Helicenes are ortho-fused polycyclic aromatic or heteroaromatic compounds with at least five rings arranged to form helically shaped, chiral molecules.[1] These non-planar structures arise from the angular annulation of benzene or other aromatic rings, conferring inherent axial chirality. For helicenes with seven or more rings, the steric hindrance results in high barriers to racemization without bond breakage.[2] The first carbohelicene, [3]helicene, was synthesized in 1918 via a Pschorr reaction, marking the beginning of helicene chemistry, though earlier reports of related structures date to 1903.[4]Helicenes exhibit extended π-conjugation along their helical backbone, leading to distinctive electronic and optical properties, including high fluorescence quantum yields, large Stokes shifts, and intense chiroptical responses such as circular dichroism and circularly polarized luminescence. Their chirality, combined with tunable substituents, enables applications in asymmetric catalysis as chiral ligands or auxiliaries, where they enhance enantioselectivity in reactions like hydrogenation and aldol additions.[5] In materials science, helicenes serve as components in organic electronics, including organic light-emitting diodes (OLEDs) and photovoltaic devices, due to their ability to form self-assembled nanostructures and exhibit charge transport properties.[3]Synthesis of helicenes has evolved from classical methods like photocyclization and oxidative coupling to modern catalytic approaches, including transition-metal-catalyzed cross-couplings and enantioselective strategies using chiral auxiliaries or organocatalysts, allowing access to enantioenriched forms with high efficiency.[7] Recent advances focus on heteroatom-doped and extended helicenes for enhanced functionality in sensing, molecular machines, and biomedical imaging, where their helical scaffolds mimic biomolecular motifs.[8] Despite challenges in scalability and stability for larger variants, ongoing research underscores helicenes' versatility across organic synthesis, supramolecular chemistry, and nanotechnology.
History
Discovery
The discovery of helicenes traces back to the early 20th century with the synthesis of the first azahelicenes in 1903 by Meisenheimer and Witte during the reduction of 2-nitronaphthalene, marking the initial identification of compounds with helical topology, though their chirality was not yet recognized.[9] The first carbohelicene, [3]helicene (dibenzo[c,g]phenanthrene), was synthesized in 1918 by Weitzenböck and Klingler via the Pschorr reaction, providing the foundational example of an all-carbon ortho-fused polycyclic aromatic system, albeit without immediate appreciation of its non-planar structure.[9][10]Interest in helicenes surged in the 1950s, when Melvin S. Newman and collaborators advanced the field significantly. In 1955, Newman coined the term "helicene" to denote ortho-condensed polycyclic aromatic hydrocarbons exhibiting helical annulation due to steric overcrowding, establishing a clear nomenclature for this class of compounds.[9] The following year, Newman and Lednicer reported the first synthesis of [11]helicene (hexahelicene) using a Hauser–Kraus-type annulation to close the central rings, and successfully resolved it into enantiomers via complexation with 2-(2,4,5,7-tetranitrofluorenylideneaminooxy)propionic acid, confirming its inherent axial chirality.[12] Concurrently, X-ray crystallographic analysis by Newman et al. on [3]helicene revealed its distorted, non-planar helical conformation, with terminal benzene rings twisted by approximately 45° due to steric repulsion, providing definitive proof of the helical topology and its role in chirality.[9]Early investigations in the 1950s and 1960s further elucidated the stereochemical behavior of smaller helicenes. For [3]helicene, Newman observed atropisomerism arising from restricted rotation around the crowded ortho-fused bonds, but with a relatively low racemization barrier of approximately 24 kcal/mol, enabling rapid interconversion of enantiomers at ambient temperatures and rendering it configurationally labile.[9] In contrast, [11]helicene displayed a higher barrier exceeding 40 kcal/mol, ensuring configurational stability and highlighting how increasing ring number enhances the persistence of helical chirality in this series.[9] These foundational studies laid the groundwork for understanding helicenes as atropisomeric scaffolds with tunable stereochemical properties.
Key Developments
In the 1970s, theoretical efforts advanced the understanding of helicene geometry, with Charles A. Coulson and colleagues employing early quantum chemical methods to model the strain energy arising from the non-planar arrangement of fused rings and to predict the helical pitch in extended carbohelicenes, highlighting how increasing ring number amplifies torsional strain while maintaining π-conjugation.[13] These models laid the groundwork for quantifying the energetic barriers to racemization and the conformational stability of higher-order helicenes.The 1980s marked a pivotal shift toward stereocontrol, exemplified by advancements in asymmetric synthesis that enabled the preparation of enantiopure helicenes for the first time. Notably, Robert H. Martin reported the synthesis of enantiopure [11]helicene in 1981 through diastereoselective photocyclization of chiral precursors, achieving high optical purity and demonstrating the feasibility of controlling helical chirality at the molecular level without relying on resolution techniques.[14] This breakthrough spurred further developments in organometallic and catalytic methods, facilitating access to configurationally stable higher helicenes for chiroptical studies.During the 1990s, research expanded to heterohelicenes, incorporating heteroatoms such as nitrogen and oxygen to modulate electronic properties and enhance solubility or reactivity. Pioneering work introduced nitrogen-doped azahelicenes and oxygen-containing oxahelicenes, which exhibited tuned aromaticity and improved ligand capabilities compared to all-carbon analogs, as demonstrated in early syntheses via modified Mallory photocyclizations.[15] These variants opened avenues for applications in coordination chemistry, where the heteroatoms influenced metal binding and stereoselectivity.The 2000s witnessed the integration of helicenes into supramolecular chemistry, leveraging their axial chirality for dynamic systems like rotors and switches. Seminal contributions included the design of helicene-tethered triptycene rotors by the Kelly group, where unidirectional rotation was induced by chemical stimuli, mimicking biological motors with barriers exceeding 20 kcal/mol.[16] Similarly, light-driven helicene-based chiroptical switches emerged, enabling reversible helical inversion and applications in responsive materials.Recent progress up to 2025 has focused on multi-helicene architectures, combining multiple helical units to amplify chiroptical responses and device performance. In 2023, reports on double helicenes with B-N covalent linkages showcased enhanced stability in organic light-emitting diodes (OLEDs), achieving narrowband red emission with external quantum efficiencies over 20% and operational lifetimes improved by factors of 2-3 due to reduced aggregation and conformational rigidity. In 2025, organocatalytic [4+2] cycloadditions enabled the enantioselective synthesis of double S-shaped quadruple helicene-like molecules with up to 96% ee.[17][18] These structures represent a high-impact evolution toward durable chiral emitters in optoelectronics.
Structure and Classification
Molecular Topology
Helicenes consist of ortho-angularly annulated benzene rings that form a polycyclic aromatic framework. The term carbohelicene refers to a molecule with n such rings fused in an angular fashion, resulting in a helical topology for n ≥ 6 due to steric repulsion between the terminal benzene rings, which prevents planarity. [3]helicene exhibits a distorted, non-planar structure with a dihedral angle of about 27° between terminal rings and undergoes facile racemization, whereas [11]helicene and higher form stable helical conformations due to increased steric hindrance.[19][20]The helical geometry is quantified by parameters such as pitch (the axial advance per turn) and diameter (the radial extent of the helix). In [11]helicene, the pitch is approximately 3.5 Å, representing the distance along the helical axis for one full turn, while the diameter is about 6 Å, corresponding to the effective width of the coiled structure. These values reflect the compact, screw-like arrangement where the terminal rings overlap significantly, with the dihedral angles between adjacent rings typically around 20–30° to accommodate the twist. As n increases, the helix elongates with more turns, maintaining a roughly constant pitch but expanding the overall length and curvature. Schematic representations of helicene show a progression from the distorted fusion in [3]helicene to progressively coiled 3D models in [11]helicene and higher, where the ends "bite" together like jaws, emphasizing the transition to full helicity.[21][19][22]This non-planarity introduces strain energy, estimated at approximately 30–40 kcal/mol for [11]- and [23]helicenes, arising primarily from the deformation of the inherently planar benzene rings and the compression in the bay regions. The strain energy increases with n, as additional rings amplify the cumulative distortion needed to avoid greater steric clashes. Computational models, such as autodesmotic reactions, quantify this by comparing the energy of the helical structure to hypothetical planar references, highlighting how the energy penalty scales with helical extension.[24][25]The helical twist also distorts the π-orbitals, reducing their lateral overlap integral compared to planar polyaromatics, which contributes to the overall non-planarity by favoring twisted conformations that balance steric relief against electronic conjugation losses. This overlap integral, often calculated via quantum mechanical methods like DFT, quantifies the diminished π-system delocalization, with values decreasing as the torsion angles increase, thereby influencing properties like bandgap and conductivity. In [11]helicene, this distortion manifests as a subtle misalignment of p-orbitals perpendicular to the mean plane, essential for the molecule's inherent chirality without central stereocenters.[26][19]
Types of Helicenes
Helicenes are broadly classified into carbohelicenes, heterohelicenes, charged and metallohelicenes, extended architectures, and polyaromatic analogs like sulflower, each exhibiting distinct structural features that influence their properties and applications.[27][3]Carbohelicenes consist exclusively of ortho-fused benzene rings forming a pure hydrocarbon framework, denoted as carbohelicene where n represents the number of rings.[27] The [3]carbohelicene adopts a distorted conformation due to steric hindrance, while [11]carbohelicene and higher homologs (n ≥ 6) distort into stable helical shapes driven by overcrowding at the termini.[28] These all-carbon structures exhibit extended π-conjugation along the helical axis, contributing to their inherent chirality and electronic delocalization without heteroatom incorporation.[29]Heterohelicenes incorporate one or more heteroatoms such as nitrogen, oxygen, or sulfur into the ortho-fused ring system, replacing carbon atoms in the backbone to modulate electronic and steric properties. Azahelicenes feature nitrogen substitution, as in aza[11]helicene where nitrogen replaces a CH group at position 10, altering the electron density and potentially enhancing solubility or reactivity compared to carbohelicenes.[30] Oxahelicenes contain oxygen atoms, often in furan-like units, while thiahelicenes incorporate sulfur, typically from thiophene moieties, leading to variations in aromaticity and intermolecular interactions.[3] Recent developments include BN-doped helicenes, such as bora[11]helicenes, which incorporate boron and nitrogen for tuned electronic properties and configurational stability (as of 2025).[25] These heteroatom inclusions allow for tunable chiroptical responses while maintaining the helical topology.Charged helicenes introduce ionic functionality to improve solubility and enable interactions with polar environments, with cationic derivatives being prominent examples.[31] Cationic [11]helicene derivatives, such as those based on triarylcarbenium scaffolds, exhibit enhanced aqueous solubility due to the positive charge, facilitating applications in biological media without compromising the helical chirality.[32] Metallohelicenes integrate metal centers, often through coordination or organometallic bonds, to form multihelicenic architectures that combine helical scaffolds with transition metal properties for catalytic or luminescent purposes. These charged and metallo variants expand the functional diversity of helicenes beyond neutral hydrocarbons.[33]Extended helicenes feature multiple helical units or macrocyclic assemblies, amplifying the topological complexity and conjugation length.[34] Double and triple helicenes involve two or three fused helical segments, such as consecutively fused expanded [23]helicene subunits, resulting in larger, more contorted structures with enhanced steric strain.[7] Carbohelicene macrocycles form cyclic arrays of helical motifs, with recent 2025 developments demonstrating geometry-tunable variants through selective ring fusions that allow dynamic adjustment of cavity size and shape.[3] These extended forms exhibit superior mechanical stability and potential for host-guest chemistry.[35]Sulflower is a polyaromatic circulene, a stable octacirculene composed entirely of thiophene units, forming a flower-like, sulfur-rich heterocycle with C_{16}S_8 stoichiometry and no hydrogen atoms. This structure mimics the extended π-system of higher polyaromatics but adopts a planar, radially symmetric conformation, serving as a model for sulfur-doped carbon sulfides with potential in organic electronics.[36] Other analogs, such as oxiflowers, extend this concept by varying heteroatom placement, bridging the gap between circulenes and helical polyaromatics.[37]
Properties
Physical and Chemical Properties
Helicenes, particularly unsubstituted carbohelicenes, exhibit low solubility in non-polar solvents such as hexane due to their rigid, extended π-conjugated structures. Solubility is significantly improved by introducing alkyl or heteroatom substituents, which disrupt planarity and enhance interactions with solvents like chloroform or toluene.[38] For instance, boron-doped [39]helicenes display excellent solubility in hexane, highlighting the role of heteroatoms in modulating this property.[38]Carbohelicenes demonstrate high thermal stability, with decomposition temperatures often exceeding 400°C, as seen in double hetero[3]helicenes with a Td of 414°C.[40] Racemization barriers for [11]- and [23]helicenes are approximately 35–42 kcal/mol, reflecting their resistance to thermal inversion while maintaining helical integrity up to elevated temperatures.[41] These barriers arise from the energy required to flatten the helical backbone, a process that is concerted for shorter helicenes like [11] and [23].[42]In terms of reactivity, helicenes are electron-rich and undergo electrophilic aromatic substitution preferentially at the terminal rings, where electron density is highest, as observed in nitration and Vilsmeier-Haack reactions of [11]helicene derivatives.[43] Due to inherent strain in the ortho-annulated framework, they resist electrophilic addition reactions that would disrupt aromaticity.[43]In the solid state, helicenes often form helical dimers through π-π interactions between overlapping aromatic rings, leading to columnar packing motifs that enhance charge transport potential.[44] For [11]helicene, the melting point is 231–233°C, and the density is around 1.3 g/cm³, contributing to their robust crystalline structures.[45] These packing features are more pronounced in mid-length helicenes ([3]–[23]), where interlocking pairs stabilize the lattice via edge-to-face and face-to-face π-π contacts.[44]
Optical and Electronic Properties
Helicenes display characteristic UV-Vis absorption spectra arising from their extended π-conjugation, with absorption bands typically in the UV to near-UV region. The absorption maxima exhibit a bathochromic shift as the number of fused rings increases, reflecting the lengthening of the effective conjugation length despite the helical distortion. For instance, [11]helicene shows a λ_max of approximately 340 nm, while [23]helicene displays a red-shifted λ_max around 360 nm.[46] This trend is supported by both experimental measurements and density functional theory (DFT) calculations, where the lowest-energy transitions correspond to HOMO-LUMO excitations.[46]The fluorescence properties of helicenes are notable for their moderate quantum yields, generally ranging from 0.1 to 0.5 in solution, depending on the ring size and substituents, with emission often occurring in the blue to green region. Enantiopure helicenes exhibit circularly polarized luminescence (CPL), a key chiroptical feature stemming from their inherent helical chirality, with luminescence dissymmetry factors (g_lum) reaching up to 0.01 at the emission maximum. These values are among the higher reported for organic helicenes and highlight their potential for chiral optoelectronic applications, as documented in comprehensive reviews of chiroptical data across various helicene classes.[47]Electronic properties of helicenes are characterized by HOMO-LUMO energy gaps of 3.5–4.0 eV, determined through DFT computations such as B3LYP or CAM-B3LYP functionals, which reveal that the helical geometry reduces effective conjugation compared to planar polycyclic aromatic hydrocarbons (PAHs) like phenacenes. This distortion leads to wider bandgaps than expected for linear acenes of similar size. Redox behavior shows reversible oxidation potentials in the range of +1.0 to +1.5 V vs. saturated calomel electrode (SCE), typically accessed via cyclic voltammetry in aprotic solvents, with the exact value modulated by peripheral substituents or incorporation of heteroatoms like nitrogen, which can lower the oxidation potential due to increased electron density.[48][49]
Synthesis
Traditional Methods
The traditional synthesis of helicenes relied on non-stereoselective strategies that built the helical polycyclic aromatic framework through cyclization and dehydrogenation of linear or angular polyarene precursors, often requiring harsh conditions and multi-step sequences. These methods, developed largely in the early to mid-20th century, enabled the preparation of small carbohelicenes but were limited by modest overall yields and the inability to control helical chirality, producing only racemic mixtures.[9]A cornerstone of early helicene synthesis is the photocyclization of diaryl olefins, exemplified by the Mallory reaction introduced in the 1960s. This involves irradiating stilbene derivatives or related biaryl alkenes with ultraviolet light in the presence of an iodine or propionic acid catalyst to induce trans-to-cis isomerization, followed by electrocyclization and oxidative dehydrogenation to form the central ring of the helicene scaffold. For instance, [11]helicene can be obtained from an appropriate stilbene precursor in yields of 20–50%, making this route particularly suitable for [3]–[23]helicenes due to its simplicity and tolerance of functional groups on the aromatic rings.[50][9]Thermal cyclodehydrogenation provided another foundational pathway, typically involving the high-temperature pyrolysis (200–300 °C) of polyarylated acyclic or partially saturated precursors to drive C–H bond cleavage and ring fusion. This approach was commonly applied to synthesize [3]–[23]helicenes from diphenylnaphthalene or tetraphenylbenzene derivatives, where dehydrogenation occurs under inert atmospheres or in molten salts, yielding the fully aromatic helical structure after workup. Representative examples include the conversion of 1,1'-binaphthyl-based precursors to [3]helicene in moderate yields, though optimization via flash vacuum pyrolysis improved efficiency for smaller systems.[9]Intramolecular Diels–Alder reactions offered a convergent alternative for angular annulation, where a tethered diene and dienophile undergo cycloaddition to form a bridged intermediate, followed by retro-Diels–Alder elimination or dehydrogenation to aromatize the rings and establish the helicene topology. This method was notably used in pre-1990s syntheses for [11] and [23]helicenes, leveraging furan or pyrone dienophiles for clean extrusion of small molecules like CO or ethylene during aromatization, with overall yields often in the 10–30% range for the key steps.[9]Early landmark syntheses underscored the challenges of these routes; for example, the 1918 work by Weitzenböck and Klingler produced [3]helicene in ca. 5% overall yield via a Pschorr reaction involving diazotization and cyclization, highlighting the inefficiencies of pre-photochemical methods.[4] Despite their pioneering role, traditional approaches suffered from inherent limitations, including negligible stereoselectivity that yielded only racemates, difficulties in scaling beyond gram quantities due to low solubility and side reactions, and impracticality for larger helicenes (n > 7) owing to escalating steric strain and reduced reactivity in cyclization steps.[51]
Modern Synthetic Approaches
Modern synthetic approaches to helicenes have evolved significantly since the 1990s, leveraging transition metalcatalysis to enhance efficiency, stereocontrol, and scalability over earlier methods. These strategies often involve the construction of linear polyaryl precursors followed by cyclization, enabling the synthesis of larger and more functionalized helicenes with improved yields and selectivity.[51]Metal-catalyzed aryl-aryl couplings, particularly using nickel or palladium catalysts, have become cornerstone techniques for assembling the extended aromatic frameworks required for helicene precursors. For instance, palladium-catalyzed double C-H arylation of Z,Z-bis(bromostilbene)s enables the formation of [3]- and [11]helicenes in moderate to good yields (up to 70%), bypassing the need for preactivated halides and reducing synthetic steps. This is typically followed by oxidative cyclization to enforce the helical topology, with Pd(II) salts like Pd(OAc)₂ facilitating the key dehydrogenative aromatization under aerobic conditions. Nickel variants, such as Ni(cod)₂ with chiral ligands like (R)-QUINAP, extend this to enantioselective variants, achieving dibenzohelicenes with good enantioselectivities (up to 80% ee), though yields remain moderate (around 50%). These methods highlight the versatility of cross-coupling in accessing diverse substitution patterns while minimizing waste compared to traditional photocyclization.[52][51]Enantioselective syntheses have advanced through the integration of chiral auxiliaries and catalysts, particularly in the 2010s, allowing direct access to enantioenriched helicenes without post-synthesis resolution. BINOL-derived ligands, such as cationic phosphonites in gold(I) catalysis, enable highly selective alkyne hydroarylation cascades for [11]helicene derivatives, delivering products with >90% ee and yields exceeding 80% in optimized cases. These approaches exploit axial chirality transfer from the ligand to the emerging helical scaffold, often via intramolecular cyclizations of biaryl-alkyne substrates. For example, Rh(I) complexes with (S)-Segphos ligands achieve 1,1’-bitriphenylene-based helicenoids with up to 93% ee, demonstrating broad applicability to carbo- and hetero-variants. Such methods prioritize stereocontrol, with enantiomeric excesses routinely surpassing 90% for [11]helicene scaffolds.[53][51]On-surface synthesis represents a cutting-edge paradigm for constructing extended helicenes, utilizing scanning tunneling microscopy (STM) to guide polymerization on metal substrates like Au(111). In the 2020s, deposition of racemic heptahelicene monomers followed by tip-induced covalent linking yields heterochiral one-dimensional helicene oligomers, with lengths up to several nanometers and precise control over helical handedness alternation. This bottom-up approach, often under ultrahigh vacuum, enables the creation of unprecedented architectures unattainable in solution, such as aligned helical chains for potential molecular electronics. Yields are inherently low due to the surface-mediated nature, but the atomic precision offers unique insights into helicene assembly dynamics.The construction of multi-helicene systems, featuring multiple interlocked or fused helical units, has benefited from cascade reactions in recent years. Rhodium-catalyzed [2+2+2] cycloadditions of triynes, as developed in the late 2010s and refined into the 2020s, produce double [3]helicenes with yields around 45% and ee values up to 73%, leveraging alkyne trimerization to forge multiple rings in a single step. A 2025 review highlights extensions to quadruple S-shaped helicenoids via organocatalytic [4+2] cycloadditions, achieving up to 96% ee, though overall yields hover at 40-50% due to the complexity of the scaffolds. These cascades emphasize efficiency in building stereodefined multi-helices for advanced materials.[54]Heterohelicene routes, incorporating nitrogen or other heteroatoms, frequently employ directed ortho-metalation (DOM) for precise regiocontrol in aza-variants. Lithium-directed ortho-lithiation of carbazole derivatives, followed by nucleophilic substitution and intramolecular C-H borylation, yields B,N-doped double [23]azahelicenes with up to 70% efficiency over multi-step sequences. This method exploits the directing ability of nitrogen for selective C-H activation, enabling incorporation of up to four heteroatoms while maintaining helical integrity. Yields typically range from 50-70%, with the approach proving scalable for PDI-fused heterohelicenes used in optoelectronics.
Stereochemistry
Chirality and Configuration
Helicenes exhibit chirality arising from their non-superimposable mirror-image helical structures, a form of atropisomerism due to restricted rotation along the ortho-fused aromatic rings that prevents planarization. This axial chirality results in two enantiomers designated as P (plus, right-handed helix) or M (minus, left-handed helix) according to the helicity rule proposed by Cahn, Ingold, and Prelog, where the configuration is determined by viewing the helix along its axis and assigning priority based on the direction of twist.[55]The barrier to helicity inversion, which interconverts P and M enantiomers via thermal racemization, depends on the number of fused rings (n) and structural strain. For [11]helicene, the free energy of activation (ΔG‡) is approximately 36.2 kcal/mol, corresponding to a half-life of 48 minutes at 205°C. This barrier generally increases with n for carbohelicenes up to n ≈ 9 due to enhanced steric interactions between terminal rings in the transition state, though very large or expanded helicenes may exhibit lower barriers owing to increased flexibility.[56][57]Chiroptical properties of helicenes are characterized by distinct Cotton effects in circular dichroism (CD) spectra, reflecting their helical handedness. The P enantiomer of [11]helicene displays a positive Cotton effect in the 300–400 nm region, associated with the π–π* transitions of the extended conjugated system, while the M enantiomer shows the opposite negative signal. These effects are intensified in multiple helicenes due to cumulative excitonic coupling between component units.[55][58]The helical sense in helicenes can be predicted and induced from point chirality in synthetic precursors by adapting Cahn-Ingold-Prelog priority rules to the helical axis, where the chiral center's configuration dictates the preferred twist direction through steric or electronic interactions during cyclization. For instance, an (R)-configured auxiliary often leads to a P-helicene via directed folding that minimizes steric clashes.[55]Racemization in helicenes follows first-orderkinetics, governed by the unimolecular inversion through a planar transition state. For [11]helicene, with its high ΔG‡, the half-life at room temperature is extremely long, ensuring configurational stability under ambient conditions and enabling isolation of enantiopure forms.[56][59]
Enantiomer Resolution and Stability
Enantiopure helicenes are typically obtained through chromatographic resolution or asymmetric synthesis, as these methods allow for the isolation and direct preparation of single helical enantiomers, denoted as P or M configurations. High-performance liquid chromatography (HPLC) using chiral stationary phases (CSPs), such as cellulose derivatives like Chiralcel OD or Chiralpak IA, has been a cornerstone for resolving racemic mixtures of helicenes. For instance, [11]helicene and its derivatives achieve enantiomeric excesses (ee) exceeding 99% via preparative CSP-HPLC, enabling scalable isolation of pure enantiomers with baseline separation under optimized conditions like hexane/isopropanol mobile phases.[60][61]Asymmetric synthesis provides an alternative route to enantiopure helicenes by leveraging chiral catalysts to induce helical chirality during construction. In the 2020s, rhodium(III)-catalyzed C-H activation/annulation reactions using chiral Cp*Rh(III) complexes and chiral phosphoric acids have enabled the enantioselective formation of azonia[11]helicenes from isoquinoline derivatives and alkynes, yielding up to 96% ee and near-quantitative yields.[62] These methods highlight the efficiency of transition-metal catalysis in controlling helical handedness without post-synthetic resolution.The configurational stability of enantiopure helicenes is profoundly influenced by substituents, which modulate the racemization barrier (ΔG‡) through steric and electronic effects. Alkyl groups positioned at the inner helix, such as methyl or tert-butyl at C1/C16 of [11]helicene, elevate ΔG‡ by approximately 5 kcal/mol compared to unsubstituted analogs, enhancing resistance to thermal inversion by increasing steric repulsion in the transition state.[57] This substituent tuning is crucial for applications requiring persistent chirality.Functionalized helicenes exhibit dynamic helical switching, where external stimuli like light or heat trigger reversible inversion between P and M forms. In heterohelicenes bearing photochromic dithienylethene units, blue light irradiation induces ring closure, extending the helical pitch, while red light or mild heat (100°C) reverses the process via ring opening; such systems demonstrate rapid switching reversible by mild heating to 100 °C.[63]Enantiopure helicenes maintain their configuration indefinitely under standard storage at -20°C in inert atmospheres, showing no detectable racemization over extended periods. Thermal racemization becomes appreciable only above 150°C, where diaza[11]helicene derivatives exhibit half-lives of hours or longer, depending on substituents, allowing safe handling at elevated temperatures for short durations.[64]
Applications
In Materials Science
Helicenes play a significant role in liquid crystal materials, particularly through their ability to induce chirality in nematic hosts. Enantiopure [11]helicene derivatives, when doped into cyanobiphenyl nematic liquid crystals such as E7, promote chiral nematic (cholesteric) phases via helical twisting power (HTP) values up to +68 μm⁻¹, enabling the formation of ordered helical structures that amplify molecular chirality at the mesoscale.[65]Supramolecular assemblies involving helicenes often feature helicate structures formed through coordination with transition metals, exemplifying their utility in creating chiral architectures for sensing applications. For instance, Cu(I) complexes with helicene-capped pentadentate phosphole-pyridine ligands self-assemble into configurationally stable double-stranded helicates, where the helical chirality influences ligand-ligand charge transfer and chiroptical properties, enabling selective detection in sensor devices.[66] These metal-helicene helicates exhibit enhanced stereochemical control, with the intrinsic helicity of the ligands propagating to the supramolecular level for applications in enantioselective recognition.Recent advances in on-surface chemistry have highlighted helicenes in the formation of self-assembled monolayers on highly oriented pyrolytic graphite (HOPG), where enantiopure derivatives form ordered structures via van der Waals interactions, enabling chiral patterns on surfaces.[67] Such assemblies leverage the helicene's axial chirality for potential applications in chiral separation and sensing on carbon substrates.Regarding mechanical properties, incorporation of helicene into polymeric matrices improves solubility and processability, leading to robust films. For example, a helicene-based semiconducting polymer exhibits enhanced solubility (>10 mg mL⁻¹ in chlorobenzene) due to the twisted backbone reducing π-π stacking, resulting in smooth, uniform films with high elastic modulus (4.6 GPa) and fracture strength (0.2 GPa).[68] This mitigates aggregation in conjugated polymers, supporting applications in solar cells.
In Optoelectronics and Catalysis
Helicenes have emerged as promising emitters in organic light-emitting diodes (OLEDs) due to their non-planar helical architecture, which suppresses aggregation-caused quenching and enhances device stability. In 2024, B,N-embedded hetero[69]helicene (BN[69]H) derivatives were employed as dopants in OLEDs, achieving a maximum external quantum efficiency (EQE) of 35.5% with a narrow emission bandwidth of 48 nm at 578 nm.[70] This high performance stems from the rigid helical scaffold, which promotes efficient exciton utilization and reduces intermolecular π-π interactions that lead to quenching in planar polycyclic aromatic hydrocarbons.Enantiopure helicenes enable circularly polarized OLEDs (CP-OLEDs) by generating polarized electroluminescence through their intrinsic chirality. Devices incorporating (M)- and (P)-BN[69]H as emitters exhibited intense circularly polarized emission with an electroluminescence dissymmetry factor (|g<sub>EL</sub>|) of 6.2 × 10<sup>-3</sup>, among the highest for helicene-based systems, alongside photoluminescence dissymmetry factors (|g<sub>lum</sub>|) up to 5.8 × 10<sup>-3</sup>.[70] These values highlight the role of the helical twist in amplifying chiroptical responses, facilitating applications in 3D displays and optical encryption.In sensing applications, chiral helicenes serve as hosts for enantioselective recognition of biomolecules via fluorescence modulation. The diol-functionalized hexahelicene HELIXOL acts as a fluorescent sensor for α-amino acid derivatives and amino alcohols, where binding induces enantioselective quenching of its emission due to hydrogen-bonding interactions between the helicene's phenolic groups and the guest's amino functionalities.[71] For instance, (P)-HELIXOL shows differential quenching efficiencies for D- and L-enantiomers, enabling naked-eye detection and quantitative chiral analysis with high selectivity over other amino acids.[72]Helicene-based ligands have advanced asymmetric catalysis, particularly in rhodium-mediated reactions, by providing a sterically demanding chiral environment analogous to DuPhos phosphines. In the 2000s, enantiopure phosphine-substituted helicenes, such as (M,M,S,i)-L4, were developed as ligands for Rh-catalyzed hydrogenation of dimethyl itaconate, delivering the product in 96% enantiomeric excess (ee) under 90 atm H<sub>2</sub> at room temperature.[73] These ligands induce high enantioselectivity through their helical backbone, which enforces a matched stereochemical configuration with the metal center, outperforming flexible phosphine analogs in substrate scope for α,β-unsaturated esters.[73]Photoresponsive helicenes function as molecular switches by undergoing reversible conformational or structural changes under light irradiation, enabling multi-state systems for logic operations. Helicene-chromene hybrids, such as those featuring a [3]helicene linked to a photochromic chromene unit, exhibit switchable chiroptical properties via UV/visible light-induced ring closure/opening, operating as INHIBIT logic gates for optical data storage.[74] Advances in 2025 have extended these to dithienylethene-helicene macrocycles, achieving reversible photoswitching with preserved helicity and amplified chirality for multi-state systems.[75]