Fact-checked by Grok 2 weeks ago

Styrene

Styrene is an with the C₈H₈ and the CH₂=CHC₆H₅, commonly known as vinylbenzene or ethenylbenzene. It appears as a clear, colorless to yellowish oily liquid with a sweet, floral , has a molecular weight of 104.15 g/mol, a of 0.906 g/cm³ at 25°C, and boils at 145.2°C. Styrene is flammable, volatile, and highly reactive due to its attached to a ring, making it a key in . The majority of styrene—approximately 90%—is produced industrially through the catalytic dehydrogenation of , where (C₆H₅CH₂CH₃) is heated with at 550-650°C over catalysts, yielding styrene and in an endothermic, . An alternative route is the oxidation of in the propylene oxide/styrene monomer (PO/SM) , which accounts for about 10% of but is less common overall due to complexity. Global was approximately 29 million metric tons as of 2024, primarily from petroleum-derived feedstocks, supporting its role as a foundational chemical in the . Styrene's primary application is as a for , with around 60% used to produce (PS), a versatile found in , , disposable cups, and consumer goods due to its lightweight, rigid, and insulating properties. It is also copolymerized to form materials like (ABS) for automotive parts and electronics, styrene-butadiene rubber (SBR) for tires, and unsaturated resins for composites in and boating. These applications leverage styrene's ability to form durable, impact-resistant polymers essential to modern manufacturing. Exposure to styrene poses health risks, primarily affecting the ; acute causes irritation to the eyes, , and throat, headaches, , and , while chronic exposure may lead to neurological effects like weakness, depression, and changes. The International Agency for Research on Cancer (IARC) classifies styrene as possibly carcinogenic to s (Group 2B) based on limited evidence in humans and sufficient evidence in animals showing and other tumors, while the U.S. National Toxicology Program (NTP) lists it as reasonably anticipated to be a . evidence remains limited, primarily from occupational exposures. Occupational safety standards limit airborne exposure to 100 over an 8-hour workday, with ongoing research into its environmental persistence and .

Chemical and Physical Properties

Structure and Nomenclature

Styrene is an with the molecular formula C₈H₈. Its molecular structure consists of a ring (C₆H₅-) covalently bonded to a (-CH=CH₂), forming phenylethene, where the carbon-carbon of the vinyl moiety is directly attached to the aromatic ring. This arrangement imparts characteristic properties to the molecule, with the enabling reactivity at the unsaturated site. The International Union of Pure and Applied Chemistry (IUPAC) recommends the systematic name ethenylbenzene for this compound, reflecting the ethenyl (vinyl) substituent on the benzene parent structure; alternatively, vinylbenzene is also accepted in general nomenclature. The common name "styrene" is a retained IUPAC name, alongside synonyms such as styrol, cinnamene, and phenylethylene. The term "styrene" originates from "styrol," derived from storax balsam, a resin from trees of the Liquidambar genus (family Altingiaceae), where the compound was first isolated in the early 19th century. Key identifiers for styrene include the (CAS) 100-42-5, a molecular weight of 104.15 g/, and the (SMILES) notation C=CC1=CC=CC=C1.

Physical Characteristics

Styrene appears as a colorless to pale yellowish oily liquid at standard and pressure, exhibiting a distinctive sweet, balsamic that can become sharp or disagreeable in impure samples due to the presence of aldehydes. This threshold in air is approximately 0.1 , allowing detection at low concentrations. Key thermodynamic properties include a of 0.9059 g/cm³ at 20 °C, a of -30.6 °C, and a of 145.2 °C at 1 atm. The is 31 °C (closed cup), underscoring its classification as a with potential for vapor-air mixtures to ignite under ambient conditions. Styrene's measures 5 mmHg at 20 °C, reflecting moderate that facilitates and contributes to its handling requirements in settings. The heat of vaporization is 38.2 kJ/mol, a value relevant for processes involving changes such as or storage. In terms of solubility, styrene shows limited affinity for , with a solubility of 0.03 g/100 mL at 20 °C, attributable in part to its nonpolar structure. However, it is fully miscible with common organic solvents, including , , and acetone, enabling its use in solvent-based applications. The refractive index, an optical property indicative of its interaction with , is 1.5469 (n_D^{20}), consistent with its aromatic and alkenic .
PropertyValueConditionsSource
Density0.9059 g/cm³20 °CO'Neil et al. (2001) via ATSDR
-30.6 °C-SIRC (2020)
145.2 °C1 SIRC (2020)
31 °CClosed cup SDS (2024)
Water solubility0.03 g/100 mL20 °CO'Neil et al. (2001) via ATSDR
(n_D)1.546920 °CO'Neil et al. (2001) via ATSDR
5 mmHg20 °CSIRC (2020)
Heat of vaporization38.2 kJ/molCAMEO Chemicals (NOAA)

Chemical Reactivity

Styrene exhibits high reactivity primarily due to the conjugated π-system between the group's and the ring, which stabilizes reactive intermediates such as carbocations and radicals formed during and free radical processes. This conjugation lowers the energy barrier for addition reactions at the exocyclic , making styrene more reactive than isolated alkenes toward electrophiles and radicals. Key reactions of styrene include electrophilic addition of halogens, such as , across the vinyl double bond to form 1,2-dibromo-1-phenylethane (C₆H₅CHBrCH₂Br). Hydrogenation of the double bond yields (C₆H₅CH₂CH₃), a process that is rapid and well-studied under catalytic conditions. Oxidation can produce (C₆H₅CHO) or (C₆H₅COOH), depending on reaction conditions and oxidants, with selective pathways favoring the former under controlled mild oxidation. Styrene has a strong tendency to undergo free , initiated by from peroxides or light, followed by through successive of units to the growing chain . The involves ( + ), (chain + → extended chain ), and eventual termination, without delving into detailed . This process forms , represented by the equation: n \ce{C6H5CH=CH2} \rightarrow \left[ \ce{-CH2-CH(C6H5)-} \right]_n Under inert conditions, styrene remains stable, but it is sensitive to , , and oxygen, which can trigger unwanted by generating initiating radicals. Inhibitors like tert-butylcatechol are added to commercial styrene to enhance stability during storage and handling.

Natural Occurrence and Sources

In Nature

Styrene occurs naturally in trace amounts in certain plants and foods. The decarboxylation of trans-cinnamic acid, a key intermediate derived from phenylalanine via the phenylpropanoid biosynthetic pathway, has been demonstrated in plant cell cultures, yielding styrene at room temperature without additional catalysts. Notable examples include cinnamon (Cinnamomum cassia), where styrene forms as a minor volatile component often associated with microbial activity, and coffee beans (Coffea spp.), in which it is present at low levels as a natural constituent. In cinnamon essential oils, concentrations can reach up to 0.02% (average 0.020%, range 0.016–0.024%), contributing to the aroma profile, though levels vary by origin and processing. Microbial production of styrene also occurs in natural settings, primarily through metabolic pathways in fungi and that convert precursors like or related s. Fungi such as species, including P. camemberti and P. expansum, biosynthesize styrene during growth on organic substrates like tree or dairy media, with rates up to 52.5 μg/h per 10 g of . This involves enzymatic steps similar to those in , rather than oxidation pathways. , including certain environmental strains, can similarly generate styrene via , often as a byproduct of breakdown or synthesis in and contexts. As a volatile compound, styrene is detected in tobacco smoke, where it arises from the thermal degradation of natural plant components during combustion. It also appears in processed foods like roasted nuts, such as cashew kernels (), with higher concentrations in roasted samples compared to raw ones due to heat-induced formation from or precursors. Environmentally, styrene is present in petroleum-related deposits, notably as a trace component in derived from coal processing.

Commercial Precursors

The primary commercial precursor for styrene production is (C₆H₅CH₂CH₃), which is synthesized through the of with . This process typically employs zeolite-based catalysts in either vapor-phase or liquid-phase reactors to achieve high selectivity toward ethylbenzene, minimizing byproducts such as diethylbenzene. Ethylbenzene accounts for over 99% of global styrene feedstock, serving as the key intermediate in the dominant dehydrogenation route. Alternative feedstocks include , a from that contains styrene directly and can be extracted via processes like selective and distillation. Emerging routes utilize , often coupled with , for side-chain to produce styrene in a single step, offering potential cost advantages due to the lower price of these inputs compared to and . These alternatives represent a smaller share of production but are gaining interest for and feedstock flexibility. Recent developments include exploration of renewable precursors, such as bio-based derived from , to reduce reliance on feedstocks. Benzene, a core component for ethylbenzene, is primarily sourced from catalytic reforming of petroleum naphtha, which converts low-octane hydrocarbons into high-aromatic reformate. Ethylene, the alkylating agent, is derived mainly from steam cracking of hydrocarbons like ethane, propane, or naphtha in high-temperature furnaces. Global styrene production, reliant on these precursors, reached approximately 42 million tonnes in 2024.

Historical Development

Discovery and Early Uses

Styrene, a colorless oily liquid, was first isolated in pure form in 1839 by apothecary through the distillation of storax, a resin obtained from the bark of the tree. named the volatile substance "styrol" and noted its tendency to thicken into a gelatinous, rubber-like material upon exposure to air, light, and heat, though he did not fully understand the transformation process. Shortly thereafter, in 1845, English chemist John Blyth and German chemist August Wilhelm von Hofmann independently observed that styrol underwent a similar solidification when exposed to , terming the resulting product "metastyrol" and recognizing it as a distinct polymeric form, an early observation of photopolymerization. Further studies in the mid-19th century advanced the chemical characterization of the compound; in 1866, French chemist determined its empirical formula as C₈H₈ and described the conversion of styrol to metastyrol as a reaction, providing one of the first explicit recognitions of such a process in . During the , styrene's applications were derived primarily from its natural occurrence in , which had long been employed in medicinal preparations, perfumes, and . The balsam, containing styrene as a minor component, found use in small-scale formulations for varnishes and resins, particularly for imparting luster to wood and metal surfaces, as well as in for experiments. These pre-industrial uses remained limited to artisanal and laboratory contexts, with no large-scale or commercial production emerging until .

Industrial Scale-Up

The commercialization of styrene production began in the 1930s, driven by the need for synthetic rubber alternatives in Germany. IG Farbenindustrie AG pioneered the first large-scale synthesis through the catalytic dehydrogenation of ethylbenzene, establishing facilities that supplied styrene for Buna-S rubber production by 1936. IG Farben began commercial production of polystyrene in 1930, marking the first large-scale use of polymerized styrene. This process marked a shift from laboratory-scale experiments to industrial viability, with initial capacities supporting the manufacture of tires and other wartime materials. The demand surged during World War II, as natural rubber shortages prompted massive expansion; IG Farben's output was integral to Germany's synthetic rubber program, producing Buna-S (styrene-butadiene rubber) on a scale that exceeded 100,000 tons annually by the war's end. Post-war reconstruction and economic recovery fueled a production boom, particularly in the United States during the . U.S. facilities, initially ramped up for wartime needs, transitioned to peacetime applications like and expanded , with annual output surpassing 500,000 tons by the mid- to meet growing consumer demand. Key innovations included early on processes, such as Ivan Ostromislensky's 1927 U.S. No. 1,643,673 for styrene , which laid foundational techniques for commercial production adopted in and scaled . This era saw widespread adoption of dehydrogenation technologies, enabling efficient scaling and establishing styrene as a cornerstone of the . By the 2020s, styrene production has evolved toward greater amid environmental pressures and market growth. Innovations like the Lummus/UOP Process and advanced catalysts have reduced energy consumption by up to 82% in some conversions, minimizing CO₂ emissions while maintaining high yields from ethylbenzene dehydrogenation. As of 2024, global capacity exceeded 42 million tons per year, with dominating at over 65% of total output; alone accounted for more than 55% of this capacity, driven by integrated complexes and export-oriented expansion.

Production Methods

Industrial Processes

The dominant industrial method for styrene production is the catalytic dehydrogenation of , which accounts for over 90% of global output. In this process, is vaporized and mixed with before being passed over an iron oxide-based catalyst, often promoted with , at temperatures of 600–650°C and . The primary reaction is endothermic and reversible: \ce{C6H5CH2CH3 ⇌ C6H5CH=CH2 + H2} This equilibrium-limited reaction achieves single-pass conversions of approximately 60–65%, with steam serving to lower the of reactants, shift the toward products, and suppress formation on the catalyst. The reaction mixture is then cooled, and products are separated via , yielding styrene with selectivities exceeding 90%. is produced as a valuable , while minor side products such as and are recycled to upstream ethylbenzene synthesis. An alternative commercial route is the propylene oxide/styrene monomer (POSM) process, which co-produces and . is first oxidized with molecular oxygen to form ethylbenzene , which is then cleaved in the presence of a soluble catalyst and reacted with . This yields and in a 2:1 molar ratio, with overall yields around 90–95%. The operates at milder conditions (100–150°C) compared to dehydrogenation and avoids byproduct, but it requires careful handling of the intermediate to prevent . POSM accounts for about 10% of global production and is favored in integrated facilities where demand aligns with output. Styrene can also be extracted as a byproduct from (pygas), a C5–C9 from operations containing 0.5–1% styrene. Recovery involves initial to isolate the C8 aromatic cut, followed by using polar solvents like N-methylpyrrolidone to separate styrene from and xylenes. This method contributes a small (less than 5%) to total supply but enhances overall by valorizing pygas streams. Emerging routes aim to diversify feedstocks and improve sustainability; one involves side-chain of with over metal or catalysts (e.g., Cs-modified X ) at 400–500°C, producing styrene directly with selectivities up to 90%. Another is the oxidative coupling of and , leveraging ethane's abundance from , though it remains at pilot scale with catalysts like supported oxides achieving modest yields (20–30%). Additionally, bio-based routes using renewable feedstocks such as bio-ethanol-derived and bio- are gaining traction, with pilot-scale demonstrations achieving comparable yields to conventional methods as of 2025. Process economics for styrene production are dominated by energy costs, particularly in the dehydrogenation route, which consumes 10–15 per of styrene due to high-temperature and generation (-to-ethylbenzene ratio of 10–15:1). Byproduct , such as to ethylbenzene , recovers 95–98% of aromatics and offsets costs, but the emits CO2 equivalent to 1.5–2 per of styrene from fuel . enhancements in the include reactors with Pd-based hydrogen-permeable membranes integrated into dehydrogenation, which remove to boost conversions to 80–90%, reduce usage by 20–30%, and lower by enabling lower operating temperatures. These innovations, demonstrated in pilot plants, also minimize buildup and extend catalyst life, aligning with decarbonization goals.

Laboratory Preparation

One common laboratory method for preparing styrene involves the acid-catalyzed of , a process that eliminates water to form the double bond. This classic approach typically employs concentrated as the catalyst at temperatures around 180°C, proceeding via an E1 mechanism where the alcohol is protonated, followed by loss of water and to yield styrene. The is represented by : \text{C}_6\text{H}_5\text{CH(OH)CH}_3 \xrightarrow{\text{H}_2\text{SO}_4, 180^\circ\text{C}} \text{C}_6\text{H}_5\text{CH=CH}_2 + \text{H}_2\text{O} Yields in laboratory settings often reach 70-90%, depending on reaction conditions and catalyst concentration, though side products like diphenylethane can form if temperatures exceed 200°C. Another versatile synthetic route utilizes the , which couples with methylenetriphenylphosphorane (Ph₃P=CH₂), a non-stabilized . This method provides excellent stereocontrol and is particularly useful for substituted styrenes, with the ylide generated from methyltriphenylphosphonium bromide and a strong base like . The reaction proceeds via oxaphosphetane intermediate collapse, delivering styrene in yields typically exceeding 80% after aqueous workup and . It is advantageous in laboratory scale for its mild conditions ( to in or THF) and avoidance of harsh acids. Additional laboratory routes include the palladium-catalyzed , where couples with a such as vinyl bromide in the presence of a Pd(0) precatalyst like Pd(OAc)₂, a (e.g., PPh₃), and a base (e.g., Et₃N) to form styrene via migratory insertion and β-hydride elimination. This C-H activation variant enables direct arene vinylation under relatively mild conditions (80-120°C in DMF), with yields of 60-85% reported for unsubstituted cases, though can be an issue without directing groups. Decarboxylative elimination from derivatives offers another pathway, involving heating trans- with a catalyst in (PEG) or deep eutectic solvents to extrude CO₂ and generate styrene. This metal-mediated process operates at 150-200°C, achieving 70-95% yields for electron-rich derivatives, and is noted for its use of biorenewable precursors. Regardless of the synthetic route, laboratory-purified styrene is obtained via at reduced pressure (e.g., 10-20 mmHg) to lower the to 40-50°C and minimize thermal , often after adding inhibitors like tert-butylcatechol to stabilize the . This step removes unreacted starting materials, byproducts, and polymerization inhibitors, yielding colorless styrene with purity >95% suitable for small-scale reactions.

Applications and Uses

Polymerization Reactions

Styrene, a vinyl monomer, undergoes chain-growth polymerization primarily through free radical, anionic, cationic, and coordination mechanisms to form polystyrene and its copolymers. These reactions exploit the reactivity of the vinyl group in styrene, enabling the synthesis of polymers with tailored microstructures and properties. Free radical polymerization is the most common industrial method for producing polystyrene from styrene. Initiation typically occurs via the thermal decomposition of peroxides, such as benzoyl peroxide, which generates primary radicals that add to the styrene monomer to form a chain-initiating radical. Propagation proceeds through successive addition of styrene monomers to the growing radical chain, while termination involves combination or disproportionation of two radicals. The kinetics follow the standard free radical mechanism, with the propagation rate given by: \text{rate} = k_p [M] [R^\bullet] where k_p is the propagation rate constant, [M] is the monomer concentration, and [R•] is the concentration of propagating radicals. Anionic polymerization of styrene yields atactic polystyrene with narrow molecular weight distributions and is particularly suited for producing well-defined architectures. This process uses strong bases like n-butyllithium (n-BuLi) as initiators in polar solvents such as tetrahydrofuran (THF), where the initiator deprotonates or adds to styrene to form a carbanionic chain end that propagates by nucleophilic addition. The "living" nature of anionic polymerization, characterized by the absence of termination or transfer reactions, allows for the sequential addition of different monomers to synthesize block copolymers, such as polystyrene-block-polybutadiene. Cationic polymerization of styrene employs Lewis acids like BF₃ or AlCl₃ to generate carbocations from the , leading to growth via , though it is less common due to challenges in controlling molecular weight. Coordination polymerization, notably using Ziegler-Natta catalysts such as CpTiCl₃/MAO systems, enables the synthesis of syndiotactic with high stereoregularity. These catalysts coordinate to the styrene , facilitating stereospecific insertion into the metal-carbon . Thermal of styrene occurs without added initiators at elevated temperatures of 100–200°C, initiated by spontaneous formation from the , and is used in bulk processes despite lower control over polydispersity. The primary product of styrene is (PS), a glassy with a temperature (T_g) of approximately 100°C, imparting rigidity and transparency suitable for applications in foam insulation and packaging materials. Atactic PS from free radical or anionic routes is amorphous, while syndiotactic PS from coordination methods exhibits higher crystallinity and melting point. Copolymers such as styrene- rubber (SBR), containing about 25% styrene and 75% , are produced via free radical and provide enhanced elasticity for .

Other Industrial Applications

Beyond its dominant role in polymerization, finds application as a in various industrial formulations due to its effective solvency for resins and polymers, particularly in paints, coatings, and inks, where it aids in dissolving and dispersing components for improved application properties and durability. Styrene also serves as a key intermediate in the synthesis of certain pharmaceuticals and agrochemicals, often through derivatives like , which undergoes reactions such as ring-opening to form chiral building blocks for active compounds. In adhesive production, styrene contributes to latex (SBL), an that provides strong bonding, water resistance, and flexibility in applications like , , and construction adhesives. Additionally, styrene acts as a reactive and cross-linking agent in unsaturated polyester resins (UPR), facilitating the curing process to form durable composites used in boat hulls, automotive parts, and building materials by copolymerizing with the resin's unsaturated sites.

Health, Safety, and Environmental Concerns

Health Effects and Exposure

Styrene exposure primarily occurs through in occupational settings, such as during the production of plastics and resins, where workers may encounter airborne vapors due to its volatility. The (OSHA) has established a (PEL) of 100 as an 8-hour time-weighted average (TWA), with a ceiling limit of 200 and a peak limit of 600 for no more than 5 minutes in any 3-hour period, to protect against adverse health effects. Dermal absorption through the skin is possible but occurs at a low rate, averaging approximately 1 μg/cm² per minute, and is generally considered minimal compared to unless prolonged contact with the liquid form happens. Oral is rare and typically limited to accidental or environmental scenarios. Acute to styrene via can cause to the eyes, , , and upper , manifesting as burning sensations, tearing, and coughing at concentrations as low as 100 . At higher levels exceeding 100 , central (CNS) depression may occur, leading to symptoms such as , , , , and impaired coordination; these effects are reversible upon cessation of but highlight the need for immediate and protective measures in workplaces. Gastrointestinal disturbances, including , have also been reported in cases of high acute or . Chronic occupational exposure to styrene at levels below the OSHA PEL but above background has been associated with neurotoxic effects, including cognitive deficits, reduced reaction times, memory impairment, and color vision disturbances, as observed in longitudinal studies of reinforced plastics workers. Ototoxicity is another key concern, with evidence of hearing loss, particularly in the high-frequency range, linked to cumulative exposure and potentiated by concurrent noise; animal models confirm cochlear damage starting from the middle turn of the organ of Corti. Regarding carcinogenicity, the International Agency for Research on Cancer (IARC) classifies styrene as Group 2A (probably carcinogenic to humans) based on limited evidence in humans and sufficient evidence in experimental animals, while its primary metabolite, styrene-7,8-oxide, is classified as Group 2A (probably carcinogenic to humans) due to its genotoxic epoxide structure that forms DNA adducts, potentially contributing to links with leukemia and lymphoma in exposed cohorts. In December 2024, the U.S. Environmental Protection Agency (EPA) initiated prioritization of styrene as a high-priority substance for risk evaluation under the Toxic Substances Control Act (TSCA), recognizing it as a probable human carcinogen. In the 2020s, the U.S. Environmental Protection Agency's (EPA) Integrated Risk Information System () assessment, updated in 2020, evaluates styrene's potential for reproductive and developmental toxicity, noting inconclusive but suggestive evidence from human epidemiological studies of increased risks such as spontaneous abortions and , alongside clearer effects in animal models like reduced fetal weight and skeletal variations at maternally toxic doses. of styrene exposure commonly involves measuring urinary metabolites, particularly (MA) and phenylglyoxylic acid (PGA), which reflect recent uptake; levels of MA + PGA above 3,000 mg/g correspond to workplace exposures around 100 , providing a reliable, non-invasive tool for assessing compliance and health risks in exposed populations.

Environmental Concerns

Styrene is volatile and primarily partitions to air, where it degrades rapidly through reaction with hydroxyl radicals, with an atmospheric half-life of about 1-2 days. In water and soil, it undergoes biodegradation by microorganisms, with half-lives ranging from days to weeks under aerobic conditions, though slower in anaerobic environments. Styrene has low bioaccumulation potential, with bioconcentration factors (BCF) below 100 in aquatic organisms, indicating it does not persist or accumulate significantly in food chains. Ecotoxicity is generally low at environmental concentrations, but spills can cause acute effects to aquatic life, such as reduced mobility in fish at levels above 10 mg/L. Ongoing monitoring focuses on releases from industrial sites and waste management to prevent localized impacts.

Polymerization Hazards and Mitigation

Styrene monomer is prone to autopolymerization, a free radical that can initiate spontaneously under certain conditions, leading to an capable of . This reaction generates significant heat, approximately 70 kJ/mol, causing temperatures to rise rapidly—potentially exceeding 200°C in uncontrolled scenarios—and resulting in pressure buildup from and formation, which may rupture storage vessels. Triggers include elevated temperatures above °C, exposure to light, and contaminants such as peroxides, metal salts, strong acids, or oxygen. Historical incidents underscore the severity of these hazards, with runaways contributing to over 33% of 30 analyzed thermal incidents in specific units from 1988 to 2013, including explosions and fires. In the and beyond, several cases involved tank ruptures due to uncontrolled , such as those linked to inadequate cooling or depletion during storage, highlighting the risk in bulk handling. To mitigate these risks, styrene is stabilized with polymerization inhibitors, most commonly (TBC) added at concentrations of 10-15 , which interrupts the free radical chain by scavenging reactive species. Storage practices include maintaining temperatures below 25°C through or cooling systems to slow reaction , and employing blanketing with controlled oxygen levels (6-10% v/v) to prevent oxidative initiation while avoiding fully inert atmospheres that could exacerbate other hazards. , such as rapid cooling via sprays or venting systems designed for , are essential to dissipate heat during incipient runaways. Regulatory frameworks classify styrene as a UN 2055 hazardous , a Class 3 (packing group III), with specific labeling for hazards to alert handlers of the stabilized nature and potential for violent reaction if uninhibited. Guidelines from the emphasize regular monitoring of levels, with spectroscopic methods like UV-Vis or Raman enabling real-time detection of TBC depletion down to 0.1 , supporting proactive adjustments in industrial settings.

References

  1. [1]
    Styrene | C6H5CHCH2 | CID 7501 - PubChem - NIH
    Styrene is used predominately in the production of polystyrene plastics and resins. Styrene is also used as an intermediate in the synthesis of materials used ...
  2. [2]
    Styrene | 100-42-5 - ChemicalBook
    Sep 25, 2025 · Styrene monomer has been extensively used in the manufacture of chemical intermediates, fi lling components, plastics, resins, and stabilizing ...Chemical properties · Polymerization Reaction · Production
  3. [3]
    [PDF] Styrene - U.S. Environmental Protection Agency
    Styrene is primarily used in the production of polystyrene plastics and resins. Acute (short-term) exposure to styrene in humans results in mucous membrane ...<|separator|>
  4. [4]
    Styrene Production Process - Vaisala
    Styrene is mostly manufactured by dehydrogenation. In the dehydrogenation process, fresh and recycled ethylbenzene is mixed with steam and heated.
  5. [5]
    Styrene Production Process: Key Routes and Industrial Efficiency
    Oct 3, 2025 · The most dominant route for styrene production is the catalytic dehydrogenation of ethylbenzene, widely used across North America, Europe, and ...
  6. [6]
    Styrene - Properties, uses, and applications - Repsol
    Styrene is used in polystyrene packaging, car components, household appliances, EPS, styrenic rubbers, ABS/SAN copolymers, and unsaturated polyester resins.
  7. [7]
  8. [8]
    What is styrene monomer and how is it used? - Shell Global
    The majority of styrene is used in the production of polystyrene for items such as medical devices, household appliances, drinks cups, food containers and ...
  9. [9]
    Styrene | Public Health Statement | ATSDR - CDC
    Styrene is a colorless liquid, used in manufacturing, plastics, and rubber. It's found in air, soil, and water, and may cause nervous system issues in workers.
  10. [10]
    Styrene - the NIST WebBook
    Styrene · Formula: C8H · Molecular weight: 104.1491 · IUPAC Standard InChI: InChI=1S/C8H8/c1-2-8-6-4-3-5-7-8/h2-7H,1H2 Copy · IUPAC Standard InChIKey: ...
  11. [11]
    Blue Book chapter P-1 - IUPAC nomenclature
    The name 'styrene' is a retained name acceptable in general IUPAC nomenclature as being clear and unambiguous along with the substitutive name 'vinylbenzene'.
  12. [12]
    Styrene - American Chemical Society
    Jun 9, 2015 · Its name comes from styrax (or storax) balsam, the resin of the Liquidambar genus of trees that grow in many places worldwide. M. Bonastre first ...Missing: synonyms IUPAC
  13. [13]
    Table 4-2, Physical and Chemical Properties of Styrene - NCBI
    Density at 20 °C · 0.9059, O'Neil et al. 2001 ; Odor, If pure, sweet and pleasant; commonly contains aldehydes which provide it with a penetrating, sharp, and ...
  14. [14]
    [PDF] Styrene: Chemical Identity & Physical Properties - SIRC
    Styrene: Chemical Identity & Physical Properties. GENERAL INFORMATION ... Chemical formula. C8H8. Chemical structure. Molecular weight. 104.15 g/mol. Color.
  15. [15]
  16. [16]
    100-42-5(Styrene) Product Description - ChemicalBook
    Melting point: -31 °C (lit.) ; Boiling point: 145-146 °C (lit.) ; Density, 0.906 g/mL at 25 °C ; vapor density, 3.6 (vs air) ; vapor pressure, 12.4 mm Hg ( 37.7 °C).Missing: C6H5CH= CH2 vaporization<|control11|><|separator|>
  17. [17]
    Synthetic Methods in Polymer Chemistry - csbsju
    Resonance delocalization in its reactive intermediates makes styrene amenable to almost any method of polymerization, including anionic, cationic, radical, and ...
  18. [18]
    Polar additions to the styrene and 2-butene systems. I. Distribution ...
    Polar additions to the styrene and 2-butene systems. I. Distribution stereochemistry of bromination products in acetic acid.
  19. [19]
    Monitoring Heterogeneously Catalyzed Hydrogenation Reactions at ...
    Sep 11, 2019 · Styrene. The hydrogenation of styrene into ethylbenzene is a fast and well-studied reaction. (36,37) In our setup it takes place inside the ...<|control11|><|separator|>
  20. [20]
    Selective Oxidation of Styrene to Benzaldehyde Using Nanobubbles
    Oct 27, 2025 · The most challenging problem for the selective oxidation of styrene to benzaldehyde is excessive oxidation producing benzoic acid. (47) This ...
  21. [21]
    (PDF) A Theoretical and Experimental Study for Screening Inhibitors ...
    Oct 1, 2019 · Styrene monomer is essential for producing polymer products like electrical insulators but is prone to undesirable polymerization due to heat, ...
  22. [22]
    [PDF] Styrene Monomer: Safe Handling Guide | Plastics Europe
    Being rather volatile and having a flash point of 31oC, styrene is classified as a flammable substance, which in use may form flammable/explosive vapour-air ...
  23. [23]
    A plastic pathway | Nature Chemical Biology
    Aug 17, 2011 · Styrene can be produced naturally by yeast and plants, but ... decarboxylation, with trans-cinnamic acid (tCA) as the intermediate.
  24. [24]
    Synthesis of styrenes through the biocatalytic decarboxylation of ...
    When trans-cinnamic acid was incubated with plant cell cultures at room temperature, styrene was obtained. 4-Hydroxy-3-methoxystyrene (2a), 3-nitrostyrene ...
  25. [25]
    Safety and efficacy of feed additives consisting of essential oils ... - NIH
    In addition, cinnamon bark oil contains low concentrations of styrene (on average: 0.020%, range: 0.016–0.024%). The occurrence of styrene in cinnamon bark oil ...
  26. [26]
    Sustainable bio-production of styrene from forest waste
    A strain of Penicillium expansum produced styrene when grown on tree bark media. •. 52.5 μg/h styrene production rate from fungus grown on 10 g bark media.
  27. [27]
    Metabolism of phenylalanine and biosynthesis of styrene in ...
    Feb 12, 2007 · We examined the biosynthetic pathway leading to styrene formation by Penicillium camemberti using labelled compounds. As styrene is strongly ...
  28. [28]
    Effects of pH and Cultivation Time on the Formation of Styrene ... - NIH
    Apr 4, 2019 · Styrene can be formed by the microbial metabolism of bacteria and fungi. In our previous study, styrene was determined as a spoilage marker ...
  29. [29]
    Styrene | ToxFAQs™ | ATSDR - CDC
    Exposure to styrene is most likely to occur from breathing indoor air that is contaminated with styrene vapors from building materials, tobacco smoke, and use ...
  30. [30]
    [PDF] The influence of roasting conditions on volatile flavour compounds ...
    Jan 2, 2019 · Two aromatic hydrocarbons were detected in both raw and roasted cashew nuts: toluene and styrene were found in significantly higher amounts in ...
  31. [31]
  32. [32]
    One-Step Production of Ethylbenzene from Ethane and Benzene on ...
    Jun 10, 2024 · A one-step process for the production of ethylbenzene from ethane and benzene was used to overcome the issues of high energy consumption and long process flow.<|separator|>
  33. [33]
    Sulzer successfully starts up world's largest pygas-to-styrene ...
    Mar 21, 2023 · The new facility upgrades pygas, a by-product obtained from three steam crackers in Zhoushan's plant, by extracting valuable styrene for use in the ...
  34. [34]
    Side-chain alkylation of toluene with methanol to produce styrene
    Producing styrene by toluene side-chain alkylation with methanol has been deemed as a potential one-step alternative to the conventional two-step route.
  35. [35]
    Direct conversion of toluene into styrene with high selectivity over a ...
    Feb 1, 2021 · Alkali ion-exchanged Al-rich X zeolites are the most representative catalysts for converting toluene to styrene. The alkali cations serve as ...
  36. [36]
    Catalytic reforming boosts octane for gasoline blending - EIA
    Apr 9, 2013 · Because reformate contains significant amounts of benzene, toluene, and xylene, it also is an important source of feedstock for the ...
  37. [37]
    Understanding Naphtha & Ethane Cracking Processes - Hose Master
    The majority of ethylene is produced using a process called “steam cracking”, a thermal process where hydrocarbons are broken down, or “cracked” into smaller ...
  38. [38]
    Styrene Market Size, Share, Growth, and Forecast, 2035
    The global Styrene market stood at approximately 42 million tonnes in 2024 and is expected to grow at a CAGR of 4.37% during the forecast period until 2035.
  39. [39]
    [PDF] Introduction to Styrenic Polymers
    A few years later, in 1845, Blyth and Hofmann [2] observed that `metastyrol' was formed when styrene was exposed to sunlight, while it remained unchanged in ...
  40. [40]
    STYRENE - Molecule of the Month - September 2021 (HTML version)
    It goes back to the first time that styrene was isolated in 1839. A German named Eduard Simon got a volatile liquid from 'storax', the resin of the American ...
  41. [41]
    Age-Old Resins of the Mediterranean Region and Their Uses - jstor
    The resin has been much used as a metal varnish and gives good lustre in thin coats. It has also been used as a leather and paper varnish, being often employed ...Missing: styrene | Show results with:styrene
  42. [42]
    Styrene - Some Industrial Chemicals - NCBI Bookshelf - NIH
    Packaging is the single largest use in which styrene-containing resins, particularly foams, are used as fillers and cushioning. Construction applications ...
  43. [43]
    First car tyre made from synthetic rubber - K 2025
    The first car tire made from synthetic rubber (SBR) was presented by IG Farben on 15.2.1936, as part of a Nazi emergency plan.Missing: WWII | Show results with:WWII
  44. [44]
    U.S. Synthetic Rubber Program - National Historic Chemical Landmark
    The U.S. Synthetic Rubber Program was a government-sponsored effort to produce synthetic rubber during WWII after natural rubber supply was cut off. It was a ...Missing: SBR | Show results with:SBR
  45. [45]
    Polystyrene Derivative - an overview | ScienceDirect Topics
    Polystyrenes are aromatic polymers made synthetically from styrene monomers. I. G. Farben produced the first commercially available polymers in 1931 [121].
  46. [46]
    [PDF] Styrene - DSIR
    Post world war period witnessed a boom in styrene demand due to its application in the manufacture of synthetic rubber. This led to a dramatic increase in ...Missing: 1950s | Show results with:1950s
  47. [47]
    Origin and Early Development of Rubber-Toughened Plastics
    Ostromislensky (32), while employed by Naugatuck Chemical Company, patented a process for making rubber-toughened polystyrene by polymer izing a solution of ...
  48. [48]
    New Catalyst Makes Styrene Manufacturing Cheaper, Greener
    Feb 26, 2021 · Specifically, the conversion process that incorporates the new catalyst uses 82% less energy – and reduces carbon dioxide emissions by 79%. “ ...Missing: 2020s | Show results with:2020s
  49. [49]
    [PDF] The Emergence and Evolution of Atom Efficient and/or ... - FUPRESS
    Apr 15, 2025 · Global styrene production was estimated to be about 37 million tons in 2022. Styrene is a low-cost monomer used in many polymer / co-polymer ...
  50. [50]
    [PDF] PowerPoint 演示文稿 - Singapore Chemical Industry Council
    May 21, 2025 · The styrene industry chain entered a high capacity-expansion phase. 2014-2024 China Styrene Industry Chain Capacity Growth. -10%. 0%. 10%. 20%.
  51. [51]
    Evaluation of the Parameters and Conditions of Process in the ...
    About 90% of the total production of styrene is based on the catalytic dehydrogenation of ethylbenzene (EB). This reaction is reversible and endothermic, and ...
  52. [52]
    The nature of the iron oxide-based catalyst for dehydrogenation of ...
    The active surface is a potassium iron oxide with a 1 : I atomic ratio of K : Fe, whereby iron is only in its trivalent state.
  53. [53]
    US3474153A - Dehydrogenation of ethylbenzene to styrene
    ... ethylbenzene feed to the first reaction zone. The high conversion process generally results in about 60% to 65% conversion of ethylbenzene to styrene. This ...
  54. [54]
    [PDF] Styrene Production from Ethylbenzene
    147.4 kJ/mol liquid (25 °C). 103.4 kJ/mol. Heat of vaporization, ∆HV. (25 °C). 421.7 J/g. (145 °C). 356.7 J/g. Heat of polymerization, ∆Hp (25 °C). –69.8 kJ/mol.
  55. [55]
    Technology (M): Lyondellbasell PO/SM - ppPLUS
    Mar 13, 2025 · The LyondellBasell PO/SM (Propylene Oxide-Styrene Monomer) process is a proprietary technology used for the co-production of Propylene Oxide (PO) and Styrene ...
  56. [56]
    Best practices for pygas-based styrene extraction
    Best practices for pygas-based styrene extraction. Pyrolysis gasoline (pygas) is a by-produced fraction of hydrocarbons generated from a steam cracker.
  57. [57]
    SNOW: Styrene from Ethane and Benzene - ScienceDirect.com
    Styrene is one of the most important monomers for the production of polymers, resins, and rubbers. Styrene production exceeds 25 million MT/y.
  58. [58]
    [PDF] Energy efficiency analysis of styrene production by adiabatic ...
    Styrene production uses high-temperature steam, consuming much energy. Heat integration showed savings in external heating/cooling, and steam recovery can ...
  59. [59]
    CFD analysis of a Pd-based membrane reactor to carry out ...
    Enhancing styrene production: CFD analysis of a Pd-based membrane reactor to carry out ethylbenzene dehydrogenation · 1. Introduction · 2. Model development · 3.
  60. [60]
    [PDF] Rhenium- and molybdenum-catalyzed dehydration reactions
    Classically, the dehydration reaction is performed using strong Brønsted acids, such as sulfuric5 or phosphoric acid,6 p-toluenesulfonic acid,7 or solid acids ...
  61. [61]
    Lab notes and data for “Preparation of the Styrenes” - Chegg
    Jan 28, 2021 · The reaction for the given mechanism is: OH H H2SO4 1-phenylethan-1-01 styrene REACTANT PRODUCT Dehydration of alcohol in the presence of acid ...<|separator|>
  62. [62]
    Illustrated Glossary of Organic Chemistry - Wittig reaction
    The phosphonium ylide adds to benzaldehyde to produce styrene (PhCH=CH2), along with triphenylphosphine oxide (Ph3P=O) as a by-product. The four-membered ring ...
  63. [63]
    Wittig Reaction - Organic Chemistry Portal
    The Wittig Reaction allows the preparation of an alkene by the reaction of an aldehyde or ketone with the ylide generated from a phosphonium salt.Missing: styrene methylenetriphenylphosphorane
  64. [64]
    Palladium-catalyzed vinylic hydrogen substitution reactions with aryl ...
    Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and styryl halides. Click to copy article linkArticle link copied! R. F. Heck ...
  65. [65]
    Preparation of functional styrenes from biosourced carboxylic acids ...
    For the first time, decarboxylation of α-amino acids to the corresponding amines was successfully performed with good to high yields and extended to the ...Missing: elimination | Show results with:elimination
  66. [66]
    Styrene purification process - US4003800A - Google Patents
    Vacuum distillation is commonly used to keep tower temperatures low and to minimize the polymerization of styrene. In addition, to further reduce styrene ...Missing: laboratory yields
  67. [67]
    Efficient Dehydration of 1-Phenylethanol to Styrene by Copper(II ...
    Oct 11, 2024 · The compound is a catalyst for the dehydration of 1-phenylethanol to styrene (+H 2 O) in over 95% yield at 1 mol % catalyst loading.
  68. [68]
    A Renaissance in Living Cationic Polymerization | Chemical Reviews
    Cationic polymerization has a long history, and the first research into cationic polymerization on record was conducted in the late 18th century.
  69. [69]
    Kinetics of polymerization of styrene initiated by substituted benzoyl ...
    Kinetics of polymerization of styrene initiated by substituted benzoyl peroxides. ... mechanism for radical production in the presence of vinyl monomers.
  70. [70]
    [PDF] Kinetics of Free Radical Polymerization of Styrene to Complete ...
    The study examines styrene polymerization with thermal and peroxide initiation, deriving a kinetic model and simulating polymer reactor systems.
  71. [71]
    Kinetics of Anionic Polymerization of Styrene in Tetrahydrofuran
    Synthesis of poly(4‐hydroxystyrene)‐based block copolymers containing acid‐sensitive blocks by living anionic polymerization. Journal of Polymer Science ...
  72. [72]
    Precise Synthesis of Functional Block Copolymers by Living Anionic ...
    Jan 24, 2017 · The anionic polymerization of styrene derivatives bearing carbazole moieties, such as N-ethyl-2-vinylcarbazole (NE2VCz) and N-ethyl-3 ...
  73. [73]
    Discovery of Syndiotactic Polystyrene: Its Synthesis and Impact
    May 26, 2020 · This Editorial discusses two seminal Macromolecules publications by Ishihara et al. concerning the synthesis of syndiotactic polystyrene.
  74. [74]
    Thermal polymerization of styrene at high conversions and ...
    Uninhibited styrene polymerizes even at room temperature, although slowly. Industrial thermal processes usually operate in the temperature range 100-200°C. A ...
  75. [75]
    Comparison of Physical Aging and Glass Transition in Glassy ...
    2 Mar 2025 · At 65 °C, the PS layer is in its glassy state (T g bulk = 100 °C for PS), while the PnBMA layer is still in the equilibrium state above its Tg ( ...
  76. [76]
    Styrene-Butadiene Rubber - an overview | ScienceDirect Topics
    STYRENE-BUTADIENE RUBBER (SBR). A copolymer of about 25% styrene and 75% butadiene. A general purpose synthetic rubber which is competitive in properties with ...
  77. [77]
    PRODUCTION, IMPORT/EXPORT, USE, AND DISPOSAL - NCBI - NIH
    Styrene is mainly produced by dehydrogenation of ethylbenzene. It's used in plastics, resins, and rubber. Disposal includes landfill and incineration.
  78. [78]
  79. [79]
    Styrene-Butadiene Latex | SB Latex Copolymers
    Styrene-butadiene (SB) latex is a common type of emulsion polymer used in a number of industrial and commercial applications.
  80. [80]
    Advancements in monomers and reinforcements of unsaturated ...
    Apr 1, 2025 · Styrene, a commonly used crosslinking agent, is highly reactive and cost-effective. The proportion of styrene in UPR significantly impacts the ...
  81. [81]
    styrene, 100-42-5 - The Good Scents Company
    Functional use(s) - cosmetic agents. Has an balsamic type odor.
  82. [82]
  83. [83]
    Human exposure to styrene. VI. Percutaneous absorption ... - PubMed
    The results obtained show that the rate of absorption of styrene through the skin is very low, averaging 1 +/- 0.5 micrograms/cm2 X min. This rate seems to be ...
  84. [84]
  85. [85]
    [PDF] ATSDR Styrene ToxFAQs
    Styrene is a component of tobacco smoke. Avoid smoking in enclosed spaces like inside the home or car in order to limit exposure to children and other family ...
  86. [86]
    HEALTH EFFECTS - Toxicological Profile for Styrene - NCBI Bookshelf
    Exposure to 50 ppm styrene for 13 weeks resulted in atrophy of the nasal olfactory epithelium and dilatation, hypertrophy and hyperplasia of Bowman's gland ( ...DISCUSSION OF HEALTH... · TOXICITIES MEDIATED... · BIOMARKERS OF...
  87. [87]
  88. [88]
    Determination of mandelic acid and phenylglyoxylic acid in the urine ...
    This paper describes a sensitive biological monitoring method for assessing exposure to styrene. Two major metabolites of styrene, mandelic acid (MA) and ...Missing: routes | Show results with:routes
  89. [89]
    Probing into Styrene Polymerization Runaway Hazards: Effects of ...
    May 3, 2019 · In this work, the thermal runaway hazards of the ethylbenzene–styrene system with different monomer mass fractions were calorimetrically investigated.<|control11|><|separator|>
  90. [90]
    [PDF] STYRENE - CAMEO Chemicals
    9.12 Latent Heat of Vaporization: 156 Btu/lb = 86.8 cal/g = 3.63 X 105 J/kg ... 9.19 Reid Vapor Pressure: 0.27 psia. NOTES. JUNE 1999. Page 2. STYRENE. STY.Missing: melting refractive
  91. [91]
    ICSC 0073 - STYRENE - International Chemical Safety Cards (ICSCs)
    Attacks rubber, copper and copper alloys. Flash point: 31°C c.c. The substance can be absorbed into the body by inhalation of its vapour. The substance is ...
  92. [92]
    Probing into Styrene Polymerization Runaway Hazards: Effects of ...
    May 3, 2019 · In this work, the thermal runaway hazards of the ethylbenzene–styrene system with different monomer mass fractions were calorimetrically investigated.
  93. [93]
    (PDF) Probing into Styrene Polymerization Runaway Hazards
    May 3, 2019 · Polymerization reactions have caused a number of serious incidents in the past; they are prone to reaction runaways because of their ...Missing: historical | Show results with:historical
  94. [94]
    [PDF] Styrene HSSE and Product Stewardship - Shell Global
    inhibitor – typically 10-15 ppm pTBC (para – tert. - Butylcatechol). The prevention of polymer build-up during storage is of great concern. At ambient ...
  95. [95]
    [PDF] Styrene Monomer Product Stewardship Summary
    If nitrogen or inert gas blanketing is used, there must be sufficient oxygen levels (6 – 10% v/v O2) in the gas mixture to inhibit polymerization but at the ...Missing: refrigeration | Show results with:refrigeration
  96. [96]
    Determination of 4-tert-butyl pyrocatechol content in styrene with ...
    Apr 30, 2025 · A novel analytical method was developed for the determination of 4-tert-butylpyrocatechol (TBC) in styrene. Various measurement methods were compared.
  97. [97]
    Review of quantitative and qualitative methods for monitoring ...
    Mar 31, 2023 · This paper provides a broad overview of the available methods, both for laboratory and industrial use, that allow qualitative and quantitative analysis of the ...