Fact-checked by Grok 2 weeks ago

Radical polymerization

Radical polymerization is a process in which free radicals serve as the reactive intermediates, enabling the addition of monomers—typically those containing carbon-carbon double bonds—to a growing chain. The involves four primary steps: , where radicals are generated from initiators such as peroxides or azo compounds via heat, light, or reactions; , the rapid addition of monomers to the radical chain end; termination, through radical coupling or to halt chain growth; and , which redistributes radicals to other molecules, influencing molecular weight. This versatile method tolerates a wide range of functional groups and solvents, making it suitable for , , , and polymerizations. As the most widely used industrial polymerization technique, radical polymerization accounts for approximately 50% of global polymer production, yielding key materials such as (LDPE), (PS), and poly() (PVC). Its cost-effectiveness and ability to process diverse vinyl and acrylic monomers have driven applications in plastics, coatings, adhesives, and rubbers, including specialized products like acrylic rubber and (PTFE). In recent advances, controlled radical polymerization techniques, such as (ATRP) and reversible addition-fragmentation chain transfer (), enable precise control over polymer architecture, molecular weight, and polydispersity, expanding uses into biomedical hydrogels, , and .

Introduction

Definition and Scope

Radical polymerization is a process in which the kinetic-chain carriers are free radicals, typically carbon-centered radicals at the growing chain end. In this method, polymerization proceeds through the successive addition of these radicals to monomers, such as those containing carbon-carbon double bonds, leading to the formation of a chain. The process is characterized by a slow rate of initiation, a fast rate of propagation, and a rapid rate of termination, enabling high-speed reactions that account for approximately half of production. Key features of radical polymerization include its tolerance for a wide range of functional groups and solvents, including water, which allows for versatile applications without stringent purification requirements. It is particularly applicable to monomers like styrene (yielding polystyrene), methyl methacrylate (for polymethyl methacrylate), acrylonitrile, vinyl acetate, and vinyl chloride (producing polyvinyl chloride). The reaction is highly exothermic, with enthalpy changes typically ranging from 8 to 20 kcal/mol per monomer addition, necessitating careful heat management in industrial settings. Due to the planar nature of the radical intermediate, which provides no inherent stereocontrol, the resulting polymers are generally atactic, lacking regular stereochemistry along the chain. In comparison to ionic polymerizations, such as anionic or cationic methods, radical polymerization offers greater simplicity and robustness, as it does not require strict conditions or sensitivity to impurities. While ionic approaches can achieve higher control over molecular weight and —often producing stereoregular polymers—radical polymerization exhibits higher exothermicity and typically yields less ordered atactic structures. This contributes to its widespread industrial adoption for despite limitations in precision.

Historical Development

The earliest documented observation of radical polymerization occurred in 1835 when French chemist Henri Victor Regnault exposed gas to and noted the formation of a white, solid polymeric material at the bottom of the container. This accidental discovery marked the first recognition of vinyl chloride's tendency to polymerize under light, though Regnault did not pursue its implications further. In the 1920s, German chemist advanced the understanding of through his seminal work on s, proposing in his 1920 paper "Über Polymerisation" that high-molecular-weight substances form via the sequential addition of monomer units into long covalent chains rather than mere associations of small molecules. Staudinger's model laid the groundwork for viewing as a controlled linkage process, influencing subsequent research on reaction mechanisms. The 1930s brought key milestones in elucidating the radical nature of the process, with American chemist Paul J. Flory describing the of vinyl in 1937 as a involving free s, where , , and termination steps govern the overall rate. Concurrently, German chemist Gerhard V. Schulz contributed to this framework through his 1939 studies on the , confirming bimolecular termination and providing experimental validation for Flory's theoretical predictions. These insights solidified the free as the dominant paradigm for understanding . During in the 1940s, the urgent need for spurred the development of techniques, particularly for rubber (SBR), which was produced on an industrial scale in the United States starting in 1942 to replace supplies disrupted by wartime blockades. This method, refined from earlier Buna-S processes, enabled high-yield production of durable elastomers via aqueous emulsions, demonstrating radical polymerization's scalability for strategic materials. Industrial adoption of radical polymerization expanded in the 1940s with batch processes for commodity plastics like and (PVC), which were conducted in stirred autoclaves to manage exothermic reactions and achieve consistent molecular weights. By the 1970s, the shift to continuous reactors, such as tubular and stirred-tank systems, improved efficiency and product uniformity for large-scale operations, reducing labor and energy costs in PVC and production. The post-1980s era saw a pivot toward controlled radical methods to achieve precise polymer architectures, exemplified by the introduction of (ATRP) in 1995 by Matyjaszewski and colleagues, who demonstrated living polymerization using catalysts to mediate halogen transfer and minimize termination. Building on this, in 1998, John Chiefari and coworkers at developed reversible addition-fragmentation (RAFT) polymerization, employing thiocarbonylthio compounds as chain transfer agents to enable controlled growth with low polydispersity across diverse monomers. These innovations marked a from conventional free processes to more versatile techniques for .

Fundamental Mechanism

Initiation

Initiation is the initial step in radical polymerization, wherein primary radicals are generated from an initiator and subsequently add to a molecule, forming the first propagating chain carrier (or active center). This process establishes the number of polymer chains and influences the overall polymerization rate and molecular weight distribution. The efficiency of initiation depends on the successful escape of primary radicals from the initiator's decomposition site to interact with the , avoiding recombination or other deactivation pathways. Several methods exist for generating initiating radicals, categorized by the energy source or chemical pathway employed. Thermal initiation commonly uses , such as benzoyl peroxide (BPO), which decompose at temperatures around 80–100°C to produce radicals via homolytic bond scission. Photochemical initiation involves photoinitiators like benzoin ethers, which cleave under (UV) light to generate radicals suitable for at ambient conditions. initiation employs pairs such as (e.g., ) with reducing agents like ferrous ions or amines, enabling radical formation at lower temperatures (often below 50°C) through mechanisms. Radiation-induced initiation, typically using gamma rays or electron beams, directly ionizes the or to produce radicals without added chemical initiators, offering precise control in specialized applications. The of initiators, particularly peroxides, proceeds through homolytic cleavage of the weak O–O ( ≈ 150 kJ/mol), yielding two alkoxy or aryloxy radicals that can further fragment or abstract . This unimolecular process follows , with the rate constant described by the : k_d = A \exp\left(-\frac{E_a}{RT}\right) where k_d is the decomposition rate constant, A is the , E_a is the (typically 100–170 kJ/mol for peroxides), R is the , and T is the absolute temperature. Initiator efficiency (f), defined as the fraction of primary radicals that successfully initiate chains rather than recombining or undergoing side reactions, ranges from 0.3 to 0.8. Factors reducing f include the —wherein geminate radicals recombine within the solvent cage before diffusing apart—and competing reactions like induced by propagating radicals. Higher concentrations and lower viscosities generally improve f by facilitating radical escape.

Propagation

In the propagation step of radical polymerization, a growing polymer chain radical adds to the double bond of a monomer molecule, forming a new bond and generating a new radical at the end of the extended chain. This process, represented as \ce{M_n^\bullet + CH2=CHX -> M_{n+1}^\bullet}, where \ce{M_n^\bullet} is the chain radical and X is the substituent on the monomer, repeats rapidly, allowing the chain to grow by hundreds or thousands of monomer units while maintaining a head-to-tail regioregularity. Each propagation addition is highly exothermic, with the enthalpy change \Delta H_p typically around -70 to -80 kJ/mol per unit; for styrene, this value is approximately -71 kJ/mol, contributing significantly to the overall heat release in the polymerization process. The of is governed by the k_p, which generally falls in the range of $10^2 to $10^4 L/mol·s depending on the and , and becomes diffusion-controlled at high conversions where chain mobility decreases. The for is R_p = k_p [\ce{M^\bullet}][\ce{M}], where [\ce{M^\bullet}] is the concentration of propagating radicals and [\ce{M}] is the concentration, highlighting the second-order dependence on these species. Factors influencing the propagation rate primarily stem from structure, such as the electronic nature of substituents that affect and energies; for instance, electron-rich monomers like acrylates exhibit higher k_p values compared to electron-deficient ones due to favorable polar interactions in the step. Additionally, the reactivity of the propagating remains largely unchanged along the chain length because each new end possesses similar to the previous one, avoiding variations in rate with chain size.

Termination

Termination in radical polymerization primarily occurs through bimolecular reactions between two propagating radicals, effectively without generating new radicals. The two dominant mechanisms are (), in which the radicals form a new carbon-carbon to yield a single polymer molecule, and , involving that produces one with a saturated end and another with an unsaturated end. These processes follow second-order kinetics with respect to radical concentration, expressed by the rate equation: R_t = 2 k_t [M^\bullet]^2 where R_t is the rate of termination, k_t is the termination rate constant (typically $10^7 to $10^9 L mol^{-1} s^{-1}), and [M^\bullet] denotes the concentration of propagating radicals. The relative prevalence of combination versus disproportionation varies by monomer structure; in styrene polymerization, combination predominates (approximately 80-90%), whereas in methacrylates such as , disproportionation is favored (about 70-75% at ambient temperatures). This mechanistic preference directly impacts polymer end-group functionality, with yielding difunctional chains suitable for further reactions and introducing variability in chain-end chemistry. Overall, termination depletes the population of active chains, thereby controlling the extent of polymerization and contributing to the molecular weight distribution, often resulting in polydispersity indices greater than 1.5 in conventional systems.

Chain Transfer

In radical polymerization, chain transfer occurs when a propagating polymer radical abstracts an atom, typically hydrogen, from a chain transfer agent (CTA), thereby terminating the growth of the current chain while generating a new radical capable of initiating another chain. This process relocates the radical activity without net loss of radicals, distinguishing it from termination, which destroys radicals altogether. The mechanism is represented as: \ce{R_n^\bullet + S-H -> R_n-H + S^\bullet} where R_n^\bullet is the propagating radical, S-H is the transfer agent, R_n-H is the dead polymer chain, and S^\bullet is the new radical. The rate of chain transfer is given by R_{tr} = k_{tr} [R_n^\bullet][S], where k_{tr} is the rate constant for transfer and [S] is the concentration of the transfer agent. Chain transfer agents include the itself, solvents, or deliberately added compounds. Transfer to often involves abstraction of labile , such as the allylic in α-methylstyrene, which exhibits a transfer constant C_{tr} = k_{tr}/k_p (where k_p is the ) of approximately 0.041 at 50°C. Solvents like serve as CTAs through donation, with C_{tr} \approx 3.4 \times 10^{-4} for . Added CTAs, such as thiols (e.g., n-dodecanethiol), are highly effective due to the weak S-H bond, yielding C_{tr} values ranging from 10 to 20 in many systems, and up to 21 for certain thiols (e.g., 1-butanethiol) in at 60°C. These quantify the relative likelihood of transfer versus , with values typically much less than 1 for and solvents but significantly higher for efficient CTAs. The primary effect of is to limit molecular weight by increasing the number of chains formed per initiating , as each event caps one chain and starts another. This introduces specific end-groups from the (e.g., thiol-derived groups) and can promote branching if occurs to chains via abstraction from the backbone. In systems with high rates, such as those using thiols, molecular weights are intentionally kept low, often below 10,000 g/mol, to produce oligomers or telechelic . competes with and termination but primarily influences chain length distribution without substantially affecting the overall rate. High is exploited in telomerization, a process where excess (telogen) reacts with (taxogen) to yield low-molecular-weight telomers with controlled functionality, such as in the production of fluorinated from vinylidene fluoride and mercaptans. This intentional use of efficient CTAs like or thiols enables precise control over chain length, typically achieving degrees of of 1–20, while maintaining the mechanism's versatility.

Polymerization Techniques

Conventional Free Radical Methods

Conventional free radical polymerization methods encompass , , , and techniques, which rely on or photochemical without mechanisms for radical deactivation, leading to polymers with broad molecular weight distributions. These approaches are widely used industrially due to their and , though they often require careful management of heat and to prevent uncontrolled reactions. Bulk polymerization involves heating pure monomer with a free radical initiator, such as benzoyl peroxide for styrene, to produce high-purity polymers without solvent residues. This method offers advantages like straightforward process design and maximal monomer concentration for high reaction rates, but it suffers from rapid viscosity increase as conversion progresses, complicating mixing and heat dissipation. A key challenge is the Trommsdorff-Norrish effect (also known as autoacceleration or gel effect), where rising viscosity reduces termination rates more than propagation, causing sudden rate acceleration, potential thermal runaway, and irregular polymer properties; this was first observed in the bulk polymerization of methyl methacrylate. Bulk methods are commonly applied to monomers like methyl methacrylate for cast sheets, but are limited to low conversions (typically below 30-50%) to mitigate these issues. Solution polymerization dissolves the and initiator in an organic , such as for styrene to form , enabling better control over reaction conditions. The reduces , facilitating stirring and efficient removal through or , which is particularly beneficial for exothermic polymerizations. However, it dilutes the concentration, lowering the overall rate and yield per volume, and necessitates downstream recovery, which can be energy-intensive and may introduce impurities if to occurs—, for instance, exhibits moderate activity. This technique is favored for producing soluble polymers like in applications requiring uniform molecular weights, though choice must minimize degradation of bulk properties during removal. Emulsion polymerization disperses water-insoluble monomer droplets (e.g., styrene or ) in water using above the , with water-soluble initiators like persulfates generating radicals that enter micelles to initiate , forming stable particles. predominates, where oligoradicals enter micelles swollen with , leading to compartmentalization that suppresses termination and yields high molecular weight polymers (often >10^6 g/mol) at rapid rates. The Smith-Ewart theory describes the kinetics, predicting the number of particles proportional to concentration (N_p ∝ [S]^{0.6}) and outlining three cases based on average radicals per particle (ñ): Case II (ñ = 0.5) applies to many systems, explaining the linear rate with conversion in Interval II. This method produces latexes for products like (PVC) via batch processes and styrene-butadiene rubber (SBR) through semibatch feeding to control composition, offering advantages in heat dissipation and easy product isolation but requiring removal. Suspension polymerization suspends droplets (typically 0.1-2 mm) in with mechanical agitation and stabilizers like , where oil-soluble initiators polymerize within droplets to form solid beads without micellar involvement. Stabilizers prevent coalescence, yielding spherical particles larger than those from (10-1000 μm vs. 50-500 nm), which simplifies recovery via and filtration without emulsion breaking. For production using styrene/, this process provides uniform beads for ion-exchange resins or , with advantages including effective through the aqueous phase and no residual , though it demands precise control of agitation to maintain droplet size distribution. Compared to , suspension yields coarser products but avoids latex stability issues, making it suitable for bead-based applications.

Controlled Radical Polymerization

Controlled radical polymerization encompasses a class of techniques designed to confer "living" or controlled characteristics upon free radical polymerization processes. These methods operate through the establishment of a rapid and reversible between a low concentration of active propagating radicals and a large excess of dormant polymer chains, which temporarily sequesters the radicals to suppress irreversible termination reactions. This minimizes side reactions, enabling the production of s with predictable molecular weights and narrow polydispersity indices (PDI) typically ranging from 1.1 to 1.5, a significant improvement over conventional free radical polymerization where PDI values often exceed 2. The primary techniques in controlled radical polymerization include nitroxide-mediated polymerization (NMP), (ATRP), and reversible addition-fragmentation (RAFT) polymerization. NMP, pioneered in the 1980s with early demonstrations of nitroxide trapping in radical systems, achieved practical control in the early 1990s using stable nitroxides such as 2,2,6,6-tetramethylpiperidin-1-oxyl () as mediators; these species reversibly couple with carbon-centered s to form alkoxyamines, effectively deactivating the chains. , introduced in 1995, employs a catalyst, typically (I) complexes with bipyridine or similar ligands, in conjunction with alkyl halide initiators to facilitate reversible halogen atom transfer between the metal center and the propagating , maintaining the dormant species as alkyl halides. polymerization, developed in 1998, utilizes thiocarbonylthio compounds as agents (CTAs) to mediate across a broad range of monomers without requiring metal catalysts. Mechanistically, these techniques rely on a general dormant/active that supports controlled . In the active , a propagating P_n^\bullet adds (M) to form P_m^\bullet, which then reversibly deactivates to an inactive species, such as an alkoxyamine in NMP, an alkyl in ATRP, or an adduct radical in RAFT. For RAFT specifically, the process involves an addition-fragmentation : the propagating radical adds to the CTA (R-S-C(=S)-Z) to form an radical, which fragments to release a new propagating radical (R^\bullet) and a thiocarbonylthio-macromolecule; this macro-CTA then participates in further reversible transfers, ensuring equal chain growth opportunities. These methods offer distinct advantages, including the ability to synthesize well-defined block copolymers through sequential monomer addition, leveraging the retained dormant chain ends for reactivation, and the preservation of end-group functionality for post-polymerization modifications such as or conjugation. However, RAFT polymerization can introduce challenges, as the sulfur-containing end-groups often impart a yellow-to-red color and a characteristic to the resulting polymers, necessitating additional purification steps for certain applications.

Kinetics and Molecular Weight Control

Rate Laws and Mechanisms

The kinetics of radical polymerization are governed by the rates of , , and termination steps, leading to specific rate laws that describe the overall polymerization rate R_p, defined as the rate of consumption -\frac{d[M]}{dt}. Under typical conditions, the concentration of propagating s [M•] is low and varies rapidly, necessitating the steady-state approximation, which assumes that the rate of radical formation equals the rate of radical destruction, such that \frac{d[M•]}{dt} = 0. This approximation yields the radical concentration as [M•] = \left( \frac{R_i}{2 k_t} \right)^{1/2}, where R_i is the initiation rate, k_t is the termination rate constant, and the factor of 2 accounts for the bimolecular nature of termination. The propagation rate is then R_p = k_p [M•][M], where k_p is the propagation rate constant and [M] is the monomer concentration. Substituting the steady-state expression for [M•] gives the fundamental rate law R_p = k_p \left( \frac{R_i}{2 k_t} \right)^{1/2} [M]. This shows that R_p is first-order in [M] and half-order in R_i, reflecting the square-root dependence on radical concentration. For thermal initiation with a decomposable initiator, the initiation rate is R_i = 2 f k_d [I], where f is the initiator efficiency (typically 0.3–0.8, accounting for radicals lost to side reactions), k_d is the initiator decomposition rate constant, and [I] is the initiator concentration. The decomposition rate k_d follows Arrhenius behavior, increasing exponentially with temperature, which accelerates initiation and thus R_p. Combining this with the propagation rate law yields the overall expression R_p = \frac{k_p}{(2 k_t)^{1/2}} (f k_d [I])^{1/2} [M], assuming negligible chain transfer. This half-order dependence on [I] is a hallmark of radical polymerization kinetics. In photopolymerization, initiation occurs via photoinitiator absorption of light, making R_i proportional to light intensity I_0, typically R_i = 2 f \phi \epsilon [PI] I_0, where \phi is the , \epsilon is the absorptivity, and [PI] is the concentration; higher I_0 thus increases R_p via enhanced radical generation, though excessive intensity can promote side reactions. Deviations from the ideal rate law arise at higher conversions due to physical effects. Autoacceleration, or the Trommsdorff-Norrish effect, manifests as a sudden increase in R_p because rising polymer concentration elevates solution , creating local microenvironments with higher effective [M] near reaction sites. More critically, the gel effect reduces k_t as radical termination becomes diffusion-limited in the viscous medium, amplifying [M•] and thus R_p beyond the steady-state prediction; this can lead to uncontrolled exotherms in bulk polymerizations.

Molecular Weight and Distribution

In radical polymerization, the number-average degree of polymerization (\overline{DP}_n), which represents the average number of units per , is governed by the relative rates of and -stopping processes. Specifically, \overline{DP}_n is given by the of the rate (R_p) to the combined rates of termination (R_t) and (R_{tr}): \overline{DP}_n = \frac{R_p}{R_t + R_{tr}}. This expression arises because each step adds one unit to a growing , while termination or transfer events halt that 's , determining its final length. Substituting the kinetic expressions for these rates yields a more detailed form: R_p = k_p [M][M^\bullet], R_t = 2k_t [M^\bullet]^2, and R_{tr} = \sum k_{tr} [S][M^\bullet], where k_p is the propagation rate constant, [M] the monomer concentration, [M^\bullet] the concentration of propagating radicals, k_t the termination rate constant, and the sum encompasses chain transfer rate constants k_{tr} and concentrations [S] of transfer agents (such as monomer, solvent, or added agents). Thus, \overline{DP}_n \approx \frac{k_p [M]}{2 k_t [M^\bullet] + \sum k_{tr} [S]}. Polymerization conditions, including initiator concentration (which influences [M^\bullet]), monomer concentration, and temperature (affecting rate constants), directly control \overline{DP}_n by modulating these terms; higher [M] or lower [M^\bullet] typically increases chain length. The polydispersity index (PDI, defined as \overline{M}_w / \overline{M}_n) in conventional radical polymerization typically ranges from 1.5 to 2, reflecting the breadth of the molecular weight distribution due to the stochastic nature of termination events. This value arises from the equal probability of chain growth versus stopping at each step, leading to a most probable (Flory-Schulz) distribution where the weight fraction of chains with degree of polymerization x is w_x = x (1 - p)^{x-1} p^{x-1} (with p the probability of propagation); for high conversion, PDI approaches 1.5 if termination occurs exclusively by combination or 2 by disproportionation. In contrast, controlled radical methods achieve lower PDI values (often 1.1–1.5) by minimizing termination and enabling more uniform chain growth. When chain transfer dominates over termination (R_{tr} \gg R_t), the molecular weight is primarily controlled by transfer agents, as described by the Mayo equation: \frac{1}{\overline{DP}_n} = \frac{1}{\overline{DP}_{n0}} + C_{tr} \frac{[S]}{[M]}, where \overline{DP}_{n0} is the degree of polymerization without transfer, and C_{tr} = k_{tr}/k_p is the transfer constant. This linear relationship allows precise tuning of chain length by adjusting [S], such as adding thiols (with high C_{tr} \approx 10^{-2} to 1) to limit \overline{DP}_n in industrial processes like polystyrene production. The Flory-Schulz distribution persists in transfer-dominated systems, maintaining the characteristic PDI of 2, but with a shifted average length.

Thermodynamics and Process Considerations

Thermodynamic Principles

Radical polymerization is governed by the thermodynamic favorability of the step, which is highly exothermic with an change (ΔH) typically ranging from -20 to -100 kJ/mol per unit. This exothermicity stems from the net energy release during bond reorganization, where the C=C in the (bond energy ≈ 610 kJ/mol) is effectively broken, and two C-C σ bonds (total ≈ 710 kJ/mol) are formed in the growing chain. The process is accompanied by a negative change (ΔS ≈ -100 to -120 J mol⁻¹ K⁻¹), arising from the loss of translational and rotational as discrete molecules are constrained within the chain. The ceiling temperature (T_c) represents the equilibrium point where the forward propagation rate equals the reverse depropagation rate, rendering net polymerization zero at that monomer concentration. This temperature is derived from the condition ΔG = 0, yielding
T_c = \frac{\Delta H}{\Delta S}
where ΔH and ΔS are the enthalpy and entropy changes for propagation. For instance, bulk styrene exhibits a T_c of 310°C, while methacrylates generally have T_c values around 220°C. Above T_c or at low monomer concentrations, depropagation dominates, limiting polymer formation.
Although is thermodynamically irreversible under standard conditions (e.g., ), reversibility becomes relevant at elevated temperatures approaching T_c or for monomers with inherent , such as α-methylstyrene (T_c ≈ 60°C). In these cases, the equilibrium monomer concentration [M]_e increases, potentially leading to if the system is heated sufficiently.

Heat Transfer and Reactor Design

Radical polymerization reactions are highly exothermic, with typical enthalpies of polymerization ranging from -30 to -100 kJ/mol, releasing significant heat that can lead to if not properly managed. This exothermicity poses substantial challenges in process control, as uncontrolled heat accumulation can cause rapid temperature increases, potentially exceeding the of the reaction mixture and resulting in buildup or vessel rupture. Autoacceleration, also known as the or Trommsdorff-Norrish effect, further exacerbates local heating by increasing the rate at higher conversions due to diffusion-limited termination of radicals in the viscous medium, creating hotspots that intensify difficulties. Effective cooling strategies are essential to dissipate the generated during radical polymerization. In batch s, jacket cooling with circulating provides direct heat removal from the reactor walls, maintaining isothermal or near-isothermal conditions for small-scale operations. For continuous processes, tubular flow reactors facilitate heat management through high surface-to-volume ratios and the introduction of cold feeds at multiple points, which absorb and distribute heat along the reactor length. In , inert diluents such as hydrocarbons or alcohols serve as heat sinks, lowering the reaction and enhancing convective to cooling surfaces, thereby preventing localized overheating. Reactor configurations are selected based on production scale, control needs, and heat management requirements. Batch reactors offer simplicity and flexibility for small-scale or specialty production, allowing straightforward charging and discharge while relying on cooling for removal. Semi-batch reactors improve by enabling controlled addition, which limits the instantaneous release and mitigates exotherm peaks, commonly used in processes requiring precise composition gradients. Continuous stirred-tank reactors (CSTRs) achieve steady-state operation with uniform mixing, supporting large-scale production by balancing continuous feed and withdrawal to stabilize and . Safety in radical polymerization design hinges on predicting and limiting potential temperature excursions. The adiabatic temperature rise, ΔT_ad, which estimates the maximum temperature increase under no heat removal, is calculated as ΔT_ad = (-ΔH [M]_0) / (C_p ρ), where ΔH is the of polymerization, [M]_0 the initial concentration, C_p the , and ρ the ; this metric guides reactor sizing and cooling capacity to prevent . Monomer feed strategies, such as starved-feed in semi-batch modes, further enhance safety by restricting the monomer inventory at any time, thereby capping the heat release rate and keeping ΔT below critical thresholds.

Polymer Structure and Reactivity

Stereochemistry

In radical polymerization, the propagating intermediate is typically planar at the carbon-centered site, enabling rapid rotation and minimal steric bias during monomer addition. This lack of results in polymers with predominantly atactic , where the configuration of adjacent chiral centers is randomly distributed along the chain. Tacticity describes the stereochemical arrangement of repeat units in a chain and is classified into three primary types: isotactic, where all adjacent stereocenters have the same relative configuration (meso s, m); syndiotactic, featuring alternating configurations (racemic diads, r); and atactic, characterized by a random mixture of m and r diads. Conventional free radical polymerization of typical vinyl monomers, such as styrene (yielding , PS) and methyl methacrylate (yielding poly(methyl methacrylate), PMMA), produces largely atactic polymers due to the absence of external stereocontrol. For PS, the meso diad probability (P_m) is approximately 0.36 across a wide range, indicating a slight syndiotactic (P_r ≈ 0.64) but overall random microstructure. In contrast, PMMA exhibits a stronger inherent syndiotactic preference, with P_m ≈ 0.17 at 30 °C (P_r ≈ 0.83), though still considered atactic relative to highly stereoregular forms produced by other methods. Exceptions to this random stereochemistry arise under specific conditions that subtly influence radical approach or conformation. Lowering the polymerization temperature can enhance syndiotacticity in monomers like or by restricting rotational freedom and amplifying steric interactions. In alternating copolymers formed via , such as styrene-maleic anhydride, the obligatory 1:1 sequencing imposes constraints on the propagating . The of radical polymers is commonly assessed using () spectroscopy, particularly ^{1}H and ^{13}C NMR, which resolve triad sequences—mm (isotactic), mr (heterotactic), and rr (syndiotactic)—based on differences in the backbone or side-chain protons. For instance, in PMMA, the methylene ^{13}C signals at approximately 54-55 distinguish these triads, allowing quantification of P_m and P_r from peak intensities via the relations P_m = [mm] / ([mm] + [mr]/2 + [rr]), and deviations from Bernoullian statistics indicate penultimate or antepenultimate effects. This analysis is crucial because tacticity profoundly impacts polymer properties: atactic configurations typically yield amorphous materials with lower temperatures and no crystallinity (e.g., atactic melts near 0 °C), whereas even modest syndiotactic content enhances chain packing, , and mechanical strength compared to fully isotactic crystalline forms.

Copolymerization Reactivity

In radical copolymerization, the reactivity ratios r_1 and r_2 quantify the relative reactivities of propagating radicals toward the two monomers. Specifically, r_1 = k_{11} / k_{12}, where k_{11} is the rate constant for addition of monomer 1 to a radical chain ending in monomer 1, and k_{12} is the rate constant for addition of monomer 2 to the same radical; r_2 = k_{22} / k_{21} is defined analogously. These parameters indicate the preference of each radical type for homopropagation versus cross-propagation during chain growth. The nature of the copolymer sequence distribution depends on the values of r_1 and r_2. An ideal copolymerization occurs when r_1 = r_2 = 1, resulting in random monomer incorporation proportional to the feed without bias toward blocks or alternation. Alternating tendencies arise when both r_1 < 1 and r_2 < 1, favoring cross-addition due to lower homopropagation rates. An azeotropic , where the instantaneous copolymer composition matches the monomer feed and remains constant throughout polymerization, exists when r_1 r_2 = 1. The instantaneous copolymer composition is described by the Mayo-Lewis copolymer equation: \frac{d[M_1]}{d[M_2]} = \frac{r_1 [M_1]/[M_2] + 1}{1/r_2 + [M_1]/[M_2]} This differential equation predicts how the polymer chain composition drifts with conversion based on initial feed ratios and reactivity ratios. For example, in the free radical copolymerization of (monomer 1) and (monomer 2) in bulk at 60°C, r_1 = 0.4 and r_2 = 6.0, leading to gradient copolymers enriched in acrylate units early in the reaction and shifting toward styrene-rich sequences at higher conversions. To predict reactivity ratios for dissimilar monomers without extensive experimentation, the Alfrey-Price Q-e scheme assigns parameters reflecting resonance stabilization (Q) and polar effects (e) to each monomer and radical. The scheme expresses r_1 = (Q_1 / Q_2) \exp[-e_1 (e_1 - e_2)] and r_2 = (Q_2 / Q_1) \exp[-e_2 (e_2 - e_1)], with as the reference monomer having Q = 1.0 and e = -0.8. This empirical model accounts for electronic influences on addition rates, enabling estimation of copolymer behavior for new monomer pairs.

Applications and Industrial Use

Key Polymers and Processes

Poly(styrene) (PS) is one of the most widely produced polymers via radical polymerization, typically conducted through bulk or suspension processes using initiators such as benzoyl peroxide (BPO). In bulk polymerization, the reaction occurs without solvent, relying solely on the monomer, initiator, and optional chain-transfer agents, which allows for high monomer concentration but requires careful control to manage the exothermic reaction. Suspension polymerization disperses styrene droplets in water with stabilizers, enabling easier heat dissipation and producing bead-like particles suitable for further processing. PS finds extensive use in expanded foam for insulation and packaging due to its lightweight and insulating properties, as well as in rigid packaging applications for its clarity and moldability. Poly(methyl methacrylate) (PMMA) is synthesized by radical polymerization in solution or bulk methods, employing peroxide initiators like benzoyl peroxide to achieve high molecular weights. The bulk process involves direct polymerization of the monomer, while solution polymerization uses solvents to moderate viscosity and facilitate heat transfer. The resulting atactic structure of PMMA, characterized by irregular stereochemistry from the free radical mechanism, renders it amorphous and contributes to its exceptional optical clarity, with transmittance exceeding 90% in the visible spectrum, making it ideal for lenses and glazing. Poly(vinyl chloride) (PVC) is predominantly manufactured through suspension polymerization, where vinyl chloride monomer is dispersed in water with suspending agents and polymerized using redox initiators, such as combinations of organic peroxides and reducing agents, to generate radicals at moderate temperatures around 50–60°C. This process yields porous resin particles that are easily separated and processed into various forms. However, PVC is prone to dehydrochlorination during polymerization and subsequent processing, where elimination of HCl leads to chain unsaturation, discoloration, and potential autocatalytic degradation, necessitating stabilizers and precise temperature control to mitigate these risks. Polyacrylonitrile (PAN) is produced via aqueous suspension radical polymerization for applications in textile fibers, where the process involves dispersing acrylonitrile in water with stabilizers and initiators like persulfates to form a polymer that is dissolved in solvents and wet-spun into fibers with high tensile strength. Low-density polyethylene (LDPE) is obtained through high-pressure radical polymerization of ethylene, pioneered by Imperial Chemical Industries (ICI) in the 1930s at pressures of 1000–3000 bar and temperatures of 150–300°C using peroxide initiators, resulting in branched chains that impart flexibility. Styrene-butadiene rubber (SBR) is synthesized primarily by emulsion polymerization, involving the copolymerization of styrene and butadiene in aqueous micelles with redox initiators at low temperatures (around 5–10°C for cold emulsion) to produce a latex that is coagulated into rubber, offering good abrasion resistance and processability. This process accounts for a significant portion of the synthetic rubber market, with SBR comprising approximately 40% of global production due to its versatility in tire manufacturing. Acrylic rubber, a copolymer of acrylates such as ethyl acrylate with cure-site monomers, is produced via emulsion radical polymerization in aqueous systems using redox initiators, resulting in elastomers with excellent oil and heat resistance for seals and gaskets in automotive applications. Polytetrafluoroethylene (PTFE) is manufactured through suspension radical polymerization of tetrafluoroethylene (TFE) in water with fluorosurfactants and persulfate initiators under moderate pressure and temperature, yielding granular or fine powder resins known for their non-stick and chemical-resistant properties in coatings and seals.

Commercial Significance

Radical polymerization underpins a substantial portion of the global plastics industry, with production exceeding 200 million metric tons annually as of 2023, accounting for approximately 50% of all industrial polymers. This dominance stems from its role in manufacturing key commodity plastics such as , , , and acrylates, which together represent over half of thermoplastic production volumes. The process's economic viability arises from its low-cost initiators, tolerance for impurities, and scalability to large-scale reactors, enabling efficient output for mass-market applications. Key economic drivers include the accessibility of radical methods, which facilitate high-volume production at competitive prices; for instance, the global PS market was valued at around USD 34 billion in 2023, driven by demand in consumer goods and packaging. Similarly, PVC production reached approximately 50 million tons in 2023, with pipes and fittings comprising about 44% of its applications, particularly in construction where durable, cost-effective piping systems account for a significant share of infrastructure projects. These factors have solidified radical polymerization's position as a cornerstone of the USD 800 billion-plus polymers sector, supporting industries from automotive to electronics through reliable supply chains. In societal terms, radical polymerization-derived materials permeate daily life, with PS foam serving as a primary insulator and protective layer in food packaging, safeguarding perishable goods during transport and storage due to its lightweight and thermal properties. Acrylate-based adhesives, produced via radical processes, enable versatile bonding in consumer products like tapes and labels, enhancing durability in packaging and assembly applications. However, environmental concerns have escalated, as PS contributes notably to microplastic pollution; fragmented PS particles persist in ecosystems, entering food chains and posing risks to marine life and human health through bioaccumulation. The industry's evolution reflects broader technological and societal shifts, originating in the 1940s with wartime imperatives for synthetic rubber via emulsion radical polymerization, which addressed natural rubber shortages during World War II. By the 2020s, focus has pivoted toward recyclability, with innovations in chemical depolymerization enabling recovery of monomers from PS and PVC waste streams to mitigate plastic pollution and support circular economies, aligning production with sustainability regulations.

Recent Developments

Advances in Initiators and Catalysts

Significant advances in initiators for radical polymerization since 2010 have focused on photoinitiators that enable mild, visible-light-mediated processes, enhancing control and reducing energy requirements. A notable example is the use of fac-Ir(ppy)₃ as a photoredox catalyst in , which allows for the controlled polymerization of methacrylates under visible light irradiation at room temperature. This system operates with low catalyst loadings (ppm levels) and provides temporal control by toggling the light source, yielding polymers with narrow molecular weight distributions (Đ < 1.2) and high chain-end fidelity. Similarly, organocatalysts have emerged to eliminate metal components entirely, addressing toxicity concerns in biomedical applications. For instance, phenothiazine derivatives serve as organic photoredox catalysts in metal-free , enabling the synthesis of well-defined polymers from acrylates and methacrylates in the presence of visible light, with molecular weights up to 10,000 g/mol and low dispersities. Improvements in catalysts have centered on enhancing efficiency through ligand design, particularly for copper-based ATRP systems. Modified bipyridine ligands, such as 4,4'-di(5-nonyl)-2,2'-bipyridine (dNbpy), combined with reducing agents in activators regenerated by electron transfer (ARGET) ATRP, have reduced copper loadings to 50-100 ppm while maintaining control over polymerization kinetics and polymer architecture. These ligands accelerate the activation-deactivation equilibrium, allowing for the production of high-molecular-weight polymers (M_n > 100,000 g/mol) with minimal catalyst residue. Organometallic-mediated radical polymerization (OMRP) represents another precision-oriented advancement, utilizing main-group or transition-metal complexes like to reversibly trap radicals, achieving low polydispersities (Đ ≈ 1.1-1.3) for a range of monomers including and styrene. OMRP's tunability via metal-ligand interactions provides superior control in challenging media, such as aqueous or polar solvents. Hybrid systems integrating photochemistry with chain transfer agents have further expanded spatiotemporal control. Photoinduced reversible addition-fragmentation chain transfer (Photo-RAFT) polymerization, developed around 2015, combines visible light with dithioester chain transfer agents (CTAs) and organic photocatalysts like eosin Y, enabling oxygen-tolerant, living polymerization of acrylamides with precise spatial patterning via masked irradiation. This approach yields block copolymers with quantitative chain-end retention and dispersities below 1.2, facilitating applications in 3D printing and microstructured materials. Overall, these innovations reduce toxicity by minimizing metal use—such as in metal-free O-ATRP—and improve energy efficiency through low-intensity light sources, as demonstrated in aqueous photoATRP for synthesizing biocompatible polymers like poly(ethylene glycol) acrylates with Cu levels below 10 ppm. In 2025, further advances include stereoregular radical polymers enabling selective spin transfer over long distances for spintronic applications. Verdazyl radical polymers have demonstrated high spin mixing conductance, advancing organic spintronics.

Sustainable and Green Approaches

Efforts to enhance the sustainability of radical polymerization have increasingly focused on water-based systems, which minimize the use of volatile organic solvents and reduce environmental impact. Miniemulsion polymerization, a variant of emulsion techniques, disperses hydrophobic monomers in aqueous media using surfactants and co-stabilizers to form stable submicron droplets, enabling efficient polymerization with significantly lower solvent consumption compared to traditional organic-phase methods. This approach has been particularly effective for producing latexes and coatings, achieving high monomer conversion rates while aligning with green chemistry principles by promoting water as the primary reaction medium. Additionally, research in the 2020s has explored enzyme-mimetic initiators, such as peroxidase mimics, to catalyze reversible deactivation radical polymerization (RDRP) under mild aqueous conditions, mimicking biological radical control mechanisms to yield well-defined polymers with reduced energy input and byproduct formation. The integration of bio-based monomers represents a key strategy for replacing petroleum-derived feedstocks with renewable alternatives, thereby lowering the of radical polymerization processes. Itaconic acid, derived from of carbohydrates, undergoes free polymerization to form poly(itaconic acid) and its copolymers, which exhibit biodegradability and pH-responsive properties suitable for applications like superabsorbent polymers and systems. For instance, dialkyl itaconates copolymerize with conventional monomers via mechanisms, demonstrating reactivity ratios that support tunable compositions for sustainable materials. Similarly, such as β-myrcene and d-limonene serve as green substitutes for styrene in radical copolymerizations, yielding elastomers and porous monoliths with mechanical properties comparable to synthetic analogs while being sourced from plant essential oils. These bio-based polymers degrade more readily in natural environments, addressing the persistence of traditional styrenic materials. Recycling integration through radical depolymerization has emerged as a vital approach to close the loop in polymer lifecycles, enabling monomer recovery from post-consumer waste. For poly(methyl methacrylate) (PMMA), thermal radical unzipping initiates chain-end radicals that propagate depolymerization, reverting the polymer to nearly pure methyl methacrylate monomer at temperatures as low as 250°C when using controlled radical polymerization precursors. This process achieves up to 92% monomer yield in bulk conditions without solvents, facilitating scalable chemical recycling that outperforms conventional pyrolysis in selectivity and energy efficiency. Energy reduction strategies in radical polymerization emphasize flow chemistry combined with LED photopolymerization, which utilize for precise and lower power consumption than traditional UV or thermal systems. In continuous flow reactors, LED-driven photoinduced RDRP enables rapid polymerization of acrylates and methacrylates with spatiotemporal control, reducing reaction times to hours (e.g., 90% in 3 hours) compared to batch methods, with improved efficiency through efficient light penetration and heat dissipation. This setup supports metal-free conditions, minimizing waste and enabling on-demand production scales. In contexts, controlled radical has been applied to upcycle polystyrene waste, functionalizing it with anhydrides or bio-monomers to create value-added composites, as demonstrated in European initiatives from 2022 onward that enhance recyclate purity for applications.

References

  1. [1]
    Radical chemistry in polymer science: an overview and recent ... - NIH
    In this review, we discuss radical chemistry in polymer science from four interconnected aspects. ... 1 Key features of radical polymerization: Radical ...
  2. [2]
    Radical Polimerization - an overview | ScienceDirect Topics
    Radical polymerization is defined as a type of chain-growth polymerization where the growth of the polymer occurs exclusively through reactions between the ...Missing: paper | Show results with:paper
  3. [3]
    Atom transfer radical polymerization | Nature Reviews Methods ...
    Jan 9, 2025 · Atom transfer radical polymerization (ATRP) is a synthetic method that harnesses radical polymerization to yield well-defined polymers of ...Missing: paper | Show results with:paper
  4. [4]
  5. [5]
  6. [6]
    Polymerization Reactions
    At first glance we might expect the product of the free-radical polymerization of ethylene to be a straight-chain polymer. ... atactic polymer with no commercial ...
  7. [7]
    Polyvinyl Chloride - The Plastics Historical Society
    Dec 6, 2016 · Vinyl chloride was first produced in 1835 by HV Regnault who also observed its polymerisation by sunlight. Ostromislenski investigated it ...
  8. [8]
    Technical progresses for PVC production - ScienceDirect.com
    This article covers the recent developments of manufacturing processes of PVC. In the early 1930s industrial process of PVC has been developed in Germany and ...
  9. [9]
    Celebrating 100 years of “polymer science”: Hermann Staudinger's ...
    Abstract. The first translation into English of Hermann Staudinger's seminal 1920 paper 'On polymerisation', to mark its 100 year anniversary.
  10. [10]
    Hermann Staudinger Foundation of Polymer Science - Landmark
    In a paper entitled "Über Polymerisation," Staudinger presented several reactions that form high molecular weight molecules by linking together a large ...Missing: seminal | Show results with:seminal
  11. [11]
    Origins and Development of Initiation of Free Radical Polymerization ...
    Feb 11, 2010 · In 1937 Flory described the kinetics of the vinyl polymerization as a chain reaction with participation of free radicals and postulated that two ...
  12. [12]
    U.S. Synthetic Rubber Program - National Historic Chemical Landmark
    They discovered in 1929 that Buna S (butadiene and styrene polymerized in an emulsion), when compounded with carbon black, was significantly more durable than ...Missing: SBR | Show results with:SBR
  13. [13]
    Emulsion Styrene Butadiene Rubber (E-SBR) - AZoM
    Feb 12, 2003 · In the 1930's, the first emulsion polymerized styrene- butadiene rubber known as Buna S was prepared by I. G. Farbenindustrie in Germany. The ...
  14. [14]
    Polymerization Processes, 2. Modeling of Processes and Reactors
    Oct 15, 2011 · Various types of polymerization reactors have been proposed both for batch and continuous processes. A commonly used industrial reactor for a ...
  15. [15]
    Controlled/"Living" Radical Polymerization. Halogen Atom Transfer ...
    Article November 1, 1995. Controlled/"Living" Radical Polymerization ... The First Dive into the Mechanism and Kinetics of ATRP. Macromolecules 2020 ...Missing: original | Show results with:original
  16. [16]
  17. [17]
    Process hazard and decomposition mechanism of benzoyl peroxide ...
    Feb 15, 2023 · The primary radicals generated during thermal decomposition of BPO have high reactivity and can be applied to initiate free radical reactions.
  18. [18]
    Photochemistry of benzoin ethers. Type I cleavage by low energy ...
    A latent photoreaction predominates within water-soluble calixarenes: photochemistry of benzoin alkyl ethers. Chemical Communications 2005, 21 (32) , 4056.Missing: initiation | Show results with:initiation
  19. [19]
    Persulfate - Redox Tech
    Persulfate is a strong oxidant that has been widely used for initiating emulsion polymerization reactions, clarifying swimming pools, hair bleaching, micro- ...
  20. [20]
    γ-Ray Induced Polymerization - Stanford
    Mar 22, 2012 · Polymerization using γ-rays proceeds via the same mechanism as the free radical initiation mechanism described previously, but without the use of chemical ...<|control11|><|separator|>
  21. [21]
    [PDF] MECHANISMS OF DECOMPOSITION OF INITIATORS
    unimolecular decay of molecules on gas pressure. When the pressure is growing, the rate constant of decomposition kd → k∞, and k∞ = k#e. −E/RT. (1.1).
  22. [22]
    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).
  23. [23]
    Cage Effect - an overview | ScienceDirect Topics
    Owing to the cage effect, the rate of chain initiation is generally less than twice the rate of production of primary radicals.
  24. [24]
    [PDF] Mechanism and Kinetics of Free Radical Chain Polymerization
    Jan 29, 2001 · □ Mechanism of Radical Chain Polymerization. ◇ Fundamental steps (initiation, propagation, termination). ◇ Kinetic chain length and ...
  25. [25]
    Probing into Styrene Polymerization Runaway Hazards: Effects of ...
    May 3, 2019 · First of all, the styrene polymerization reaction is relatively highly exothermic with a heat generation at around 71 kJ·mol–1. At the same time ...
  26. [26]
    Chapter 5 Propagation - ScienceDirect.com
    The heat of styrene polymerization and of its ringsubstituted derivatives is considerably lower, by about 20 kJ mol-'.
  27. [27]
    The mechanism of the propagation step in free-radical ...
    In addition, the studies reviewed in this paper indicate that a number of factors, including polar interactions, radical stabilisation effects, direct ...
  28. [28]
    CHAPTER 1: Kinetics and Thermodynamics of Radical Polymerization
    The reactions between two carbon-centered radicals generally give a mixture of disproportionation and combination. Which termination mode dominates depends ...
  29. [29]
    Radical Polymerization of Acrylates, Methacrylates, and Styrene
    The changes in reaction enthalpy and entropy are both negative values for free radical polymerization. Thus, there is a ceiling temperature (Tc) above which the ...
  30. [30]
    Detailed analysis of termination kinetics in radical polymerization
    In the field of free radical polymerization, exploring the relationship between polymeric radicals and the termination coefficient is an intriguing subject.
  31. [31]
    Rate Constants in Free Radical Polymerization. III. Styrene1
    Evaluation of the Rate Coefficient of Bimolecular Chain Termination kt. ... Propagation and Termination Constants in Free Radical Polymerization. 1999https ...
  32. [32]
    The termination reaction in radical polymerizations. II ...
    It has been found that polystyrene radicals undergo combination at 60°. In the case of methyl methacrylate, disproportionation occurs about 6 times as ...
  33. [33]
    Termination Mechanism in the Radical Polymerization of Methyl ...
    A novel method to determine the termination mechanism of radical polymerization, ie, the selectivity between disproportionation (Disp) and combination (Comb), ...
  34. [34]
    Control of the Termination Mechanism in Radical Polymerization by ...
    Nov 14, 2016 · The termination mechanism of radical polymerization, that is, disproportionation (Disp) versus combination (Comb), determines the chain length and end-group ...Abstract · Introduction · Results and Discussion · Conclusion
  35. [35]
    None
    Below is a merged summary of "Chain Transfer in Free Radical Polymerization" from George Odian's *Principles of Polymerization* (4th ed., Wiley, 2004), consolidating all information from the provided segments into a comprehensive response. Where information is consistent across summaries, it is streamlined; where it varies or adds detail, it is included with attribution to specific aspects. Due to the volume and complexity of the data, I will use a combination of narrative text and a table in CSV format to present the details efficiently and densely.
  36. [36]
    The Mechanism of Vinyl Polymerizations 1 - ACS Publications
    Catalytic Chain Transfer Mediated Autopolymerization of Divinylbenzene ... Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1937 ...
  37. [37]
    From telomerization to living radical polymerization - Boutevin - 2000
    Sep 15, 2000 · A brief description of the various methods of telomerization by the radical and redox methods is given. Chain lengths and functionality are ...
  38. [38]
    Fundamentals of Emulsion Polymerization | Biomacromolecules
    Jun 16, 2020 · This overview discusses the fundamentals of emulsion polymerization and related processes with the object of providing understanding to researchers.
  39. [39]
    Scale-Up of Polymerization Process:  A Practical Example
    ### Summary of Bulk Polymerization in Radical Polymerization
  40. [40]
    Polymers - MSU chemistry
    Radical polymerization gives a statistical copolymer. However, the product of cationic polymerization is largely polystyrene, and anionic polymerization favors ...Missing: ionic | Show results with:ionic
  41. [41]
    Auto-Acceleration in Free-Radical Polymerizations Under ...
    Dec 10, 1981 · This effect which was reported for the first time in 1939 by NORRISH and BROOKMAN (1) to occur in the bulk polymerization of methyl methacrylate ...Missing: original paper
  42. [42]
  43. [43]
    Solution Polymerization - an overview | ScienceDirect Topics
    In solution polymerization, free radical cross-linking polymerization mechanisms can result in the formation of a macrogel even with dilute monomer solutions.
  44. [44]
    Radical chemistry in polymer science: an overview and recent ...
    Radical polymerization, which IUPAC defines as 'A chain polymerization in which the kinetic-chain carriers are radicals' [13], is the most widely used reaction ...
  45. [45]
    Kinetics of Emulsion Polymerization | The Journal of Chemical Physics
    As a basis for understanding emulsion polymerization, the kinetics of free radical reactions in isolated loci is discussed.Missing: original | Show results with:original
  46. [46]
    An Updated Review on Suspension Polymerization - ACS Publications
    The presence of suspending agents (e.g., stabilizers) hinder the coalescence of monomer droplets and the adhesion of partially polymerized particles during the ...
  47. [47]
    Atom Transfer Radical Polymerization | Chemical Reviews
    According to SciFinder Scholar, 7 papers were published on ATRP in 1995, 47 in 1996, 111 in 1997, 150 in 1998, 318 in 1999, and more than 300 in 2000. In ...Introduction · Catalysts · General Comments on the... · III. Materials Made by ATRPMissing: original | Show results with:original
  48. [48]
    Controlled radical polymerization - ScienceDirect.com
    Fundamentals of controlled/“living” radical polymerization are given together with a discussion of selected initiating/catalytic systems.
  49. [49]
    Genesis of the CSIRO polymer group and the discovery and ...
    Oct 12, 2005 · Genesis of the CSIRO polymer group and the discovery and significance of nitroxide-mediated living radical polymerization. David H. Solomon,.
  50. [50]
    Living Free-Radical Polymerization by Reversible Addition ...
    Communication to the Editor July 22, 1998. Living Free-Radical Polymerization by Reversible Addition−Fragmentation Chain Transfer: The RAFT Process. Click to ...
  51. [51]
  52. [52]
    Controlled Radical Polymerization: State-of-the-Art in 2011
    Mar 20, 2012 · In radical polymerization, termination is often a diffusion-controlled process and differences between termination rate constants for different ...
  53. [53]
    H 2 O 2 Enables Convenient Removal of RAFT End-Groups from ...
    Dec 23, 2016 · RAFT-synthesized polymers are typically colored and malodorous due to the presence of the sulfur-based RAFT end-group(s).Missing: Disadvantages odor
  54. [54]
    Principles of Polymerization | Wiley Online Books
    Principles of Polymerization ; Author(s):. George Odian, ; First published:9 January 2004 ; Print ISBN:9780471274001 | ; Online ISBN:9780471478751 | ...
  55. [55]
    [PDF] FREE-RADICAL POLYMERIZATION
    Free-radical polymerization involves initiation, propagation, and termination steps. The kinetic chain length is the average monomers consumed by each radical.
  56. [56]
    Effect of Light Intensity on the Free-Radical Photopolymerization ...
    Jul 10, 2020 · Indeed, the rate of side reactions between primary radicals, which might occur during irradiation with high intensities, is negligible.
  57. [57]
    A Critical Experimental Examination of the Gel Effect in Free Radical ...
    Over the past two decades, a substantial case has been made associating and/or attributing the “gel” effect or autoacceleration in free radical polymerizationIntroduction · Background · Results and Discussion · Acknowledgment
  58. [58]
    None
    ### Summary of Free-Radical Polymerization Content
  59. [59]
    Polydispersity Index - an overview | ScienceDirect Topics
    Step-growth polymerization reactions typically yield values of Mw/Mn of around 2.0 and chain-growth polymerization yield Mw/Mn values in the range of 1.5–20.
  60. [60]
    [PDF] Polymers: Molecular Weight and its Distribution
    Jan 15, 2001 · The theory of step-growth (condensation) polymerization for flexible chain polymers led to the Flory-Schulz distribution. In the theory, p ...
  61. [61]
    50th Anniversary Perspective: RAFT Polymerization—A User Guide
    This Perspective summarizes the features and limitations of reversible addition–fragmentation chain transfer (RAFT) polymerization, highlighting its strengths ...The RAFT Process · Monomer Class and RAFT Agent · Process of Polymerization
  62. [62]
    Chain Transfer in the Polymerization of Styrene. II. The Reaction of ...
    Mayo. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1948, 70, 7, 2373–2378. Click to copy citation ...
  63. [63]
    Degree of polymerization and chain transfer in methyl methacrylate
    The chain-transfer constants of twenty-six solvents in the polymerization of methyl methacrylate have been calculated, using Mayo's equation for chain-transfer
  64. [64]
    [PDF] Handbook of Radical Polymerization
    ... Odian, Principles of Polymerization, 3rd ed., Wiley, New York, 1991, p. 198. 313. A. Kuriyama and T. Otsu, Polym. J. 16 (1984). 314. T. Otsu and A. Kuriyama ...
  65. [65]
    Kinetics and temperature evolution during the bulk polymerization of ...
    This high reaction exotherm, combined with the auto-acceleration effect, can lead to thermal runaway and increase the reaction temperature above the boiling ...
  66. [66]
    Autoacceleration in Free Radical Polymerization. 1. Conversion
    Propagating radical termination at high conversion in emulsion polymerization of MMA. Rate coefficient determination from ESR data.Missing: heat challenges exotherm
  67. [67]
    Inhibition of Free Radical Polymerization: A Review - PMC - NIH
    Jan 17, 2023 · The thermal polymerization of MMA is inhibited when the solutions of quinones and amines in MMA was preliminarily irradiated with visible spectrum.<|control11|><|separator|>
  68. [68]
    Polymerization Reactor - an overview | ScienceDirect Topics
    Thus, cooling of the polymerization mixture is effected by the introduction of cold monomer at several side-feed points along the multizone reactor. Moreover, ...
  69. [69]
    Inert Diluent - an overview | ScienceDirect Topics
    The inert diluent conveys the heat of polymerization from the growing polymer particles to the water-cooled jacket of the reactor. In addition, it also has to ...
  70. [70]
    Improved control through a semi-batch process in RAFT-mediated ...
    Sep 24, 2015 · RAFT-mediated polymerization is compatible with virtually all monomer classes that are accessible via conventional radical polymerization.
  71. [71]
    Analytical Steady‐State Model for the Free Radical Solution ...
    May 27, 2023 · In this study, an analytical solution was obtained for the model of free radical solution copolymerization in a continuous stirred tank reactor ...
  72. [72]
    Optimal Operating Policies for the Nitroxide-Mediated Radical ...
    In this work, optimal operating policies for the industrial-level semibatch living free radical polymerization of styrene are proposed.
  73. [73]
    [PDF] The Mechanism of Stereoregulation in Free-Radical Polymerization ...
    Mar 21, 2012 · Polymers produced by radical polymerisation typically have a tacticity in the range of r = 0.7 to m = 0.7 depending on the type of monomer and ...
  74. [74]
  75. [75]
    Mechanistic considerations on styrene–maleic anhydride ...
    In addition to that new developments in terms of living radical copolymerization of styrene and maleic anhydride are discussed from a mechanistic point of view.
  76. [76]
    Tacticity control approached by electric-field assisted free radical ...
    Jun 30, 2023 · The proportions of isotactic, syndiotactic and atactic triads in ... Okamoto, Stereochemistry of the Radical Polymerization of Vinyl ...
  77. [77]
    The Copolymerization of Styrene and Methyl Methacrylate
    Copolymerization. I. A Basis for Comparing the Behavior of Monomers in Copolymerization; The Copolymerization of Styrene and Methyl Methacrylate. Click to copy ...
  78. [78]
    80 years of the Mayo Lewis equation. A comprehensive review on ...
    Thus, the ratio of M1 to M2 units in the chain must equal the ratio of the respective average sequence lengths: F 1 i n s F 2 i n s = N ¯ 1 N ¯ 2 Substitution ...
  79. [79]
    Reactivity Ratio Estimation in Radical Copolymerization
    An error-in-variables-model (EVM) framework is presented for the optimal estimation of reactivity ratios in copolymerization systems.
  80. [80]
    [PDF] 6.6.3 Polystyrene 6.6.3.1 General Styrene readily polymerizes to ...
    Styrene polymerizes to polystyrene via free radical chain mechanism, using heat or initiators. Bulk, solution, suspension, and emulsion techniques are used.
  81. [81]
    [PDF] Polymer Structures Chapter 4 - nanoHUB
    Optical properties: crystalline -> scatter light (Bragg) amorphous -> transparent. Most covalent molecules absorb light outside visible spectrum,. e.g. PMMA ( ...
  82. [82]
    Suspension polymerization of vinyl chloride at low temperature with ...
    Suspension polymerization of vinyl chloride was carried out at temperatures of -30° to +30°C. with the use of a monomer-soluble oxidation-reduction ...
  83. [83]
    [PDF] Inhibiting the deterioration of plasticized poly (vinyl chloride) a ...
    Attempts to process PVC in the raw form using heat and pressure, result in severe degradation of the polymer (Nass, 1977).
  84. [84]
    Synthesis of Exfoliated Polyacrylonitrile/Na−MMT Nanocomposites ...
    Polyacrylonitrile/Na−MMT nanocomposites were synthesized through an emulsion polymerization of acrylonitrile (AN) using ...
  85. [85]
    [PDF] Introduction to Macromolecular Chemistry - Merten Lab
    ▫ High pressure PE = LDPE: commerciallized by ICI in 1939. ▫ Radical polymerization in bulk at 1400-3500 bar, 130-330 °C. ▫ Initiation: oxygen, organic ...
  86. [86]
    [PDF] Market Study: Styrene-Butadiene Rubber (SBR) - Ceresana
    SBR is produced by emulsion polymerization (E-SBR) or by solution polymerization (S-SBR). Styrene- butadiene latex, which is used for example as coating in ...
  87. [87]
    Synthetic Rubber Market Size, Share & Growth | Report [2032]
    Styrene butadiene rubber (SBR) is the leading segment and accounted for the largest market share. The growth is attributed to its cost-effective properties and ...Missing: percentage | Show results with:percentage<|separator|>
  88. [88]
    Radical Polymerization in Industry - ResearchGate
    FRP is the most common type of polymerization in the industry, accounting for around 40−45% of all industrial polymers 22 due to its simplicity and tolerance to ...
  89. [89]
  90. [90]
    Polystyrene Market| Globalization & Industrialization Drive Growth
    Polystyrene Market size was valued at the US $ 33.73 Bn. in 2023 and expected to reach US$ 44.36 Bn. by 2030. at a CAGR of 3.99% during the forecast period.
  91. [91]
    PolyVinyl Chloride Market Size, Share, Analysis & Forecast 2035
    The largest market share of PVC goes with high proportion in Pipes and Fittings, about 44% in the year 2024. ... PVC Market in Asia Ends August Lower as China ...
  92. [92]
    Acrylate Market Size & Share | Industry Forecast Analysis, 2032
    The market size for acrylate was valued at around USD 11.3 billion in 2022 and is estimated to reach USD 17.8 billion by 2032, owing to the increasing demand ...
  93. [93]
    Polystyrene - Chemical Safety Facts
    Polystyrene (solid and foam) is widely used to protect consumer products. CD and DVD cases, foam packaging peanuts for shipping, food packaging, meat/poultry ...Key Points/overview · Uses & Benefits · Safety Information
  94. [94]
    Acrylic Adhesives Market Size & Share Analysis - Growth Trends
    Nov 5, 2024 · The acrylic adhesives market is estimated at USD 15.89 billion in 2024, projected to reach USD 20.14 billion by 2028, with a 6.10% CAGR. ...
  95. [95]
    Polystyrene microplastic particles in the food chain: Characteristics ...
    Sep 20, 2023 · However, the widespread use of polystyrene microplastics (PS-MPs) in the food chain has raised global concerns regarding marine and human ...
  96. [96]
    Origins and Development of Initiation of Free Radical Polymerization ...
    Aug 7, 2025 · In the 1940s the polymerization by redox processes was found independently and simultaneously at IG Farben in Germany and ICI in Great Britain. ...
  97. [97]
    Polymer Recycling and Upcycling: Recent Developments toward a ...
    Oct 2, 2023 · Recent developments include direct recycling, chemical upcycling, and creating sustainable monomers/polymers, with mechanical recycling ...
  98. [98]
    Organocatalyzed atom transfer radical polymerization driven by ...
    ATRP has historically relied on transition-metal catalysts to mediate this equilibrium and polymerize monomers with diverse functionality into macromolecules ...
  99. [99]
    Exploring Bio‐Based Monomers in Emulsion and Miniemulsion ...
    Mar 5, 2025 · This review examines recent advancements in bio-based monomers as sustainable alternatives to petroleum-based monomers in emulsion and ...
  100. [100]
    Enzyme Catalysis for Reversible Deactivation Radical Polymerization
    Jun 27, 2022 · In this Minireview we discuss the key roles enzymes play in RDRP, including their ATRPase, initiase, deoxygenation, and photoenzyme activities.Missing: mimetic initiators 2020s
  101. [101]
    Reactivity Ratios of Biobased Dialkyl Itaconate Radical ...
    Nov 26, 2024 · Itaconates available from renewable resources constitute a group of monomers that are used in several types of polymerizations.Introduction · Materials and Experimental... · Results and Discussion · Conclusions
  102. [102]
    Terpene based acrylates as substitutes for petrochemical compounds
    May 14, 2025 · Terpene-based acrylic monomers provide a sustainable and environmentally friendly alternative to conventional acrylic monomers.
  103. [103]
    Bulk depolymerization of poly(methyl methacrylate) via chain-end ...
    Sep 14, 2023 · The depolymerization of PMMA prepared by these RDRP techniques can achieve up to 92% reversion to monomer at temperatures 250°C lower than those ...
  104. [104]
    Metal-Free ATRP Catalyzed by Visible Light in Continuous Flow
    Sep 7, 2020 · This work combines three essential components toward a “greener” chemistry: miniaturization, renewable energy, and metal-free catalysis that can ...
  105. [105]
    Progress and Challenges in Polystyrene Recycling and Upcycling
    May 17, 2024 · By the end of 2023, the worldwide demand for PS exceeds 15 million tons (accounting for ~5 % of all plastics), with an anticipated increase of 1 ...