Fact-checked by Grok 2 weeks ago

Tropolone

Tropolone is a heterocyclic with the molecular formula C₇H₆O₂, consisting of a seven-membered non-benzenoid aromatic that incorporates a group at position 1 and a hydroxy at position 2, enabling enol-keto tautomerism and conferring unique chelating properties. Its IUPAC name is 2-hydroxycyclohepta-2,4,6-trien-1-one, and it serves as the parent structure for a class of natural products known as tropolones, which feature additional substituents on the ring. Physically, tropolone appears as a white to pale yellow crystalline solid with a of 50–52 °C and a of 80–84 °C at 0.1 mmHg. It exhibits good solubility in organic solvents such as , , and (up to 30 mg/mL in ), but limited in . Chemically, its arises from a 6π-electron system, allowing it to participate in reactions typical of both enols and ketones, including metal with divalent cations like Cu²⁺ and Fe³⁺ due to the formation of stable five-membered rings. Tropolone occurs naturally as a in various organisms, including such as Burkholderia plantarii, where it functions as a inhibiting rice seedling growth. It is also produced by certain fungi and plants, often as part of more complex tropolonoids; approximately 200 such derivatives have been identified across these sources since the mid-20th century. Biosynthetically, tropolones are typically derived from pathways involving ring expansion and oxidation of precursor cyclic structures. Tropolone and its derivatives display diverse biological activities, including antibacterial, , antiviral, and antitumor effects, largely attributed to their ability to inhibit metalloenzymes by chelating cofactors like . Notably, tropolone acts as a potent inhibitor of mushroom (IC₅₀ = 0.4 μM), with potential applications in pigmentation disorders and formulations. In and synthesis, tropolone serves as a building block for pharmaceuticals, such as the gout treatment , and as a in for constructing complex tropolonoids.

Structure and Properties

Molecular Structure

Tropolone exhibits a seven-membered cyclic structure with the molecular formula C<sub>7</sub>H<sub>6</sub>O<sub>2</sub>, characterized by a of three double bonds in a ring, a hydroxyl group at position 2, and a at position 1. This arrangement positions the hydroxyl and carbonyl groups in proximity, facilitating potential intramolecular interactions. The systematic name, 2-hydroxycyclohepta-2,4,6-trien-1-one, underscores the -keto positioning central to its molecular architecture, where the form dominates due to stabilization by the adjacent carbonyl. X-ray crystallographic analyses of tropolone and its derivatives indicate s and angles consistent with partial aromatic character, including an average C-C of approximately 1.40 across the ring, reflecting delocalized π-electrons rather than strict alternation. Specific measurements, such as the C(1)-C(2) bond at 1.452 in the neutral form, further support this non-alternating geometry. Tropolone serves as a hydroxylated derivative of tropone (C<sub>7</sub>H<sub>6</sub>O), the deoxygenated analog consisting of cyclohepta-2,4,6-trien-1-one, which acts as a key structural precursor in elucidating tropolone's conjugated framework.

Physical and Spectroscopic Properties

Tropolone is a pale yellow crystalline solid with a of 50–52 °C and a of 243 °C at . It exhibits moderate solubility in water (approximately 41 g/L at 25 °C) and high solubility in organic solvents such as (30 g/L), (10 g/L), , and . The UV-Vis of tropolone in nonpolar solvents displays intense bands at 220–250 nm (ε_max ≈ 30,000 M⁻¹ cm⁻¹) and a less intense band at 340–375 nm (ε_max ≈ 8,000 M⁻¹ cm⁻¹), corresponding to π-π* transitions in the conjugated tropylium-like ring system; these features shift in polar or protic media due to solvent interactions with the enolic . In the infrared spectrum, tropolone shows a characteristic broadened C=O stretching vibration at approximately 1600 cm⁻¹, shifted and widened by intramolecular ing in the form, along with a broad O-H stretching band centered around 3200 cm⁻¹ extending to lower frequencies due to the strong chelated . The ¹H NMR spectrum in deuterated solvents reveals seven ring protons with signals typically between 6.8 and 7.9 , exhibiting the expected multiplicity for the unsymmetrical structure, while the enolic OH proton appears as a broad at ~16 , a diagnostic indicator of the intramolecular ing stabilizing the . The ¹³C NMR spectrum consists of four signals reflecting rapid at , with the carbonyl/enolic carbons at δ ≈ 172 (relative to TMS), the symmetric C-3/C-5 at ~135 , C-4 at ~130 , and C-2/C-6 at ~125 , consistent with the time-averaged C_{2v} of the delocalized system.

Tautomerism and Aromaticity

Tropolone undergoes - , primarily existing in the form named 2-hydroxycyclohepta-2,4,6-trien-1-one, which is stabilized by a strong intramolecular between the hydroxyl proton and the carbonyl oxygen. The less prevalent form, 2-oxocyclohept-3,5,7-trien-1-one, lies higher in energy, with computational studies estimating the difference at approximately 6 kcal/mol favoring the in both gas and aqueous phases. The aromatic character of tropolone resides in its form, which features a delocalized 6π-electron system across the seven-membered ring, adhering to for monocyclic, planar, and fully conjugated species with 4n + 2 π electrons (where n = 1). This is supported by resonance structures that distribute the electron density from the enolized OH and adjacent C=O groups, effectively mimicking a tropylium-like anion-cation pair within the neutral framework. Computational evidence reinforces this , with nucleus-independent (NICS) analyses yielding negative values (e.g., around -18 at the ring center) that signify diatropicity and magnetic shielding consistent with aromatic stabilization in the form, whereas the tautomers display near-zero or positive NICS indicative of reduced or absent . Additionally, tropolone's reactivity patterns, such as preferential electrophilic attack at positions and to the group akin to , further corroborate its aromatic delocalization. Tropolone serves as a neutral counterpart to the (C₇H₇⁺), a canonical 6π-aromatic species, illustrating how substitution enables charge-neutral in seven-membered rings through analogous π-delocalization.

Synthesis

Historical Development

The tropolone structure was first proposed in 1945 by Michael J. S. Dewar to account for the unusual chemical behavior of stipitatic acid, a isolated from the Penicillium stipitatum. In 1950, Robert E. Lutz and J. S. Gillespie Jr. provided key evidence for the tropolone framework during degradation studies of puberulic acid, another fungal from Penicillium puberulum; oxidative cleavage and yielded products consistent with a seven-membered hydroxy-ketone ring, confirmed by comparison with known derivatives. Independent total syntheses of tropolone in 1950 by three groups solidified its structure: William von E. Doering and Leon H. Knox reported the synthesis via oxidation of cyclohepta-1,3,5-triene with ; J. W. Cook, A. R. Gibb, R. A. , and A. R. Somerville used bromination followed by dehydrobromination of cycloheptanone; and Tetsuo Nozoe, Shuichi Seto, Yoshio Kitahara, Masao Kunori, and Yuya Nakayama achieved synthesis by bromination and dehydrobromination of cycloheptanone, with structural verification through , spectroscopic, and degradative matches to natural samples. The name "" for the parent ketone (cyclohepta-2,4,6-trien-1-one) was introduced by Doering and Frances L. Detert in 1951 upon its synthesis via addition to and thermal rearrangement, highlighting its aromatic-like stability; "" followed as the enol form with the conventional "-olone" for α-hydroxy ketones. Initial efforts to elucidate tropolone's were complicated by its atypical reactivity, including resistance to typical -ketone interconversions and strong metal-chelating ability, which deviated from expected aliphatic behavior; by 1955, these properties were ascribed to delocalized π-electron involving zwitterionic contributors, as detailed in comprehensive reviews integrating synthetic, degradative, and spectroscopic data.

Modern Synthetic Methods

One prominent modern approach to tropolone synthesis involves the oxidation of derivatives to install the requisite , facilitating ization to the . serves as an effective oxidant in this transformation, converting to in yields of 50–70%, with subsequent or base-promoted enolization yielding tropolone. Recent variants employ of dioxole-fused s under mild conditions (air exposure at ), achieving up to 71% yield for α-tropolones while enabling scalability for analogs. oxidation has also been utilized for larger-scale preparations of substituted tropolones, offering robust conditions for contexts. Cycloaddition strategies provide efficient routes to tropolone scaffolds, particularly through of fulvenes acting as dienes with activated dienophiles like acetylenedicarboxylate. The Nozaki method from the 1960s exemplifies this, involving the of fulvene with dimethyl acetylenedicarboxylate to form a bicyclic , followed by and to tropolone (overall yield ~40% originally). Improved modern protocols, incorporating optimized and solvent systems, have elevated yields to over 80% for functionalized derivatives, enhancing applicability to complex tropolonoids. Ring expansion reactions from cyclohexadienone precursors represent another key modern tactic, leveraging for migratory insertion at the carbonyl to expand the six-membered ring to seven. This method typically proceeds in 50–60% yield under aprotic conditions, with the resulting readily converted to tropolone via enolization. Photochemical variants, employing UV (λ ~350 nm) in the presence of sensitizers like , offer greater selectivity for substituted systems, avoiding side reactions and achieving comparable efficiency in post-2000 applications. Asymmetric synthesis of chiral tropolones has advanced significantly since 2000, often integrating chiral or catalysts into or ring expansion steps for stereocontrol. For instance, enantioselective [4+3] using chiral oxyallyl cations derived from enable access to tropolone precursors with >90% , followed by . In ring expansion contexts, sulfur ylide-mediated of chiral cyclohexadienones, as in Banwell's synthesis of (-)-, delivers the tropolone moiety in high enantiopurity ( >95%) over multi-step sequences.

Chemical Reactivity

General Reactions

Tropolone exhibits amphoteric reactivity due to its enol-keto tautomerism and the presence of both a hydroxyl group and a carbonyl functionality within a seven-membered aromatic ring, enabling both electrophilic and nucleophilic reactions. The intramolecular between the hydroxyl and carbonyl groups stabilizes the conjugate base, enhancing acidity with a of approximately 6.7, which is lower than that of typical ( ~10) and allows under mildly basic conditions. Electrophilic aromatic substitution preferentially occurs at the electron-rich positions 3 and 5 of the tropolone ring, facilitated by its aromatic character. Halogenation with (Br₂) in acetic acid or aqueous media yields 3-bromotropolone as the primary monobrominated product, with further bromination possible at position 5 under excess reagent to form 3,5-dibromotropolone. Nitration using (HNO₃) in glacial acetic acid introduces a nitro group predominantly at position 5, producing 5-nitrotropolone, reflecting the directing influence of the hydroxyl group and the ring's distribution. Nucleophilic addition targets the electrophilic , often followed by tautomerization to restore . For instance, reaction with Grignard reagents such as methylmagnesium bromide adds to the carbonyl, forming a tertiary intermediate that, upon and tautomerization, yields tropolone alcohols like 2-hydroxy-2-methylcyclohepta-2,4,6-trien-1-one derivatives. The hydroxyl group undergoes typical derivatization reactions akin to enols or . Esterification proceeds readily with acid chlorides, such as , in the presence of a base like , forming tropolone esters (e.g., acetyltropolone) that maintain the ring's planarity. Ether formation occurs via with alkyl halides (e.g., methyl ) under basic conditions, such as with , to generate tropolone alkyl ethers, though care is required to avoid rearrangement due to the acidic proton.

Metal Chelation and Coordination Chemistry

Tropolone functions as a bidentate ligand in coordination chemistry, coordinating to metal ions through its hydroxyl oxygen and the oxygen of the keto tautomer, thereby forming stable five-membered chelate rings. This O,O'-bidentate binding mode is observed across a range of transition and main-group metals, including Fe³⁺, Cu²⁺, and Al³⁺, due to the ligand's ability to delocalize electron density across its seven-membered ring, enhancing chelate stability. The resulting complexes often exhibit high thermodynamic stability, with formation constants (log β) exceeding 10 for many first-row transition metal ions under aqueous conditions at ionic strength 0.1 M and 25°C; for instance, the overall stability constant for the tris(tropolonato)copper(II) complex is approximately log β₃ = 18.5, reflecting strong chelation driven by both enthalpic and entropic contributions from the rigid ring structure. Representative complexes, such as tris(tropolonato)iron(III), adopt an octahedral geometry with the iron center surrounded by six oxygen atoms from three bidentate tropolonate ligands, leading to a propeller-like arrangement that minimizes steric repulsion. In this structure, the –O bond lengths are short, averaging around 1.92 , indicative of strong covalent interactions and transfer from the to the metal, as confirmed by . Similar octahedral coordination is seen in Al³⁺ and other trivalent metal complexes, where the tropolonate ligands impose planarity on the chelate rings, contributing to the overall rigidity and aromatic character of the coordinated system; for ²⁺, the geometry is often distorted square-planar or pseudo-octahedral in bis complexes, with Cu–O bonds near 1.93 . These structural features underscore tropolone's versatility in stabilizing high-oxidation-state metals through π-donation from the enolized . In , tropolone-metal complexes are valued for their intense colors arising from ligand-to-metal charge transfer transitions, enabling sensitive spectrophotometric detection of metal ions. For example, the Fe³⁺-tropolone complex forms a deeply species extractable into organic , allowing quantitative determination of iron at micromolar levels via measurements in the visible range, often around 500–550 nm depending on the and . This application extends to other metals like Cu²⁺ and Al³⁺, where the chelates facilitate separation and preconcentration in solvent extraction protocols, with detection limits as low as 10⁻⁶ M due to the high molar absorptivities (ε > 10⁴ L mol⁻¹ cm⁻¹). Tropolone complexes also exhibit redox properties that support electron transfer processes, particularly in catalytic applications involving variable oxidation states. Vanadium-tropolone species, such as oxovanadium(IV) complexes, participate in oxidation reactions by cycling between V(IV) and V(V) states, facilitating the transfer of oxygen atoms from peroxides to substrates like sulfides or alkenes; for instance, these complexes catalyze sulfoxidation with turnover numbers exceeding 1000 using H₂O₂ as the oxidant, owing to the ligand's ability to stabilize high-valent intermediates and modulate potentials (E₁/₂ ≈ 0.5 V vs. NHE). This activity highlights tropolone's role in mimicking enzymatic active sites, such as those in vanadium haloperoxidases.

Natural Occurrence

In Plants and Fungi

Tropolones occur naturally in various coniferous trees, particularly in the heartwood of species within the family, where they contribute to the wood's durability. For instance, β-thujaplicin, a tropolone derivative, is prominently found in the heartwood of western red cedar (Thuja plicata), with concentrations reaching up to 0.5% by weight in mature trees. Similar compounds are present in other conifers, such as Cupressus lusitanica, where β-thujaplicin levels vary based on environmental and genetic factors, enhancing the wood's resistance to biological degradation. In fungi, tropolones are produced as secondary metabolites, with notable examples including stipitatic acid isolated from Talaromyces stipitatus (formerly Penicillium stipitatum) in 1942. Puberulic acid, another tropolone, was identified from Penicillium puberulum around the same era, marking early discoveries of these compounds in fungal cultures during the . These fungal tropolones often feature hydroxyl substitutions that influence their bioactivity. The biosynthesis of tropolones in plants and fungi involves distinct yet convergent pathways centered on ring expansion mechanisms. In coniferous trees, thujaplicins like β-thujaplicin derive from the , starting with (GPP) and proceeding through cyclization, oxidation, and rearrangement to form the seven-membered tropolone ring. In fungi, such as T. stipitatus, stipitatic acid is synthesized via nonreducing synthases (NR-PKS) that elongate shikimate-derived precursors like 3,5-dihydroxybenzoic acid, followed by and ring expansion to yield the tropolone core. These pathways highlight the evolutionary adaptation of tropolones across eukaryotes, with machinery prominent in fungal production. Ecologically, tropolones in plant heartwood serve critical defensive roles, exhibiting strong antifungal activity against wood-decaying fungi like Poria placenta and , thereby conferring natural durability to coniferous timber. Their insect-repellent properties further deter borers and other pests, as demonstrated by the toxicity of β-thujaplicin and related troponoids to species like the adzuki bean weevil, reducing infestation in forest ecosystems. In fungi, these metabolites likely play similar protective functions in and decaying matter, inhibiting competitor growth.

In Bacteria and Other Microorganisms

Tropolones are produced by diverse bacterial species, notably in the genera Pseudomonas and Streptomyces, where they function primarily in iron acquisition and antimicrobial competition within microbial communities. In Pseudomonas strains such as P. donghuensis and plant-associated isolates like Pseudomonas sp. Ps652, 7-hydroxytropolone (HYS) and 3,7-dihydroxytropolone (3,7-dHT) are key metabolites that chelate iron with high affinity, enabling efficient nutrient scavenging under iron-limited conditions. These compounds exhibit siderophore activity, as demonstrated by their ability to solubilize iron in chrome azurol S assays and promote bacterial growth in iron-deficient media. In species, including S. cyaneofuscatus and S. luteogriseus, tropolones such as 3,7-dihydroxytropolone are synthesized at yields reaching 380 mg/L in optimized cultures, contributing to ecological fitness through broad-spectrum antagonism against phytopathogens like Streptomyces scabies. Other , such as Burkholderia plantarii and marine Phaeobacter inhibens, produce tropolones like basic tropolone and tropodithietic acid, which enhance formation and inhibit competitor growth in polymicrobial environments. These bacterial tropolones play a chelating role in iron uptake, supporting survival in nutrient-scarce niches. Bacterial tropolone biosynthesis typically intercepts the phenylacetic acid (PAA) catabolic pathway, where enzymes like enoyl-CoA hydratases and thioesterases divert intermediates toward tropone ring formation, followed by cytochrome P450-mediated hydroxylations. This contrasts with fungal polyketide routes by integrating primary metabolic degradation with secondary modifications via dedicated gene clusters, such as the tpo locus in Pseudomonas or trl cluster in Streptomyces. In certain marine bacteria like Phaeobacter, hybrid non-ribosomal peptide synthetase-polyketide synthase (NRPS-PKS) systems assemble sulfur-containing tropolones, highlighting pathway diversity across prokaryotes. Production levels in bacterial cultures are generally modest, often in the range of 200–400 mg/L for optimized strains, translating to low cellular concentrations (approximately 1–10 μg/g dry weight), yet they accumulate significantly in biofilms to disrupt in rival microbes, such as inhibiting autoinducer signaling in Vibrio species. This localized enrichment underscores tropolones' role in microbial warfare and niche colonization.

Derivatives and Applications

Natural Derivatives

Hinokitiol, also known as β-thujaplicin, is a naturally occurring tropolone derivative characterized by an isopropyl substituent at the 4-position of the tropolone core, with the molecular formula C<sub>10</sub>H<sub>12</sub>O<sub>2</sub>. It was first isolated in 1935 from the heartwood of the hinoki tree (Chamaecyparis obtusa), a species native to East Asia, where it contributes to the wood's natural resistance against microbial decay. Subsequent isolations have confirmed its presence in related cupressaceous trees, such as Chamaecyparis taiwanensis and Thuja plicata. Colchicine represents a more complex tropolone alkaloid, featuring a seven-membered tropolone ring fused to a trimethoxyphenyl moiety within its overall structure (C<sub>22</sub>H<sub>25</sub>NO<sub>6</sub>). This compound has been isolated from the corms and seeds of Colchicum autumnale, commonly known as autumn crocus, a plant native to Europe and North Africa. Historically extracted for medicinal purposes since ancient times, colchicine's tropolone unit is integral to its recognition as a key natural analog in alkaloid chemistry. Stipitatic acid is a carboxylated tropolone derivative produced by certain fungi, notably Talaromyces stipitatus (previously classified as Penicillium stipitatum), featuring a group at the 3-position adjacent to the hydroxyl. Isolated initially in the 1940s from fungal cultures, it undergoes to yield the parent tropolone structure, highlighting its role as a biosynthetic precursor in fungal pathways. The compound's isolation from T. stipitatus mycelia has provided insights into tropolone formation via oxidative processes from intermediates.

Synthetic Derivatives and Biological Activities

Synthetic tropolone derivatives have been developed through targeted modifications to enhance their chelating properties and biological profiles. For instance, 3-isopropyltropolone, a synthetic analog of the natural (β-thujaplicin), has been developed to enhance properties such as , improving cellular uptake and permeability compared to unsubstituted tropolones, as demonstrated by systematic studies on alkyl-substituted variants. Similarly, halogenated derivatives such as 5-chlorotropolone are synthesized by direct of tropolone, enhancing metal-binding affinity due to the electron-withdrawing effects of the atom, which strengthens coordination in catalytic applications like palladium-mediated cross-coupling reactions. Recent studies (2025) have explored O-derivatization of tropolones to improve their pharmacological profiles. These synthetic tropolones exhibit diverse biological activities, primarily leveraging their metal-chelating capabilities to disrupt essential microbial and cellular processes. In antibacterial applications, analogs demonstrate potent activity against , including methicillin-resistant Staphylococcus aureus (MRSA), with minimum inhibitory concentrations (<sub>80</sub>) in the range of 1–2 μg/mL, outperforming some conventional antibiotics in resistant strains. For anticancer effects, β-thujaplicin and its derivatives inhibit tumor cell proliferation through mechanisms involving DNA damage and induction, with IC<sub>50</sub> values of approximately 1.5–2 μM in certain cell lines and 150–320 μM in cell lines, though specific inhibition remains under investigation in related troponoids. Antiviral properties are evident in derivatives targeting (HCV), where compounds like 3,7-dibromotropolone bind to the NS3 domain, inhibiting with low micromolar potency in subgenomic replicon assays. Beyond , synthetic tropolones find industrial utility in preservation and processes. Copper-tropolone complexes serve as effective wood preservatives, penetrating timber deeply in ammoniacal solutions to protect against fungal decay by chelating essential metals like iron, thus preventing oxidative damage. In , β-isopropyltropolone acts as a selective extractant for transition metals such as iron(III), facilitating efficient separation from aqueous solutions through reversible complexation in organic phases.

References

  1. [1]
    Tropolone | C7H6O2 | CID 10789 - PubChem - NIH
    Tropolone is a cyclic ketone that is cyclohepta-2,4,6-trien-1-one substituted by a hydroxy group at position 2. It is a toxin produced by the agricultural ...
  2. [2]
    Synthesis of Naturally Occurring Tropones and Tropolones - NIH
    This review discusses the types of naturally occurring tropones and tropolones and outlines important methods developed for the synthesis of tropone and ...
  3. [3]
    Tropolone CAS#: 533-75-5 - ChemicalBook
    Melting point, 50-52 °C (lit.) ; Boiling point, 80-84 °C/0.1 mmHg (lit.) ; Density, 1.1483 (rough estimate) ; vapor pressure, 3.4Pa at 25℃ ; refractive index ...Missing: physical | Show results with:physical
  4. [4]
    Tropolone | CAS NO.:533-75-5 - GlpBio
    Rating 5.0 (36) · Free 365-day returnsTropolone is a terpene that has been found in cupressaceous tree bark and has diverse biological activities.
  5. [5]
    Tropolone - an overview | ScienceDirect Topics
    Tropolones are seven-membered aromatic rings that are key components in several bioactive natural products, such as colchicine and stipitatic acid.
  6. [6]
    Tropolone - an overview | ScienceDirect Topics
    Tropones and tropolones refer to non-benzenoid seven-membered aromatic compounds with a carbonyl group (Scheme 1), which are also called troponoids or ...
  7. [7]
    Tropolone natural products - RSC Publishing
    Dec 17, 2018 · This review provides a comprehensive overview of newly discovered natural products containing a tropolonoid motif covering 2008 up to 2018
  8. [8]
    Tropolone | C7H6O2 - ChemSpider
    Tropolone has the molecular formula C7H6O2, and is also known as 2,4,6-Cycloheptatrien-1-one, 2-hydroxy- and 2-Hydroxy-2,4,6-cycloheptatrien-1-one.
  9. [9]
    1222 - Robertson : The Crystal Structure of Cu$ric TroPolone 273 ...
    The interatomic distances given in Fig. 5 are calculated on this basis. The average value of the carbon-carbon bond length in the tropolone ring is 1-40 A.
  10. [10]
  11. [11]
    Tropones and Tropolones | Chemical Reviews - ACS Publications
    Troponoid Atropisomerism: Studies on the Configurational Stability of Tropone-Amide Chiral Axes. Organic Letters 2019, 21 (7) , 2412-2415.
  12. [12]
  13. [13]
    Tropolone CAS 533-75-5 | 108637 - Merck Millipore
    Boiling point, 243.4 °C (1013 hPa) ; Density, 1.321 g/cm3 (20 °C) ; Flash point, 112 °C ; Melting Point, 49 - 52 °C ; Vapor pressure, 0.034 hPa (25 °C).Missing: atmospheric | Show results with:atmospheric
  14. [14]
    None
    ### Summary of UV-Vis Absorption Data for Tropolone
  15. [15]
    Spectroscopy and photophysics of tropolone in condensed media
    The UV–visible absorption, emission and emission excitation spectra of TRN in 3-methylpentane (3MP) at room temperature are shown in Fig. 2, and in a 3MP ...Missing: maxima | Show results with:maxima
  16. [16]
    Evidence for the O−H⋅⋅⋅O=C Resonance‐Assisted Hydrogen ...
    Aug 15, 2024 · ... signal in the 1H NMR spectrum (δOH=15.3–17.0 ppm). The XRD reveals a longer H-bond length for the O−H⋅⋅⋅O=C IMHB in the tropolone ...
  17. [17]
    Tropolone(533-75-5) 1H NMR spectrum - ChemicalBook
    ChemicalBook Provide Tropolone(533-75-5) 1H NMR,IR2,MS,IR3,IR,1H NMR,Raman,ESR,13C NMR,Spectrum.Missing: 13C data
  18. [18]
    13C-Nuclear Magnetic Resonance Spectrum of Tropolone
    The 13C-n.m.r. spectrum of tropolone has been obtained and it clearly shows that there is a rapid tautomeric equilibrium at room temperature.
  19. [19]
    Computational Study Of Keto-Enol Equilibria Of Tropolone In Gas ...
    Aug 6, 2025 · (3) = −6.0 ± 2.4 kcal/mol is obtained from the total energy difference between 2 and 3. The two keto tautomers are less stable than their enol ...
  20. [20]
    Computational study of tautomerism and aromaticity in mono‐ and ...
    Jul 10, 2012 · Keto-enol tautomerism in mono- and dithio-substituted analogs of tropolone was investigated using electronic structure computations.
  21. [21]
    Defining rules of aromaticity: a unified approach to the Hückel, Clar ...
    The approach shows why and how the Hückel rule works, how the Randić conjugated circuits result from the analysis of canonical structures, and also how the Clar ...
  22. [22]
    [PDF] and novel carbo-aromatic compounds
    May 18, 2015 · that tropolone is slightly more aromatic than tropone (TRE = 0.172 β ... criterion (NICS values at the center of the rings: -17.8 ppm for ...<|control11|><|separator|>
  23. [23]
    SYNTHESIS OF TROPOLONE | Journal of the American Chemical ...
    This article is cited by 65 publications. Frank-Gerrit Klärner, Maitland Jones, Jr. and Ronald M. Magid. William von Eggers Doering's Many Research ...Missing: authors | Show results with:authors
  24. [24]
    Synthesis of α-Tropolones through Autoxidation of Dioxole-Fused ...
    Jan 10, 2022 · Herein, we describe the formation of tropolones through the autoxidation of Büchner reaction-derived cycloheptatrienes.Introduction · Results and Discussion · Conclusions · Experimental Section
  25. [25]
    Carbocycloaddition Strategies for Troponoid Synthesis - PMC - NIH
    The following review article provides an overview of strategies for troponoids wherein the 7-membered carbocycle is generated through a cycloaddition reaction.
  26. [26]
    Formation Constants of Metal Complexes of Tropolone and its ...
    Thermodynamic proton—ligand stability constants of tropolone in mixtures of water and dioxane. ... Metal-ligand stability constants of hydroxamic acids.
  27. [27]
    Formation Constants of some Metal-Tropolone Complexes - Nature
    The formation constants of several metal derivatives of tropolone have been determined in order to compare the co-ordination tendency of tropolone with that of ...
  28. [28]
    Stability of Tropolone Chelates of the Bi-and Tervalent Metal Ions
    Metal-Chelate Compounds," Prentice-Hall, Englewood. Cliffs, N. J., (1952) p. 191; J. Lewis and R. G. Wilkins, ed., "Modern Coordination Chemistry," Interscience.
  29. [29]
    The crystal structure of tris(tropolonato)iron(III)(“ferric tropolone”)
    ### Summary of Crystal Structure of Tris(tropolonato)iron(III)
  30. [30]
    Chelate Chemistry. I. Tropolone and Aminotroponimine Derivatives ...
    Reaction of Grignard Reagents with Dianionic Hexacoordinated Silicon Complexes: Organosilicon Compounds from Silica Gel. Angewandte Chemie International ...
  31. [31]
    Tropolone as anionic and neutral ligand in lead(II) and bismuth(III ...
    Jan 5, 2017 · The tropolonate anion (trop−) acts as the bidentate oxygen donor ligand forming a five-membered chelate ring with metal ions.
  32. [32]
    (PDF) Metal complexes with tropolones - Academia.edu
    FAQs. The stability constants for tropolone complexes follow the order UO2 > Cu > Ni > Zn with increasing atomic number, indicating a correlation with ionic ...
  33. [33]
    Chelate Chemistry. V. Metal Chelates Based on Tropolone and Its ...
    Coordination chemistry of Be 2+ ions with chelating oxygen donor ligands: further insights using electrospray mass spectrometry. Zeitschrift für ...
  34. [34]
    Biosynthetic Pathway of β-Thujaplicin in the Cupressus lusitanica ...
    The biosynthesis pathway from glucose to GPP and a skeletal rearrangement of GPP to β-thujaplicin are proposed, based on the NMR analysis of β-thujaplicin ...
  35. [35]
    Classic fungal natural products in the genomic age - RSC Publishing
    Mar 14, 2018 · Raistrick, Birkenshaw and Chambers first reported the isolation of stipitatic acid 4, isolated from Penicillium stipitatum (later renamed as ...
  36. [36]
    In vitro and in vivo antimalarial activity of puberulic acid and its new ...
    Nov 10, 2010 · In general, the natural and synthetic tropolones have been reported to show antibacterial, antifungal, insecticidal, antiviral and antitumor ...
  37. [37]
    Production of ゚-Thujaplicin in Cupressus lusitanica Suspension ...
    The fungal elicitor-in- duced biosynthesis of B-thujaplicin was promoted by the feedings of malate, pyruvate, fumarate, succinate, and acetate. These results ...
  38. [38]
    Chemical mechanisms involved during the biosynthesis of tropolones
    Tropolone biosynthesis involves four routes from precursors, with key steps including ring expansion of an alkylated 6-membered ring.
  39. [39]
    Influence of Tropolone on Poria placenta Wood Degradation - PMC
    In a previous study, we have shown that tropolone strongly inhibits the growth of several wood rot fungi on malt-agar, the antifungal activity of tropolone ...Missing: ecological repellent
  40. [40]
    Repellency and toxicity of troponoid compounds against the adzuki ...
    In this study, we compared the repellency and toxicity of six troponoid compounds – tropiliden, tropone, tropolone, α-thujaplicin, β-thujaplicin (hinokitiol) ...
  41. [41]
    Critical Review on the Use of Extractives of Naturally Durable ... - NIH
    Jan 18, 2024 · This critical review summarizes the relevant literature on naturally durable woods, their extractives, and their potential use as bio-inspired wood protectants.
  42. [42]
    7-Hydroxytropolone produced and utilized as an iron-scavenger by ...
    Aug 19, 2016 · Pseudomonas donghuensis can excrete large quantities of iron chelating substances in iron-restricted environments.
  43. [43]
    Understanding the biosynthesis, metabolic regulation, and anti ...
    Aug 29, 2024 · Tropolone SMs have been documented in plants, fungi, and bacteria and possess an extensive range of biological activities, including ...
  44. [44]
    Biosynthesis of Tropolones in Streptomyces spp. - NIH
    May 31, 2018 · Four of these chemicals, including 1,4,6-cycloheptatriene-1-carboxylic acid, tropone ... Doering WVE, Detert FL. 1951. Cycloheptatrienylium ...
  45. [45]
    Bacterial Tropone Natural Products and Derivatives - PubMed Central
    Tropones play important roles in the terrestrial and marine environment where they act as antibiotics, algaecides, or quorum sensing signals, while their ...
  46. [46]
    Hinokitiol | C10H12O2 | CID 3611 - PubChem - NIH
    Molecular Formula. C10H12O ; Synonyms. hinokitiol; 499-44-5; beta-Thujaplicin; 4-Isopropyltropolone; 2,4,6-Cycloheptatrien-1-one, 2-hydroxy-4-(1-methylethyl)-.
  47. [47]
    Hinokitiol - Molecule of the Month - March 2024 - HTML-only version
    In 1935, he isolated a new compound which has the formula C10H12O2, from the Taiwanese hinoki tree (Chamaecyparis taiwanensis), naming the compound hinokitiol, ...
  48. [48]
    Chemical structure of hinokitiol. - ResearchGate
    It was first isolated from the heartwood of Chamaecyparis taiwanensis, Chamaecyparis obtusa, Calocedrus formosana, and Thuja plicata, which are native to ...Missing: formula | Show results with:formula
  49. [49]
    Colchicine of Colchicum autumnale, A Traditional Anti-Inflammatory ...
    Nov 2, 2022 · Colchicine is a tropolane alkaloid, widely used for its anti-gout and anti-inflammatory activities (4-6). Colchicine by binding to the tubulins ...
  50. [50]
    (PDF) Colchicine: A Review About Chemical Structure and Clinical ...
    Aug 8, 2025 · Colchicine is an alkaloids extracted from the colchicum autumnale plant, and it contains a molecular formula C22H25NO6, and it consists of three rings.
  51. [51]
    Genetic, molecular, and biochemical basis of fungal tropolone ... - NIH
    Stipitatic acid 1 and other fungal tropolones continue to stimulate interest—for example, puberulic acid 2 (5-hydroxy stipitatic acid, also discovered by ...
  52. [52]
    70-year-old chemical mystery solved: How tropolone ... - ScienceDaily
    Apr 17, 2012 · New theoretical models developed to understand tropolones now underpin our understanding of structure and bonding in organic chemistry. However ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  53. [53]
  54. [54]
    Novel α-substituted tropolones promote potent and selective ...
    Derivatization of the tropolone ring with lipophilic substituents may provide an opportunity to enhance intracellular delivery [22] of the compound while ...
  55. [55]
    496. Tropolones. Part III. Halogenated derivatives ... - RSC Publishing
    Halogenated derivatives of tropolone. J. W. Cook, A. R. M. Gibb and R. A. Raphael. Abstract. The first page of this article is displayed as the abstract. 496 ...Missing: catalysts | Show results with:catalysts
  56. [56]
    Tropolone as Neutral Compound and Ligand in Palladium Complexes
    The first approach is based on the palladium-catalysed cross-coupling reaction of halogenated azines with organozinc derivatives of ferrocenes (the Negishi ...<|separator|>
  57. [57]
    Synthesis and Evaluation of Troponoids as a New Class of Antibiotics
    Nov 8, 2018 · The MIC80 was defined as the concentration of an antibacterial agent that inhibited bacteria growth ≥80% compared to untreated control cultures.
  58. [58]
    Hinokitiol Induces DNA Damage and Autophagy followed by Cell ...
    Here, we reveal the novel mechanisms by which hinokitiol exerts its potent anticancer effects on several lung adenocarcinoma cell lines as well as EGFR-TKI- ...
  59. [59]
  60. [60]
    Tropolone and its derivatives as inhibitors of the helicase activity of ...
    In this report, we demonstrate the interaction of the non-structural protein 3 (NS3) of hepatitis C virus (HCV) with alkaloide tropolone (2-hydroxy-2,4 ...Missing: polymerase | Show results with:polymerase
  61. [61]
    US7427316B2 - Tropolone complexes as wood preservatives
    Complexes of tropolone and copper and/or zinc were solubilized in ammoniacal solution providing preservative solutions that fully penetrate wood.Missing: ecological roles repellent
  62. [62]
    Kinetics and mechanism of extraction of iron(III) with .beta.
    Mar 1, 1971 · Kinetics and mechanism of extraction of iron(III) with .beta.-isopropyltropolone ... Kinetic Studies of Solvent Extraction of Metal Complexes.