Benzophenone
Benzophenone, systematically named diphenylmethanone, is an organic compound with the molecular formula C13H10O, characterized by a carbonyl group bonded to two phenyl rings.[1] It exists as a white, crystalline solid with a melting point of 47.5–49.0 °C and is sparingly soluble in water but miscible with organic solvents. Industrially synthesized primarily through the copper-catalyzed aerial oxidation of diphenylmethane, benzophenone functions as a versatile photoinitiator in UV-curable inks, coatings, and adhesives, initiating free-radical polymerization upon light exposure, and as a UV stabilizer in plastics to mitigate photodegradation. Animal studies demonstrating carcinogenicity have prompted the International Agency for Research on Cancer to classify it as possibly carcinogenic to humans (Group 2B), informing regulatory scrutiny of its presence as a contaminant or intentional additive in consumer goods.[1][2][3]Physical and Chemical Properties
Molecular Structure and Bonding
Benzophenone has the molecular formula C₁₃H₁₀O and the systematic name diphenylmethanone.[1] It features a central carbonyl group (C=O) where the carbon atom is bonded to two phenyl rings via sigma bonds.[4] The carbonyl carbon is sp² hybridized, resulting in a trigonal planar geometry around it with bond angles near 120°.[5] The C=O bond exhibits partial double bond character, typically with a length of about 1.21 Å in similar ketones, while the adjacent C-C bonds to the phenyl ipso carbons are shortened to approximately 1.48 Å due to conjugation with the aromatic π-systems.[5] This conjugation allows for delocalization of the carbonyl π-electrons into the phenyl rings, influencing the electronic properties.[6] Despite the favoring of planarity by π-conjugation, steric repulsion between the ortho hydrogens of the two phenyl rings leads to a non-planar conformation, with dihedral angles between the phenyl planes and the carbonyl plane typically around 30°-40° in the gas phase and varying from 42° to 68° in crystal polymorphs.[7] [8] The phenyl rings themselves remain planar, maintaining their aromatic character with alternating single and double bonds.[1]Physical Characteristics
Benzophenone appears as a white crystalline solid or prismatic crystals at room temperature.[1][9] It possesses a characteristic geranium-like or rose-like odor.[10][11] The compound has a melting point of 48.5 °C and a boiling point of 305 °C, decomposing above 320 °C.[1][12] Its density is 1.11 g/cm³ at 20 °C.[11] Benzophenone exhibits low solubility in water, approximately 140 mg/L at 25 °C, but is readily soluble in organic solvents such as ethanol, ether, and chloroform.[10][12] Benzophenone exists in multiple polymorphic forms, including alpha (monoclinic), beta, and gamma (orthorhombic), with the stable form melting near 48 °C.[1][13]
Spectroscopic and Thermodynamic Data
Benzophenone exhibits characteristic infrared absorption for the carbonyl group at approximately 1663 cm⁻¹, reflecting the conjugation with the adjacent phenyl rings that lowers the stretching frequency compared to unconjugated ketones (typically 1710–1715 cm⁻¹).[14] Aromatic C-H stretching appears around 3000–3100 cm⁻¹, with C=C stretching in the 1450–1600 cm⁻¹ region.[15] In ¹H NMR spectroscopy (CDCl₃ solvent), the aromatic protons resonate as a multiplet between 7.46 and 7.79 ppm, corresponding to the ten equivalent hydrogens on the two phenyl rings, with ortho protons to the carbonyl slightly deshielded.[16] ¹³C NMR shows the carbonyl carbon at about 196.7 ppm, ipso carbons at 137–138 ppm, and other aromatic carbons in the 128–133 ppm range.[17] UV-Vis spectroscopy reveals two main bands: a strong π→π* transition at λ_max ≈ 252 nm (ε ≈ 1.4 × 10⁴ M⁻¹ cm⁻¹) due to the conjugated system, and a weaker n→π* band at λ_max ≈ 366 nm (ε ≈ 100 M⁻¹ cm⁻¹), the latter being responsible for its photochemical activity.[18][19]| Property | Value | Conditions/Source |
|---|---|---|
| Melting point | 47.8 °C | Triple point at 321.03 ± 0.05 K[20][21] |
| Boiling point | 305.4 °C | At standard pressure[20] |
| Heat capacity (solid) | Measured from 5 K to 440 K via adiabatic calorimetry; increases from near 0 J/mol·K at low T to ~200 J/mol·K near melting | Derived thermodynamic functions available[22][21] |
History and Natural Occurrence
Discovery in Nature
Benzophenone, the unsubstituted parent compound (C₆H₅)₂CO, occurs rarely in nature, with confirmed identifications limited to specific plant species. It has been isolated from the rhizomes of Iris adriatica, an endemic iris species native to Croatia, where it was identified among secondary metabolites including xanthones and isoflavonoids in analyses conducted in 2017.[25] This represents one of the few verified natural sources of the unsubstituted form, as most naturally occurring benzophenones feature additional substitutions such as polyisoprenyl or hydroxyl groups.[26] The compound has also been reported in Hemidesmus indicus, known as Indian sarsaparilla, a vine used in traditional Ayurvedic medicine, though specific isolation details and dates for this source remain less documented in primary literature.[26] Earlier reports from the 1970s suggested trace presence in grapes and mangoes, potentially contributing to natural flavors, but these claims conflict with some analyses questioning its natural occurrence in fruits absent substitution.[27] Such detections, when verified, typically involve minute concentrations, underscoring benzophenone's primary association with synthetic production rather than abundant biosynthesis in ecosystems.[2]Early Synthesis and Industrial Adoption
The earliest reported synthesis of benzophenone involved the dry distillation of calcium benzoate, a method documented by Eugène-Melchior Péligot in 1834, yielding the ketone through thermal decomposition of the symmetrical calcium salt of benzoic acid.[28] This technique, later corroborated by Chancel in 1849, represented an initial laboratory-scale preparation exploiting the general reactivity of calcium carboxylates to form diaryl ketones upon heating.[28] Advancements in the late 19th century introduced more efficient routes, notably the Friedel-Crafts acylation of benzene with benzoyl chloride in the presence of anhydrous aluminum chloride, developed by Charles Friedel and James M. Crafts in 1884, which achieved higher yields (up to 66%) and greater control over reaction conditions.[28][2] Alternative early methods included the reaction of benzene with phosgene and aluminum chloride, also pioneered by Friedel and Crafts around 1877–1884, facilitating access to the compound for further derivatization.[28] Industrial adoption of benzophenone emerged in the late 19th century, driven by its utility as a perfume fixative to prevent UV-induced degradation of scents and colors in fragrances and soaps.[10] By the 1920s, commercial production for fragrance applications had scaled to approximately 100,000 pounds annually, reflecting broader integration into the perfume industry via scalable Friedel-Crafts processes.[27] Early 20th-century expansion included its role in synthesizing pharmaceuticals and insecticides, capitalizing on the compound's reactivity as a ketone intermediate.[10]Production and Synthesis
Industrial Synthesis Routes
The primary industrial synthesis of benzophenone employs Friedel–Crafts acylation, reacting benzene with benzoyl chloride in the presence of anhydrous aluminum chloride as a Lewis acid catalyst.[2] This process utilizes excess benzene to suppress polyacylation side products, proceeding exothermically at temperatures typically controlled below 50°C to manage heat and ensure selectivity, with yields reaching approximately 66%.[2] The reaction generates hydrogen chloride as a byproduct, necessitating corrosion-resistant equipment and downstream purification via distillation or crystallization to isolate the product.[29] This route leverages readily available precursors, as benzoyl chloride derives from benzoic acid chlorination, making it economically viable for large-scale production despite the environmental challenges of aluminum chloride sludge disposal.[30] An alternative acylation variant uses phosgene (COCl2) as the acylating agent, where two equivalents of benzene react with one equivalent of phosgene under anhydrous aluminum chloride catalysis to directly form benzophenone and two moles of hydrogen chloride.[31] This method avoids pre-formed benzoyl chloride but introduces handling risks associated with phosgene's toxicity, limiting its adoption compared to the benzoyl chloride process; it is conducted similarly under anhydrous conditions with excess arene to favor mono-substitution.[31] Benzophenone can also be manufactured via catalytic oxidation of diphenylmethane ((C6H5)2CH2), often employing copper catalysts and air as the oxidant, which selectively cleaves the benzylic C–H bond to the ketone without over-oxidation to carboxylic acids under controlled conditions.[2] Diphenylmethane itself is produced from benzene and formaldehyde, providing an integrated pathway; this route offers advantages in reducing Lewis acid waste but requires precise temperature and oxygen control to achieve high selectivity, with yields varying based on catalyst efficiency.[2] Such oxidative processes represent efforts to develop greener alternatives to traditional acylation amid regulatory pressures on inorganic salt generation.[29]Laboratory Preparations
Benzophenone can be synthesized in the laboratory via Friedel-Crafts acylation, in which benzene reacts with benzoyl chloride in the presence of anhydrous aluminum chloride as a Lewis acid catalyst to produce benzophenone and hydrogen chloride.[28] This method, first reported in 1884, proceeds through electrophilic aromatic substitution where the acylium ion (PhCO⁺) generated from benzoyl chloride attacks the benzene ring.[28] The reaction requires careful control to minimize side products like diphenylmethane derivatives, typically employing excess benzene as solvent and maintaining anhydrous conditions to avoid catalyst deactivation.[32] An alternative laboratory route involves the formation and subsequent hydrolysis of benzophenone dichloride ((C₆H₅)₂CCl₂). In the first step, benzene and carbon tetrachloride react in the presence of aluminum chloride at 5–10°C to generate the dichloride intermediate via sequential Friedel-Crafts alkylation and chlorination.[28] The mixture is then hydrolyzed with water under steam distillation conditions, yielding benzophenone with reported efficiencies of 80–89% based on benzene (490–550 g from scaled inputs of 550 mL each of benzene and CCl₄ with 455 g AlCl₃).[28] This procedure emphasizes efficient stirring and temperature control to prevent tar formation and aluminum chloride caking.[28] A historical method, suitable for small-scale laboratory demonstration, entails the dry distillation of calcium benzoate, which decomposes to benzophenone and calcium carbonate at elevated temperatures around 300–400°C.[28] First documented in 1834, this pyrolysis approach produces benzophenone as the primary ketone product from the symmetric calcium salt, though yields are generally lower and require distillation for purification due to potential charring.[28] Modern adaptations may involve mixing with calcium formate or acetate for related aromatic ketones, but pure calcium benzoate suffices for benzophenone.[28]Applications and Uses
UV Absorption and Stabilization
Benzophenone displays characteristic UV absorption bands centered around 250 nm and 330–360 nm in solvents such as ethanol and cyclohexane, with broader absorption extending from approximately 280–340 nm due to its conjugated π-system involving the carbonyl group and phenyl rings.[19][33] This spectral profile enables it to capture UVB and UVA radiation effectively, with molar absorptivity (ε) values indicating strong absorption in the log ε range of 3–4 across these wavelengths as measured in standard spectroscopic databases.[18] In photostabilization applications, benzophenone functions primarily as a UV absorber in polymers and coatings by intercepting harmful UV photons before they induce chain scission or radical formation in the host material.[34] The absorbed energy is dissipated harmlessly as thermal vibration or non-radiative decay, inhibiting photooxidative degradation processes that reduce molecular weight and mechanical integrity.[35][36] This mechanism contrasts with reactive stabilizers like hindered amine light stabilizers (HALS), which scavenge radicals post-formation; benzophenone's filtrative action provides upfront protection, particularly in polystyrene and polyolefin formulations exposed to outdoor weathering.[37][38] Derivatives of benzophenone, such as 2-hydroxy-4-methoxybenzophenone, extend this utility with enhanced solubility and absorption up to 400 nm, but the parent compound itself is incorporated at concentrations of 0.1–2% by weight in PVC films and adhesives to extend service life under solar exposure by factors of 2–5 compared to unstabilized controls.[39][38] Compatibility with non-polar polymers arises from its aromatic structure, though migration or volatilization during processing can limit long-term efficacy in high-temperature applications above 200°C.[40] Empirical studies confirm reduced yellowing and embrittlement in stabilized samples after accelerated UV testing equivalent to 1–2 years of natural exposure.[36]Industrial and Material Science Uses
Benzophenone functions as a Type II photoinitiator in ultraviolet (UV)-curing formulations for inks, coatings, and adhesives, where it absorbs UV radiation (primarily in the 290-400 nm range) to generate free radicals via hydrogen abstraction from co-initiators or substrates, thereby initiating rapid polymerization under ambient conditions.[41][42] This application is prevalent in the printing industry for offset and flexographic inks, as well as in industrial coatings for metal and wood substrates, enabling solvent-free processes that reduce volatile organic compound emissions.[43][44] In polymer and plastics manufacturing, benzophenone acts as an ultraviolet absorber and stabilizer, dissipating absorbed energy as heat to prevent chain scission, discoloration, and loss of mechanical properties in materials exposed to sunlight or artificial light.[45][46] It exhibits strong compatibility with polyolefins, polyvinyl chloride (PVC), styrenics, polycarbonates, acrylic resins, and engineering thermoplastics, typically incorporated at concentrations of 0.1-2% by weight to extend service life in outdoor applications such as greenhouse films, automotive parts, and packaging.[31][47] Its broad absorption spectrum, centered on UVA wavelengths, provides effective photostabilization without significant yellowing of the host polymer.[38][40] Benzophenone derivatives, such as 2-hydroxybenzophenone, are grafted onto polymer backbones via photochemical reactions to enhance intramolecular UV protection in coatings and films, reducing migration and improving long-term durability.[48] In adhesive formulations, it contributes to both initiation in UV-curable systems and stabilization against environmental degradation, supporting applications in electronics encapsulation and pressure-sensitive tapes.[2] These roles leverage benzophenone's thermal stability (melting point 47.5°C, boiling point 305.9°C) and solubility in organic solvents, facilitating uniform dispersion in viscous media.[49]Fragrance and Pharmaceutical Roles
Benzophenone serves as a fixative and modifier in fragrance compositions, imparting a sweet-woody, floral-rosy, and powdery-geranium odor profile that enhances the longevity and depth of scents in perfumes, colognes, and soaps.[50][51] Its low vapor pressure at room temperature allows it to anchor volatile top notes, particularly in rose and geranium accords, while contributing a subtle metallic undertone without dominating the overall bouquet.[52][53] In addition to direct olfactory contributions, benzophenone stabilizes fragrances against photodegradation by absorbing ultraviolet radiation, preventing the loss of color and scent in products exposed to light.[54] This dual role as both an aroma enhancer and protective agent makes it valuable in low-cost formulations, though usage levels are typically trace to avoid overpowering other notes.[55] In pharmaceutical applications, benzophenone functions primarily as a synthetic intermediate rather than an active ingredient in final formulations. It serves as a scaffold for constructing drug frameworks, such as diphenylmethanol derivatives, which are precursors to various therapeutic agents.[31] The compound's structure enables facile modifications, including reduction to benzhydrols via selective hydrogenation, which are employed in the commercial synthesis of pharmaceuticals targeting diverse indications.[56] Benzophenone imines, derived from it, act as ammonia equivalents in Buchwald-Hartwig amination reactions, facilitating the introduction of amine functionalities in drug candidates.[57] Its prevalence in medicinal chemistry stems from natural occurrences in bioactive molecules, underscoring its utility in analog design for potency optimization, though direct incorporation into dosage forms is limited due to potential toxicity concerns.[58] Regulatory scrutiny, including classifications as a possible carcinogen by some agencies, influences its residual presence in synthesized APIs.[2]Chemical Reactivity
Electrophilic and Nucleophilic Reactions
Benzophenone, as an aryl ketone lacking alpha hydrogens, resists enolization but readily undergoes nucleophilic addition at the electrophilic carbonyl carbon. Reduction with sodium borohydride (NaBH₄) in protic solvents like methanol or ethanol proceeds via hydride transfer, yielding diphenylmethanol (benzhydrol) in high yields, typically 80-95% under mild conditions at room temperature.[59] Similarly, organomagnesium reagents such as phenylmagnesium bromide add to the carbonyl, forming triphenylmethanol after hydrolysis, a classic demonstration of 1,2-addition in sterically hindered ketones.[60] These reactions highlight the carbonyl's susceptibility to nucleophiles, though steric bulk from the phenyl groups reduces reactivity compared to aliphatic ketones like acetone.[61] In the presence of strong nucleophilic bases such as sodium amide (NaNH₂), benzophenone participates in the Haller-Bauer cleavage, a C-C bond scission unique to non-enolizable ketones. Treatment with NaNH₂ in refluxing benzene or toluene produces benzamide and benzene, with yields up to 70% reported in early studies.[62] The mechanism involves initial nucleophilic addition of the amide anion to form a tetrahedral intermediate, followed by proton abstraction (likely from solvent or ring) and expulsion of the phenyl carbanion, which protonates to benzene; this pathway is favored over deprotonation due to the absence of alpha hydrogens.[63] Electrophilic aromatic substitution on benzophenone targets the phenyl rings, which are deactivated and meta-directed by the electron-withdrawing carbonyl group. Halogenation, such as bromination, requires harsh conditions like excess Br₂ with Lewis acids (e.g., FeBr₃), yielding meta-bromobenzophenone as the predominant isomer due to the directing effect.[64] Nitration similarly occurs at the meta position under mixed acid conditions, though poly-substitution is minimized by the strong deactivation.[65] These transformations proceed via the standard Wheland intermediate, with the positive charge stabilized at the meta position relative to the carbonyl.[66]Photochemical and Radical Processes
Benzophenone exhibits strong absorption in the ultraviolet region due to its n-π* transition centered around 365 nm, facilitating efficient excitation upon irradiation.[67] Following absorption, the singlet excited state undergoes rapid intersystem crossing to the triplet state with near-unity quantum yield, a process established through decades of spectroscopic studies.[68] The triplet state lifetime in non-polar solvents like benzene is approximately 1–9 microseconds at room temperature, shortening in protic environments due to enhanced quenching.[69] The dominant photochemical pathway involves hydrogen abstraction by the triplet benzophenone from suitable donors, such as alcohols or amines, yielding the benzophenone ketyl radical (Ph₂C•OH) and a complementary carbon-centered radical.[70] This biradical mechanism, confirmed via flash photolysis and electron paramagnetic resonance, exhibits selectivity favoring weaker C–H bonds, including tertiary or allylic hydrogens, with activation energies correlating to bond dissociation energies.[71] [72] In the absence of donors, triplet benzophenone can undergo self-quenching or dimerization to benzopinacol, where two ketyl radicals couple after intra- or intermolecular hydrogen transfer.[73] These radical processes underpin benzophenone's role as a Type II photoinitiator in free radical polymerization, where co-initiators like amines donate hydrogen to generate initiating radicals for acrylate or vinyl monomers.[74] Photolysis rates and polymerization efficiencies depend on the triplet's reactivity, with quantum yields for hydrogen abstraction often exceeding 0.5 in optimal media.[75] Applications extend to holographic lithography and lipid radical studies, where magnetic fields modulate abstraction rates by influencing radical pair recombination.[76] [77]Derivatives and Analogs
Key Commercial Benzophenones
Benzophenone derivatives, particularly those numbered as UV absorbers, dominate commercial applications due to their ability to dissipate UV energy as heat, protecting polymers, coatings, and personal care products from photodegradation. These compounds typically feature hydroxyl or alkoxy substitutions on the aromatic rings, enhancing solubility and spectral coverage. Key examples include benzophenone-3 (2-hydroxy-4-methoxybenzophenone, CAS 131-57-7), which absorbs UVB (290–320 nm) and short UVA rays, making it a staple in over-the-counter sunscreens at concentrations up to 6% by weight as approved by regulatory bodies like the FDA.[78][79] Benzophenone-3 is produced via condensation of 4-methoxyphenol with benzoic acid derivatives, with global production volumes exceeding thousands of metric tons annually for cosmetic use.[80] Benzophenone-4 (2-hydroxy-4-methoxybenzophenone-5-sulfonic acid, CAS 4065-45-6), also known as sulisobenzone, offers water solubility through its sulfonic acid group, enabling incorporation into aqueous formulations such as water-resistant sunscreens and hair sprays at levels up to 5–10%.[81][82] It provides broad-spectrum UV protection (UVB and UVA II) and is often combined with other filters for enhanced efficacy. In industrial contexts, benzophenone-12 (2-hydroxy-4-(1-octyloxycarbonyl)ethoxybenzophenone) serves as a high-performance stabilizer in polyolefins, PVC, and engineering plastics, preventing yellowing and embrittlement during outdoor exposure; commercial grades like Lowilite 22 achieve this via ester linkages that improve compatibility and migration resistance in polyethylenes and polypropylenes.[83]| Derivative | Chemical Name | CAS Number | Primary Commercial Role |
|---|---|---|---|
| Benzophenone-1 | 2,4-Dihydroxybenzophenone | 131-56-6 | UV stabilizer in plastics and coatings; metabolite of other benzophenones in sunscreens[78] |
| Benzophenone-2 | 2,2',4,4'-Tetrahydroxybenzophenone | 131-55-5 | UV absorber in acrylics, polyesters, and cosmetic formulations for extended UVA protection[84] |
| Benzophenone-3 | 2-Hydroxy-4-methoxybenzophenone (oxybenzone) | 131-57-7 | Sunscreen active ingredient; photoinitiator in inks and adhesives[79] |
| Benzophenone-4 | 2-Hydroxy-4-methoxybenzophenone-5-sulfonic acid (sulisobenzone) | 4065-45-6 | Water-soluble UV filter in lotions and shampoos[81] |
| Benzophenone-12 | 2-Hydroxy-4-(1-octyloxycarbonyl)ethoxybenzophenone | 1843-05-6 | Polymer stabilizer in films and fibers for outdoor durability[83] |
Structural Modifications
Substitutions on the phenyl rings of benzophenone, primarily at ortho, meta, or para positions, represent the predominant structural modifications, altering electronic distribution, steric effects, and intermolecular interactions without disrupting the core diaryl ketone framework.[58] These changes typically involve introducing electron-donating groups (e.g., hydroxyl, methoxy, amino) or electron-withdrawing groups (e.g., halogens, nitro), which modulate the carbonyl's electrophilicity and the molecule's conjugation.[85] For example, para-halogenation, as in 4,4'-dichlorobenzophenone or 4-fluorobenzophenone derivatives, stabilizes radical anions formed during electrochemical reduction and enhances lipophilicity for applications in polymer synthesis.[85] [86] Hydroxyl and alkoxy substitutions, often at positions 2 and 4, enable intramolecular hydrogen bonding and extend π-conjugation, shifting UV absorption maxima to longer wavelengths (e.g., 2-hydroxy-4-methoxybenzophenone absorbs at ~290-360 nm).[58] Such modifications reduce phototoxicity compared to unsubstituted benzophenone while preserving photosensitizing potential, as ortho-hydroxyl groups quench triplet states via hydrogen abstraction.[87] Amino-substituted variants, particularly 4-aminobenzophenones, exhibit enhanced binding to biological targets like HIV reverse transcriptase, with IC₅₀ values as low as 0.5 nM against wild-type strains due to improved hydrogen-bonding capabilities.[88] Structure-activity relationship (SAR) analyses indicate that para-substitutions generally yield superior potency in medicinal analogs, with meta-chloro or fluoro groups boosting anti-inflammatory or anti-microtubule effects by optimizing steric fit and electronics (e.g., 4-fluorobenzophenone IC₅₀ = 0.5 nM in HT-29 cells).[58] [86] In natural benzophenones from plants and fungi, additional prenyl or polyprenyl chains at ortho/para sites diversify bioactivity, such as antiparasitic effects (e.g., tenellone A, IC₅₀ = 1.8 μM against Toxoplasma gondii).[58] These modifications are synthesized via Friedel-Crafts acylation of substituted aromatics or cross-coupling reactions, allowing precise control over substituent placement.[58]Human Health Effects
Toxicological Profiles
Benzophenone demonstrates low acute toxicity via oral exposure, with LD50 values in rats reported as 1,900 mg/kg body weight in one study and exceeding 10,000 mg/kg in another.[89] Dermal LD50 in rabbits also indicates low toxicity, exceeding 2,000 mg/kg.[3] No significant irritation or sensitization potential has been consistently observed in standard assays, though it may cause mild eye irritation.[90] In subchronic exposure studies, the liver emerges as the principal target organ across species. In a 14-week dietary administration to F344/N rats and B6C3F1 mice at concentrations up to 20,000 ppm (equivalent to 75–4,200 mg/kg/day), dose-dependent increases in liver weights, centrilobular hypertrophy, and cytoplasmic vacuolization occurred at ≥1,250 ppm in mice and ≥5,000 ppm in rats, accompanied by elevated serum enzymes such as alanine aminotransferase and sorbitol dehydrogenase.[89] Kidney effects, including tubular dilatation and increased weights, were noted in rats at higher doses (≥2,500 ppm males), but not microscopically in mice despite weight increases.[89] Genotoxicity assessments show benzophenone is not mutagenic in Salmonella typhimurium strains (TA98, TA100, TA1535, TA1537) with or without metabolic activation, nor does it induce micronuclei in mouse bone marrow erythrocytes in vivo.[89] The International Agency for Research on Cancer (IARC) classifies benzophenone as possibly carcinogenic to humans (Group 2B), citing sufficient evidence of hepatocarcinogenicity in experimental animals, though mechanistic data link effects to non-genotoxic modes such as cytochrome P450 induction and sustained hepatocyte proliferation rather than direct DNA damage.[91][3] Reproductive and developmental toxicity studies in rats and rabbits reveal no adverse effects on fertility, gestation, or offspring viability at doses below those inducing maternal toxicity; a two-generation rat study confirmed reduced body weights and increased liver enzymes in parental animals and F1/F2 offspring only at maternally toxic levels (≥312 mg/kg/day).[90] No specific endocrine-disrupting effects attributable to benzophenone itself have been established in these mammalian models, distinguishing it from certain hydroxylated derivatives.[90] Overall, toxicity thresholds align with hepatic overload from high systemic exposure, with no-observed-adverse-effect levels (NOAELs) around 100–300 mg/kg/day in repeated-dose studies.[89][90]Exposure Routes and Risk Assessments
Humans are primarily exposed to benzophenone through dermal contact with consumer products such as printing inks, adhesives, and fragrances where it serves as a photoinitiator or fixative, as well as through occupational handling in manufacturing settings involving these materials.[92] Inhalation of vapors or particulates occurs mainly in industrial environments during processing or use of benzophenone-containing formulations, while oral exposure is possible via indirect migration from food packaging or contaminated water, though levels are typically low.[92] [2] Benzophenone is readily absorbed through the skin, gastrointestinal tract, and respiratory system, with in vivo dermal absorption studies in rats demonstrating penetration rates supporting systemic distribution.[92] Risk assessments indicate low acute toxicity, with oral LD50 values exceeding 5000 mg/kg in rats, classifying it as practically non-toxic via single exposures.[89] Chronic effects in animal studies include liver and kidney damage from repeated oral or dermal administration, alongside reproductive toxicity such as reduced fertility in rats at doses around 300 mg/kg/day.[90] The International Agency for Research on Cancer (IARC) classifies benzophenone as Group 2B ("possibly carcinogenic to humans") based on sufficient evidence of liver and lung tumors in rodents, though mechanisms appear non-genotoxic and relevance to human exposure remains uncertain due to species differences and lack of human epidemiological data.[93] [92] Regulatory evaluations, such as Canada's screening assessment, conclude that benzophenone does not pose an unacceptable risk to human health at current environmental or consumer exposure levels, which are estimated below thresholds for adverse effects (e.g., systemic exposure from dermal sources <1 μg/kg bw/day).[92] However, occupational exposures may warrant controls, as no specific permissible exposure limits exist from bodies like OSHA or ACGIH, and prolonged contact could lead to organ accumulation given its metabolism to benzhydrol and excretion via urine.[94] California's Proposition 65 requires warnings for carcinogenicity based on animal data, reflecting precautionary approaches despite limited human evidence.[91] Overall, risks are mitigated by low volatility, limited bioaccumulation in humans, and typical dilute use concentrations.[92]Regulatory Status
Approvals and Limits by Agencies
The U.S. Food and Drug Administration (FDA) amended its food additive regulations on October 9, 2018, to revoke authorization for benzophenone as a synthetic flavoring agent and adjuvant, determining it no longer met safety standards due to insufficient toxicological data supporting its use in food.[95] Benzophenone is prohibited in printing inks for direct food contact materials such as cardboard or paperboard packaging.[96] The U.S. Environmental Protection Agency (EPA) includes benzophenone on the Toxic Substances Control Act (TSCA) Inventory, subjecting it to chemical inventory reporting and risk management requirements, though no agency-specific numerical limits for environmental emissions or water quality criteria were established as of 2023.[97] In the European Union, benzophenone (CAS 119-61-9) is registered under the REACH regulation with a tonnage band exceeding 1,000 tonnes per annum, and it is classified under CLP as acutely toxic to aquatic life (Aquatic Acute 1) and causing serious long-term damage to aquatic environments (Aquatic Chronic 1).[98] For food contact materials, EU Commission Directive 2002/72/EC sets a specific migration limit of 0.6 mg/kg from plastics into food simulants.[2] Canada's Environmental Protection Agency flags benzophenone under the Toxic Substances Control Act equivalent with restrictions on significant new activities, including prohibitions on certain manufacturing volumes exceeding 10 tonnes annually without notification, effective January 30, 2021.[99]| Agency | Key Approval/Status | Specific Limits/Conditions |
|---|---|---|
| FDA (U.S.) | Revoked for food additives (2018); prohibited in food contact inks | None authorized for direct food exposure[95] |
| EPA (U.S.) | TSCA Inventory-listed | No numerical environmental discharge limits specified[97] |
| ECHA (EU) | REACH-registered; CLP-classified as aquatic toxin | Migration limit 0.6 mg/kg in plastic food contact materials[98][2] |
| Health Canada | CEPA-managed with new activity notifications | Manufacturing >10 tonnes/year requires assessment[99] |