Fact-checked by Grok 2 weeks ago

Chain transfer

Chain transfer is a fundamental process in free , where the propagating at the end of a growing chain reacts with another —such as a , , another chain, or a dedicated chain agent (CTA)—to the activity, thereby terminating the growth of the original chain and initiating on a new species. This reaction, often occurring via or other transfers, results in the formation of a "dead" chain with a specific end group and a new active that continues the process./Chapter_08:_Reactions_of_Alkenes/8.7:_Polymerization/Free_Radical_Polymerization) Unlike chain termination, chain does not reduce the overall concentration of radicals in the system, allowing to proceed at a similar rate while influencing chain length distribution. The mechanism of chain transfer typically involves the abstraction of an atom (most commonly ) from a labile site on the by the propagating , generating a more stable on the agent that can then add to a molecule./Chapter_08:_Reactions_of_Alkenes/8.7:_Polymerization/Free_Radical_Polymerization) Chain can be intermolecular, occurring between distinct molecules, or intramolecular, such as where the abstracts a from within its own chain, leading to short-chain branching. Common CTAs include thiols (e.g., mercaptans), halogenated compounds like (CCl₄), and certain solvents, which are selected for their weak bonds that facilitate efficient without significantly altering the . Chain transfer plays a critical role in controlling molecular weight, polydispersity, and , often intentionally employed to produce polymers with desired properties such as lower or specific end functionalities for further reactions. For instance, in the high-pressure free of to produce (LDPE), chain transfer to the polymer backbone via intramolecular reactions promotes extensive branching, resulting in an amorphous, flexible material./Chapter_08:_Reactions_of_Alkenes/8.7:_Polymerization/Free_Radical_Polymerization) In advanced techniques like reversible addition-fragmentation chain transfer (RAFT) , specialized CTAs enable precise control over chain length and allow the synthesis of complex structures like block copolymers, enhancing applications in and .

Principles and Mechanism

Definition and Role in Polymerization

Chain-growth polymerization involves three primary stages: , where an active is generated from a or initiator; , in which the active repeatedly adds units to extend the ; and termination, where the active is deactivated, halting growth. This process is common in radical, cationic, and anionic mechanisms, producing polymers with controlled architectures. Chain transfer is a key side reaction in , wherein the active center—such as a —from a growing chain is transferred to another , thereby stopping elongation of the original chain and potentially starting a new one. This transfer generates a new active species without fully terminating the overall . The role of chain transfer is to limit the length of individual chains, thereby influencing the molecular weight distribution and enabling the production of polymers with tailored properties, such as reduced or branched structures. It occurs in free radical, cationic, and anionic polymerizations, though it has been most extensively studied in free radical systems. By reducing the average , chain transfer prevents the formation of excessively high molecular weights that could lead to gelation or processing difficulties. In industrial applications, chain transfer agents are deliberately added to achieve precise control over polymer characteristics.

General Reaction Mechanism

In free radical polymerization, chain transfer occurs when a propagating polymer radical, denoted as \ce{P_n^\bullet}, abstracts an atom—typically a hydrogen atom—from a transfer agent \ce{X-H}, resulting in a dead polymer chain \ce{P_n-H} and a new radical \ce{X^\bullet}. This process terminates the growth of the original chain while generating a new radical site capable of initiating polymerization. The reaction is represented as: \ce{P_n^\bullet + X-H -> P_n-H + X^\bullet} The new radical \ce{X^\bullet} then adds to a monomer molecule \ce{M}, forming a new propagating radical \ce{P_1^\bullet} and restarting the propagation cycle: \ce{X^\bullet + M -> P_1^\bullet} This reinitiation step ensures that chain transfer does not halt the overall polymerization but redistributes the active centers. Chain transfer integrates into the broader cycle by competing directly with and termination reactions. The propagating \ce{P_n^\bullet} can either add another unit (propagation, rate constant k_p), undergo bimolecular termination with another radical (rate constant k_t), or participate in (rate constant k_{tr}). The relative rates determine the extent of chain length control, with transfer acting as a side reaction that limits chain growth without consuming the radical pool entirely. The for the chain transfer step, involving hydrogen abstraction, is generally higher than that for , which proceeds via addition to the ; typical differences range from 10 to 25 /, making transfer less favorable at lower s unless the transfer agent is designed for enhanced reactivity. This energetic barrier positions chain transfer as a secondary in standard free systems, though its contribution increases with temperature due to the positive difference in activation energies. While free mechanisms dominate, variations exist in other polymerization types; for instance, in , transfer often involves hydride shifts from the chain or to the growing , deactivating one site and activating another. However, radical-mediated transfer remains the primary focus in most conventional chain-growth es./02%3A_Synthetic_Methods_in_Polymer_Chemistry/2.04%3A_Cationic_Polymerization)

Types of Chain Transfer

To Monomer

Chain transfer to occurs through the abstraction of a labile from the molecule by a propagating , typically involving allylic C-H bonds that are weakened due to stabilization of the resulting . This generates a dead chain and a new monomer-derived , often resonance-stabilized, which can reinitiate but with reduced reactivity. In monomers like alpha-methylstyrene and , the allylic hydrogens are particularly susceptible to abstraction, forming delocalized radicals that enhance the transfer efficiency compared to non-allylic systems. The prevalence of this transfer varies significantly among monomers, reflecting differences in bond dissociation energies and stability. For styrene, the chain transfer constant C_m is approximately $10^{-5}, indicating moderate activity due to its allylic-like hydrogens on the beta carbon. In contrast, exhibits a much lower C_m \approx 10^{-7}, as its structural features provide less stabilization for the abstracted . These values, derived from experimental , highlight how allylic weakening of C-H bonds in certain monomers elevates transfer rates, while others remain negligible. The mid-chain radicals produced from transfer to monomer often reinitiate slowly owing to their lower reactivity toward addition, leading to potential branching when they eventually propagate or to temporary delays in the overall rate. This slow reinitiation can introduce structural irregularities, such as short branches, particularly in systems with higher C_m. Experimentally, the chain transfer constant C_m is determined using the Mayo plot, which involves graphing \log(1/\overline{DP}) against $1/[M], where \overline{DP} is the and [M] is the concentration; the slope provides C_m under conditions where other transfer processes are minimized. This method relies on varying concentration in bulk polymerizations to isolate the intrinsic transfer to from initiation and termination effects.

To Solvent or Chain Transfer Agent

Chain transfer to solvent or intentionally added chain transfer agents (CTAs) provides a controlled means to regulate polymer molecular weight in processes, distinct from inherent transfer to which typically offers limited . s such as hydrocarbons like exhibit low chain transfer activity, with a transfer C_s \approx 10^{-5} in styrene , resulting in minimal on length unless used in high concentrations. In contrast, chlorinated solvents like enable more effective transfer through chlorine atom abstraction by the propagating , yielding C_s \approx 10^{-3} for styrene, though this can introduce halogen end groups into the . Designed CTAs, such as mercaptans (thiols), halocarbons, and allyl compounds, are purposefully incorporated to achieve precise molecular weight reduction via efficient hydrogen or atom abstraction from weak bonds like S-H in thiols or C-Cl in halocarbons. For example, n-dodecyl mercaptan serves as a prominent thiol CTA in styrene polymerization, with C_s \approx 10-15 at 50°C, facilitating rapid transfer without significantly altering the polymerization rate. Allyl compounds, activated by their double bonds, promote transfer through allylic hydrogen abstraction or fragmentation, offering alternatives to thiols for applications requiring odorless agents. These mechanisms ensure the growing radical terminates while generating a new radical from the CTA, maintaining overall propagation efficiency. The high transfer efficiency of these agents allows their use at low concentrations, typically 0.1-5 wt% relative to , to substantially lower molecular weight while avoiding inhibition of the rate, unlike less selective methods. This controllability enhances processability by reducing without compromising . In industrial contexts, such CTAs are essential in the of rubber (SBR), where incremental addition of mercaptans like tert-dodecyl mercaptan targets specific molecular weights and viscosities for and rubber applications.

To Polymer or Initiator

Chain transfer to polymer involves reactions where a propagating radical abstracts a hydrogen atom from an existing polymer chain, generating a new radical site while terminating the original chain's growth. This process is divided into intramolecular and intermolecular variants. Intramolecular chain transfer, commonly known as backbiting, occurs when the propagating radical abstracts a hydrogen from a methylene group within the same polymer chain, typically via a 5- or 6-membered ring transition state. This 1,5- or 1,6-hydrogen shift yields a tertiary midchain radical, which can reinitiate propagation, resulting in short-chain branches of 1 to 5 carbon units. In acrylate polymerizations, backbiting predominates at elevated temperatures, forming tertiary radicals that constitute up to 90% of active species and contribute to short-chain branching. The chain transfer constant to polymer, C_p, is typically on the order of $10^{-4} for many vinyl monomers, such as styrene and methyl methacrylate, reflecting its relatively low efficiency compared to other transfer pathways. Intermolecular chain transfer to , though rarer due to lower probability of radical encounters, involves from a different polymer molecule, leading to and long-chain branches. This broadens the molecular weight distribution and increases polydispersity by creating interconnected polymer networks. In produced via free- es, intramolecular backbiting is a of short-chain branches like ethyl and butyl groups, enhancing crystallinity and mechanical properties, while intermolecular events contribute to long-chain branching that affects melt . Chain transfer to initiator typically involves reactions between propagating radicals and initiator-derived species, such as radicals or decomposition fragments. In systems using (AIBN), transfer occurs to cyanoisopropyl radicals generated during decomposition, or to cage products like tetramethylsuccinonitrile, which possess abstractable hydrogens. The mechanism entails direct hydrogen abstraction or addition-elimination, terminating the growing chain and potentially generating a new initiating radical. This process is often negligible under standard conditions due to low initiator concentrations but becomes significant in high-initiator experiments, where it can reduce overall molecular weight and broaden polydispersity by inefficiently consuming initiator. In such cases, it competes with primary radical termination, altering the without substantially affecting polymer microstructure.

Kinetic and Quantitative Aspects

Chain Transfer Constant

The chain transfer constant, denoted as C, quantifies the efficiency of chain transfer relative to in processes. It is a dimensionless defined as the of the rate constant for the chain transfer (k_{tr}) to the rate constant for the (k_p): C = \frac{k_{tr}}{k_p}. This indicates the likelihood that a propagating radical will undergo transfer instead of adding another unit. For transfers to specific , the constant is subscripted accordingly; for example, the chain transfer constant to monomer is C_M = k_{tr,M} / k_p, and to solvent or a dedicated chain transfer agent, C_S = k_{tr,S} / k_p. Typical values of C vary widely depending on the nature of the transfer agent and . Chain transfer to is generally inefficient, with C_M ranging from $10^{-5} to $10^{-3}; for instance, C_M \approx 5.3 \times 10^{-5} for styrene and $5.2 \times 10^{-5} for at 50°C. In contrast, efficient chain transfer agents like thiols exhibit much higher constants, such as C_S \approx 21 for n-butyl mercaptan in styrene at 60°C, enabling effective control over length even at low concentrations. Chain transfer constants are determined experimentally through methods that analyze polymerization outcomes under varying conditions. The Mayo method is a standard approach, relying on the relationship between the degree of polymerization and transfer agent concentration. By conducting polymerizations at different ratios of transfer agent [S] to monomer [M] and measuring the number-average degree of polymerization \overline{DP}_n, a plot of $1/\overline{DP}_n versus [S]/[M] yields a straight line with slope equal to C_S: \frac{1}{\overline{DP}_n} = \frac{1}{\overline{DP}_0} + C_S \frac{[S]}{[M]}, where \overline{DP}_0 is the degree of polymerization in the absence of transfer agent. This method assumes steady-state kinetics and negligible other transfer pathways. Alternatively, constants can be derived from molecular weight distribution data using Stockmayer equations, which model polydispersity as influenced by transfer probability and allow extraction of C from the breadth of the distribution. Several factors influence the magnitude of the chain transfer constant. Temperature plays a key role, as the activation energy for transfer (E_{tr}) is typically higher than for propagation (E_p), resulting in C increasing with temperature according to the Arrhenius relation: C = A \exp\left( -\frac{E_{tr} - E_p}{RT} \right), where A is the pre-exponential factor, R is the gas constant, and T is temperature in Kelvin. This temperature sensitivity makes chain transfer more prominent at elevated temperatures. Solvent choice also affects C by modulating radical stability; polar or electron-donating solvents can stabilize radicals, enhancing k_{tr} for certain transfers and thus increasing C. The overall probability of chain transfer in a system, relative to , provides a comprehensive measure of transfer efficiency across multiple agents. This probability \nu_{tr} is the ratio of the total chain transfer rate R_{tr} to the rate R_p: \nu_{tr} = \frac{R_{tr}}{R_p}. The rate is R_p = k_p [M] [R^\bullet], where [M] is the concentration and [R^\bullet] is the concentration of propagating radicals. The total transfer rate, summing over all transfer agents i (e.g., , , additives) with concentrations [X_i], is R_{tr} = \sum_i k_{tr,i} [X_i] [R^\bullet]. Dividing these rates eliminates [R^\bullet] and yields \nu_{tr} = \frac{\sum_i k_{tr,i} [X_i]}{k_p [M]} = \frac{\sum_i C_i [X_i]}{[M]}. This expression shows that the effective transfer probability scales with the weighted sum of individual C_i values normalized by concentration, directly influencing the average chain length as \overline{DP}_n \approx 1 / \nu_{tr} when chain transfer dominates the chain stopping process.

Effects on Molecular Weight and Distribution

Chain transfer reactions in free significantly reduce the number-average (\bar{DP}_n) by providing an additional pathway for chain cessation beyond bimolecular termination. In the absence of chain transfer, \bar{DP}_n \approx \frac{k_p [M]}{(2 k_t R_i)^{0.5}}, where k_p is the rate constant, [M] is the concentration, k_t is the termination rate constant, and R_i is the rate (denoted as DP_0). With chain transfer, this expression modifies to \bar{DP}_n \approx \frac{k_p [M]}{(2 k_t R_i)^{0.5} + \sum k_{tr,i} [X_i]} = \frac{DP_0}{1 + \nu_{tr} DP_0}, where \nu_{tr} = \sum C_i \frac{[X_i]}{[M]}. Thus, \bar{DP}_n decreases as the transfer probability increases, allowing precise control over molecular weight by adjusting transfer agent concentration relative to . The polydispersity index (PDI, or ) is also affected. Without transfer, the ideal PDI is 1.5 for termination by or 2 for . Chain transfer to small molecules (constant probability, as covered in the "Types of Chain Transfer" section) typically shifts the PDI towards 2. In contrast, chain transfer to polymer (chain-length dependent) can broaden the to PDI > 2 due to unequal transfer probabilities across chain lengths, leading to branching. Certain processes like intramolecular can influence PDI depending on uniformity, but generally promote broader distributions through short-chain branching. Representative examples illustrate these effects. In of styrene, addition of n-butyl mercaptan as a chain transfer agent reduces the number-average molecular weight (\bar{M}_n) to approximately 85,000 g/ from higher values (typically >200,000 g/ without transfer), demonstrating the proportional impact of transfer agent concentration. High levels of efficient transfer agents, such as thiols with C_s > 10, can further suppress \bar{M}_n to yield oligomers (<10,000 g/), useful for low-molecular-weight applications. In the overall kinetic scheme, chain transfer interacts with termination to determine properties, but becomes the dominant factor at high transfer agent concentrations or elevated temperatures, where transfer rates increase exponentially with differences. Using chain transfer constants (e.g., C_s for solvents or agents), these effects can be predicted and tuned without altering or kinetics significantly.

Advanced and Controlled Methods

Reversible Addition-Fragmentation Chain Transfer (RAFT)

Reversible Addition-Fragmentation Chain Transfer () polymerization is a controlled technique that employs thiocarbonylthio compounds as chain transfer agents (CTAs) to achieve reversible deactivation of growing chains, enabling the of polymers with well-defined architectures and narrow molecular weight distributions. Developed by researchers at the () in 1998, RAFT builds on conventional by introducing an between active propagating radicals and dormant , which minimizes termination events and allows for living-like behavior. Common CTAs include dithioesters, trithiocarbonates, xanthates, and dithiocarbamates, where the Z-group stabilizes the intermediate radical and the R-group influences reinitiation efficiency. The overall polymerization rate in RAFT is similar to that of conventional , given by R_p = k_p [M] \sqrt{\frac{f k_d [I]_0}{k_t}}, but modulated by a K = \frac{k_{\text{add}} k_{\text{frag}}}{k_{-\text{add}}} that governs the addition-fragmentation . The mechanism proceeds through three main steps: , fragmentation, and reinitiation. In the addition step, a propagating P_n^\bullet adds to the (R-S-C(=S)Z), forming a reversible . This intermediate then undergoes fragmentation, releasing a new R^\bullet and forming the dormant P_n-S-C(=S)Z, which transfers control to another . The released R^\bullet rapidly reinitiates by adding to a (M), forming a new propagating P_m^\bullet. This cycle establishes a rapid between active and dormant chains, ensuring most chains grow simultaneously and maintaining high end-group fidelity. RAFT offers significant advantages, including the production of polymers with low polydispersity indices (PDI ≈ 1.1–1.5), predictable molecular weights, and the ability to synthesize complex structures such as block copolymers through sequential addition, often achieving near-quantitative conversions. It exhibits broad tolerance to functional groups and is compatible with a wide range of s, including methacrylates, acrylates, and styrenes, without requiring stringent oxygen-free conditions typical of other controlled methods. However, RAFT can suffer from an initial inhibition period due to slow reinitiation by the R-group , particularly with certain CTAs, and the sulfur-containing agents often impart an undesirable to the resulting polymers.

Catalytic Chain Transfer Polymerization (CCTP)

Catalytic chain transfer polymerization (CCTP) is a free polymerization technique that employs low concentrations of complexes, primarily (II) species such as porphyrins or salen derivatives, to achieve highly efficient chain transfer, enabling the synthesis of low molecular weight with controlled architectures. In this process, the (II) complex interacts with the growing polymeric through a degenerative transfer , where the metal abstracts a from the β-position relative to the center, typically from the α-methyl group in methacrylates. This step generates a dead polymer chain terminated with a (macromonomer) and a transient (III)- species. The Co(III)-H then rapidly transfers the to a molecule, regenerating the active Co(II) and initiating a new growing chain. This operates without significantly retarding the overall rate and is particularly effective for acrylates and methacrylates due to favorable kinetics for abstraction and reinitiation. The efficiency of CCTP is characterized by exceptionally high chain transfer constants (C_CT), ranging from 10^4 to 10^6 for methacrylates, which allow the use of concentrations as low as relative to . These values reflect the rapid rate of transfer compared to , enabling precise control over molecular weight at low loadings; for instance, C_CT values around 3 × 10^4 have been reported for using porphyrin complexes like bis[(difluoroboryl)dimethylglyoximato]. The process is typically conducted at temperatures between 50 and 120°C, with optimal performance for methacrylates around 60–80°C to balance transfer efficiency and avoid side reactions like deactivation. The resulting products are α-telechelic oligomers, where the α-end derives from the initiator (often saturated) and the ω-end is a functionality suitable for further reactions, yielding polymers with number-average degrees of (DP_n) inversely proportional to the [Co]/[M] ratio, approximately DP_n ≈ 1/(C_CT × [Co]/[M]). CCTP was pioneered in the late by Enikolopov et al. with initial demonstrations using chelates, but significant development and application to controlled occurred in the through the work of Haddleton, , and collaborators, who optimized porphyrin-based systems for methacrylates and explored aqueous and variants. The frequency is directly tied to the [Co]/[M] ratio, allowing tunable lengths from a few units to hundreds, as the frequency of events scales with concentration. This method's utility lies in producing vinyl-terminated macromonomers for applications in adhesives, coatings, and high-solids formulations, where the end-group enables copolymerization or ; it also integrates with living radical techniques like to form hybrid structures with enhanced functionality, such as block copolymers or telechelics for .

Historical and Contemporary Developments

Early Discoveries and Key Studies

The concept of chain transfer in was first formally incorporated into kinetic models by Paul J. Flory in 1937, who described it as a process limiting chain growth in vinyl polymerizations through abstraction from non-monomeric species. Early empirical observations of molecular weight reduction in polymerizing systems, such as , were noted in studies around the mid-1930s, including those by on degradative processes during thermal treatments. These initial insights laid groundwork for understanding how environmental factors influenced chain lengths, though without explicit mechanisms. During the , systematic investigations into solvent effects on molecular weight emerged, particularly in styrene polymerization, where solvents like and were shown to act as donors, reducing average chain lengths via transfer reactions. This period coincided with intensified research driven by efforts to develop , where chain transfer agents such as mercaptans were deliberately added to recipes to control and molecular weight distribution in industrial-scale production. Key contributions came from Frank R. Mayo and colleagues, who in 1943 quantified solvent transfer reactivity through measurements of , establishing a basis for comparing transfer efficiencies across media. A pivotal advancement occurred in 1947–1948, when R. A. Gregg and F. R. introduced the chain transfer constant (C_s) and the associated Mayo plot, plotting against solvent-to-monomer ratio to isolate transfer contributions in styrene systems. This graphical method, refined in their 1948 studies on mercaptan transfer, enabled precise determination of transfer rates without interference from termination, using low-initiator "catalyzed" polymerizations. Edwin J. Hart contributed to understanding initiator-derived transfer in the late 1940s, examining how fragments from peroxides participated in chain interruptions during aqueous polymerizations. By the , Gregg extended these measurements to and other monomers, measuring C_s values for dozens of solvents at 60°C and correlating them with molecular structure. Experimental techniques advanced significantly in this era, with inhibitors like employed to quench mid-reaction, allowing isolation of effects for end-group analysis via and . Radiochemical labeling, using isotopes such as S-35 in thiols or C-14 in solvents, facilitated direct tracing of sites through degradation and scintillation counting of end groups, providing quantitative validation of kinetic models. In the , electron spin (ESR) began confirming intramolecular as a mechanism in acrylates, detecting midchain radicals formed by hydrogen abstraction within the growing chain. These developments marked a transition from empirical adjustments in production to a rigorous kinetic framework, emphasizing 's role in polydispersity control.

Recent Applications and Research

In the polymer industry, chain transfer agents (CTAs) continue to play a crucial role in (PVC) production, particularly for synthesizing low molecular weight polymers used in plastisols. These plastisols, applied in coatings and adhesives, benefit from CTAs like halogenated hydrocarbons or trithiocarbonates that control molecular weight to achieve viscosities suitable for processing, reducing (VOC) emissions in end-use applications. For instance, PVC incorporating CTAs has enabled the development of polymers with reduced or zero VOC requirements, enhancing environmental compliance in and coverings. Reversible addition-fragmentation chain transfer (RAFT) polymerization has found significant application in fabricating nanoparticles for , leveraging precise control over architecture to create responsive carriers. Recent advancements include photoinitiated RAFT processes yielding reduction-responsive protein- nanoparticles that encapsulate therapeutics with high loading efficiency, enabling targeted release in biological environments. These systems, often involving block copolymers, improve and reduce side effects in cancer therapies. Catalytic chain transfer (CCTP) remains vital for producing waxes, which serve as lubricants, dispersants, and processing aids in plastics and coatings. Using or catalysts, CCTP yields low molecular weight, branched polyethylenes with tunable properties; for example, benzocycloalkyl systems produce waxes with molecular weights below 10 kg/mol, offering superior for industrial formulations. Zinc-enhanced metallocene further optimizes chain transfer, achieving high yields and narrow distributions for sustainable wax production. Research in the 2010s advanced oxygen-tolerant polymerization, addressing the need for ambient-condition synthesis without rigorous . Photoinduced electron/energy transfer- (PET-RAFT) systems, utilizing photocatalysts like or organic dyes, enable in open air, with rates up to 80% conversion in hours; this has expanded RAFT to aqueous and biological media. Enzyme-mediated variants, such as with , further enhance tolerance, achieving dispersities below 1.2. In the 2020s, studies on bio-based CTAs derived from have gained traction for sustainable synthesis. Coordinative chain transfer of like yields elastomeric materials with high renewability; for example, β-pinene copolymers via exhibit tensile strengths comparable to petroleum-based analogs while reducing carbon footprints. These efforts align with goals by utilizing abundant plant-derived monomers. Computational modeling using (DFT) has elucidated chain transfer barriers in , aiding CTA design. DFT calculations reveal activation energies for transfer to monomers like or acrylates, typically 5–15 kJ/mol (1–4 kcal/mol) higher than , consistent with small transfer constants and enabling prediction of molecular weight distributions. Recent applications include modeling self-initiated systems, where barriers inform strategies for high-purity polymers. Emerging trends integrate chain transfer with to produce functional polymers, combining end-groups with azide-alkyne cycloadditions for precise grafting. This hybrid approach yields brush or star architectures with pendant functionalities like fluorophores, enhancing properties for sensors and biomaterials; yields exceed 95% in modular syntheses. Sustainability drives development of metal-free CTAs, such as organocatalytic systems avoiding transition metals, which minimize and enable recyclable processes with efficiencies over 90%. Challenges in scaling controlled methods like and persist, including maintaining low dispersities (Đ < 1.3) at industrial volumes and managing in exothermic reactions. Environmental concerns with sulfur-based RAFT CTAs involve and potential residues, prompting shifts to sulfur-free alternatives that reduce impacts by up to 50%. As of 2025, chain transfer integrates with resins, where RAFT-enabled photopolymerization produces tunable microstructures; mixing CTAs adjusts mechanical properties, yielding objects with elongations from 10-200%. Publications explore AI-optimized chain transfer constants (C), using to predict values from molecular descriptors, accelerating CTA screening for targeted polydispersities.

References

  1. [1]
    Radical chemistry in polymer science: an overview and ... - PMC - NIH
    When chain transfer happens, the originally growing chain halts while a new chain launches from the radical species formed from the chain transfer agent. In the ...
  2. [2]
    [PDF] Chapter 1 Free-Radical Polymerization
    A chain transfer agent, S is a molecule containing a weak bond that can be broken to lead to radical transfer, similarly as in the case of monomer above (e.g. ...
  3. [3]
  4. [4]
    Chain Transfer - an overview | ScienceDirect Topics
    Chain transfer is the reaction of a propagating radical with a nonradical substrate (X–Y, Scheme 50) to produce a dead polymer chain and a new radical (Y•) ...
  5. [5]
    Polymers - MSU chemistry
    Chain transfer reactions are especially prevalent in the high pressure radical polymerization of ethylene, which is the method used to make LDPE (low ...
  6. [6]
    Chain transfer agents and its use in polymer synthesis
    Chain transfer agents are useful for controlling the molecular weights of polymers, for reducing gelation when polymerizations and copolymerizations involving ...<|control11|><|separator|>
  7. [7]
    Chain Transfer Agent - an overview | ScienceDirect Topics
    Transfer agents find widespread use in both industrial and laboratory polymer syntheses. They are used to control: 1. the molecular weight of polymers. 2 ...
  8. [8]
    [PDF] Mechanism and Kinetics of Free Radical Chain Polymerization
    Jan 29, 2001 · Free radical chain polymerization involves initiation, propagation, and termination steps. Termination can be by coupling or disproportionation ...
  9. [9]
    Introduction to polymers: 4.3.3 Termination and transfer | OpenLearn
    The third method of termination is chain transfer in which a radical abstracts a hydrogen atom from a neighbouring molecule. In the case of polystyrene the ...Missing: definition | Show results with:definition
  10. [10]
    Chain transfer to monomer and polymer in the radical ...
    ... activation energy for chain transfer is on the order of 20–25 kJ mol−1. These plots were linear between the number and weight-average degrees of polymerization ...
  11. [11]
    Chain transfer in ethylene polymerization. V. The effect of temperature
    This value is equal to the difference in activation energy between the transfer step (hydrogen abstraction) and the propagation step (addition to the monomer ...<|separator|>
  12. [12]
    [PDF] Handbook of Radical Polymerization
    ... Chain Transfer to Monomer in Ethylene Polymerization Chain transfer to ... allylic hydrogen does not add to monomer because of its stability. Addition ...
  13. [13]
    [PDF] Free-Radical Polymerization of Olefin Monomers
    Monomer Chain-Transfer Constants 92. Monomer. Methyl acrylate. Methyl methacrylate. Styrene. Vinyl acetate. Table 11-15. Cм × 104. 0.07-0.4. 0.07-0.5. 0.6. 20.Missing: C_m | Show results with:C_m
  14. [14]
    [PDF] Chain transfer to solvent in BN 2‐vinylnaphthalene radical ...
    Herein, we report characterization of solvent effects in the polymerization of BN 2-vinylnaphthalene (BN2VN), as well as the synthesis of high molecular weight.
  15. [15]
    [PDF] Effect of Substituent of Chain Transfer agent in the Free Radical ...
    The calculated chain transfer constant of cumyl chloride shows higher value (0.0463) than that of carbon tetrachloride (0.0011) in the styrene polymerization.
  16. [16]
  17. [17]
    CHAPTER 1: Kinetics and Thermodynamics of Radical Polymerization
    A radical polymerization process is basically constituted by just four types of reactions, which are initiation, propagation, transfer and termination.
  18. [18]
    Chain transfer activity of some activated allylic compounds
    Various olefins that are activated towards free radical addition and contain a homolytic leaving group in the allylic position are effective chain transfer.
  19. [19]
    Optimization of the chain transfer agent incremental addition in SBR ...
    Nov 21, 2011 · The incremental addition of chain transfer agent (CTA) in an emulsion styrene‐butadiene polymerization reaction was optimized, ...
  20. [20]
    Radical Polymerization of Acrylates, Methacrylates, and Styrene
    The kinetics of the radical polymerization process of reacting acrylic monomers, methacrylic monomers, and styrene is reviewed.<|control11|><|separator|>
  21. [21]
    [PDF] CHAIN TRANSFER IN POLYMERIZATIONS—MOLECULAR ...
    Chain transfer to polymer results in the formation of a radical site on a polymer chain. The polymerization of the monomer at this site leads to the formation ...
  22. [22]
    Chain Transfer Reaction - an overview | ScienceDirect Topics
    The activation energy for the most favorable CTM mechanism in MA (MA-1) was found to be much higher (by ~20 kJ mol−1) than those for the most favorable ones in ...
  23. [23]
    Ab initio calculation of the rate coefficient for short-chain branching ...
    Short-chain branching in polyethylene, which involves a six-centre transition state, is studied by ab initio quantum mechanics up to the QCISD(T) level.
  24. [24]
    Further studies on the thermal decomposition of AIBN—implications ...
    The products formed during thermal decomposition of AIBN-α-13C at 60° in benzene and dimethyl sulphoxide have been identified and quantified by 13C-NMR.
  25. [25]
    Effects of Initiator Loss by Chain Transfer in Free-Radical ... - jstor
    initiator loss by chain transfer to initiator are more pronounced as shown in figures. 3 and 4. Increases in the instantaneous values for Pn become more ...
  26. [26]
  27. [27]
    Co(II)-Mediated Catalytic Chain Transfer Polymerization (CCTP ...
    ... defined as the ratio of the rate constant for the chain transfer reaction to the rate constant for propagation (1):. C s = k t r s / k p, 1. The chain transfer ...
  28. [28]
    [PDF] kp [ ] 2(fkd kt [ ] 2 - MIT OpenCourseWare
    CTA (chain-transfer-agents). CS x 104. CS x 104. For styrene. Vinyl acetate ... n-butyl mercaptan. 210,000. 480,000. SH. 10.569, Synthesis of Polymers, Fall ...
  29. [29]
    Determination of the Chain‐Transfer Constant from the Mayo ...
    The aim of this study was to investigate the kinetics of the free-radical homopolymerization of. MMA and S in the presence of a CTA as the mercaptan DDT at two ...
  30. [30]
    Polydispersity Index - an overview | ScienceDirect Topics
    The polydispersity index is the ratio of the two molecular weights just described: P I = M w ¯ M n ¯ . Note that PI>1, except in the case where all the ...
  31. [31]
    Chain Growth Polymerization - an overview | ScienceDirect Topics
    Chain transfer requires higher activation energy than propagation but becomes a competitive pathway only at higher temperatures. Bimolecular radical coupling/ ...Missing: E_tr | Show results with:E_tr
  32. [32]
    Chain Transfer in the Polymerization of Styrene - ACS Publications
    Inverse Temperature Dependence of Chain Transfer Rate Constant for a Chromium Metalloradical in Polymerization of MMA. Macromolecules 2003, 36 (26) , 9716 ...Missing: E_tr | Show results with:E_tr
  33. [33]
    Living Free-Radical Polymerization by Reversible Addition ...
    Synthesis of Polyester-Polystyrene Hybrid Gel by PET-RAFT Polymerization in “RAFT ... Macromolecules. Cite this: Macromolecules 1998, 31, 16, 5559–5562. Click to ...
  34. [34]
    50th Anniversary Perspective: RAFT Polymerization—A User Guide
    Since its first report in 1998, RAFT polymerization has gathered a tremendous momentum from both academia and industry. Nearly 20 years ...The RAFT Process · Monomer Class and RAFT Agent · Process of Polymerization
  35. [35]
    Toward Living Radical Polymerization - ACS Publications
    This Account focuses on nitroxide-mediated polymerization (NMP) and polymerization with reversible addition−fragmentation chain transfer (RAFT), two of the ...
  36. [36]
    Advances and prospects of RAFT polymerization-derived ...
    However, the presence of sulfur in RAFT agents is a source of color, odor, and potential degradation, which could be a disadvantage for many applications ...Missing: period | Show results with:period
  37. [37]
    [PDF] Shining a light on catalytic chain transfer - Pure
    Jan 1, 2002 · ... methyl methacrylate. (MMA) and styrene. For MMA a ... chain transfer coefficients, CT, for methyl methacrylate is given in Table 2.1.Missing: C_m | Show results with:C_m
  38. [38]
    [PDF] Overview of cobalt-mediated radical polymerization - HAL
    Mar 29, 2021 · According to the commonly accepted mechanism [52], a cobalt(II) complex reacts with a polymeric radical 21 forming an alkyl-cobalt(III) complex ...
  39. [39]
    Co(II)-Mediated Catalytic Chain Transfer Polymerization (CCTP ...
    May 28, 2025 · CCTP was conducted in both batch and with three different flow reactors to examine the influence of operational conditions on the synthesis of ...Introduction · Experimental Section · Results and Discussion · Conclusions
  40. [40]
    [PDF] Catalytic chain transfer polymerisation (cctp) of methyl methacrylate
    Abstract: Catalytic chain transfer polymerisation of methyl methacrylate has been investigated by a range of cobalt(I1) complexes. The effect of catalyst ...Missing: development 1990s
  41. [41]
    High fidelity vinyl terminated polymers by combining RAFT and ...
    In this work, we combine reversible addition fragmentation transfer polymerization (RAFT) with cobalt catalytic chain transfer (CCT) to obtain ω-unsaturated ...
  42. [42]
    [PDF] Hermann Staudinger - Nobel Lecture
    Classification of macromolecular substances. I. Substances occurring in nature. 1. Hydrocarbons - rubber, guttapercha, balata. 2. Polysaccharides - celluloses, ...Missing: transfer | Show results with:transfer
  43. [43]
    [PDF] Research in Synthetic Rubber - National Science Foundation
    The synthetic rubber research program was transferred to the Rubber Reserve. Company when the Office of Rubber Director was abolished in September 1944.
  44. [44]
    Chain transfer in the polymerisation of styrene iii. The reactivities of ...
    Chain transfer in the polymerisation of styrene iii. The reactivities of hydrocarbons toward the styrene radical. R. A. Gregg and F. R. Mayo, Discuss.Missing: C_s 1950s
  45. [45]
    The Use of S35 in the Measurement of Transfer Constants
    When Mayo falls short ( C tr ≫ 1): the use of cumulative chain length distribution data in the determination of chain transfer constants ( C tr ) for ...
  46. [46]
    Electron spin resonance spectra of polymethyl methacrylate and ...
    Abstract. It is shown that there exists a difference in the electron spin resonance spectra of commercial-grade and laboratory-purified samples of polymethyl ...Missing: backbiting | Show results with:backbiting
  47. [47]
    New developments in emulsion?PVC polymerisation to produce ...
    Aug 6, 2025 · PVC polymerisation to produce polymers with the potential of reduced or zero VOC requirements when used in plastisol applications. March 2007 ...
  48. [48]
    Polyvinyl chloride applications along with methods for managing its ...
    Dec 13, 2024 · Common ingredients for PVC synthesis include water, VCM-soluble initiator, dispersant, and chain transfer agent.
  49. [49]
    Self-Assembled Protein–Polymer Nanoparticles via Photoinitiated ...
    Feb 13, 2025 · In this study, we report the synthesis of reduction-responsive protein–polymer nanoparticles by a photoinitiated polymerization-induced self-assembly (photo- ...
  50. [50]
    Polymerization-Induced Self-Assembly for Efficient Fabrication of ...
    Apr 11, 2023 · Fabrication of functional nano-objects through RAFT dispersion polymerization and influences of morphology on drug delivery. ACS Appl Mater ...
  51. [51]
    Synthesis of linear low-density polyethylene waxes through chain ...
    Nov 1, 2024 · LLDPE waxes are synthesized using benzocycloalkyl nickel catalysts, yielding branched waxes with low molecular weights and diverse molecular ...
  52. [52]
    Metallocene Polyethylene Wax Synthesis - ACS Publications
    The catalysts demonstrated very high ethylene polymerization activities, almost 3 times higher than sMAO-(CpnBu)2ZrCl2 (industrial standard zirconocene catalyst) ...
  53. [53]
    Advanced Polyethylene Wax via Zinc‐Enhanced Metallocene ...
    Oct 28, 2025 · Zinc‐mediated chain transfer mechanisms and process optimization (Zn/Zr ratios, catalyst loading, temperature, pressure) are investigated ...
  54. [54]
    Red Light-Driven, Oxygen-Tolerant RAFT Polymerization Enabled ...
    Photoinduced reversible-deactivation radical polymerization (photoRDRP) techniques enable the synthesis of well-defined polymers under mild conditions.
  55. [55]
    Aqueous photo-RAFT polymerization under ambient conditions
    May 24, 2024 · Oxygen tolerant RAFT polymerization of MSEMA, which is reactive toward the reactive oxygen species generated in PET-RAFT, was also successful.
  56. [56]
    Review Advances in enzyme-catalysis-mediated RAFT polymerization
    Jul 21, 2021 · As is well-known, the intolerance to oxygen is one of the core limitations for all controlled, radical polymerization processes, including RAFT ...
  57. [57]
    Coordinative Chain Transfer Polymerization of Sustainable Terpene ...
    The present investigation involves the coordinative chain transfer polymerization (CCTP) of biobased terpenes in order to obtain sustainable polymers.
  58. [58]
    Coordinative Chain Transfer Polymerization of Sustainable Terpene ...
    Jul 17, 2022 · The present investigation involves the coordinative chain transfer polymerization (CCTP) of biobased terpenes in order to obtain sustainable ...
  59. [59]
    Bio-based polymyrcene with highly ordered structure via solvent free ...
    This work describes the first reversible addition fragmentation chain transfer (RAFT) polymerization of myrcene, a simple monoterpene.
  60. [60]
    An investigation on chain transfer to monomers and initiators ...
    Sep 21, 2022 · Chain transfer constant can be measured by conventional Mayo equation at low concentration with the plot between degree of polymerization (DP) ...
  61. [61]
    DFT Study on the Propagation Kinetics of Free-Radical ...
    We explored chain length effects by examining addition reactions of monomeric, dimeric, trimeric, and tetrameric radicals to the monomers. We have also modeled ...
  62. [62]
    Density Functional Theory Study of Free-Radical Polymerization of ...
    Aug 6, 2025 · This study deals with modeling the propagation and the chain transfer reactions in the free radical polymerization of ethylene, methyl ...
  63. [63]
    Click Chemistry for Biofunctional Polymers: From Observing to ...
    Dec 2, 2024 · In this review, we discuss various types of click reactions─including azide–alkyne cycloadditions, nucleophilic and radical thiol click reactions.
  64. [64]
    Click chemistry strategies for the accelerated synthesis of functional ...
    Apr 29, 2021 · Click chemistry is one of the most powerful strategies for constructing polymeric soft materials with precise control over architecture and functionality.
  65. [65]
    Metal-free atom transfer radical polymerization with ppm catalyst ...
    Jan 18, 2021 · Organocatalytic atom transfer radical polymerization (O-ATRP) is recently emerging as an appealing method for the synthesis of metal-free ...
  66. [66]
    Progress and Perspectives Beyond Traditional RAFT Polymerization
    Aug 26, 2020 · Challenges still remain in the scaling and translation of these techniques into industrial settings, as well as fully exploiting the ...
  67. [67]
    Toward Green Atom Transfer Radical Polymerization: Current Status ...
    Feb 17, 2022 · This review focuses on the relation between ATRP technology and the 12 principles of green chemistry, which are paramount guidelines in sustainable research ...
  68. [68]
    Sulfur-free and Surfactant-free RAFT-Mediated Hybrid Emulsion ...
    Sep 4, 2022 · However, the challenges of the conventional organic sulfur-based RAFT agents causing color and undesired odor problems still attract attention.
  69. [69]
  70. [70]
    Tuning the Mechanical Properties of 3D‐printed Objects by Mixing ...
    Aug 9, 2024 · In this study, a facile way is introduced to modulate the mechanical properties of 3D printed objects by mixing two chain transfer agents (CTAs) ...Abstract · Introduction · Results and Discussion · Conclusion
  71. [71]
    Basic concepts and tools of artificial intelligence in polymer science
    Apr 28, 2025 · 8 AI-guided closed-loop optimization of reversible addition–fragmentation chain transfer (RAFT) polymerization using Bayesian optimization.Missing: constants | Show results with:constants