Properties of C8H8O3 (Vanillin):
Alternative NamesMethyl vanillin Vanillic aldehyde Elemental composition of C8H8O3
Related compounds
Sample reactions for C8H8O3
Vanillin (C8H8O3): Chemical CompoundScientific Review Article | Chemistry Reference Series
AbstractVanillin, systematically named 4-hydroxy-3-methoxybenzaldehyde, is a phenolic aldehyde with the molecular formula C8H8O3 and a molar mass of 152.15 g/mol. This organic compound serves as the principal organoleptic component of natural vanilla, constituting approximately 2% by dry weight of cured vanilla pods. Vanillin manifests as white to off-white crystalline needles with a characteristic sweet, balsamic aroma. The compound exhibits a melting point of 81 °C and a boiling point of 285 °C. Its chemical structure incorporates three functional groups: aldehyde, hydroxyl, and ether, which collectively govern its reactivity and physical properties. Industrial production predominantly utilizes synthetic routes from petrochemical precursors, with global production exceeding 16,000 metric tons annually. Vanillin finds extensive application as a flavoring agent in food products, fragrances in perfumery, and as a chemical intermediate in pharmaceutical synthesis. IntroductionVanillin represents one of the most significant flavor compounds in the chemical industry, with annual consumption exceeding that of any other single flavoring agent. This phenolic aldehyde belongs to the benzaldehyde derivative class and functions as the primary sensory component responsible for the characteristic aroma and flavor of vanilla. The compound's discovery and isolation in 1858 by Théodore Nicolas Gobley marked a pivotal advancement in flavor chemistry. Subsequent structural elucidation by Ferdinand Tiemann and Wilhelm Haarmann in 1874 enabled the development of synthetic production methods that now dominate commercial supply. Vanillin's molecular architecture features an aromatic benzene ring substituted with hydroxyl, methoxy, and formyl groups at positions 4, 3, and 1 respectively, creating a distinctive electronic configuration that influences its chemical behavior and sensory properties. The compound's significance extends beyond culinary applications to include roles in chemical synthesis, pharmaceutical manufacturing, and materials science. Molecular Structure and BondingMolecular Geometry and Electronic StructureVanillin crystallizes in the monoclinic crystal system with space group P21/c and unit cell parameters a = 12.091 Å, b = 5.585 Å, c = 15.480 Å, and β = 105.67°. The molecular geometry exhibits approximate planarity with slight deviations from perfect coplanarity due to steric interactions between functional groups. The benzene ring maintains standard aromatic characteristics with bond lengths averaging 1.39 Å. VSEPR theory predicts sp2 hybridization for all ring carbon atoms and the aldehyde carbon, with bond angles approximating 120° throughout the molecular framework. The methoxy group adopts a conformation where the methyl group lies nearly coplanar with the aromatic ring, minimizing steric hindrance. Resonance structures demonstrate charge delocalization between the phenolic oxygen and the aromatic system, with significant contribution from quinoidal forms that stabilize the molecular architecture. The electronic structure features highest occupied molecular orbitals localized on the phenolic oxygen and aromatic system, while the lowest unoccupied molecular orbitals concentrate on the aldehyde functionality. Chemical Bonding and Intermolecular ForcesCovalent bonding in vanillin follows typical aromatic patterns with carbon-carbon bond lengths of 1.39 Å and carbon-oxygen bonds measuring 1.36 Å for phenolic C-O, 1.42 Å for methoxy C-O, and 1.21 Å for the aldehyde C=O bond. The molecular dipole moment measures 4.17 D in benzene solution, reflecting significant charge separation due to the electron-donating methoxy and hydroxy groups contrasted with the electron-withdrawing aldehyde functionality. Intermolecular forces include strong hydrogen bonding capacity through both donor (phenolic OH) and acceptor (carbonyl oxygen, ether oxygen) sites. The phenolic hydrogen forms hydrogen bonds with bond energies approximately 25 kJ/mol, while the carbonyl oxygen accepts hydrogen bonds with energies around 17 kJ/mol. Van der Waals forces contribute significantly to crystal packing, with calculated dispersion forces of 8 kJ/mol between aromatic rings. The compound demonstrates moderate polarity with a calculated log P value of 1.21, indicating balanced hydrophilic-lipophilic character. Physical PropertiesPhase Behavior and Thermodynamic PropertiesVanillin presents as white crystalline needles with orthorhombic bipyramidal morphology when recrystallized from aqueous solution. The compound exhibits a sharp melting point at 81.0 °C with enthalpy of fusion measuring 22.7 kJ/mol. Boiling occurs at 285.0 °C under standard atmospheric pressure with enthalpy of vaporization of 55.3 kJ/mol. Sublimation becomes significant above 70 °C with sublimation enthalpy of 48.9 kJ/mol. The density of crystalline vanillin measures 1.056 g/cm3 at 25 °C, while the liquid density at the melting point is 0.987 g/cm3. The refractive index of crystalline vanillin is 1.574 at 589 nm wavelength. Specific heat capacity for solid vanillin is 1.23 J/g·K at 25 °C, increasing to 1.87 J/g·K for the liquid phase at 85 °C. Thermal decomposition commences above 150 °C with gradual liberation of carbon monoxide and formaldehyde. Spectroscopic CharacteristicsInfrared spectroscopy reveals characteristic vibrations at 3325 cm-1 (O-H stretch), 3085 cm-1 (aromatic C-H stretch), 2840 cm-1 (aldehyde C-H stretch), 1665 cm-1 (C=O stretch), 1590 cm-1 and 1510 cm-1 (aromatic C=C stretches), and 1260 cm-1 (C-O stretch). Proton NMR spectroscopy (400 MHz, DMSO-d6) displays signals at δ 9.75 ppm (singlet, 1H, aldehyde), δ 7.40 ppm (multiplet, 3H, aromatic), δ 6.85 ppm (doublet, 1H, aromatic), and δ 3.80 ppm (singlet, 3H, methoxy). Carbon-13 NMR shows resonances at δ 191.2 ppm (aldehyde), δ 152.8 ppm (C4), δ 148.1 ppm (C3), δ 129.5 ppm (C1), δ 124.3 ppm (C6), δ 115.2 ppm (C5), δ 108.7 ppm (C2), and δ 55.6 ppm (methoxy). UV-Vis spectroscopy demonstrates absorption maxima at 230 nm (ε = 12,400 M-1cm-1) and 278 nm (ε = 9,200 M-1cm-1) in ethanol solution. Mass spectrometry exhibits molecular ion peak at m/z 152 with major fragments at m/z 151 [M-H]+, m/z 123 [M-CHO]+, and m/z 93 [M-CH3-CO]+. Chemical Properties and ReactivityReaction Mechanisms and KineticsVanillin demonstrates diverse reactivity patterns centered on its three functional groups. The aldehyde functionality undergoes typical carbonyl reactions including nucleophilic addition, oxidation to carboxylic acid, and reductive amination. Oxidation with silver oxide or potassium permanganate yields vanillic acid with second-order rate constant k = 3.4 × 10-3 M-1s-1 at 25 °C. The phenolic hydroxyl group exhibits acidity and participates in electrophilic aromatic substitution, O-alkylation, and formation of phenolic esters. Ether cleavage with hydroiodic acid proceeds with activation energy 85 kJ/mol to yield catechol derivatives. The aromatic system undergoes electrophilic substitution preferentially at the 5-position, with bromination occurring at rate k = 2.1 M-1s-1 in acetic acid. Hydrogenation of the aromatic ring requires severe conditions (100 atm H2, 150 °C) with platinum catalyst. Vanillin displays stability in air at room temperature but gradually oxidizes upon prolonged exposure to atmospheric oxygen. Acid-Base and Redox PropertiesThe phenolic hydroxyl group confers weak acidity with pKa = 7.78 in water at 25 °C, indicating moderate proton donation capability. Protonation occurs exclusively at the carbonyl oxygen with pKa of the conjugate acid estimated at -2.3. Redox properties include reduction potential E1/2 = -1.23 V vs. SCE for one-electron reduction of the carbonyl group in acetonitrile. Oxidation potentials measure Epa = +1.05 V vs. SCE for phenolic oxidation. The compound functions as a radical scavenger with second-order rate constant for reaction with DPPH radical of 1.8 × 103 M-1s-1. Buffering capacity is negligible except in the pH range 6.8-8.8 where the phenol-phenolate equilibrium operates. Vanillin remains stable in aqueous solution between pH 3 and 8, with decomposition occurring outside this range through aldol condensation and oxidative pathways. Synthesis and Preparation MethodsLaboratory Synthesis RoutesSeveral efficient laboratory syntheses of vanillin have been developed. The Reimer-Tiemann reaction represents the earliest synthetic approach, involving electrophilic formylation of guaiacol under basic conditions. This method proceeds through dichlorocarbene intermediate generated from chloroform and potassium hydroxide, yielding vanillin in 35-40% yield after steam distillation purification. A more efficient laboratory synthesis utilizes Duff reaction conditions, where hexamethylenetetramine serves as the formylating agent for guaiacol in trifluoroacetic acid solvent, achieving 65% yield. Oxidation of isoeugenol with nitrobenzene or potassium persulfate provides vanillin through quinone methide intermediates with 70% isolation yield. Modern laboratory preparations often employ catalytic oxidation of creosol with cobalt(III) acetate in acetic acid solution, producing vanillin in 85% yield with excellent purity. Purification typically involves recrystallization from hot water or toluene-hexane mixtures, yielding analytical-grade material with melting point 81-82 °C. Industrial Production MethodsIndustrial vanillin production employs predominantly petrochemical-based routes due to economic considerations. The Rhodia process, accounting for approximately 85% of global production, utilizes a two-step sequence from guaiacol. First, electrophilic carboxylation occurs using glyoxylic acid in acidic medium to produce vanillylmandelic acid. Subsequent oxidative decarboxylation with copper(II) catalysts at 90-100 °C yields vanillin with overall efficiency of 75%. Lignin-derived vanillin production represents 15% of manufacturing capacity, utilizing sulfite pulping liquor as raw material. Alkaline oxidation of lignosulfonates with oxygen at 150-160 °C under pressure generates vanillin through cleavage of coniferyl alcohol units, with subsequent extraction and purification yielding technical-grade material. Biotechnological production methods employing ferulic acid conversion using Pseudomonas or Amycolatopsis strains have been developed but remain economically noncompetitive, with production costs exceeding $700/kg compared to $15/kg for synthetic vanillin. Analytical Methods and CharacterizationIdentification and QuantificationVanillin identification employs multiple analytical techniques with gas chromatography-mass spectrometry serving as the primary confirmatory method. Reverse-phase high performance liquid chromatography with UV detection at 280 nm provides quantitative analysis with detection limit 0.1 μg/mL and linear range 0.5-500 μg/mL. Capillary electrophoresis with UV detection offers alternative quantification with separation efficiency exceeding 100,000 theoretical plates. Spectrophotometric methods utilizing diazotization or complex formation with iron(III) chloride enable rapid screening with detection limit 5 μg/mL. Thin-layer chromatography on silica gel with toluene-ethyl acetate-formic acid mobile phase (5:4:1) provides Rf value 0.45 with visualization by UV quenching or vanillin-HCl reagent producing pink coloration. Fourier-transform infrared spectroscopy confirms identity through fingerprint region 1500-500 cm-1 with characteristic carbonyl stretch at 1665 cm-1. Purity Assessment and Quality ControlVanillin purity specification typically requires minimum 99.5% by HPLC area normalization. Common impurities include vanillic acid (maximum 0.1%), guaiacol (maximum 0.05%), and acetovanillone (maximum 0.2%). Residual solvent content is controlled with toluene limit 100 ppm and methanol limit 500 ppm. Heavy metal contamination is restricted to less than 10 ppm lead and 5 ppm arsenic. Quality assessment includes melting point determination (range 81-82 °C), specific optical rotation (not more than ±0.05°), and moisture content by Karl Fischer titration (maximum 0.5%). Spectrophotometric purity requires absorbance ratio A278/A230 between 0.74 and 0.76 in ethanol solution. Storage stability testing demonstrates less than 0.5% decomposition after 24 months at room temperature in sealed containers protected from light. Applications and UsesIndustrial and Commercial ApplicationsVanillin serves as the dominant flavor compound in the food industry, with annual consumption exceeding 16,000 metric tons worldwide. The ice cream industry utilizes 60% of production, while chocolate manufacturing accounts for 15% and baked goods 10%. Fragrance applications consume 10% of production in perfumes, soaps, and detergents where it functions as a warm, sweet note in oriental and gourmand fragrance compositions. Pharmaceutical formulations employ vanillin as a flavor masking agent for unpleasant-tasting drugs, particularly in pediatric suspensions and chewable tablets. Chemical manufacturing utilizes vanillin as a precursor to pharmaceuticals including L-dopa, trimethoprim, and vanillylmandelic acid. Emerging applications include use as an antimicrobial agent in food packaging materials and as a corrosion inhibitor for mild steel in acidic environments. The global market value exceeds $300 million annually with growth rate of 3-5% per year. Research Applications and Emerging UsesResearch applications exploit vanillin's chemical functionality for sophisticated synthetic transformations. As a chiral auxiliary, vanillin derivatives facilitate asymmetric synthesis of amino acids and pharmaceutical intermediates. In materials science, vanillin serves as a renewable monomer for epoxy resins and polyesters with enhanced biodegradability. Electrochemical studies utilize vanillin as a model compound for investigating electrode kinetics and adsorption phenomena. Catalytic research employs vanillin as a substrate for developing new oxidation catalysts and hydrogenation systems. Analytical chemistry applications include use as a derivatization agent for amine detection and as a standard for chromatographic method development. Emerging applications investigate vanillin's potential as an antioxidant in polymer stabilization and as a precursor to liquid crystalline materials with mesogenic properties. Patent activity remains strong with 45 new patents filed annually related to vanillin production and applications. Historical Development and DiscoveryThe history of vanillin spans both natural product isolation and synthetic chemistry development. Natural vanilla use dates to pre-Columbian Mesoamerican civilizations where the Totonac people cultivated Vanilla planifolia. European introduction occurred following Spanish conquest of Mexico in the 16th century. Scientific investigation began with vanillin isolation by Théodore Nicolas Gobley in 1858 through ethanol extraction and recrystallization from vanilla pods. Structural elucidation came in 1874 when Ferdinand Tiemann and Wilhelm Haarmann deduced the molecular formula and functional group arrangement. The first industrial synthesis developed concurrently through the Reimer-Tiemann reaction discovered in 1876. Early 20th century production shifted to eugenol from clove oil as starting material. The 1930s witnessed development of lignin-based production from sulfite pulping waste, which dominated production until the 1970s. Petrochemical-based synthesis emerged in the 1970s with development of the glyoxylic acid process. Recent decades have seen advancement of biotechnological approaches using microbial transformation of ferulic acid. ConclusionVanillin represents a chemically significant molecule that bridges natural product chemistry and industrial synthesis. Its distinctive molecular architecture incorporating phenolic, ether, and aldehyde functionalities creates unique physicochemical properties that underlie its widespread applications. The compound's commercial importance continues to drive innovation in production methodologies, particularly toward more sustainable and efficient processes. Fundamental chemical characteristics including acid-base behavior, redox properties, and spectroscopic features are thoroughly characterized and provide textbook examples of substituted benzaldehyde chemistry. Ongoing research explores new applications in materials science, catalysis, and fine chemical synthesis that exploit vanillin's multifunctional character. The compound remains a paradigm for understanding relationships between molecular structure and organoleptic properties in flavor chemistry. Future developments will likely focus on green chemistry approaches to vanillin production and expansion of its applications beyond traditional uses in food and fragrance industries. | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
Chemical Compound Properties DatabaseThis database contains physical properties and alternative names for thousands of chemical compounds. In chemical formula you may use:
The database includes melting points, boiling points, densities, and alternative names collected from various chemical sources. What are compound properties?Chemical compound properties include physical characteristics such as melting point, boiling point, and density, which are important for chemical identification and applications. Alternative names help identify the same compound when referenced by different naming conventions.How to use this tool?Enter a chemical formula (like H2O) or compound name (like water) to look up available properties and alternative names. The tool will search through the database and display any available physical properties and known alternative names for the compound. | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
