS PubMed.
- Y. Yokoyama, D. Sakamaki, A. Ito, K. Tanaka and M. Shiro, A Triphenylamine Double-Decker: From a Delocalized Radical Cation to a Diradical Dication with an Excited Triplet State, Angew. Chem., Int. Ed., 2012, 51, 9403–9406 CrossRef CAS PubMed.
- Y. Ni, T. Y. Gopalakrishna, H. Phan, T. Kim, T. S. Herng, Y. Han, T. Tao, J. Ding, D. Kim and J. Wu, 3D global aromaticity in a fully conjugated diradicaloid cage at different oxidation states, Nat. Chem., 2020, 12, 242–248 CrossRef CAS PubMed.
- M. Stępień and L. Latos-Grażyński, Tetraphenyl-p-benziporphyrin:A Carbaporphyrinoid with Two Linked Carbon Atoms in the Coordination Core, J. Am. Chem. Soc., 2002, 124, 3838–3839 CrossRef PubMed.
- M. Stępień, L. Latos-Grażyński, N. Sprutta, P. Chwalisz and L. Szterenberg, Expanded Porphyrin with a Split Personality: A Hückel–Möbius Aromaticity Switch, Angew. Chem., Int. Ed., 2007, 46, 7869–7873 CrossRef PubMed.
- Z. Li, T. Y. Gopalakrishna, Y. Han, Y. Gu, L. Yuan, W. Zeng, D. Casanova and J. Wu, [6]Cyclo- para -phenylmethine: An Analog of Benzene Showing Global Aromaticity and Open-Shell Diradical Character, J. Am. Chem. Soc., 2019, 141, 16266–16270 CrossRef CAS PubMed.
- Z. Li, X. Hou, Y. Han, W. Fan, Y. Ni, Q. Zhou, J. Zhu, S. Wu, K.-W. Huang and J. Wu, [8]Cyclo-para-phenylmethine as a Super-Cyclooctatetraene: Dynamic Behavior, Global Aromaticity, and Open-Shell Diradical Character in the Neutral and Dicationic States, Angew. Chem., Int. Ed., 2022, 61, e202210697 CrossRef CAS PubMed.
- S. Wu, Y. Han, Y. Ni, X. Hou, H. Wei, Z. Li and J. Wu, Unveiling Möbius/Hückel Topology and Aromaticity in A Core-Expanded [10]Annulene at Different Oxidation States, Angew. Chem., Int. Ed., 2024, 63, e202320144 CrossRef CAS PubMed.
- S. Dong, T. Y. Gopalakrishna, Y. Han and C. Chi, Cyclobis(7,8−(para –quinodimethane)–4,4′–triphenylamine) and Its Cationic Species Showing Annulene–Like Global (Anti)Aromaticity, Angew. Chem., Int. Ed., 2019, 58, 11742–11746 CrossRef CAS PubMed.
- Y. Ma, Y. Han, X. Hou, S. Wu and C. Chi, Facile Synthesis and Global Aromaticity of Aza-Superbenzene and Aza-Supernaphthalene at Different Oxidation States, Angew. Chem., Int. Ed., 2024, 63, e202407990 CrossRef CAS PubMed.
- Y. Ni, M. E. Sandoval-Salinas, T. Tanaka, H. Phan, T. S. Herng, T. Y. Gopalakrishna, J. Ding, A. Osuka, D. Casanova and J. Wu, [n]Cyclo-para-biphenylmethine Polyradicaloids: [n]Annulene Analogs and Unusual Valence Tautomerization, Chem, 2019, 5, 108–121 CAS.
- J. C. Buttrick and B. T. King, Kekulenes, cycloarenes, and heterocycloarenes: addressing electronic structure and aromaticity through experiments and calculations, Chem. Soc. Rev., 2017, 46, 7–20 RSC.
- M. Stępień, An Aromatic Riddle: Decoupling Annulene Conjugation in Coronoid Macrocycles, Chem, 2018, 4, 1481–1483 Search PubMed.
- F. Diederich and H. A. Staab, Benzenoid versus Annulenoid Aromaticity: Synthesis and Properties of Kekulene, Angew. Chem., Int. Ed. Engl., 1978, 17, 372–374 CrossRef.
- M. A. Majewski, Y. Hong, T. Lis, J. Gregoliński, P. J. Chmielewski, J. Cybińska, D. Kim and M. Stępień, Octulene: A Hyperbolic Molecular Belt that Binds Chloride Anions, Angew. Chem., Int. Ed., 2016, 55, 14072–14076 CrossRef CAS PubMed.
- W. Fan, Y. Han, S. Dong, G. Li, X. Lu and J. Wu, Facile Synthesis of Aryl-Substituted Cycloarenes via Bismuth(III) Triflate-Catalyzed Cyclization of Vinyl Ethers, CCS Chem., 2021, 3, 1445–1452 CrossRef CAS.
- X. Lu, T. Y. Gopalakrishna, H. Phan, T. S. Herng, Q. Jiang, C. Liu, G. Li, J. Ding and J. Wu, Global Aromaticity in Macrocyclic Cyclopenta-Fused Tetraphenanthrenylene Tetraradicaloid and Its Charged Species, Angew. Chem., Int. Ed., 2018, 57, 13052–13056 CrossRef CAS PubMed.
- H. Gregolińska, M. Majewski, P. J. Chmielewski, J. Gregoliński, A. Chien, J. Zhou, Y.-L. Wu, Y. J. Bae, M. R. Wasielewski, P. M. Zimmerman and M. Stępień, Fully Conjugated [4]Chrysaorene. Redox-Coupled Anion Binding in a Tetraradicaloid Macrocycle, J. Am. Chem. Soc., 2018, 140, 14474–14480 CrossRef PubMed.
- S. Das, T. S. Herng, J. L. Zafra, P. M. Burrezo, M. Kitano, M. Ishida, T. Y. Gopalakrishna, P. Hu, A. Osuka, J. Casado, J. Ding, D. Casanova and J. Wu, Fully Fused Quinoidal/Aromatic Carbazole Macrocycles with Poly-radical Characters, J. Am. Chem. Soc., 2016, 138, 7782–7790 CrossRef CAS PubMed.
- C. Liu, M. E. Sandoval-Salinas, Y. Hong, T. Y. Gopalakrishna, H. Phan, N. Aratani, T. S. Herng, J. Ding, H. Yamada, D. Kim, D. Casanova and J. Wu, Macrocyclic Polyradicaloids with Unusual Super-ring Structure and Global Aromaticity, Chem, 2018, 4, 1586–1595 CAS.
- B. Prajapati, D.-K. Dang, P. J. Chmielewski, M. A. Majewski, T. Lis, C. J. Gómez-García, P. M. Zimmerman and M. Stępień, An Open-Shell Coronoid with Hybrid Chichibabin–Schlenk Conjugation, Angew. Chem., Int. Ed., 2021, 60, 22496–22504 CrossRef CAS PubMed.
- N. Zhang, W. Li, J. Zhu, T. Wang, R. Zhang, K. Chi, Y. Liu, Y. Zhao and X. Lu, Periphery Fusion Strategy of a Carbazole-Based Macrocycle toward Coplanar N-Heterocycloarene for High-Mobility Single-Crystal Transistors, Adv. Mater., 2023, 35, 2300094 CrossRef CAS PubMed.
- D. Myśliwiec and M. Stępień, The Fold-In Approach to Bowl-Shaped Aromatic Compounds: Synthesis of Chrysaoroles, Angew. Chem., Int. Ed., 2013, 52, 1713–1717 CrossRef PubMed.
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.