Acetophenone is the organic compound with the chemical formula C₆H₅C(O)CH₃, representing the simplest aromatic ketone where a phenyl group is attached to the carbonyl carbon of an acetyl moiety.[1] It appears as a colorless to pale yellow viscous liquid with a sweet, pungent, orange-blossom-like odor, possessing a molecular weight of 120.15 g/mol, a boiling point of 202.1 °C, a melting point of 19.4 °C, and a density of 1.028 g/cm³ at 20 °C.[1] Slightly soluble in water (approximately 6,130 mg/L at 25 °C), it is fully miscible with organic solvents and exhibits flammability with a flash point of 77 °C.[1]Primarily produced as a byproduct of cumene oxidation or through the Friedel-Crafts acylation of benzene with acetyl chloride using aluminum chloride as a catalyst, acetophenone serves as a versatile intermediate in organic synthesis.[2] Its major applications include acting as a fragrance ingredient in perfumes, soaps, and detergents due to its floral scent; a synthetic flavoring agent in foods, approved by the FDA; and a catalyst in the polymerization of olefins, as well as a photosensitizer and specialty solvent for resins and plastics.[1][2] In the pharmaceutical and chemical industries, it functions as a precursor for odorants, dyes, pharmaceuticals, and even riot control agents.[2]Environmentally, acetophenone is moderately volatile and biodegradable in water and soil, with atmospheric half-lives of about 6 days via reaction with hydroxyl radicals, though it poses risks as an irritant to eyes and skin and a potential narcotic at high concentrations.[1][2] It occurs naturally in some plants, honey, and essential oils but is also detected in vehicle exhaust and coalcombustion emissions.[1] Health assessments indicate low acute toxicity (oral LD50 in rats: 2.2 g/kg), with no evidence of carcinogenicity or mutagenicity, though exposure limits are set at 10 ppm (TLV) to mitigate irritation and neurological effects.[1][2]
Properties
Physical properties
Acetophenone has the molecular formula C₆H₅C(O)CH₃, corresponding to C₈H₈O, and a molecular weight of 120.15 g/mol.[1]It appears as a colorless to pale yellow viscous liquid that solidifies into crystals at low temperatures, exhibiting a sweet, pungent odor reminiscent of orange blossoms.[1][3]Key physical constants include a boiling point of 202.1 °C at 760 mmHg, a melting point of 19.4 °C, a density of 1.028 g/cm³ at 20 °C, and a refractive index of 1.5372 at 20 °C.[1]
Acetophenone is the simplest aromatic ketone, characterized by a benzene ring directly attached to a carbonyl group, which is further bonded to a methyl group, giving the molecular formula C₈H₈O and systematic name 1-phenylethanone.[1]The key functional group is the ketone carbonyl (C=O), which shows a characteristic infrared stretching absorption at approximately 1685 cm⁻¹, shifted lower due to conjugation with the aromatic ring.[4] In ultraviolet-visible spectroscopy, it exhibits absorption at around 242 nm (ε ≈ 12,000 in ethanol), corresponding to the primary π-π* transition involving the conjugated system.[1]As a ketone, acetophenone displays typical reactivity at the carbonyl carbon, undergoing nucleophilic addition reactions; for instance, it reacts with Grignard reagents such as methylmagnesium bromide to form a tertiary alcohol upon subsequent hydrolysis. The presence of α-hydrogens on the methyl group imparts acidity, facilitating enolization under basic conditions and enabling participation in aldol condensation reactions with other carbonyl compounds.Acetophenone is stable toward hydrolysis under neutral or mildly acidic conditions but is susceptible to oxidative cleavage by strong agents like potassium permanganate (KMnO₄), which cleaves the C-C bond adjacent to the carbonyl, yielding benzoic acid as the primary product.[5]Spectroscopic identification includes ¹H NMR signals with the methyl protons appearing as a singlet at δ 2.6 ppm and the aromatic protons as a multiplet between δ 7.5 and 8.0 ppm.[6] In electron ionization mass spectrometry, the base peak occurs at m/z 105, corresponding to the benzoyl cation (C₆H₅CO⁺) formed by loss of the methyl radical from the molecular ion at m/z 120.[7]
Synthesis
Industrial production
Acetophenone is primarily produced industrially as a byproduct of the cumene oxidation process (Hock process) for phenol and acetone production, contributing over 90% to overall supply.[8] In this process, acetophenone forms during the acid-catalyzed decomposition of cumene hydroperoxide. Alternative routes include the Friedel-Crafts acylation of benzene with either acetyl chloride or acetic anhydride in the presence of aluminum chloride (AlCl3) as a Lewis acid catalyst.[9] This electrophilic aromatic substitution reaction proceeds under anhydrous conditions to prevent hydrolysis of the acylating agent, typically at temperatures between 40°C and 60°C, achieving high yields approaching 95%.[10] The process is conducted in batch or continuous reactors, with careful control of the exothermic reaction to maintain selectivity and minimize polyacylation byproducts.[11]Another industrial route involves the selective oxidation of ethylbenzene using air or oxygen in the liquid phase, often catalyzed by cobalt or manganese salts, to form acetophenone as the main product.[12] This method operates at moderate temperatures (around 100-130°C) and pressures, with acetophenone yields up to 85-90% based on optimized catalyst systems.[13]Global production of acetophenone reached approximately 92,000 metric tons in 2024, with major manufacturing hubs in China, the United States, and Europe due to integrated petrochemical complexes supplying downstream industries like resins and pharmaceuticals.[14] The crude product from either route is purified via vacuum distillation at reduced pressure (typically 10-50 mmHg) to separate acetophenone (boiling point ~202°C at atmospheric pressure) from unreacted benzene, catalyst residues, and high-boiling impurities, achieving purities exceeding 99%.[15]Historically, acetophenone production relied on coal tar-derived benzene before the 1950s, but shifted post-World War II to abundant petrochemical feedstocks like those from catalytic reforming and steam cracking, enabling scalable and cost-effective synthesis aligned with the growth of the global olefins and aromatics industry.[8]
Laboratory preparation
Acetophenone is commonly prepared in the laboratory via the Friedel-Crafts acylation of benzene with acetic anhydride or acetyl chloride in the presence of a Lewis acid catalyst such as anhydrous aluminum chloride (AlCl₃).[16] The reaction proceeds through electrophilic aromatic substitution, where the acylium ion (CH₃CO⁺) generated from the acylating agent attacks the benzene ring. In a typical procedure, benzene (approximately 0.02 mol) is mixed with AlCl₃ (0.075 mol) in a flask equipped with a reflux condenser and dropping funnel, followed by the dropwise addition of acetic anhydride (0.04 mol) over 30 minutes while controlling the exothermic reaction. The mixture is then refluxed for 1 hour, cooled, and quenched by pouring into a mixture of ice and concentrated HCl to hydrolyze the aluminum complex. The organic layer is extracted with diethyl ether, washed with NaOH solution to remove acidic impurities, dried over magnesium sulfate, and purified by vacuum distillation, yielding acetophenone in 70-85% isolated yield.[17] Purity is assessed using thin-layer chromatography (TLC) with a suitable eluent like ethyl acetate/hexane or gas chromatography-mass spectrometry (GC-MS), showing a molecular ion at m/z 120.[17]An alternative laboratory route involves the selective oxidation of styrene using a palladium catalyst in a Wacker-type process, which converts the terminal alkene to the methyl ketone with high efficiency under mild conditions.[18] For example, styrene is oxidized with hydrogen peroxide or oxygen in the presence of PdCl₂ and a co-catalyst like CuCl₂ in an aqueous or ionic liquid medium at room temperature to 60°C, producing acetophenone in yields up to 88% after extraction and distillation.[19] This method avoids harsh Lewis acids and is suitable for small-scale syntheses, with product characterization via GC-MS confirming the structure. Another oxidation approach uses chromic acid to convert ethylbenzene to acetophenone by benzylic oxidation, though it requires careful control to prevent over-oxidation to benzoic acid; ethylbenzene is treated with CrO₃ in sulfuric acid, followed by extraction and distillation, affording 50-70% yields.[20]Modern variants enhance efficiency through techniques like microwave-assisted Friedel-Crafts acylation, where benzene, acetic anhydride, and AlCl₃ are irradiated at 100-150 W for 5-10 minutes, reducing reaction time from hours to minutes while maintaining 80-90% yields and minimizing side products.[21] For substituted analogs, palladium-catalyzed couplings, such as the reaction of aryl halides with acetylzinc reagents or acyl stannanes, provide versatile access with yields of 70-95%, often under Suzuki-like conditions using Pd(PPh₃)₄ and a base.[22] These methods emphasize purity and scalability for research applications.Safety considerations in laboratory preparations include handling AlCl₃ in a dry atmosphere to prevent violent reactions with moisture, which generates HCl gas; reactions should be conducted in a fume hood with appropriate ventilation and acid-neutralizing traps. Oxidants like chromic acid require disposal as hazardous waste due to chromium toxicity.[16]
Applications
Industrial and commercial uses
Acetophenone serves as a key ingredient in the fragrance industry, imparting a sweet, orange blossom-like odor that enhances floral compositions such as hawthorn, mimosa, and heliotrope accords in perfumes, soaps, detergents, creams, and lotions.[1] In the flavor sector, it acts as a flavoring agent in foods, nonalcoholic beverages, and tobacco products, contributing almond- or cherry-like notes at low concentrations, typically up to 28 ppm in categories like soft candy and nonalcoholic beverages; it holds Generally Recognized as Safe (GRAS) status from the FDA and FEMA for these applications.[23][24]As a solvent, acetophenone is utilized in the manufacturing of plastics, resins, and inks due to its ability to dissolve nitrocellulose, cellulose acetate, vinyl resins, and alkyd resins, providing stability and a pleasant odor in formulations.[1][24] Additionally, its derivatives, such as α-hydroxyacetophenones, serve as photoinitiators in UV-curing systems for inks, coatings, and adhesives, enabling rapid polymerization under ultraviolet light.[25]As of 2024, the global acetophenone market was valued at approximately USD 215 million, projected to reach USD 249 million in 2025, with roughly 30% allocated to fragrances and flavors, 34% to solvents, and 35% to the pharmaceutical industry, underscoring its commercial significance in these sectors.[26][27]Historically, a derivative of acetophenone, chloroacetophenone (known as CN), was developed in the late 19th century and widely applied as a tear gas for riot control starting in the early 20th century, including during World War I and subsequent law enforcement uses.[28]
Pharmaceutical and research uses
Acetophenone serves as a key intermediate in the synthesis of various pharmaceuticals, particularly through its conversion to phenethylamine derivatives via reduction, which are building blocks for antihistamines such as diphenhydramine analogs.[9] It is also employed in the preparation of anti-inflammatory agents, including analogs of ketoprofen, where acetophenone derivatives undergo carbonylation and hydrolysis steps to form the arylpropionic acid scaffold essential for non-steroidal anti-inflammatory drugs (NSAIDs).[29] Additionally, acetophenone contributes to the synthesis of analgesics by providing the aromatic ketone moiety that can be functionalized into opioid receptor modulators or other pain-relief compounds.[9]In medicinal chemistry research, acetophenone acts as a versatile reagent for constructing bioactive scaffolds, notably through the Claisen-Schmidt condensation with aldehydes to yield chalcones, which exhibit anti-inflammatory, antimicrobial, and anticancer properties and serve as precursors to flavones with similar therapeutic potential.[30] It is frequently used as a model substrate in asymmetric catalysis studies, particularly for the hydrogenation or reduction of prochiral ketones to chiral alcohols, as demonstrated in Noyori-type ruthenium-catalyzed reactions that achieve high enantioselectivity for (R)-1-phenylethanol, a motif in many chiral pharmaceuticals. Laboratory protocols often employ acetophenone as a standard compound for ketone reductions using sodium borohydride (NaBH4) in methanolic solutions, yielding 1-phenylethanol with near-quantitative efficiency, or for enamine formation with pyrrolidine under acidic conditions to facilitate alkylation reactions in total synthesis.[31]Derivatives of acetophenone have garnered attention for their niche medicinal applications, including potential as antioxidants and antimicrobials; for instance, prenylated acetophenones isolated from plants demonstrate potent activity against Gram-positive bacteria and free radical scavenging comparable to ascorbic acid.[32] Recent 2024 investigations into natural acetophenone analogs, such as paeonol and apocynin, highlight their cardioprotective effects through anti-inflammatory and vasodilatory mechanisms in ischemia-reperfusion models.[33] For pharmaceutical applications, high-purity acetophenone is available in grades meeting United States Pharmacopeia (USP) standards to ensure compliance in drug manufacturing processes.[34]
Natural occurrence
Sources in nature
Acetophenone occurs naturally in the essential oils of plants belonging to over 24 families, where it contributes to floral and fruity aromas.[35] In tea plants (Camellia sinensis), it is a prominent volatile in flowers, comprising up to 31.1% of the essential oil extracted by simultaneous distillation-extraction, and is also present in leaves, imparting sweet, citrus-like notes derived from L-phenylalanine metabolism.[36][37] Trace amounts appear in jasmine (Jasminum sambac) flowers, enhancing the characteristic scent during blooming, and in strawberry (Fragaria spp.) fruits as part of the volatile profile influencing aroma formation.[38][39] In vetiver (Vetiveria zizanioides) roots, it constitutes approximately 1.08% of the essential oil.[40]Fungal and microbial sources include strains of Penicillium and Aspergillus, which produce acetophenone and its derivatives as secondary metabolites; for instance, freshwater isolates of Lindgomyces madisonensis yield acetophenone analogs under epigenetic modulation. Recent studies highlight its production by skin microbiota bacteria, enriched in flavivirus-infected mammalian hosts to modulate insect vector attraction.[41]Trace levels of acetophenone are detected in mammalian urine, such as in red foxes (Vulpes vulpes), where it varies with metabolic factors and contributes to scent profiles (median 4.6 μg/mg creatinine).[42] In insects, it acts in chemical ecology, including as an allelochemical influencing blood-feeding arthropod behavior.[8]Isolation from these sources typically involves steam distillation of plant materials or microbial cultures to obtain essential oils or extracts, followed by gas chromatography-mass spectrometry (GC-MS) for identification and quantification; concentrations are generally below 1% in most plant oils but can reach higher in specific floral extracts.[43][37]In nature, acetophenone serves evolutionary roles as a defense compound, deterring herbivores through emission after plant damage, or as a pollinator attractant in floral volatiles; it can also repel or draw insects depending on context, aiding plant-insect interactions.[44][8]
Biosynthesis
In plants, acetophenone is primarily biosynthesized through the phenylpropanoid pathway, initiating from L-phenylalanine derived from the shikimate pathway. L-Phenylalanine is deaminated by the enzyme phenylalanine ammonia-lyase (PAL, EC 4.3.1.5) to form trans-cinnamic acid, followed by side-chain shortening via a β-oxidative mechanism involving sequential β-hydroxylation, oxidation, and decarboxylation to yield acetophenone. This pathway has been elucidated in species such as Camellia sinensis, where feeding experiments with stable isotope-labeled precursors confirmed the incorporation of ¹³C from L-[ring-¹³C₆]phenylalanine into the aromatic ring of acetophenone, while ¹³C from [1-¹³C]cinnamic acid labeled both the carbonyl and methyl carbons, supporting the β-oxidative shortening of the side chain. Labeling with sodium [1-¹³C]acetate showed minimal incorporation into acetophenone, indicating that the acetyl moiety originates from the phenylpropanoid precursor rather than direct acetylation by acetyl-CoA.Hydroxyacetophenone derivatives, such as picein (4-hydroxyacetophenone β-D-glucoside) and pungenin, follow a similar initial route in conifers like white spruce (Picea glauca), where the acetophenone core is formed prior to glycosylation in phase II metabolism. The enzyme UDP-glycosyltransferase (e.g., PgUGT5b) catalyzes the addition of glucose to the phenolic hydroxyl group, storing the compounds constitutively at levels up to 10% of needle dry mass in resistant genotypes. Upon stress, β-glucosidases like Pgβglu-1 hydrolyze these glucosides, releasing aglycones such as piceol and pungenol for defense. Biosynthesis is upregulated by herbivore attack from spruce budworm (Choristoneura fumiferana), with resistant trees showing 1,000-fold higher β-glucosidase expression and up to 69% increases in aglycone levels post-feeding. In Norwayspruce (Picea abies), p-hydroxyacetophenone and picein accumulation is induced by abiotic stresses, including UV radiation and drought, serving as indicators of physiological strain with ratios shifting toward the aglycone under prolonged exposure.[45]In fungi, acetophenone biosynthesis mirrors the plant β-oxidative pathway, starting from L-phenylalanine via PAL-mediated deamination to cinnamic acid, followed by side-chain degradation to the ketone. This has been observed in the white-rot fungus Bjerkandera adusta, where the process contributes to secondary metabolism, though specific downstream enzymes beyond PAL remain unidentified. Unlike polyketide routes in some fungal aromatic compounds, acetophenone formation here relies on phenylpropanoid-derived precursors rather than malonyl-CoA condensation.[46]Microbial variants in bacteria often produce acetophenone as an intermediate in catabolic pathways rather than dedicated anabolic routes. In styrene-degrading bacteria such as Pseudomonas and Rhodococcus species, styrene is epoxidized by styrene monooxygenase (StyAB) to (S)-styrene oxide, which is isomerized by styrene oxide isomerase to 2-phenylethanol; subsequent oxidation by alcohol dehydrogenase and aldehyde dehydrogenase yields acetophenone en route to benzoic acid. This pathway enables transient accumulation of acetophenone during aerobic styrene catabolism. Additionally, commensal skin bacteria like Bacillus subtilis produce acetophenone as a volatile metabolite, with production enhanced by flavivirus infections (e.g., Zika or dengue), promoting mosquito attraction; however, the precise biosynthetic enzymes in these contexts are not fully characterized. Isotopic studies in bacterial systems are limited, but analogous labeling in related phenylpropanoid pathways confirms precursor origins from shikimate-derived aromatics. Regulation in bacteria may involve stress responses, such as quorum sensing or viral modulation of microbiota, but lacks the UV-specific induction seen in plants.
Biological effects
Pharmacology
Acetophenone and its derivatives exhibit various biological activities, including antioxidant properties through radical scavenging mechanisms. For instance, certain natural acetophenone derivatives, such as those isolated from Euphorbia ebracteolata and Borassus flabellifer, demonstrate potent antioxidant effects in the DPPH assay, with IC50 values of approximately 34.62 μg/mL and 20.02 μM, respectively, indicating their ability to neutralize free radicals by donating hydrogen atoms or electrons.[35]Analgesic effects have been observed in derivatives like acetophenone semicarbazone, which reduce pain responses in acetic acid-induced writhing tests in mice by inhibiting cyclooxygenase (COX) pathways, thereby decreasing prostaglandin-mediated inflammation and nociception.[47]Pharmacokinetic profiles of acetophenone derivatives, such as paeonol (2'-hydroxy-4'-methoxyacetophenone), reveal rapid absorption following oral administration, with good bioavailability estimated around 70% in some formulations, though overall low due to poor aqueous solubility. Metabolism primarily occurs via cytochrome P450 enzymes, including CYP2E1, leading to oxidation products like phenylacetic acid derivatives, while the elimination half-life is short, typically 2-4 hours, facilitating quick clearance but limiting sustained exposure.[8][48]Therapeutic potential of acetophenone derivatives includes applications in anti-inflammatory drugs, exemplified by analogs related to bupivacaine, which modulate local anesthesia and reduce inflammation through COX inhibition. A 2024 review highlights the cardioprotective role of paeonol via activation of the Nrf2/HO-1 pathway, which upregulates antioxidant enzymes like superoxide dismutase and glutathione peroxidase, mitigating oxidative stress and improving cardiac remodeling in myocardial infarction models.[49]At the molecular level, the ketone carbonyl group in acetophenone facilitates hydrogen bonding interactions with biological receptors, such as those involved in enzyme active sites or signaling pathways, enhancing binding affinity and bioactivity. Chiral analogs display enantioselective effects, where specific stereoisomers exhibit differential potency in biological assays, such as in the reduction of prochiral ketones to alcohols with varying enantiomeric excess.[50][51]Clinical data on acetophenone itself remains limited, with no large-scale human trials reported; however, preclinical studies demonstrate efficacy in pain models, including a 40% reduction in writhing responses in acetic acid-induced assays for derivatives like α-(phenylselanyl) acetophenone, supporting their potential as analgesic agents.[52]
Toxicity
Acetophenone exhibits moderate acute toxicity via oral administration, with an LD50 of 2081 mg/kg in rats according to OECD Test Guideline 401.[1] Inhalation exposure shows low acute toxicity, with an LC50 greater than 210 ppm over 8 hours in rats and 1200 mg/m³ (244 ppm) over 4 hours in mice.[1] The compound acts as a mild skin irritant in rabbits based on Draize testing, potentially causing dryness or cracking upon prolonged contact, while it produces severe eye irritation, including corneal injury in rabbits and transient effects in humans.[1]Chronic exposure data for acetophenone in humans are limited, with no established evidence of carcinogenicity (classified as Group D by the EPA). As a solvent, repeated inhalation may pose risks of neurotoxic effects such as central nervous system depression, prompting the OSHA permissible exposure limit (PEL) of 10 ppm as an 8-hour time-weighted average to mitigate irritation and narcotic symptoms.[24][53] Earlier animal studies (as of 2015 EPA assessment) indicated no reproductive or developmental toxicity at doses up to 750 mg/kg/day over 28 days in rats, with no fetal abnormalities observed; however, a 2025 ECHA RAC opinion proposes classifying acetophenone as a reproductive toxicant Category 1B (H360FD: may damage fertility and the unborn child) based on evidence of developmental toxicity in rat studies, pending final regulatory adoption.[2][54]Under EU REACH regulations, acetophenone's harmonised classification (as of November 2025) is Acute Toxicity Category 4 (oral; H302: Harmful if swallowed) and Eye Irritation Category 2 (H319: Causes serious eye irritation). A 2025 RAC opinion proposes adding Reproductive Toxicity Category 1B (H360FD) and Specific Target Organ Toxicity (single exposure) Category 3 (H336: May cause drowsiness or dizziness).[55][54] The 2025 OSHA standards reaffirm the PEL of 10 ppm to address potential solvent-related hazards.[53]In the environment, acetophenone is biodegradable, with reported half-lives of 4.5 days in lake water, 8 days in river water, and 32 days in groundwater under aerobic conditions.[1] It demonstrates low bioaccumulation potential, with a bioconcentration factor (BCF) of approximately 1 and a log Kow of 1.63, indicating minimal partitioning into fatty tissues or soil adsorption (Koc 10-270).[1]To mitigate exposure risks, personal protective equipment such as gloves, goggles, and respiratory protection should be used, alongside adequate ventilation in handling areas. In case of eye contact, immediate flushing with water for at least 15 minutes is recommended, followed by medical attention; skin contact requires washing with soap and water.[1] Given the proposed reproductive toxicity classification, additional precautions may be advised for pregnant workers or in consumer product formulations pending final EU decisions.