Fact-checked by Grok 2 weeks ago

Radical initiator

A radical initiator is a designed to generate free s through homolytic , typically triggered by heat, light, or reactions, thereby initiating radical chain processes such as and . These initiators play a crucial role in controlling the start of reactions by providing the initial radicals that propagate chain mechanisms, often at temperatures below 100°C to facilitate and industrial applications. In , they enable the formation of polymers like by adding to monomers, leading to controlled growth of macromolecular chains. Radical initiators are broadly classified into thermal, photochemical, and types based on their method. initiators, the most common, decompose upon heating to produce radicals; examples include peroxides with weak O-O bonds and azo compounds that release gas. Photochemical initiators, or photoinitiators, respond to light for applications requiring spatial or temporal control, such as in coatings and adhesives. initiators involve between components, allowing initiation at lower temperatures and in aqueous systems. They are valuable for biomedical s. Among the most widely used are , which decomposes at 70–80°C to form alkyl radicals and , and benzoyl peroxide, effective at 70–80°C via O-O bond homolysis to yield oxy radicals. These compounds are selected based on factors like , , and with solvents or monomers to optimize reaction efficiency and minimize side reactions. In free , initiators like AIBN and s dominate commercial processes due to their reliability in producing high-molecular-weight polymers for plastics, rubbers, and resins. Beyond polymerization, they facilitate reactions such as alkane halogenation and , underscoring their versatility in synthetic chemistry.

Fundamentals

Definition and Principles

A radical initiator is a that decomposes under specific conditions to generate free radicals, which serve as the starting point for chain reactions in processes such as and . These initiators are essential in radical-mediated reactions because they provide the initial reactive species necessary to drive the overall process. Free radicals are atoms, molecules, or ions with at least one in their outer shell, rendering them highly reactive and short-lived due to their tendency to participate in rapid bond-forming or breaking reactions. In the context of radical initiators, the typically produces two such radicals from a single initiator , as represented by the general : \text{Initiator} \rightarrow 2R^\bullet where R^\bullet denotes a free species. This generation of radicals marks the phase of a , which proceeds through three main stages: , where the primary radicals are formed and begin reacting with substrates; , involving sequential radical additions or abstractions that build molecular chains; and termination, where radicals combine or disproportionate to end the reaction. In industrial applications, radical initiators play a pivotal role in the synthesis of polymers via chain-growth mechanisms, such as the of monomers to produce materials like plastics and coatings, enabling efficient large-scale production under controlled conditions like or photochemical activation.

Historical Development

The discovery of free radicals laid the foundational groundwork for understanding radical initiators in . In 1900, Moses Gomberg reported the first stable organic free radical, triphenylmethyl, challenging prevailing theories on carbon valency and demonstrating the existence of trivalent carbon species. This breakthrough was further substantiated in 1929 when Friedrich Paneth and Wilhelm Hofeditz isolated short-lived methyl radicals using a mirror technique, confirming the transient nature of alkyl radicals and their role in chain reactions. During the 1920s, advanced by proposing the macromolecular hypothesis, linking high-molecular-weight substances to chain-growth mechanisms involving radicals, though initial polymerizations were not explicitly radical-initiated. The practical application of radical initiators emerged in the early with peroxy compounds. Between 1910 and 1930, the first radically synthesized polymers, such as poly(), were produced using like benzoyl peroxide as initiators, marking the shift from empirical to mechanistic processes. In 1937, formulated the kinetics of vinyl as a , emphasizing , , and termination steps, which provided a theoretical framework for initiator efficiency. accelerated development due to rubber shortages; in the 1940s, -initiated of rubber (GR-S) became central to production in the U.S., with systems combining and reducing agents enabling low-temperature for industrial scalability. Post-WWII commercialization expanded initiator varieties through patents and industrial adoption. By the late 1940s and 1950s, azo compounds gained prominence; although (AIBN) was synthesized in 1896 by Johannes Thiele and Karl Heuser, its use as a clean, non-oxygenated radical initiator in became widespread, offering advantages over peroxides in controlling molecular weight. Numerous patents, such as those for diacyl peroxides and azo derivatives, facilitated commercial production, enabling of and acrylics. From the 1980s onward, innovations focused on precision and sustainability. Takeo Otsu introduced iniferter systems in 1982, precursors to controlled , allowing reversible activation for narrow polydispersity polymers. The 1990s saw the advent of (ATRP) in 1995 by Krzysztof Matyjaszewski, using halogen-transfer agents for living-like radical processes, revolutionizing block copolymer synthesis. Into the 2020s, research has emphasized eco-friendly photoinitiators, such as low-VOC, LED-curable systems based on BODIPY derivatives, reducing energy use and environmental impact in UV-curing applications.

Chemical Types

Peroxy Compounds

Peroxy compounds, also known as , are characterized by the general formula R-O-O-R', where R and R' are organic substituents such as alkyl, aryl, or acyl groups. The defining feature is the weak O-O bond, with a typically around 150 kJ/mol (approximately 36 kcal/mol), which predisposes these compounds to homolytic cleavage under appropriate conditions. This distinguishes peroxy compounds from other radical initiators, as the peroxide linkage imparts inherent lability influenced by the nature of the substituents. The properties of peroxy compounds, particularly their thermal stability, vary significantly based on the substituents attached to the oxygen atoms. For instance, benzoyl peroxide (BPO), with the formula (C6H5CO)2O2, exhibits moderate thermal stability and undergoes decomposition at temperatures between 70°C and 80°C. In contrast, di-tert-butyl peroxide (DTBP), (CH3)3C-O-O-C(CH3)3, demonstrates higher stability due to the bulky tert-butyl groups that sterically hinder bond cleavage, decomposing around 120°C. These variations allow for tailored selection in processes requiring specific activation temperatures, with the O-O bond's weakness ensuring efficient radical generation while substituents modulate reactivity and safety profiles. Synthesis of peroxy compounds typically involves peroxidation reactions, such as the direct reaction of (H2O2) with hydrocarbons, alcohols, or carboxylic acids under acidic conditions. For example, hydroperoxides (R-O-O-H) can be prepared via of hydrocarbons in the presence of oxygen and initiators, while diacyl peroxides like BPO are formed by acylating with acid chlorides. Peroxy esters, another subclass, are synthesized by coupling peracids with alkylating agents or through esterification processes, often yielding products in the range of 75-77% under optimized conditions. These methods leverage the electrophilic nature of peroxides to incorporate organic groups, though they require precise control to avoid side reactions or instability during isolation. Peroxy compounds offer advantages such as high , particularly in aqueous environments where they can generate reactive effectively for processes like . Their tunable allows for controlled production under mild thermal conditions, enhancing versatility in synthetic applications. However, a key disadvantage is their potential for , especially when impure or exposed to contaminants, , or , necessitating stringent and handling protocols to mitigate hazards. This sensitivity arises from the labile O-O bond, which can lead to runaway reactions if not managed properly.

Azo Compounds

Azo compounds serve as a prominent class of thermal radical initiators, characterized by their general R–N=N–R', where R and R' are typically alkyl or substituted alkyl groups that influence and decomposition kinetics. The nitrogen-nitrogen undergoes homolytic cleavage upon heating, generating two alkyl s (R• and R'•) and nitrogen gas (N₂) as a neutral byproduct, with decomposition temperatures ranging from 40°C to 100°C depending on the substituents. This process provides a clean source of radicals without introducing oxygenated species, distinguishing azo initiators from other types. A key example is (AIBN), with the structure (CH₃)₂C(CN)–N=N–C(CN)(CH₃)₂, where the cyano groups stabilize the resulting 2-cyano-2-propyl radicals and tune the to approximately 10 hours at 65°C in organic solvents. AIBN is lipophilic and readily soluble in common organic solvents like or , enabling its use in non-aqueous systems, while the evolution of N₂ gas during decomposition helps minimize side reactions by diluting the reaction mixture and reducing cage recombination. Other variants, such as 2,2'-azobis(2-methylpropionamidine) dihydrochloride (AAPH), offer hydrophilic properties for aqueous applications, with a 10-hour at 56°C due to the substituents. Synthesis of azo compounds like AIBN typically involves the oxidation of the corresponding derivative, such as 2,2'-azobis(isobutyronitrile) hydrazo compound, using agents like under controlled conditions to achieve high yields and purity. This method, often conducted in aqueous or alcoholic media at mild temperatures, allows for the preparation of symmetric azo initiators with tailored decomposition profiles by selecting appropriate hydrazine precursors. Compared to peroxides, azo compounds exhibit advantages in providing a more predictable with well-defined , reducing the risk of induced and permitting operation under milder conditions. Their in a wide range of solvents facilitates versatile applications, though they remain sensitive to impurities like metals or light, which can accelerate unintended .

Redox and Photoinitiators

Redox initiators operate through reactions between an oxidant and a reductant, generating free s at ambient or low temperatures without requiring . A classic example is the persulfate-ascorbate system, where serves as the oxidant and as the reductant; the ascorbate reduces persulfate to sulfate radicals (SO₄•⁻), which then propagate chain initiation in . This mechanism enables controlled radical generation, often at temperatures as low as 30°C, making it suitable for sensitive processes. In , systems like persulfate-ascorbate facilitate efficient radical entry into micelles, enhancing polymerization rates and enabling the synthesis of stable particles under mild conditions. The system's versatility allows for formation, as demonstrated in rapid cross-linking reactions at , avoiding high-energy inputs that could degrade delicate components. Photoinitiators, in contrast, rely on light to produce radicals, offering temperature-independent initiation that is ideal for spatially controlled reactions. Type I photoinitiators, such as those in the Irgacure series (e.g., Irgacure 184 and 2959), undergo α-cleavage upon UV to directly form initiating radicals, while Type II initiators like abstract hydrogen from a co-initiator (e.g., amines) after excitation, generating radicals via a bimolecular process. These systems typically operate under UV or visible light (300–500 nm), with tunable wavelengths enabling applications in precision manufacturing like . Recent advancements have focused on visible-light photoinitiators to improve , addressing limitations of UV-based systems in biomedical contexts. For instance, derivatives, such as mono- and di-allylated forms, initiate free-radical of under 405–505 LEDs with high efficiency (up to 83% conversion), while also producing for antibacterial effects in photodynamic inactivation, supporting sustainable, non-leaching materials for . Similarly, (FMN), a water-soluble derivative, combined with , enables photocrosslinking of biocompatible hydrogels (e.g., PEGDA, silk fibroin) under 465 blue via to generate radicals, achieving >95% cell viability in encapsulation for 3D biofabrication. These developments extend to BODIPY-based systems, which drive under 405–530 in three-component setups, facilitating fluorescent 3D-printed structures with potential in biocompatible photopolymerization.

Reaction Mechanisms

Thermal Initiation

Thermal initiation is the process by which radical initiators decompose under heat to generate free s through homolytic bond cleavage, typically involving the breaking of weak bonds such as the O-O linkage in peroxides or N-N in azo compounds. This method serves as a primary mechanism for initiating radical reactions, where the applied overcomes the , producing two radical species from a single initiator . For example, the can be represented as: \ce{I ->[heat] 2 R^\bullet} where I is the initiator and R• denotes the resulting radicals. The rate of thermal decomposition follows first-order kinetics, meaning the decomposition rate depends solely on the initiator concentration and is independent of other reactants. The rate constant k_d is described by the Arrhenius equation: k_d = A e^{-E_a / RT} where A is the pre-exponential factor, E_a is the activation energy, R is the gas constant, and T is the absolute temperature. This temperature dependence allows precise control over radical generation; for instance, benzoyl peroxide (BPO), a common peroxy initiator, has an activation energy of approximately 132 kJ/mol. Half-life calculations provide a practical measure of initiator and efficiency under specific conditions, defined as t_{1/2} = \ln(2) / k_d for processes. For BPO, the is about 10 hours at 72°C, dropping significantly at higher temperatures due to the exponential nature of the Arrhenius relationship—reaching roughly hour at 92°C. This enables selection of initiators based on required reaction temperatures, with higher E_a values indicating greater thermal at ambient conditions. Several factors influence the efficiency of thermal decomposition beyond temperature. Solvents play a key role through the , where the nascent radicals form within a solvent "" and may recombine or disproportionate before diffusing apart; higher viscosity solvents increase recombination, reducing the effective radical yield and thus lowering the observed k_d. Inhibitors, such as radical scavengers, can indirectly affect the process by reacting with escaped radicals, though they do not alter the primary unimolecular decomposition step. To minimize radical pairing and maximize efficiency, reactions are often conducted in low-viscosity media that promote rapid from the cage.

Photochemical Initiation

Photochemical initiation involves the absorption of photons by a molecule, promoting it to an excited electronic state that subsequently undergoes bond cleavage to generate free radicals. This process adheres to the first law of , which states that only absorbed light can induce chemical change, and is governed by the Beer-Lambert law, describing the attenuation of light intensity through a medium as I = I_0 e^{-\epsilon c l}, where I is the transmitted intensity, I_0 the incident intensity, \epsilon the molar absorptivity, c the concentration, and l the path length. Upon excitation, the typically fragments homolytically, as represented by the general : \text{Initiator} + h\nu \rightarrow \text{R}^\bullet + \text{R}^\bullet The efficiency of this radical is quantified by the (\Phi), defined as the number of radicals produced per absorbed, often ranging from 0.1 to 1 for effective systems. Key factors influencing photochemical initiation include the wavelength of light used, with ultraviolet (UV) light (typically 250–400 nm) being common for many photoinitiators due to their strong absorption in this range, while visible light (400–700 nm) enables deeper penetration and reduced scattering in thicker samples. For instance, ketones such as benzophenone or acetophenone undergo Norrish Type I cleavage, where the excited carbonyl compound breaks the bond alpha to the carbonyl group, yielding acyl and alkyl radicals that initiate polymerization. This wavelength specificity allows tailoring of initiators to match available light sources, contrasting with broader thermal activation. A primary advantage of photochemical initiation is the precise spatial and temporal control it affords in photopolymerization processes, enabling on-demand activation through patterned light exposure without affecting unirradiated areas. This is particularly valuable in applications like and , where reaction rates can reach seconds under ambient conditions. Since 2010, developments in LED-compatible photoinitiators have addressed limitations of traditional mercury lamps, focusing on visible-light-absorbing Type I systems like silyloxy-substituted anthraquinones and dye-based hybrids that operate efficiently at 405 nm with high quantum yields (>0.5). These innovations, including copper complexes and phosphine oxides tuned for low-intensity LEDs, have enhanced energy efficiency and safety in industrial photopolymerization, with polymerization rates comparable to UV systems but improved depth of cure up to several millimeters.

Redox Initiation

Redox initiation involves the generation of free radicals through one-electron transfer reactions between an oxidant and a reductant, enabling radical formation at low temperatures, often in aqueous media suitable for biological or sensitive systems. A classic example is the persulfate-iron system, where ferrous ions reduce peroxydisulfate to produce sulfate radicals: \text{S}_2\text{O}_8^{2-} + \text{Fe}^{2+} \rightarrow \text{SO}_4^{\bullet-} + \text{Fe}^{3+} + \text{SO}_4^{2-} These primary radicals can then abstract hydrogen or add to monomers, initiating chain reactions via secondary radicals, such as in polymerization processes. Key factors influencing initiation include and the s of oxidant to reductant components, which control yield and stability. In the persulfate-iron system, optimal performance occurs at 2.0–6.0, with typically conducted around 3.0 to balance generation and side reactions; lower can suppress certain pathways, reducing overall rates. Component s, such as a of S₂O₈²⁻:Fe²⁺:reductant of 2:1:1 relative to , fine-tune the reaction by minimizing termination and maximizing flux. Similarly, exemplifies this, where Fe²⁺ reacts with H₂O₂ to generate s: \text{Fe}^{2+} + \text{H}_2\text{O}_2 \rightarrow \text{Fe}^{3+} + \text{OH}^- + \text{OH}^{\bullet} pH below 3.0 enhances production in this system, while balanced H₂O₂:Fe²⁺ s prevent excessive metal precipitation or radical scavenging. The kinetics of initiation follow bimolecular rate laws, with the rate of formation depending on the concentrations of the pair, often exhibiting fractional orders such as 0.5–0.75 for the oxidant and 0.4 for the reductant. This allows precise control of flux through adjustments in component concentrations and conditions, ensuring steady without excessive heat evolution, which is particularly advantageous for low-temperature applications like biomedical processes.

Applications

Polymerization Processes

Radical initiators are essential for free radical polymerization, a versatile process widely used in industrial polymer synthesis. The initiation step involves radicals generated from the initiator adding to a vinyl monomer, forming an initiating polymer radical that propagates by successive monomer additions. For instance, in the polymerization of styrene, a radical R• adds to the double bond to yield R-CH₂-CH•Ph, starting the chain growth. This method accommodates various formats, including bulk polymerization, where only monomer and initiator are employed to produce high-purity resins like polymethyl methacrylate; solution polymerization, which uses solvents to manage viscosity and heat; and emulsion polymerization, relying on water-soluble initiators such as persulfates to form stable latex particles for coatings and adhesives. A prominent application is the suspension polymerization of vinyl chloride to produce polyvinyl chloride (PVC), where organic peroxides like lauroyl peroxide serve as initiators to control reaction rate and molecular weight. These processes typically achieve monomer conversions of around 80%, balancing productivity and product quality. In controlled radical polymerization techniques such as reversible addition-fragmentation chain transfer (RAFT), conventional thermal initiators like azo compounds or s generate initial radicals that interact with thiocarbonylthio RAFT agents, enabling precise control over polymer architecture, including narrow molecular weight distributions for . On an industrial scale, radical initiators exhibit efficiencies ranging from 0.3 to 0.8, indicating the fraction of primary radicals that successfully initiate chains rather than undergoing side reactions. High conversions, often exceeding 80% in optimized processes, underscore their economic impact, with the global initiators market valued at approximately USD 1.5 billion in 2025. Emerging applications focus on sustainable bio-based polymers, where photoinitiators derived from natural sources facilitate light-induced of renewable monomers like derivatives, supporting eco-friendly resin production for and coatings.

Organic Synthesis and Radical Reactions

Radical initiators such as play a crucial role in facilitating anti-Markovnikov reactions, where they generate radicals that add to alkenes in the presence of HBr, yielding alkyl bromides with the halogen attached to the less substituted carbon. This process proceeds via a free chain mechanism initiated by homolytic cleavage of the peroxide O-O bond, often under mild heating or light, and is selective for HBr due to the exothermic nature of the steps involving radicals. In contrast to ionic additions, this radical pathway tolerates a wide range of functional groups without requiring , enabling efficient of branched alkanes from alkenes. AIBN serves as a versatile thermal initiator for radical cyclizations in , typically paired with (Bu₃SnH) to promote intramolecular carbon-carbon bond formation, favoring the construction of five- or six-membered rings with high . For instance, haloalkenes undergo 5-exo-trig cyclization upon heating with AIBN, generating alkyl radicals that abstract to form cyclic products, as demonstrated in the of complex terpenoids like hirsutene. Peroxides similarly initiate cyclizations in C-H functionalization reactions, where transfer (HAT) from unactivated C(sp³)-H bonds generates carbon-centered radicals for subsequent coupling, offering step-economical routes to functionalized heterocycles. These methods exhibit excellent functional group tolerance, accommodating sensitive moieties like esters and amides that might interfere with . A seminal example is the , where AIBN initiates the radical decomposition of thiohydroxamate esters derived from carboxylic acids, enabling the replacement of CO₂H with H or other groups under mild conditions, with yields often exceeding 80% in representative cases like the reduction of derivatives. This reaction has been pivotal in natural product synthesis, providing access to decarboxylated alkanes without over-reduction. In modern , influenced by advancements recognized in high-impact work since the , initiators like or azo compounds complement visible-light-driven systems to generate transient s for selective C-H arylations and couplings, enhancing enantioselectivity in asymmetric syntheses through chiral ligands. Up to 2025, innovations include thermal reductive initiation using azo compounds like ACVA with formate salts to produce CO₂ anions, enabling metal-free S_RN1 couplings on scales up to 50 g with >60% yields and broad substrate scope for heteroaryl functionalizations. These lab-scale applications, often integrated with catalysts, prioritize precision over the large-scale growth seen in , underscoring s' utility in discrete molecule assembly.

Safety and Environmental Considerations

Health and Handling Hazards

Radical initiators, particularly such as benzoyl peroxide (BPO) and (MEKP), pose significant health risks primarily through and upon exposure. Contact with BPO can cause to the eyes, , and mucous membranes, leading to symptoms like redness, dryness, peeling, burning, and allergic in sensitized individuals. Similarly, MEKP exposure results in severe burns, blisters, eye , and respiratory effects including cough, dyspnea, and potential ; ingestion may induce , , , and organ damage to the liver and kidneys. Azo initiators like (AIBN) are harmful if swallowed or inhaled, irritating the eyes, , and respiratory tract, with thermal decomposition potentially releasing toxic or related nitriles, exacerbating inhalation toxicity. Regarding carcinogenicity, BPO is classified by the International Agency for Research on Cancer (IARC) as Group 3, not classifiable as to its carcinogenicity to humans, based on limited evidence from animal studies and inadequate human data. Handling radical initiators involves risks stemming from their inherent instability and reactivity, which can lead to violent , autoignition, or under certain conditions. are strong oxidizers that react vigorously with reducing agents, contaminants, or heat, potentially causing rapid gas evolution and ; for instance, MEKP has been involved in industrial incidents where contamination triggered explosive decompositions. AIBN, while thermally stable below 70°C, can self-accelerate above this threshold, releasing gas and posing explosion risks in confined spaces. Autoignition temperatures vary among peroxides, and impurities can lower the temperature at which leads to , heightening hazards during transport or storage mishandling. Occupational exposure limits help mitigate these risks; for example, the OSHA permissible exposure limit (PEL) for BPO is 5 mg/m³ as an 8-hour time-weighted average (TWA), with skin notation indicating potential dermal absorption. For MEKP, NIOSH recommends a ceiling limit of 0.2 ppm (1.5 mg/m³), though OSHA has none established. First aid protocols emphasize immediate decontamination: irrigate eyes with water for at least 15 minutes, wash skin with soap and water, move to fresh air for inhalation exposure, and seek medical attention for ingestion or severe symptoms, avoiding induced vomiting due to corrosive potential. Recent studies on photoinitiators highlight concerns over migration from UV-cured inks in consumer products like and printed materials, where compounds such as and Irgacure series may leach into food or dust, potentially causing endocrine disruption or upon chronic low-level exposure. A 2024 analysis of paper found detectable levels of photoinitiators and their products migrating under simulated conditions, though estimated human intake via or dermal remained below thresholds posing acute risks; however, long-term effects warrant further , especially for vulnerable populations. Radical initiators also present environmental risks, particularly to ecosystems. For example, BPO is toxic to aquatic organisms, with to (EbC50 in the range of low mg/L) and , necessitating careful disposal to prevent release into waterways.

Storage, Disposal, and Regulations

Radical initiators, such as and azo compounds, require specific storage conditions to maintain and prevent unintended or reactions. These materials should be kept in cool, dry environments at or below the recommended (often 15-30°C) specified for each product based on its self-accelerating decomposition (SADT), with lower temperatures recommended to minimize risks. Storage in original, unopened containers is essential, and exposure to light, heat, moisture, or contaminants must be avoided, often necessitating dark, well-ventilated areas. from flammable materials, reducing agents, and accelerators is critical, as per guidelines from the (NFPA 400), to prevent fire hazards or explosive interactions. For certain peroxides, refrigeration may be required, while azo initiators like AIBN are generally stored at ambient temperatures but benefit from inhibitors to suppress runaway . Disposal of radical initiators must prioritize safety due to their instability, with methods tailored to the compound type and quantity. are commonly disposed of through dilution to less than 1% active oxygen concentration followed by at permitted facilities, in compliance with U.S. Environmental Protection Agency (EPA) guidelines for management. For small quantities, neutralization via reduction with agents like ferrous sulfate or can deactivate peroxides before disposal, converting them to non-reactive products. Solid peroxides may be incinerated as-is or as a water-wet to control reactivity. Azo initiators, which decompose to gas and relatively non-toxic organic fragments, allow for simpler or, in some cases, recovery of byproducts for reuse in processes, aligning with minimization practices. Regulatory frameworks govern the handling, , and use of radical initiators to mitigate risks. Under the EU's REACH regulation, organic peroxides and azo compounds are classified as hazardous substances requiring registration, evaluation, and risk assessments, though no specific blanket restrictions apply to their use as initiators. In the United States, the (DOT) categorizes many as Class 5.2 () and certain azo compounds as Class 4.1 (self-reactive substances), mandating specialized packaging, labeling, and temperature-controlled to prevent or fire. As of July 2025, the EU's Chemicals Industry Action Plan promotes principles, encouraging reductions in hazardous initiator use through safer alternatives and sustainable process designs to enhance . Best practices for managing radical initiators emphasize protective measures and rapid response protocols. Personal protective equipment (PPE), including chemical-resistant gloves, safety goggles, lab coats, and respirators, must be worn during handling to prevent skin contact or inhalation. In case of spills, immediately alert personnel, evacuate if necessary, don appropriate PPE, and contain the spill using absorbent materials without generating heat or friction, followed by neutralization or cleanup per material safety data sheets. All operations should occur in well-ventilated areas with access to spill kits and emergency eyewash stations to ensure compliance with occupational safety standards.

References

  1. [1]
    Unit 5: Radicals and Radical Reactions
    More commonly, in the laboratory, radical chemistry is initiated at temperatures below 100 oC by adding a few percent of a substance (called an initiator) which ...
  2. [2]
    2.9: Radical Polymerization - Chemistry LibreTexts
    Sep 12, 2021 · A typical example is the formation of polystyrene under the influence of the radical initiator, azoisobutyronitrile (AIBN). Radical reactions ...
  3. [3]
    [PDF] Polymerization Initiators - TCI Chemicals
    Polymerization initiators regulate initiation by heat or light. Thermal initiators generate radicals/cations with heat, while photo initiators use UV light. ...
  4. [4]
    Non-ionic Chemical Reactions - MSU chemistry
    Azobisisobutyronitrile (AIBN) is the most widely used radical initiator of this kind, decomposing slightly faster than benzoyl peroxide at 70 to 80 ºC.
  5. [5]
    Polymerization Initiators - Polymer / BOC Sciences
    The initiation systems of free radical polymerization can be summarized as azo initiators, peroxides initiators, redox initiators, multifunctional initiators ...
  6. [6]
  7. [7]
    What is free radical polymerization? types, characteristics, reaction ...
    Dec 2, 2022 · Azo compounds and peroxides are the main commercially used initiators used in free radial polymerization processes. 1-1. Reaction mechanism of ...<|control11|><|separator|>
  8. [8]
    Initiate Polymerization - an overview | ScienceDirect Topics
    Initiation is defined as the series of reactions that commences with generation of primary radicals and culminates in addition to the carbon–carbon double bond ...
  9. [9]
    CHAPTER 1: Kinetics and Thermodynamics of Radical Polymerization
    A redox-initiator consists of an oxidizing and a reducing agent. In most redox initiators, the redox reaction leads to the formation of only one radical, ...
  10. [10]
    Free Radical Inhibition Mechanisms – Implications for Measurement
    Dec 6, 2021 · Free radicals are highly reactive and unstable atoms or molecules that contain at least one unpaired electron.
  11. [11]
    Fundamentals of Emulsion Polymerization | Biomacromolecules
    Jun 16, 2020 · I is initiator, i is the number of radicals R• produced from the initiator (i = 1 or, more commonly, i = 2), M is monomer, P z • is a chain ...<|control11|><|separator|>
  12. [12]
    Free Radical Polymerization - an overview | ScienceDirect Topics
    The addition of initiators can generate free radicals or active sites on the host polymer, which can promote the growth of polymer chains. Free radical ...
  13. [13]
    Light-Controlled Radical Polymerization: Mechanisms, Methods ...
    Mar 15, 2016 · Since the radicals initiated are not instantly and reversibly consumed in the reaction system, they can continually propagate before they ...
  14. [14]
    Moses Gomberg and the Discovery of Organic Free Radicals
    Gomberg published his findings in 1900, but the existence of triphenylmethyl and other organic free radicals remained in dispute for nearly a decade. They were ...
  15. [15]
    Paneth's mirrors and the isolation of methyl radicals - Chemistry World
    Mar 25, 2025 · It was not until 1900 that the first properly observable radical (in the modern sense), triphenylmethyl, was isolated by Moses Gomberg at the ...
  16. [16]
    Origins and Development of Initiation of Free Radical Polymerization ...
    Feb 11, 2010 · The first free radically synthesized polymers were produced between 1910 and 1930 by initiation with peroxy compounds.
  17. [17]
    U.S. Synthetic Rubber Program - National Historic Chemical Landmark
    The recipe consisted of monomers butadiene (75%) and styrene (25%), potassium persulfate as a catalyst or initiator, soap as an emulsifier, water, and a ...Missing: peroxide | Show results with:peroxide
  18. [18]
    2,2'-Azobis(isobutyronitrile) - American Chemical Society
    Feb 16, 2009 · Its preparation was first reported by Thiele and Heuser in 1896, and it has been used in polymerization since at least the 1940s. Its ...
  19. [19]
    Living Polymerization—Emphasizing the Molecule in Macromolecules
    Sep 17, 2017 · Kharasch's discovery in the 1930s that the presence of a radical initiator could change the regiochemistry of hydrogen halide addition to ...
  20. [20]
    Atom Transfer Radical Polymerization | Chemical Reviews
    The development of controlled/living radical polymerization (CRP) methods has been a long-standing goal in polymer chemistry.
  21. [21]
    Application of new photoinitiating systems based on BODIPY ...
    Mar 19, 2025 · Markedly, these BODIPY derivatives efficiently induced the free radical photopolymerization of acrylates under irradiation from violet-to-green ...
  22. [22]
    [PDF] Novel Process for Organic Peroxides Synthesis - PolyPublie
    Substituting one or both hydrogen atoms of hydrogen peroxide (H2O2) with an organic compound produces an organic peroxide, including hydroperoxides, diperoxides ...
  23. [23]
    [PDF] Organic Peroxides in Radical Chemistry and Stereochemical Study ...
    ... peroxyoxalates were studied as a source of tertiary alkoxy radicals. An operationally simple method to access these radicals from tertiary alkyl.
  24. [24]
  25. [25]
  26. [26]
    The Efficiency of Radical Production from Azo-bis-isobutyronitrile
    Determination of the rate of radical formation in the decomposition of the initiator azobis-isobutyronitrile in reactions of oxidation of hydrocarbons in ...
  27. [27]
    [PDF] Perkadox AIBN - Nouryon
    Half-life data. The reactivity of an azo initiator is usually given by its half-life (t1/2) at various temperatures. For Perkadox® AIBN in chlorobenzene half- ...
  28. [28]
    CN104276981A - Method of synthesis of azo compounds
    A process is provided for synthesizing an azo compound, such as AIBN, by oxidation of a hydrazo compound using hydrogen peroxide.
  29. [29]
    Synthesis and Characterization of New Surface Active Azo Initiators ...
    The synthesis of a water soluble azo initiators from 2,2'-azodiisobutyronitrile (AIBN) was performed in three steps: reaction of dinitrile with aromatic ...
  30. [30]
    [PDF] RADICAL INITIATORS, REAGENTS AND SOLVENTS ... - IJRPC
    Azo compounds are decomposed by heat or absorbing light by the cis form to produce the corresponding alkyl radicals and nitrogen.Among the azo compounds, AIBN ...
  31. [31]
    Redox two-component initiated free radical and cationic ...
    Redox polymerizations use reducing and oxidizing agents to initiate radical or cationic polymerizations, reducing activation energies for mild temperature ...
  32. [32]
    Heterogenous Reversible Deactivation Radical Polymerization at ...
    Dec 9, 2019 · of MEA in water was highly efficient using a redox initiator, potassium persulfate/sodium ascorbate, at low temps. Simultaneous control of ...
  33. [33]
    Emerging Trends in Polymerization-Induced Self-Assembly
    For example, An's group employed a persulfate/ascorbate redox initiator system to prepare core-cross-linked nanogels at 30 °C. (111) Later, the same group ...
  34. [34]
    Radical entry mechanisms in redox-initiated emulsion polymerizations
    The mechanisms and kinetics of radical entry in emulsion polymerizations utilizing redox initiation are investigated using polymerization rate data.
  35. [35]
    Difunctional monomeric and polymeric photoinitiators: Synthesis ...
    Activities of these photoinitiators together with acetophenone (AP), benzophenone (BP) and Irgacure 2959 are investigated in photopolymerizations of hexane-1,6- ...
  36. [36]
    Water-Soluble Type I Radical Photoinitiators Dedicated to Obtaining ...
    Jun 17, 2024 · The photofragmentation mechanism of these compounds was also established by using electron paramagnetic resonance spectroscopy. The cytotoxicity ...
  37. [37]
  38. [38]
    Flavin mononucleotide in visible light photoinitiating systems for multiple-photocrosslinking and photoencapsulation strategies
    ### Summary of Flavin Mononucleotide (FMN) as Photoinitiator for Radical Polymerization
  39. [39]
    Thermal Homolysis - an overview | ScienceDirect Topics
    Thermal homolysis is defined as the process in which thermal energy causes the breaking of chemical bonds, resulting in the formation of free radicals, often ...
  40. [40]
    Benzoyl Peroxide - an overview | ScienceDirect Topics
    It decomposes with a rate constant of kd≈10−5 s−1 at 60 °C and has an activation energy Ea= 132 kJ/mol. At 65 °C it has a half lifetime of 10 hours, Table 2.
  41. [41]
    Benzoyl Peroxide | C14H10O4 | CID 7187 - PubChem
    The product has a 10 hour half-life at 71 °C. Ashford, R.D. Ashford's Dictionary of Industrial Chemicals. London, England: Wavelength Publications Ltd., 1994., ...
  42. [42]
    In situ ATR-FT-IR study of the thermal decomposition of diethyl ...
    The activation energy of the thermal decomposition of DEPDC was found to be 115 kJ/mol in heptane from 40 to 74 °C and 118 kJ/mol in scCO2 from 40 to 60 °C.
  43. [43]
    The power of light in polymer science: photochemical processes to ...
    Dec 24, 2013 · In principle, the first law of photochemistry dictates that only absorbed photons can be effective on causing chemical changes. As such, the ...
  44. [44]
    Predicting 3D printing critical parameters from Basic photochemical ...
    Critical energy is mainly controlled by quantum yield of photogenerated radicals. •. Depth of penetration is predicted accurately from Bouguer-Beer-Lambert law.
  45. [45]
    Photoinitiators for UV and visible curing of coatings: Mechanisms ...
    UV and/or visible light radiation is used to induce photochemical polymerization or crosslinking of a monomer, oligomer or prepolymer formulation.
  46. [46]
    Recent Advances in Type I Photoinitiators for Visible Light Induced ...
    Jun 27, 2022 · However, UV light shows several disadvantages such as a low curing depth due to a high absorption, scattering in most materials and, moreover, ...
  47. [47]
  48. [48]
    [PDF] Advancing Materials Chemistry through Photopolymerization
    Specifically, the major benefits of photopolymerization are described, namely temporal and spatial control, extrinsic control through manipulation of the light ...
  49. [49]
    Latest Advances in Highly Efficient Dye-Based Photoinitiating ...
    We present our latest achievements in the field of modern dye-based photoinitiating systems for the radical polymerization of acrylates.
  50. [50]
    Silyloxy-substituted anthraquinones as Type I photoinitiators for ...
    Jan 7, 2025 · We examined the ability of different silyloxyanthraquinones to initiate radical photopolymerization upon irradiation at 405 nm and found that some achieved ...Missing: post- | Show results with:post-
  51. [51]
    Iron-Based Redox Polymerization of Acrylic Acid for Direct Synthesis ...
    The polymerization is performed through a redox reaction of ferrous ion and AH2 (accelerants/reductants) with persulfate ion (initiator/oxidant). This method ...
  52. [52]
    None
    Summary of each segment:
  53. [53]
    Review article The Fenton reagents - ScienceDirect.com
    Numerous transition metal ions and their complexes in their lower oxidation states (L m M n+ ) were found to have the oxidative features of the Fenton reagent.Missing: paper | Show results with:paper
  54. [54]
    Free Radical Polymerization - Chemistry LibreTexts
    Jul 7, 2015 · When radical polymerization is desired, it must be started by using a radical initiator, such as a peroxide or certain azo compounds.
  55. [55]
    Radical chemistry in polymer science: an overview and ... - PMC - NIH
    The initiation step includes Equation 1 and Equation 2, when the thermal initiator decomposes into two small-molecule radical species and the first monomer adds ...
  56. [56]
    [PDF] Economomical Application of Initiator Mix Approach in suspension ...
    Jun 3, 2012 · In suspension PVC polymerization, initiators de- termine the rate of polymerization and significantly affect the process economics. With ...
  57. [57]
    50th Anniversary Perspective: RAFT Polymerization—A User Guide
    Two radicals (I) are introduced in a system containing ten monomers (yellow) and five RAFT agents (red R-group and blue Z–C(═S)S-group). Polymerization leads to ...<|control11|><|separator|>
  58. [58]
    Initiator Efficiency - an overview | ScienceDirect Topics
    Initiator efficiency is defined as the proportion of radicals that escape the solvent cage to form initiating radicals, represented by the symbol (ƒ).
  59. [59]
    Polymerization Initiator Market | Global Market Analysis Report - 2035
    Oct 7, 2025 · The polymerization initiator market is projected to grow from USD 1.5 billion in 2025 to USD 2.4 billion by 2035, at a CAGR of 4.9%.
  60. [60]
    Photopolymerization using bio-sourced photoinitiators
    Jul 25, 2023 · This review mainly summarizes the current progress from 2018–2022 in photopolymerization for bio-based photoinitiators/photoinitiator systems.
  61. [61]
    Hydrogen Bromide and Alkenes: The Peroxide Effect
    Jan 20, 2024 · The peroxide effect is when hydrogen bromide adds to alkenes in the presence of peroxides, causing anti-Markovnikov addition, which is the " ...
  62. [62]
    Recent Advances in Radical C–H Activation/Radical Cross-Coupling
    ### Summary of Recent Advances in Radical C-H Activation/Radical Cross-Coupling
  63. [63]
    4: Radical Reactions - Chemistry LibreTexts
    Nov 15, 2023 · We will use AIBN as our radical initiator and Bu3SnH as the precursor for our chain propagating radical, tributyltin radical.
  64. [64]
    Barton Decarboxylation - Organic Chemistry Portal
    The initiation of the Barton Decarboxylation ( Bu3Sn-H -> Bu3Sn. ) is effected with a radical initiator, and as with the Barton-McCombie Deoxygenation, the ...
  65. [65]
    Radical Reactions in Organic Synthesis: Exploring in-, on-, and with ...
    Triethylborane (Et3B) in the presence of a trace of O2 is an efficient radical initiator [50] and has been widely used for synthetic radical reactions. Et3B is ...Radical Reactions In Organic... · 6. Photocatalysis In Water... · 6.1. Photocatalyzed Radical...
  66. [66]
  67. [67]
    Benzoyl peroxide - NIOSH Pocket Guide to Chemical Hazards - CDC
    Benzoyl peroxide ; Exposure Limits. NIOSH REL. TWA 5 mg/m · OSHA PEL. TWA 5 mg/m ; Measurement Methods. NIOSH 5009. See: NMAM or OSHA Methods.
  68. [68]
    Methyl ethyl ketone peroxide - CDC
    MEKP is a colorless liquid with a characteristic odor, a strong oxidizing agent, and is shock sensitive. It can cause irritation, breathing issues, and skin ...
  69. [69]
    [PDF] AZODIISOBUTYRONITRILE CAS Number - NJ.gov
    HAZARD SUMMARY. * Azodiisobutyronitrile can affect you when breathed in and by passing through your skin. * Contact can irritate the eyes and skin.Missing: toxicity | Show results with:toxicity
  70. [70]
    [PDF] SAFETY AND HANDLING OF ORGANIC PEROXIDES
    When flammable gases or mists are released as part of the decomposition, there is always the potential danger of a fire or vapor phase explosion. Therefore, the ...
  71. [71]
    BENZOYL PEROXIDE | Occupational Safety and Health Administration
    ### OSHA PEL and Health Info for Benzoyl Peroxide
  72. [72]
    Occurrences and migration characteristics of photoinitiators in paper ...
    Research has found that currently there is no health risk to the human body from PIs ingested through dust and absorbed by skin exposed to printed magazines ( ...
  73. [73]
    [PDF] Storage of organic peroxides - Nouryon
    This brochure gives guidelines for the safe storage of organic peroxides in their original packaging. Organic peroxide storage requires two import- ant ...
  74. [74]
    Safe Storage and Handling of Organic Peroxides
    Feb 10, 2022 · Guidance for the storage of organic peroxides is provided by the National Fire Protection Association's Hazardous Materials Code (NFPA 400), as well as the ...
  75. [75]
    [PDF] Disposal of Liquid Organic Peroxides | American Chemistry Council
    Liquid organic peroxides are commonly disposed of by dilution and incineration, and should not be mixed with other waste. They can be diluted to less than 1% ...
  76. [76]
    Working Safely with Peroxides and Hydroperoxides in the Laboratory
    Jan 30, 2025 · Small amounts of peroxides can be neutralized with ferrous sulfate solutions or sodium thiosulfate under controlled conditions. Never dispose of ...
  77. [77]
    [PDF] Initiators for Thermoplastics - Nouryon
    These new peroxide formulations improve sustainability and safety during the production of PVC. Food contact-approved peroxide formulations have been developed ...
  78. [78]
    [PDF] Safe transport of organic peroxides EOPSG - United Initiators
    This brochure is a guide to give a set of recommendations regarding practical safety aspects in handling and transport operations of organic peroxides for road ...
  79. [79]
    Plan for stronger EU chemical industry - European Commission
    Jul 8, 2025 · The Commission has presented an action plan for the chemicals industry to strengthen the competitiveness and modernisation of this sector.Missing: radical initiators
  80. [80]
    Guide for Chemical Spill Response
    Carefully select suitable personal protective equipment. Make sure all skin surfaces are covered and that the gloves you use protect against the hazards posed ...
  81. [81]
    Chemical Spill Procedures | Office of Environmental Health and Safety
    Spill Response and Clean-up Procedures · Immediately alert area occupants and supervisor, and evacuate the area, if necessary. · Don personal protective equipment ...
  82. [82]
    [PDF] Spill Response and Prevention - EPA
    A key practice is creating and implementing a spill response and prevention plan, which should clearly state how to prevent spills, stop the source of a spill, ...