Azulene is a non-benzenoid aromatic hydrocarbon with the molecular formula C₁₀H₈, featuring a fused cyclopentadiene and cycloheptatriene ring system that forms a bicyclic [5.3.0] structure with 10 π electrons delocalized across the rings.[1][2] This compound is notable for its vivid blue color, arising from intramolecular charge transfer that results in a significant electric dipole moment of approximately 1.08 D, contrasting sharply with its colorless isomer naphthalene, which has zero dipole moment.[3][2]Azulene was first isolated in 1863 by British chemist Septimus Piesse from the azure-blue distillates of essential oils derived from plants such as yarrow (Achillea millefolium) and wormwood (Artemisia absinthium), though its structure was not elucidated until later.[4] The correct bicyclic structure was established in 1926 by Lavoslav Ružička through analysis of guaiazulene, a naturally occurring derivative from guaiac wood oil, and confirmed via X-ray crystallography and degradation studies.[5][4] The first total synthesis of azulene was achieved in 1937 by A. S. Pfau and Placidus Plattner, involving a multi-step process based on ring expansion of a hydroazulenone intermediate, followed by decarboxylation and dehydrogenation.[6][4]Key properties of azulene include its unusual aromaticity, where the five-membered ring behaves as a cyclopentadienyl anion and the seven-membered ring as a tropylium cation in the resonance hybrid, contributing to its stability and polar nature.[2] It exhibits a melting point of 99–100 °C and boils at 170 °C under reduced pressure, with low solubility in water but good solubility in organic solvents.[1] Azulene occurs naturally as a component of essential oils from plants such as German chamomile (Matricaria recutita), primarily in the form of derivatives like chamazulene, which provide anti-inflammatory effects.[1]In applications, azulene and its derivatives are used in cosmetics for their soothing and anti-inflammatory properties, particularly in treating skin irritations, and have been explored in medicinal chemistry for potential antineoplastic and nonlinear optical materials due to their electronic properties.[1][4]
Introduction
Overview
Azulene is a bicyclic aromatic hydrocarbon with the molecular formula C₁₀H₈, serving as a structural isomer of naphthalene but distinguished by its fused five- and seven-membered rings forming a 10 π-electron system.[7] Unlike the colorless naphthalene, azulene exhibits a distinctive deep blue color arising from an intramolecular charge-transfer transition associated with its polarized structure, which shifts absorption into the visible spectrum.[2]Compounds featuring the azulene skeleton occur naturally as pigments in various biological sources, including certain mushrooms, guaiac wood oil derived from the tree Bulnesia sarmientoi, and essential oils such as those from chamomile and yarrow.[8] These natural azulene derivatives, like guaiazulene, contribute to the blue hues observed in these materials and have been utilized in traditional applications for centuries.[9]The name "azulene" derives from the Spanish word "azul," meaning blue, and was coined in 1863 by the British chemist Septimus Piesse upon isolating the compound from distillates of essential oils.[10]
History
Azulene was first isolated in 1863 by British chemist and perfumer Septimus Piesse from the azure-blue distillate obtained during steam distillation of essential oils from plants such as chamomile, yarrow, and wormwood. Piesse named the compound "azulene," derived from the Spanish word "azul" meaning blue, to reflect its distinctive vivid color, which arises from its presence as a chromophore in various plant essential oils including those from chamomile, yarrow, and wormwood.[11]The structural elucidation of azulene progressed in the 1920s and 1930s amid efforts to characterize natural blue oils. In 1926, Lavoslav Ružička proposed an initial structure for guaiazulene, a common natural derivative, but it was Placidus A. Plattner and Alexander Pfau at the University of Geneva who definitively established the fused cyclopenta-fused cycloheptene skeleton in 1936 through degradative studies and comparative analyses of natural isolates. Their work clarified the non-benzenoid bicyclic framework, resolving earlier uncertainties about the arrangement of its 10 carbon atoms and distinguishing it from naphthalene isomers.[12]Building on this foundation, Plattner and Pfau achieved the first total synthesis of azulene in 1936 via a multi-step ring expansion route starting from cyclopentenocycloheptanone, an intermediate derived from 1,6-cyclodecadiene, followed by dehydrogenation to yield the aromatic hydrocarbon in low but confirmatory yields. This milestone enabled further exploration of azulene's properties beyond natural extracts.[6]In the 1950s, with the advent of more accessible synthetic routes like the Hafner method involving cyclopentadiene and acetylene derivatives, researchers began detailed investigations into azulene's electronic structure. Early studies highlighted its nonalternant aromaticity, satisfying Hückel's 4n+2 π-electron rule across the fused rings despite unequal bond lengths, and measured its unexpectedly large dipole moment of approximately 1.0 D—attributed to charge separation between the five- and seven-membered rings—contrasting sharply with the nonpolar naphthalene. These findings, including theoretical calculations and spectroscopic analyses, underscored azulene's unique reactivity and optical properties.
Properties
Physical Properties
Azulene has the molecular formula \ce{C10H8} and a molar mass of 128.17 g/mol.[1]It appears as dark blue crystalline solid with a melting point of 98–100 °C and a boiling point of 242 °C at standard pressure.[13] The density of the liquid is 0.987 g/cm³ at 25 °C.[13] Its relatively high boiling point can be briefly attributed to contributions from aromatic stability.Azulene shows low solubility in water, with a value of 0.015 g/L at 25 °C, rendering it effectively insoluble for most practical purposes.[13] In contrast, it is readily soluble in common organic solvents such as ethanol, diethyl ether, and benzene, facilitating its handling in non-aqueous media.The characteristic blue color of azulene arises from absorption in the visible spectrum, primarily due to the weak \ce{S0 -> S1} transition with a maximum around 580 nm (\epsilon \approx 200 M^{-1} cm^{-1}), while a stronger \ce{S0 -> S2} band appears at approximately 340 nm (\epsilon \approx 5000 M^{-1} cm^{-1}).[14] These spectroscopic features are key for identification and arise from the molecule's non-alternant structure.[15]The standard enthalpy of combustion of azulene is −5290.7 kJ/mol for the solidphase at 298 K.[16] This value underscores its thermodynamic stability relative to combustion products like CO_2 and H_2O.
Chemical Properties
Azulene demonstrates notable thermalstability, a consequence of its aromatic stabilization energy, which arises from the delocalized 10 π-electron system across its fused five- and seven-membered rings. This stability is evidenced by its high enthalpy of formation in the gas phase, Δ_f H°_m(g) = 309.2 ± 4.9 kJ/mol at 298.15 K, derived from combustioncalorimetry measurements. The standard molar enthalpy of combustion for solid azulene is Δ_c H°_m = −(5285.5 ± 4.9) kJ/mol, indicating a robust energetic barrier to decomposition under standard conditions.[16][16][16]In terms of reactivity, azulene shows a distinct regioselectivity in electrophilic aromatic substitution, with the 1- and 3-positions in the five-membered ring being most reactive due to elevated electron density at these sites. This preference is observed across various electrophiles, including nitrating and halogenating agents, where substitution at the 1-position predominates, while the 2-position in the five-membered ring and positions in the seven-membered ring exhibit lower reactivity. Such behavior underscores azulene's resistance to substitution at less activated sites, preserving its core structure under mild electrophilic conditions.[17][17]The molecule's uneven charge distribution, stemming from resonance structures that polarize the π-system, imparts a permanent electric dipole moment of 1.08 D along the long axis. This polarity, unusual for a neutral hydrocarbon like its isomer naphthalene (which has zero dipole moment), contributes to azulene's characteristic blue color through intramolecular charge transfer. Basic thermodynamic properties further characterize its behavior; for instance, the constant-pressure heat capacity of gaseous azulene at 298.15 K is 128.41 J/mol·K.[16][18]Azulene also exhibits anomalous photophysical behavior by violating Kasha's rule, emitting fluorescence from the second excited singlet state (S₂) rather than the lowest (S₁). This arises because the S₁ state is antiaromatic, promoting rapid nonradiative decay to the ground state via a conical intersection, whereas the S₂ state maintains aromatic character with minimal geometric relaxation and a large S₂–S₁ energy gap (~14,000 cm⁻¹) that hinders internal conversion.
Structure and Bonding
Molecular Geometry
Azulene possesses a bicyclic structure composed of a fused five-membered cyclopentadiene ring and a seven-membered cycloheptatriene ring, sharing two adjacent carbon atoms to form a 10-carbon perimeter. This fused ring system results in a compact, non-alternant hydrocarbon framework with the formula C₁₀H₈.[1]The molecule adopts a fully planar conformation, as confirmed by X-ray crystallography, microwave spectroscopy, and electron diffraction studies, with no significant out-of-plane deviations in the gas or solid phases.[19] This planarity facilitates efficient π-orbital overlap across the ring system. Bond lengths exhibit minimal alternation, reflecting uniform electron distribution; experimental X-ray data indicate peripheral C-C bonds averaging 1.403 Å, while the central transannular bond (between positions 9 and 10) measures 1.501 Å. Representative values include the C1-C2 bond at 1.37 Å in the five-membered ring and the C3-C4 bond at 1.40 Å in the seven-membered ring, consistent with both X-ray and high-level computational geometries.[20]In the solid state, azulene crystallizes in the monoclinic space group P2₁/c, but its structure is complicated by dynamic disorder, leading to high correlations in atomic positional parameters and challenges in refinement. Invariom modeling has been employed to accurately describe the electron density, revealing a planar molecular core with intermolecular distances influenced by van der Waals interactions and the inherent dipole moment. Lattice energy minimizations indicate close-packed layers with minimal deviations from the ideal geometry, though disorder results in averaged occupancies for certain atoms in the crystal packing.[19]
Aromaticity and Electronic Properties
Azulene possesses a 10 π-electron system delocalized over its fused five- and seven-membered rings, satisfying Hückel's rule for aromaticity with n=2 in the 4n+2 formulation. This configuration confers overall aromatic stability, despite the non-alternant hydrocarbon structure that distinguishes it from benzene or naphthalene. The molecule's planarity and cyclic conjugation further support this aromatic character, enabling efficient π-overlap across the bicyclic framework.[21]The electronic structure of azulene exhibits unequal π-electron distribution, arising from resonance forms that depict the five-membered ring as a cyclopentadienyl anion (6 π electrons) and the seven-membered ring as a tropylium cation (6 π electrons), with charge separation leading to higher electron density in the five-membered ring and relative deficiency in the seven-membered ring. This polarization results in a permanent dipole moment of approximately 1.0 D, with the negative end at the five-membered ring. The non-uniform distribution underlies azulene's reactivity preferences and photophysical anomalies.[22]Frontier molecular orbitals in azulene display reduced symmetry due to its non-alternant nature, with the highest occupied molecular orbital (HOMO) localized more on the five-membered ring and the lowest unoccupied molecular orbital (LUMO) on the seven-membered ring, leading to minimal overlap and a relatively narrow HOMO-LUMO gap compared to alternant hydrocarbons like naphthalene. This gap, typically around 2.1 eV from optical transitions, facilitates accessible excited states and contributes to the molecule's blue color.[23][24]A hallmark of azulene's photophysics is its anti-Kasha emission, where fluorescence predominantly occurs from the second excited singlet state (S₂) to the ground state (S₀) rather than from the lowest excited singlet state (S₁), violating Kasha's rule. This behavior stems from the forbidden nature of the S₁ → S₀ transition, attributed to symmetry selection rules and a large energy gap (∼14,000 cm⁻¹) between S₁ and S₂, which minimizes internal conversion from S₂ to S₁.[21]Recent studies through 2025 have elucidated the excited-state dynamics of azulene, emphasizing the roles of energy gaps and vibronic coupling in modulating anti-Kasha emission. Ultrafast spectroscopy reveals rapid S₂ relaxation pathways influenced by solvent interactions and substituent effects, with vibronic couplings driving efficient nonradiative decay from higher states while preserving S₂ fluorescence quantum yields up to 0.3 in solution. These investigations highlight how perturbations in the S₁-S₂ gap can enable dual emission or suppress anti-Kasha behavior in derivatives, informing design of novel fluorophores.[25]
Synthesis
Early Syntheses
The first total synthesis of azulene was achieved in 1937 by Placidus A. Plattner and Alexander S. Pfau through a ring expansion strategy starting from indane. The key step involved the addition of diazoacetic ester to indane, which inserted a carbon unit into the five-membered ring, forming a carboxylic acid derivative after hydrolysis. Subsequent decarboxylation and dehydrogenation yielded azulene, though the overall process was multi-step and suffered from low efficiency. This pioneering route established the bicyclic [5.3.0] framework but was limited by modest yields due to side reactions during the carbene insertion and incomplete aromatization requiring harsh conditions like selenium dioxide oxidation.Purification posed significant challenges in these early efforts, as the vivid blue color of azulene facilitated visual detection but complicated isolation from colorless byproducts and polymeric impurities formed during dehydrogenation. The product often required repeated chromatography or sublimation under reduced pressure to achieve purity, highlighting the technical hurdles in handling such a reactive, colored hydrocarbon. Despite these issues, the Plattner-Pfau synthesis provided the first unambiguous confirmation of azulene's structure, aligning with spectroscopic and degradative studies from natural sources.In the 1950s, alternative routes emerged, including the Ziegler-Hafner method, which utilized fulvene intermediates in a cyclization approach. This involved generating 6-dimethylaminofulvene from cyclopentadiene and a pyridinium salt, followed by reaction with acetylene under basic conditions to form a seven-membered ring precursor, and final aromatization via dehydrogenation. Yields for this sequence reached 50-60% in optimized variants, a marked improvement over prior methods, though it still demanded careful control to avoid fulvene polymerization. The strategy exemplified the growing use of Diels-Alder-inspired cycloadditions with fulvenes as dienes, enabling substitution patterns on the five-membered ring.[26]These early syntheses underscored the synthetic challenges of constructing azulene's non-alternant aromatic system, with common limitations including poor regioselectivity in ring expansions and the need for vigorous dehydrogenation steps that reduced overall efficiency to below 20% in most cases. For instance, the diazoacetic ester addition in the Plattner-Pfau route could be represented conceptually as indane undergoing carbene insertion to yield a tropylium-like intermediate, followed by CO₂ loss and H₂ abstraction, though practical execution often introduced skeletal rearrangements. Such approaches paved the way for later refinements but remained labor-intensive for preparative scales.
Modern Methods
Since the mid-20th century, synthetic approaches to azulene have evolved toward more efficient, scalable methods that leverage catalysis and mild conditions to overcome the limitations of classical multi-step routes. These modern techniques, developed from the 1960s onward, emphasize one-pot processes, transition-metal catalysis, and assisted cycloadditions to enable the construction of the azulene core and its fused variants with improved yields and selectivity.[4]One notable advancement is the one-pot annulation via pyridine-mediated cycloadditions, which utilizes Zincke aldehyde derivatives from pyridinium salts reacting with cyclopentadienyl anions. This process involves ring opening of the pyridinium salt to form a Zincke iminium intermediate, followed by condensation with cyclopentadiene under basic conditions (e.g., sodium ethoxide), and subsequent 10π electrocyclic cyclization with amine elimination to afford the azulene framework. Recent implementations, including microwave-assisted variants, have achieved yields up to 64% for substituted azulenes like 6-methylazulene, highlighting the method's convenience for accessing functionalized derivatives.[4]Transition-metal-catalyzed routes have significantly expanded azulene synthesis, particularly through palladium-mediated couplings that facilitate ring fusion. For instance, Pd-catalyzed cross-coupling reactions, such as Suzuki-Miyaura and Heck-type annulations, enable the attachment of aryl or heteroaryl groups to azulene's five- or seven-membered rings, followed by intramolecular cyclization to form fused systems. These methods, reviewed in 2021, typically employ PdCl₂(dppf) or similar catalysts with bases like K₃PO₄, yielding aryl-substituted azulenes in 70-95% efficiency and allowing precise control over substitution patterns for extended polycyclic architectures.[27]Recent bottom-up syntheses (2022-2025) for azulene-embedded nanographenes have integrated Suzuki-Miyaura coupling with acid-mediated cyclization to construct complex fused structures. In a 2025 approach, 2-chloroazulenes bearing 1,3-ester groups undergo Pd-catalyzed Suzuki-Miyaura coupling with arylboronic acids (e.g., 1-naphthalenboronic acid) using PdCl₂(dppf) and K₃PO₄, affording biaryl intermediates in 71-97% yields. Subsequent Brønsted acid-mediated intramolecular cyclization with methanesulfonic acid (MeSO₃H) at 100°C promotes ring closure, yielding azuleno[2,1-a]phenalenones in 71-99% yields, which exhibit unique optical and electrochemical properties suitable for materials applications.[28]High-yield variants of Diels-Alder-type cycloadditions, enhanced by microwave assistance, provide rapid access to azulene scaffolds. A key example is the microwave-promoted [6+4]-cycloaddition of 6-aminofulvene with α-pyrones, which proceeds under solvent-free or low-solvent conditions to form azulene-indole adducts, followed by CO₂ extrusion and aromatization. Developed in 2001 but representative of ongoing microwave optimizations, this method accelerates reaction times to minutes while delivering yields exceeding 80% for antineoplastic azulene-indoles, demonstrating scalability for pharmaceutical precursors.[29]Brønsted acid cyclizations have emerged as versatile tools for assembling fused azulene systems from prefunctionalized precursors. In 2022, [8+2] cycloaddition of 2H-cycloheptafuran-2-ones with enamines yielded benzazulene derivatives, which were then subjected to Brønsted acid-mediated (e.g., 100% H₃PO₄) intramolecular cyclization via Knoevenagel condensation with dimethyl malonate, followed by dehydration and deprotection, affording unsubstituted benzazulenes in high yields as the most efficient route reported. Similarly, a 2023 method employs polyphosphoric acid (PPA) at 140°C to cyclize 1,3-diethoxycarbonyl-2-arylaminoazulenes through protonation and Friedel-Crafts acylation with decarboxylation, producing azuleno[2,1-b]quinolones in 85-100% yields. These acid-promoted strategies underscore the role of strong Brønsted acids in driving regioselective ring fusions under mild thermal conditions.[30][31]
Derivatives
Natural Derivatives
Azulene derivatives occur naturally in various biological sources, predominantly as sesquiterpenoids with alkyl substitutions that modify their blue pigmentation and volatility. These compounds contribute to the coloration and aromatic profiles of essential oils extracted from plants and fungi.Guaiazulene, a sesquiterpene featuring methyl and isopropyl groups on the azulene core, is primarily isolated from the essential oil of guaiac wood derived from Guaiacum officinale trees.[32] Chamazulene, a related sesquiterpene with ethyl substitution, is obtained from German chamomile (Matricaria recutita L.) via steam distillation, during which the precursor matricin converts to chamazulene, imparting anti-inflammatory properties.[33][1] These oils are obtained via steam distillation, a process that efficiently captures volatile components while converting precursors in chamomile to related azulenes.[34] Guaiazulene imparts a deep blue hue and woody, rose-like aroma, making it valuable in perfumery to enhance fragrance depth and stability.[35]In fungal sources, azulene derivatives are present in certain mushrooms, including species within the order Agaricales such as Omphalotus, where they arise from the oxidation of sesquiterpenes like lactarofulvenol and contribute to the vivid blue-green latex exuded by injured fruiting bodies.[36] These compounds can be isolated through solvent extraction followed by chromatographic separation from the fungal biomass.[33]Marine algae also harbor azulene-based metabolites, notably hydroazulene diterpenes isolated from brown algae of the genusDictyota, such as Dictyota volubilis, using solvent partitioning and purification techniques to yield structurally diverse variants with oxygenated functionalities.[37]Among structural variations, alkylated azulenes like vetivazulene—a dimethyl-isopropyl derivative—are extracted from vetiver root oil (Chrysopogon zizanioides) via steam distillation, yielding a component that adds earthy notes to the oil's complex profile.[38]
Synthetic Derivatives
Hydroxyazulenes represent key synthetic derivatives of azulene, prepared through targeted modifications to introduce hydroxyl groups at specific ring positions, altering their electronic and acidic properties. 1-Hydroxyazulene, synthesized by the reduction of 1-acetoxyazulene with lithium aluminum hydride, exists as an unstable green oil that resists keto-enol tautomerism due to the high energy barrier for rearrangement.[39] Its relatively low acidity reflects behavior comparable to non-aromatic alcohols rather than typical phenols.[40] In contrast, 2-hydroxyazulene, obtained via hydrolysis of 2-methoxyazulene with hydrobromic acid, is stable and undergoes keto-enol tautomerism, with the enol form predominating. This isomer exhibits unusual acidity due to enhanced stabilization of the conjugate base by the azulene π-system.Ring-fused azulenes have seen significant advances in synthesis between 2022 and 2025, yielding linear polycyclic hydrocarbons with embedded azulene motifs that exhibit small HOMO-LUMO bandgaps, often below 2 eV, enabling applications in optoelectronics. For instance, azuacenes—fused systems combining azulene with linear acenes in a 6-7-5 ring topology—have been prepared through stepwise cyclization and coupling reactions, demonstrating bandgaps as low as 1.75 eV alongside improved ambient stability compared to alternant polycyclic aromatic hydrocarbons.[41] Similarly, a nonalternant pentacene isomer incorporating two azulene units achieves an optical bandgap of 2.046 eV and extended half-life under air exposure, attributed to the nonalternant topology disrupting reactive edge sites.[42] Azulene-fused linear polycyclic aromatic hydrocarbons, synthesized in four steps from commercial azulene, further highlight small bandgaps (around 1.8 eV), high thermalstability up to 300°C, and reversible color changes under acid/base stimuli.[43]Azulene-based conjugated polymers, accessed through 2,6-functionalization, have emerged as a 2023 innovation, leveraging the less reactive positions in the five-membered ring for polymer backbone construction. These polymers, synthesized via direct arylation polycondensation of 2,6-dihaloazulene monomers with comonomers like carbazole or thiophene, display extended conjugation with bandgaps of 1.9–2.2 eV and proton-responsive fluorescence quenching, arising from azulenium ion formation.[44] A C-H activation strategy enables tuning of dipole orientation in 2,6-azulene copolymers, yielding unipolar n-type organic semiconductors with electron mobilities up to 0.5 cm² V⁻¹ s⁻¹ and balanced HOMO/LUMO levels for transistor applications.[45]Functional groups at positions 1 and 3 of azulene serve as critical handles for tuning reactivity in synthetic derivatives, exploiting the high electron density at these nucleophilic sites in the seven-membered ring. Electrophilic aromatic substitution readily introduces substituents like halogens or alkyl groups at 1,3-positions, modulating the dipole moment and redox potential; for example, 1,3-dichloroazulene derivatives exhibit shifted absorption maxima by 50 nm compared to unsubstituted azulene.[46] Palladium-catalyzed cross-coupling reactions, such as Suzuki-Miyaura or Sonogashira, further enable precise installation of aryl or alkynyl groups at these positions under mild conditions, enhancing solubility and conjugation length while preserving aromaticity.[27] This regioselective functionalization has facilitated derivatives with tailored electrophilicity, as seen in 1,3-bis(ethynyl)azulenes used as building blocks for extended π-systems.
Coordination and Organometallic Chemistry
Azulene as a Ligand
Azulene serves as a versatile ligand in coordination chemistry owing to its nonalternant aromatic structure, which enables diverse binding modes to transition metals. Its fused five- and seven-membered rings allow for ambidentate coordination, where the ligand can interact with metal centers through either ring system, often exhibiting hapticities ranging from η² to η⁶ depending on the metal fragment and reaction conditions.[47] This flexibility arises from azulene's zwitterionic character, with the five-membered ring acting as an electron-rich donor and the seven-membered ring as an electron-deficient acceptor, facilitating adaptation to the electronic demands of the metal.[48]A prominent binding mode is η⁶-coordination through the seven-membered ring, which mimics the π-system of benzene and donates six electrons to the metal, as seen in complexes with early transition metals and group 6 elements. In such interactions, the seven-membered ring's conjugated π-electrons engage the metal center, leading to stable sandwich-like or half-sandwich structures, while the five-membered ring remains uncoordinated or participates minimally.[47] This mode is particularly favored in mononuclear complexes where the metal requires a neutral, aromatic η⁶-ligand, enhancing the stability through back-donation from the metal to the ligand's π*-orbitals.[49]The ambidentate behavior of azulene allows it to bind selectively through the five-membered ring in η⁵-fashion, resembling cyclopentadienyl ligands, or through the seven-membered ring in η⁴- or η⁶-modes, enabling haptotropic shifts between rings in response to steric or electronic perturbations. For instance, in dinuclear complexes, one metal may coordinate to the five-membered ring while another binds the seven-membered ring, promoting bridging architectures and isomerism.[47] This dual-site coordination is influenced by spectator ligands on the metal, which can trigger rearrangements to optimize metal-ligand overlap.[48]Azulene's inherent dipole moment of approximately 1.0 D, with partial negative charge on the five-membered ring and positive on the seven-membered ring, significantly modulates metal-ligand interactions by directing charge transfer and stabilizing polar bonds. This polarity enhances the ligand's ability to stabilize low-valent metals through donation from the electron-rich five-membered ring or acceptance into the electron-poor seven-membered ring, influencing reactivity in catalytic processes.[50] The dipole also contributes to selective binding in asymmetric environments, where the ligand's orientation aligns with the metal's electrostatic field.[51]Recent advances have leveraged azulene's coordination properties in self-assembly, notably in 2025 reports of metallacycles formed from azulene-derived dipyridyl ligands and Cp*Rh building blocks. These structures assemble via η²-coordination of the pyridyl arms to the rhodium centers, with the azulene core providing rigidity and π-stacking for dimerization, yielding photocatalytically active cages for sulfide oxidation.[52] Such assemblies highlight azulene's role in constructing functional supramolecular architectures with tunable electronic properties.[53]
Notable Complexes
One notable organometallic complex involving azulene is (azulene)Fe₂(CO)₅, a dinuclear iron compound where azulene bridges two iron centers with an overall formula of C₁₀H₈Fe₂(CO)₅.[54] The structure features five carbonyl ligands distributed across the two iron atoms, with azulene coordinating in a bidentate fashion through its five- and seven-membered rings, exhibiting variable hapticity (typically η⁵:η³) that allows for bridging.[47] This complex is synthesized via reactions of azulene with iron carbonyl precursors like Fe₂(CO)₉, leading to stable dinuclear species, though detailed synthetic conditions are often explored computationally for optimization.[47]Another prominent example is (azulene)Mo₂(CO)₆, a well-established dinuclear molybdenum complex with the formula (η⁵,η⁵-C₁₀H₈)Mo₂(CO)₆.[47] In this structure, azulene serves as a bridging ligand between the two molybdenum atoms, each bearing three terminal CO ligands, resulting in a symmetric coordination where the metal-metal bond is supported by the asymmetric nature of azulene's rings.[47] The complex is prepared by refluxing azulene with Mo(CO)₆ or related precursors in nonpolar solvents, yielding a stable product characterized by its hapticity-driven bonding.[47]Recent advancements include azulene-based rhodium metallacycles developed for sulfide sensing applications.[52] These triangular metallacycles (e.g., [CpRh]₃(azulene-derived ligand)₃) are self-assembled at room temperature from a CpRh building block and dipyridyl azulene ligands in methanol/dimethylaniline mixtures, confirmed by NMR, mass spectrometry, and X-ray crystallography showing π-π stacked dimers with interplanar distances of 3.6–3.9 Å.[52] The complexes exhibit UV-vis absorption at 330 nm and 410 nm due to π–π* transitions and demonstrate high photostability, enabling efficient visible-light-driven photooxidation of sulfides to sulfoxides with yields up to 97% in 4 hours, recyclable over three cycles.[52]Regarding stability and reactivity, these azulene complexes generally display robust thermalstability due to the bridging ligand's ability to accommodate different metal oxidation states, but they exhibit fluxional behavior in solution involving haptotropic shifts between ring coordination modes.[47]CO substitution patterns are common, with labile terminal CO ligands allowing replacement by phosphines or isonitriles under mild heating, as seen in analogous systems where dissociation leads to new dinuclear variants without disrupting the core structure.[47]
Applications
Materials and Optoelectronics
Azulene-fused polycyclic aromatic hydrocarbons (PAHs) have emerged as promising materials for small-bandgap semiconductors due to their extended π-conjugation and tunable electronic properties. In 2025, efficient syntheses of such azulene-fused systems, including azuleno[2,1-a]phenalenones, were reported using Suzuki–Miyaura coupling followed by Brønsted acid-mediated cyclization, yielding compounds with distinctive optical absorption in the near-infrared region and electrochemical stability suitable for semiconductor applications.[28] Earlier work demonstrated that linear azulene-fused PAHs exhibit optical bandgaps as low as 1.0 eV, enabling their use in organic field-effect transistors and photovoltaic devices with high charge carrier mobilities.[43] Additionally, 2025 advancements in Scholl-type oxidation of 1,2,3-triarylazulenes produced embedded azulene PAHs with enhanced planarity and reduced bandgaps, further optimizing them for low-energy optoelectronic functions.[55]Conjugated polymers incorporating azulene units offer exceptional stability and reversible responses to external stimuli, making them ideal for organic electronics such as sensors and memory devices. These polymers leverage the seven-membered ring of azulene to achieve unique connectivity that promotes stimuli-responsiveness, including pH- or redox-triggered color changes and conductivity modulation without degradation. For instance, azulene-based polythiophenes demonstrate high thermal stability up to 400°C and reversible switching in field-effect transistors upon protonation, attributed to the formation of aromatic azulenium cations that alter the bandgap reversibly.[56] This responsiveness stems from azulene's inherent dipole and non-alternant structure, enabling applications in flexible electronics where durability under operational stresses is critical.The optoelectronic properties of azaazulenes, nitrogen-substituted analogs of azulene, feature narrow bandgaps and tunable fluorescence. These compounds exhibit potential for deep-red to near-infrared absorption suitable for organic light-emitting diodes and bioimaging probes due to their non-alternant electronic structure.[57]
Biological and Medicinal Uses
Guaiazulene, a derivative of azulene found in essential oils from plants such as chamomile, exhibits significant anti-inflammatory properties by inhibiting histamine release and reducing edema in animal models.[58] In clinical applications, topical formulations containing 0.02% guaiazulene alleviate itching in atopic dermatitis patients, while 0.05% ointments promote healing of diaperdermatitis in newborns within three days and prevent nipple fissures during breastfeeding.[58] Sodium guaiazulene sulfonate, a water-soluble derivative, is employed for its wound-healing effects in treating gastritis and oral canker sores due to its ability to soothe mucosal inflammation.[59] In cosmetics, guaiazulene serves as an FDA-approved colorant and active ingredient in skincare products, leveraging its anti-inflammatory and antioxidant activities to reduce skin irritation.[58]Photodynamic therapy (PDT) incorporating azulene derivatives shows promise in cancer research as photosensitizers. When activated by red laser light (625–638 nm), azulene generates singlet oxygen, leading to reactive oxygen species (ROS) production that decreases viability of peripheral blood mononuclear cells (PBMCs) at concentrations of 15 μM and energy densities of 4.2–200 J/cm², with an inverse correlation between singlet oxygen yield and cell survival (Spearman's r = -0.610).[60] Similarly, guaiazulene at 2–5 μM under 4–8 J/cm² irradiation suppresses inflammatory markers like RANTES and PGE₂ in TNF-α-stimulated PBMCs without compromising cell viability, suggesting targeted cytotoxicity against tumor cells via ROS-mediated apoptosis.[61] Pulsed laser modes with intermittency factors of 5–9 enhance singlet oxygen formation up to threefold at low azulene doses (1–10 μM), improving immunosuppressive effects potentially applicable to photodynamic cancer treatments.[62]Azulene derivatives occur naturally in fungi, such as Lactarius indigo mushrooms and Ganoderma theaecolum, where they contribute to blue pigmentation by absorbing light in the 500–700 nm range.[63] In these organisms, azulenes function in defense mechanisms, exhibiting antimicrobial and antifungal activities that inhibit pathogens like Saccharomyces cerevisiae, thereby protecting against microbial invasions and environmental stressors.[63] This pigmentation and bioactivity likely aid in ecological roles, including UV protection and deterrence of herbivores or competitors in fungal habitats.[63]Studies on hydroxyazulene (azulenol) tautomerism reveal varying acidity among isomers, with 6-hydroxyazulene displaying the strongest acidity in vacuo—greater than phenol or naphthols—while 1-hydroxyazulene is the weakest; keto tautomers influence deprotonation and acid-base behavior, which could inform proton-transfer mechanisms in drug design.[64]Azulene and its derivatives generally exhibit low toxicity profiles. Oral LD50 values for azulene are 3 g/kg in mice and 4 g/kg in rats, while guaiazulene has an LD50 of 1.55 g/kg in rats, indicating minimal acute risk at therapeutic doses with no systemic side effects from topical use, though rare allergic reactions like cheilitis may occur.[65][58]