Fact-checked by Grok 2 weeks ago

Atom transfer radical polymerization

Atom transfer radical polymerization (ATRP) is a versatile and controlled radical polymerization technique that enables the precise synthesis of polymers with tailored molecular weights, low polydispersities, and complex architectures by establishing a rapid equilibrium between dormant alkyl halide species and propagating radicals via a transition metal catalyst, typically copper-based. Independently developed in 1995 by Krzysztof Matyjaszewski and colleagues at Carnegie Mellon University using copper catalysts, and by Mitsuo Sawamoto and colleagues using ruthenium catalysts, ATRP draws from earlier atom transfer radical addition (ATRA) reactions and has since become a cornerstone of living/controlled radical polymerization methods due to its compatibility with a wide range of vinyl monomers and functional groups. The mechanism of ATRP relies on the reversible homolytic cleavage of a carbon-halogen bond in an initiator or dormant polymer chain by a transition metal complex (e.g., Cu(I)/ligand), generating a carbon-centered radical that propagates by adding monomers while the oxidized metal species (e.g., Cu(II)/ligand) deactivates radicals to prevent irreversible termination. This equilibrium, characterized by the ATRP equilibrium constant K_\text{ATRP}, maintains low radical concentrations (typically 10^{-8} to 10^{-6} M), ensuring high chain-end fidelity and enabling the formation of block copolymers, star polymers, and graft copolymers through sequential or iterative additions. Key advantages of ATRP include its robustness under industrially relevant conditions, such as aqueous or protic media, and the ability to incorporate functional monomers without protecting groups, which has facilitated its through licenses to over 17 international companies as of 2024 for applications in coatings, adhesives, and biomedical materials. Recent advances have focused on greener variants, such as photoATRP and electroATRP, which use light or to regenerate the activator , reducing catalyst loading to parts-per-million levels and minimizing metal contamination in polymers for sensitive applications like and . These developments, informed by detailed mechanistic studies, underscore ATRP's ongoing evolution as a powerful tool for designing advanced polymeric materials with enhanced performance and .

Introduction

Overview

Atom transfer radical polymerization (ATRP) is a reversible-deactivation (RDRP) technique that employs catalysts to mediate the growth of chains, enabling precise control over architecture and properties. In ATRP, is initiated by alkyl halides, which generate radicals that add to monomers, forming growing chains; these chains are then reversibly deactivated through atom transfer with a lower-oxidation-state metal complex, such as Cu(I), maintaining a low concentration of active radicals to minimize termination. This equilibrium process, often mediated by Cu-based catalysts with ligands like bipyridine, allows for the synthesis of polymers with predetermined molecular weights, narrow molecular weight distributions (polydispersity index, PDI, typically <1.5), and high chain-end fidelity, facilitating subsequent chain extensions or functionalizations. Compared to other RDRP methods, such as (NMP) and (RAFT), ATRP offers superior versatility in monomer compatibility and reaction conditions due to its robust transition metal catalysis, though it requires careful handling of metal residues. ATRP was independently demonstrated in 1995 by the and groups, who successfully polymerized styrene and acrylates using Cu(I)/bipyridine catalysts and alkyl halide initiators, achieving living-like characteristics with linear molecular weight evolution and low PDI values around 1.1–1.3.

Historical development

Atom transfer radical polymerization (ATRP) was independently developed in 1995 by Krzysztof Matyjaszewski and colleagues at and by Mitsuo Sawamoto and colleagues at as a controlled radical polymerization technique to synthesize polymers with predetermined molecular weights and low polydispersities. The initial concept drew from earlier atom transfer processes in organic synthesis, adapting them to radical polymerization using transition metal catalysts to reversibly activate dormant alkyl halide initiators. The first communication reported the successful polymerization of and mediated by Cu(I)/bipyridine complexes in 1995, demonstrating linear growth of molecular weight with conversion and chain-end fidelity. A follow-up full paper that year expanded on the mechanism and scope, confirming the "living"/controlled nature through kinetic studies and end-group analysis. Early advancements in the late 1990s addressed initiation challenges and broadened applicability. In 1997, reverse ATRP was introduced, employing conventional radical initiators and higher-oxidation-state metal complexes (e.g., Cu(II)) to generate the active catalyst in situ, simplifying handling by reducing sensitivity to oxygen. This variant facilitated heterogeneous and emulsion polymerizations. Commercialization efforts began around this time, with Matyjaszewski founding the Controlled Radical Polymerization Consortium in 2001 to transfer knowledge to industry, followed by the establishment of ATRP Solutions Inc. in 2006 to develop and supply ATRP reagents and specialty polymers for applications in coatings, adhesives, and biomaterials. The 2000s saw significant expansions, including adaptations for aqueous and biomolecular systems to enable environmentally friendly and biomedical uses. Aqueous ATRP emerged around 2000 with miniemulsion techniques for water-insoluble monomers, using hydrophilic ligands to stabilize catalysts in dispersed media. Biomolecular applications advanced through "grafting-from" strategies on proteins and surfaces, yielding hybrid materials for drug delivery, reported in key studies from the mid-2000s. A pivotal milestone was the 2006 introduction of activators regenerated by electron transfer (ARGET) ATRP, which used reducing agents like ascorbic acid to maintain low catalyst concentrations (ppm levels), enhancing efficiency and minimizing side reactions. Post-2020 developments have focused on sustainability, integrating biocatalytic and photoinduced methods to further reduce metal use and enable mild conditions. Enzyme-mediated , such as cytochrome c-catalyzed systems, was reported in 2022, allowing oxygen-tolerant polymerization in aqueous media without synthetic reductants, leveraging biological catalysts for greener processes. Photoinduced variants, including open-air green-light in 2022, have advanced metal-free or low-metal operations using photoredox catalysis for precise temporal control and reduced energy input. These innovations have overcome key challenges like catalyst efficiency and metal contamination—issues that previously hindered industrial scalability—through techniques like and , enabling ppm-level copper usage and facile removal via precipitation or filtration, thus broadening 's adoption in sustainable manufacturing.

Fundamental Components

Monomers

Atom transfer radical polymerization (ATRP) is compatible with a wide range of vinyl monomers that can form stable propagating radicals. The primary classes include acrylates, methacrylates, styrenes, acrylamides, and vinyl chloride, while less common examples encompass norbornenes and dienes. For successful polymerization, monomers must exhibit sufficient radical stability to allow for the reversible deactivation central to ATRP, ensuring low concentrations of active radicals and minimal termination events. Additionally, they require compatibility with the metal catalysts typically employed, avoiding side reactions such as excessive chain transfer or catalyst deactivation that could compromise control over molecular weight and polydispersity. Representative examples highlight the variability in reactivity. Methyl acrylate undergoes rapid polymerization, achieving high conversions in hours, whereas styrene proceeds more slowly, often requiring days under similar conditions. Acrylonitrile polymerizes even faster than acrylates, while vinyl acetate exhibits extremely sluggish rates, rendering it impractical for most ATRP applications. ATRP's versatility extends to copolymerization, enabling the synthesis of block copolymers by sequential addition of different monomers, such as combining acrylates with styrenes to produce materials with tailored phase-separated morphologies. Early limitations included sensitivity to acidic monomers, which could protonate and deactivate metal catalysts, necessitating protective strategies or alternative conditions. Post-2020 advances have expanded compatibility to include functional monomers bearing hydroxy, amino, or acidic groups, as well as bio-based variants from renewable resources, facilitating the preparation of stimuli-responsive polymers like those exhibiting lower or upper critical solution temperatures for applications in drug delivery and bioconjugates.

Initiators

In atom transfer radical polymerization (ATRP), initiators serve as the precursors to the dormant polymer chain ends, providing the alkyl halide functionality that establishes the reversible redox equilibrium essential for controlled radical polymerization. These initiators are typically organic compounds containing a transferable halogen atom, which allows for the generation of the initial propagating radicals upon activation by the catalyst. Common initiators are alkyl halides, particularly those with bromine or chlorine as the halogen due to their suitable bond strengths for reversible homolytic cleavage. For instance, ethyl 2-bromoisobutyrate is widely used for the polymerization of acrylates and methacrylates because of its compatibility with these monomers and its ester group, which stabilizes the radical intermediate. Similarly, benzyl chloride is a standard choice for styrene polymerization, offering good solubility in organic solvents and efficient initiation kinetics. These halogens (Br and Cl) are preferred over iodine, which forms weaker bonds leading to less control, or fluorine, which is too stable for activation. The design of initiators emphasizes several key principles to ensure efficient polymerization control. The halogen must exhibit appropriate affinity for the metal catalyst, typically copper, to facilitate reversible atom transfer without side reactions. Solubility in the reaction medium is crucial to promote homogeneous initiation, while the incorporation of functional groups—such as hydroxyl, amino, or vinyl moieties—enables post-polymerization modification of chain ends for applications in biomaterials or advanced materials. Initiators often feature activating substituents on the α-carbon, like carbonyl or aryl groups, to weaken the C-X bond and enhance radical generation. Initiation proceeds via homolytic cleavage of the C-X bond in the initiator by the Cu(I) catalyst, generating an alkyl radical that adds to the monomer to form the first propagating chain and a Cu(II)-X complex. This step establishes the dormant species (P-X) that participate in the subsequent equilibrium. To achieve living polymerization characteristics, such as linear molecular weight growth and low polydispersity, an excess of initiator relative to the catalyst is employed, ensuring rapid and quantitative initiation of all chains early in the reaction. Variations in initiator structure expand ATRP's versatility for complex architectures. Macroinitiators, which are pre-synthesized polymers bearing alkyl halide end groups, allow for the formation of block copolymers or graft polymers by initiating secondary chains from the dormant ends. Multifunctional initiators, such as tetrakis(bromomethyl)benzene or cyclotriphosphazene derivatives with multiple halide arms, enable the core-first synthesis of star polymers, where each arm grows independently from the central core, yielding materials with high branching and compact structures.

Catalysts

Atom transfer radical polymerization (ATRP) primarily employs transition metal catalysts in their lower oxidation state, such as Cu(I), to activate dormant alkyl halide chain ends by abstracting the halogen atom, generating a propagating radical, while the corresponding higher oxidation state species, like Cu(II), deactivates the radical by transferring the halogen back, establishing a rapid equilibrium that controls the polymerization. Copper-based systems, particularly Cu(I)/Cu(II) complexes with halides like CuBr or CuCl, have been the most widely adopted due to their efficiency in achieving low polydispersity indices (typically <1.5) and precise molecular weight control across various monomers. Alternative transition metals have been explored for specialized applications where copper's limitations, such as toxicity or color contamination, are concerns. Iron catalysts, such as Fe(II) halides (e.g., FeBr₂), offer a biocompatible option with similar redox-mediated atom transfer, enabling controlled polymerization in aqueous media and showing high cell viability in biomedical contexts. Ruthenium complexes, like RuCl₂(PPh₃)₃, provide high activity for functional monomers, while nickel and palladium systems, such as NiBr₂ with phosphine ligands or PdCl₂, demonstrate viability in niche polymerizations but with narrower applicability due to slower rates or higher costs. The evolution of ATRP catalysts has focused on reducing metal loading to parts per million (ppm) levels to address industrial scalability and environmental concerns. Activators regenerated by electron transfer (ARGET) ATRP, introduced in 2006, utilizes Cu(I)/Cu(II) at <100 ppm with reducing agents like ascorbic acid to continuously regenerate the activator, minimizing catalyst accumulation and enabling large-scale synthesis without compromising control. Post-2015 developments shifted toward metal-free organic photoredox catalysts, such as phenothiazine or oxygen-doped anthanthrene derivatives, which mimic the redox role via single electron transfer under visible light or sunlight, achieving dispersities <1.20 at loadings as low as 50 ppb and eliminating metal residues entirely. Post-2020 advancements emphasize biocompatible catalysts for biomedical applications, with iron-based systems gaining prominence due to their low toxicity and abundance. For instance, Fe(0)-mediated supplemental activator and reducing agent (SARA) operates at <50 ppm, supporting surface-initiated polymerizations on cells or implants without viability loss, ideal for tissue engineering and drug delivery. Key factors influencing catalyst selection include oxidation stability to prevent irreversible radical termination, solubility in reaction media for homogeneous catalysis, and reduced toxicity for scaling to industrial or clinical uses, often addressed by ligand pairing to enhance these properties.

Ligands

In atom transfer radical polymerization (ATRP), ligands play a pivotal role in forming stable complexes with transition metal catalysts, primarily copper, to enable efficient control over the polymerization process. Common ligand types include nitrogen-based structures such as 2,2'-bipyridine (bpy) and N,N,N′,N′′,N′′-pentamethyldiethylenetriamine (PMDETA), which were among the earliest employed for copper-mediated ATRP, offering bidentate and tridentate coordination, respectively. Phosphorus-based ligands like triphenylphosphine (PPh₃) have also been utilized, particularly in early copper systems, though they provide lower activity compared to nitrogen donors due to weaker binding and electronic mismatches. Multidentate nitrogen ligands, such as tris(2-pyridylmethyl)amine (TPMA), enhance solubility and catalytic performance by providing tetradentate coordination, allowing lower catalyst loadings and broader monomer compatibility. The primary functions of these ligands include increasing the solubility of Cu(I) species in organic media, which is essential for homogeneous catalysis, and modulating the redox potential of the metal center to accelerate the activation of dormant alkyl halide chains. By stabilizing the Cu(I)/Cu(II) equilibrium, ligands prevent the disproportionation of Cu(I) into Cu(0) and Cu(II), maintaining high concentrations of the activator species and ensuring low radical concentrations for controlled polymerization. Additionally, ligand exchange strategies facilitate catalyst recycling post-polymerization, such as through CO₂-switchable systems that enable separation and reuse of the ligand while removing copper residues. Ligand design in ATRP emphasizes tuning steric and electronic properties to optimize compatibility with specific monomers and catalysts, thereby influencing activation rates and overall efficiency. For instance, electron-donating groups on nitrogen ligands lower the redox potential, promoting faster activation for less reactive halides, while bulky substituents enhance selectivity by sterically hindering side reactions. Pyridine-imine ligands, featuring an imine nitrogen donor, exemplify this approach, providing high activity for fast-polymerizing monomers like styrene and methyl methacrylate by balancing electronic activation and steric control in copper complexes. Recent advances focus on specialized ligands to expand ATRP applications, particularly in environmentally benign conditions. Water-soluble variants, such as TPMA, enable aqueous ATRP by improving catalyst solubility without compromising activity, facilitating the polymerization of hydrophilic monomers like acrylamides. Recent developments (as of 2019) include bio-derived ligands, such as those incorporating morpholine units from renewable sources, which support sustainable iron- or copper-catalyzed ATRP while maintaining precise control over polymer architecture. These innovations enhance catalyst performance by enabling lower metal loadings and greener processes, aligning ATRP with sustainability goals.

Solvents

In atom transfer radical polymerization (ATRP), the choice of solvent plays a crucial role in maintaining homogeneous reaction conditions, ensuring the solubility of monomers, initiators, and catalysts, and influencing the overall polymerization kinetics. Common organic solvents such as and are frequently employed for homogeneous ATRP of non-polar monomers like and , as they dissolve the transition metal catalyst complexes effectively and promote uniform radical generation and propagation. Polar solvents, including and , are preferred for polymerizing ionic or polar monomers, such as with charged functionalities, due to their ability to solvate ionic species and enhance catalyst activity. The polarity of the solvent significantly affects the ATRP equilibrium constant (K_ATRP), which governs the balance between dormant alkyl halide species and active radicals, thereby modulating polymerization rates and molecular weight control. For instance, more polar solvents like can shift the equilibrium toward activation by stabilizing the lower-oxidation-state catalyst, leading to faster propagation rates compared to less polar media like , where rates are slower but control over polydispersity remains high (typically Mw/Mn = 1.1–1.3). Additionally, solvents can coordinate directly with the catalyst, altering its redox potential and influencing deactivation rates, which is particularly evident in iron-mediated ATRP where polar solvents like act as ancillary ligands to improve solubility and reaction efficiency. Efforts toward green chemistry have driven the adoption of environmentally benign alternatives to traditional volatile organic solvents in ATRP. Supercritical carbon dioxide (scCO2) serves as a non-toxic, recyclable medium that supports the polymerization of fluoromonomers and enables easy product separation without residue, though it requires high-pressure equipment. Ionic liquids, such as 1-butyl-3-methylimidazolium-based salts, provide tunable polarity and facilitate catalyst recycling in ATRP of acrylates, offering better control (PDI < 1.3) and reduced waste compared to molecular solvents, albeit with higher costs and potential toxicity concerns. Aqueous systems represent another sustainable option, particularly for biocompatible polymers like poly(ethylene glycol) (PEG) and oligo(ethylene oxide) methyl ether acrylate (OEOMA), where water acts as a benign solvent with high heat capacity for exothermic reactions, achieving low polydispersities (Mw/Mn = 1.15–1.28) using minimal catalyst loadings (20–100 ppm Cu). Challenges in solvent selection include preventing catalyst precipitation, which is common in non-polar media due to poor solubility of metal complexes, potentially leading to inhomogeneous reactions and loss of control. Post-2020 advancements have emphasized solvent-free approaches and bio-based media to enhance sustainability; for example, bio-derived solvents like (from cellulose) and Cygnet 0.0 have replaced toxic aprotic solvents such as DMF in supplemental activator and reducing agent (SARA) , yielding up to 54-fold higher effective mass yields and enabling branched polymer synthesis from natural cores like β-cyclodextrin with reduced catalyst (75 ppm) and minimal side reactions. These developments prioritize recyclability and biocompatibility while addressing residual catalyst removal through benign extraction methods.

Mechanism and Kinetics

General mechanism

Atom transfer radical polymerization (ATRP) operates through a reversible redox process that establishes a dynamic equilibrium between dormant alkyl halide species and active propagating radicals, mediated by a transition metal catalyst such as copper in its lower oxidation state complexed with ligands. This equilibrium enables controlled chain growth while suppressing irreversible termination reactions. The process involves three primary steps: activation, propagation, and deactivation. In the activation step, the dormant species, typically an alkyl halide (R-X), undergoes homolytic cleavage of the carbon-halogen bond facilitated by the Cu(I) catalyst complex (Cu(I)L_n), generating a carbon-centered radical (R•) and the oxidized Cu(II) species (Cu(II)L_nX): \text{R-X} + \text{Cu(I)L}_n \rightarrow \text{R}^\bullet + \text{Cu(II)L}_n\text{X} This step initiates the formation of active radicals from the initiator or growing chain ends. Propagation follows, where the generated radical (R•) rapidly adds to a vinyl monomer (M), extending the polymer chain to form a new radical (RM•): \text{R}^\bullet + \text{M} \rightarrow \text{RM}^\bullet This addition reaction continues iteratively, allowing for the controlled incorporation of monomer units. Deactivation then occurs as the propagating radical (RM•) reacts with the Cu(II)L_nX complex, reforming the dormant alkyl halide (RM-X) and regenerating the Cu(I)L_n activator: \text{RM}^\bullet + \text{Cu(II)L}_n\text{X} \rightarrow \text{RM-X} + \text{Cu(I)L}_n The rapid and reversible nature of this deactivation step maintains a low steady-state concentration of radicals, typically on the order of 10^{-8} M, which minimizes bimolecular termination events such as radical combination or disproportionation. The overall cycle thus establishes a fast equilibrium that favors dormant species, ensuring high chain-end fidelity and living-like polymerization characteristics. Although termination is suppressed, minor side reactions can occur, including β-hydride elimination from the propagating radical, which leads to chain transfer and unsaturated end groups, and infrequent radical-radical combination. The persistent radical effect further contributes to control by causing an accumulation of the persistent Cu(II) species early in the reaction, which shifts the equilibrium toward deactivation and reduces radical concentrations over time. The general mechanism can be schematically represented in text form as a reversible cycle:
Dormant (R-X)  ⇌  Active Radical (R•) + Cu(II)L_nX
     ↑ Activation by Cu(I)L_n          ↓ Propagation ( + M → RM• )
     ↓ Deactivation by Cu(II)L_nX      ↑
Dormant (RM-X)  ⇌  Active Radical (RM•) + Cu(II)L_nX
This equilibrium underscores the role of the catalyst in toggling between active and dormant states throughout the polymerization.

Equilibrium and rate constants

The equilibrium in atom transfer radical polymerization (ATRP) is characterized by the constant K_{\text{ATRP}} = \frac{[R^\bullet][\text{Cu(II)X}]}{[\text{RX}][\text{Cu(I)}]} = \frac{k_{\text{act}}}{k_{\text{deact}}}, where RX represents the dormant alkyl halide species, R^\bullet the propagating radical, and Cu(I)/Cu(II)X the catalyst in its reduced and oxidized forms, respectively. This equilibrium maintains a low radical concentration to minimize termination while allowing controlled chain growth. Typical values of K_{\text{ATRP}} range from $10^{-9} to $10^{-4} M, depending on the catalyst, ligand, solvent, and temperature; these small values ensure effective control by favoring the dormant species. The activation rate constant k_{\text{act}} is typically $10^{1} to $10^{3} M^{-1} s^{-1}, reflecting the relatively slow generation of radicals, while the deactivation rate constant k_{\text{deact}} is significantly faster at $10^{7} to $10^{8} M^{-1} s^{-1} , rapidly recapturing radicals to preserve chain-end fidelity. The propagation rate constant k_{\text{p}} varies with the monomer but is essential for chain extension, typically on the order of $10^{2} to $10^{4} M^{-1} s^{-1} for common vinyl monomers like acrylates or styrenes. The overall polymerization rate is expressed as R_{\text{p}} = k_{\text{p}} [\text{M}] [R^\bullet], where [M] is the monomer concentration, and the radical concentration is given by [R^\bullet] = K_{\text{ATRP}} \frac{[\text{RX}][\text{Cu(I)}]}{[\text{Cu(II)X}]}. This formulation highlights how the equilibrium directly influences the rate by controlling [R^\bullet], with higher Cu(I) concentrations shifting the equilibrium toward activation and accelerating polymerization. These kinetic parameters exhibit temperature dependence, with k_{\text{act}} and k_{\text{deact}} showing higher activation energies (typically 20–50 kJ/mol) compared to k_{\text{p}} (around 20–30 kJ/mol), leading to a shift in K_{\text{ATRP}} that increases radical concentration and rate at elevated temperatures. Kinetic simulations incorporating these constants, such as those using differential equations or software like PREDICI, enable prediction of the polydispersity index (PDI), often achieving PDI values close to 1.1–1.2 under optimized conditions by balancing activation/deactivation ratios.

Chain-end functionality

In atom transfer radical polymerization (ATRP), the retention of halogen chain-end functionality is typically greater than 95%, attributed to the rapid deactivation of propagating radicals that maintains a low radical concentration and minimizes termination events. This high fidelity is routinely quantified using techniques such as nuclear magnetic resonance (NMR) spectroscopy and matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry, which allow precise assessment of the proportion of dormant species bearing the halogen terminus. The low radical concentration arising from the activation-deactivation equilibrium further supports this preservation, ensuring that side reactions remain limited. The preservation of these functional chain ends is crucial for advanced polymer architectures, particularly enabling the synthesis of block copolymers through sequential monomer addition in subsequent ATRP steps, where the halogen terminus acts as a macroinitiator. Additionally, the halogen ends can be quantitatively transformed into other reactive groups, such as azides via nucleophilic substitution with sodium azide, facilitating copper-catalyzed azide-alkyne cycloaddition () chemistry for further conjugation or grafting applications. Despite the generally high retention, degradation of chain-end fidelity can occur through side reactions, including β-hydride elimination, which leads to loss of the halogen group under certain conditions, particularly at elevated temperatures or with certain monomers. Strategies to mitigate such losses include the addition of excess deactivator to enhance the rate of radical trapping and suppress radical-mediated side reactions. Post-polymerization modifications allow for the quantitative removal or alteration of these chain ends to tailor polymer properties for specific uses, such as reducing the halogen terminus to a saturated alkyl group using hydrogen gas with heterogeneous catalysis, which proceeds efficiently without compromising the polymer backbone. These end-group adjustments are essential for applications requiring biocompatibility or reduced reactivity, such as in biomedical materials.

Advantages and Limitations

Advantages

Atom transfer radical polymerization (ATRP) provides exceptional control over polymer molecular weight and architecture, enabling the synthesis of macromolecules with narrow molecular weight distributions, typically polydispersity indices (PDI) of 1.1–1.5, through linear chain growth and a dynamic equilibrium that favors dormant species over active radicals. This high degree of control arises from the reversible activation-deactivation process, which minimizes termination events and ensures predictable molecular weights based on the initiator-to-monomer ratio. Furthermore, ATRP exhibits excellent tolerance to a variety of functional groups, such as alcohols, amines, and ethers, allowing the direct polymerization of functionalized monomers without the need for protecting groups. High chain-end fidelity in ATRP supports subsequent transformations, such as block copolymer extensions or end-group modifications. The versatility of ATRP extends to a broad range of vinyl monomers, including (meth)acrylates, styrenes, acrylamides, and acrylonitriles, facilitating the preparation of diverse polymer structures under mild conditions, often at room temperature or in aqueous media. This compatibility enables the synthesis of complex architectures, such as block copolymers, star polymers, and polymer brushes, using strategies like sequential monomer addition or grafting techniques. For instance, well-defined block copolymers like poly(ethylene glycol)-b-poly(N-isopropylacrylamide) have been prepared with PDIs below 1.2, demonstrating the technique's utility in creating stimuli-responsive materials. ATRP variants, such as activators regenerated by electron transfer (ARGET) and initiators for continuous activator regeneration (ICAR), achieve scalability through low catalyst loadings, often at parts-per-million (ppm) levels of copper, reducing costs and facilitating industrial production. These developments support large-scale applications, including the manufacture of nanostructured materials via flow chemistry or heterogeneous systems. Recent advancements in photoinduced ATRP (photoATRP), particularly post-2020 methods using visible or near-infrared light, enhance sustainability by minimizing energy consumption and metal usage while maintaining control in oxygen-tolerant environments.

Limitations

One significant limitation of atom transfer radical polymerization (ATRP) is the reliance on transition metal catalysts, particularly copper-based systems, which introduce toxicity concerns and require costly removal processes. High catalyst loadings, often around 1 mol% or up to 10,000 ppm in classical , can contaminate the final polymer, posing risks for biomedical and electronic applications due to copper's toxicity. Removal methods such as adsorption with functionalized silica, ion exchange resins, or electrodeposition are effective but add significant expense and complexity, with residual copper levels needing to be reduced below 100 ppm for practical use. Additionally, these catalysts are highly sensitive to oxygen and impurities, necessitating rigorous deoxygenation and purification steps that limit operational simplicity and increase the risk of side reactions. Polymerization rates in ATRP can be slow for certain monomers, such as methacrylates, due to the accumulation of deactivator species like Cu(II) complexes from unavoidable termination events. This buildup shifts the activation-deactivation equilibrium unfavorably, reducing the concentration of active radicals and prolonging reaction times, which hampers efficiency in producing high-molecular-weight polymers. Scalability remains a challenge for industrial applications. Furthermore, the process is often restricted to specific solvents, with traditional organic solvents raising volatile organic compound (VOC) emission concerns, while aqueous systems introduce additional complications like catalyst hydrolysis. Post-2020 developments in eco-friendly catalyst removal, such as centrifugation in miniemulsion achieving levels below 300 ppb, show progress but remain incomplete for broad industrial adoption, particularly in green chemistry contexts. Emerging organocatalyzed (O-ATRP) variants, using organic photoredox catalysts, offer metal-free alternatives to mitigate toxicity and removal issues, as demonstrated in studies up to 2025. also faces competition from faster reversible deactivation radical polymerization (RDRP) techniques like photoinduced electron/energy transfer-reversible addition-fragmentation chain transfer (), which offer superior oxygen tolerance and milder conditions without metal catalysts. Variants of , such as activator regeneration methods, have been developed to mitigate some of these issues.

ATRP Methods

Conventional ATRP

Conventional ATRP employs a degassed reaction system comprising a Cu(I) catalyst, a coordinating ligand such as , an alkyl halide initiator, and the vinyl monomer. The procedure begins with the preparation of the Cu(I)/ligand complex, followed by the addition of the initiator and monomer under inert atmosphere to establish the reversible activation-deactivation equilibrium central to the process. Typical molar ratios for the components are [M]:[I]:[Cu(I)] = 200:1:1, with the ligand present in a 2-4 fold excess relative to Cu(I) to ensure efficient complexation. These ratios allow for controlled propagation while maintaining a low steady-state radical concentration, typically on the order of 10^{-8} M. Polymerizations are conducted in bulk or organic solution (e.g., toluene or anisole) at temperatures ranging from 80 to 130 °C, depending on the monomer's reactivity and the desired rate. For instance, the bulk polymerization of styrene using ethyl 2-bromoisobutyrate as initiator and CuBr/bipyridine as catalyst at 110 °C reaches 80% conversion in approximately 6 hours, yielding polystyrene with a number-average molecular weight of around 15,000 g/mol and polydispersity index of 1.3. Strict deoxygenation is essential to avoid irreversible oxidation of the Cu(I) species and inhibition of radical propagation, commonly achieved via multiple freeze-pump-thaw cycles or continuous argon bubbling. Additionally, a small amount of Cu(II) (e.g., 5-10% relative to Cu(I)) may be introduced to fine-tune the activator/deactivator ratio and shift the equilibrium toward the dormant species, enhancing control over chain growth. Initial demonstrations of conventional ATRP focused on proof-of-concept syntheses of homopolymers, such as polystyrene and poly(methyl acrylate), confirming linear molecular weight evolution with conversion and high chain-end fidelity for subsequent transformations.

Reverse ATRP

Reverse ATRP, first reported in 1997, represents an alternative initiation approach in atom transfer radical polymerization that employs a transition metal catalyst in its higher oxidation state, such as Cu(II), alongside a conventional thermal radical initiator like AIBN or benzoyl peroxide (BPO). In this system, the decomposition of the initiator generates radicals that react with the Cu(II) complex to produce the activating Cu(I) species and dormant alkyl halide chains in situ, thereby establishing the reversible activation-deactivation equilibrium central to ATRP. This method leverages the same core catalytic components as conventional ATRP but inverts the initiation step to avoid direct addition of alkyl halides. A primary advantage of reverse ATRP is the utilization of air-stable Cu(II) precursors, which are far less sensitive to oxygen than the Cu(I) species used in conventional ATRP, simplifying handling and enabling more robust experimental setups. Furthermore, the approach accommodates standard radical initiators, facilitating its application to monomers that polymerize effectively at elevated temperatures where thermal initiators like decompose efficiently. In a representative homogeneous procedure for styrene polymerization via reverse ATRP, bulk conditions are used with [styrene]_0 = 8.7 M, [AIBN]_0 = 0.045 M, and [CuBr_2]_0/[dNbipy]_0 = 0.045 M (yielding a molar ratio of approximately [M]:[I]:[Cu(II)] = 193:1:1), at 110 °C; the activation equilibrium forms rapidly, and chain growth proceeds via radical propagation, which shifts the equilibrium toward higher conversion. Despite these benefits, reverse ATRP can suffer from early radical-radical termination events, such as coupling, leading to reduced initiator efficiency—particularly pronounced at higher initiator-to-catalyst ratios or for monomers like under high-temperature conditions.

Activator regeneration methods

Activator regeneration methods in atom transfer radical polymerization (ATRP) address the limitation of Cu(II) deactivator accumulation in conventional ATRP by externally replenishing the Cu(I) activator, thereby maintaining the atom transfer equilibrium and enabling efficient polymerization with minimal catalyst loading. These techniques typically start with an excess of air-stable Cu(II) species and use reducing agents or external stimuli to continuously regenerate Cu(I), preventing excessive radical termination and allowing high monomer conversions under milder conditions. This approach shifts the focus from high initial Cu(I) concentrations to dynamic regeneration, enhancing overall process sustainability. Common features of activator regeneration methods include the use of parts-per-million (ppm) levels of copper catalyst, often below 10 ppm, which reduces metal contamination in polymers and lowers costs. Many variants exhibit improved oxygen tolerance compared to traditional , as the regeneration process can counteract oxidative side reactions, and they often achieve faster polymerization rates due to sustained activator availability. For instance, these methods have enabled quantitative conversions of monomers like methyl methacrylate in hours at room temperature. Chemical regeneration methods encompass several sub-types that rely on added reductants to convert Cu(II) back to Cu(I). Activators regenerated by electron transfer (ARGET) ATRP employs organic or inorganic reducing agents, such as ascorbic acid or tin(II) 2-ethylhexanoate, to provide electrons for Cu(II) reduction without generating new radicals. Initiators for continuous activator regeneration (ICAR) ATRP uses conventional radical initiators like azobisisobutyronitrile (AIBN) to produce radicals that reduce Cu(II) via side reactions, maintaining low but steady activator levels. Supplemental activator and reducing agent (SARA) ATRP utilizes zerovalent copper, such as Cu wire or powder, which acts dually as a supplemental activator and reductant through comproportionation with Cu(II), offering robust performance in diverse media. Electrochemical methods, such as electrochemically mediated ATRP (e-ATRP), apply an external electric potential or current to selectively reduce Cu(II) to Cu(I) at the electrode surface, providing precise control over the activation rate without chemical additives. Photochemical variants, known as photoATRP, harness visible or UV light to drive Cu(II) reduction, often in conjunction with photosensitive ligands or additives, enabling spatiotemporal control and operation under ambient conditions. Since 2020, hybrid energy-input methods have emerged, combining multiple stimuli like photoelectrochemistry or mechanophotochemistry to further optimize regeneration efficiency and expand applicability to complex systems. These approaches integrate the tunability of light and electricity for enhanced precision and reduced energy consumption. Overall, activator regeneration methods offer significant benefits, including minimized deactivator buildup for better chain-end fidelity, expanded tolerance to impurities and oxygen, and suitability for a wider range of solvents and monomers, facilitating scalable industrial applications.

Other advanced methods

Activators generated by electron transfer atom transfer radical polymerization (AGET ATRP) is a variant that employs a reducing agent to generate the active Cu(I) species in situ from a stable Cu(II) complex, enabling polymerization in the presence of air without the need for deoxygenation. Introduced in 2006, this method uses ascorbic acid as a non-toxic, air-stable reducing agent, allowing for the synthesis of well-defined polymers such as polystyrene with low polydispersity indices (Đ ≈ 1.2–1.4) under ambient conditions. Ascorbic acid reduces Cu(II) to Cu(I) via electron transfer, maintaining the equilibrium between dormant and active species while minimizing side reactions from oxygen. This approach has been particularly effective in miniemulsion systems, where ascorbic acid facilitates controlled polymerization of n-butyl acrylate, yielding latex particles with molecular weights up to 20,000 g/mol. Simultaneous reverse and normal initiation (SR&NI) ATRP combines elements of both conventional and reverse ATRP initiation mechanisms to achieve better control and functionality. In this process, an alkyl halide initiator and a conventional Cu(I) activator are used alongside a radical precursor (e.g., ) and Cu(II) deactivator, allowing initiation from both dormant species and in situ generated radicals. Developed in 2001, SR&NI enables the use of highly active catalysts under reverse-like conditions while preserving chain-end fidelity for block copolymer synthesis, as demonstrated in the controlled polymerization of n-butyl acrylate and styrene with targeted molecular weights and low dispersities (Đ < 1.3). This method is advantageous for systems requiring functional initiators, producing polymers with high bromine end-group fidelity (>90%) suitable for further chain extensions. Hybrid and bimetallic ATRP systems incorporate dual metal catalysts to enable tandem reactions or enhanced control in complex polymer architectures. For instance, copper-iron bimetallic setups leverage Cu(I)/Cu(II) for primary activation and Fe(II)/Fe(III) for supplementary deactivation or reduction, facilitating of block copolymers with improved tolerance to impurities. A notable example from 2007 involves an immobilized Co(II)/Cu(II) bimetallic catalyst for polymerization, achieving high conversion (>80%) and narrow molecular weight distributions (Đ ≈ 1.2) through synergistic activity. These systems expand ATRP to multifunctional materials, such as gradient copolymers, by combining the selectivity of copper with the robustness of iron or . Metal-free ATRP employs photocatalysts to replace metals, driven by to generate radicals via photoredox . Pioneered in 2014 and advanced with phenoxazine derivatives in 2016, this approach uses strongly reducing photocatalysts like N-aryl phenoxazines, which undergo photoexcitation to donate electrons to alkyl halides, forming initiating radicals without metal residues. In 2016, phenoxazine-based catalysts enabled controlled of methacrylates under mild conditions (, open air), yielding polymers with molecular weights up to 10,000 g/mol and Đ < 1.2, ideal for biomedical applications due to their purity. This method has evolved to include a range of organic dyes, offering spatiotemporal control and compatibility with sensitive substrates. Mechano-ATRP and sono-ATRP utilize force or to activate the catalyst, providing a stimulus-responsive alternative to traditional thermal or photochemical initiation. Sono-ATRP, introduced in 2018, employs ultrasonication to generate localized high-energy conditions that reduce Cu(II) to Cu(I) via cavitation-induced , enabling aqueous polymerization of acrylates with good (Đ ≈ 1.3–1.5) at ambient temperatures. Mechano-ATRP, emerging around 2016–2019, applies stress through piezoelectric materials or high-intensity to induce activation. These methods are energy-efficient and oxygen-tolerant, suitable for polymerizations in heterogeneous media. Biocatalytic ATRP harnesses enzymes as natural catalysts for green, aqueous-phase polymerizations, with advances reported in focusing on laccase-mediated systems. , a copper-containing , facilitates ATRP by oxidizing substrates to generate radicals that reduce Cu(II) to Cu(I), enabling controlled synthesis without synthetic additives. These enzyme-driven approaches promote sustainability, tolerating oxygen and biomolecules, and have been applied to encapsulate therapeutic agents in polymersomes for , allowing polymerization of vinyl monomers like N-vinylimidazole in at neutral , with molecular weights controlled up to 5,000 g/mol and low dispersities (Đ < 1.4).

Applications

Polymers synthesized

Atom transfer radical polymerization (ATRP) enables the synthesis of well-defined homopolymers with precise control over molecular weight and narrow molecular weight distributions, typically achieving dispersities below 1.2. For instance, polystyrene (PS) and poly(methyl methacrylate) (PMMA) are commonly prepared using alkyl halide initiators and copper-based catalysts, allowing targeted molecular weights from 1,000 to over 100,000 g/mol while maintaining living character for further extensions. ATRP facilitates the construction of complex polymer architectures through strategic initiator design and sequential processes. Block copolymers, such as ABA triblocks, are synthesized by sequential monomer addition, where an initial block like is grown from a macroinitiator, followed by chain extension with a second like PMMA, yielding materials with sub-10 nm domain sizes for advanced applications. Graft copolymers are produced via surface-initiated ATRP, dense polymer brushes onto substrates like silica nanoparticles, achieving graft densities up to 0.5 chains/nm² for enhanced interfacial properties. Star polymers are formed using multifunctional initiators, such as cyclotriphosphazene-based cores, to create multi-arm structures with 6–20 arms per star, exhibiting low dispersities (<1.3) and compact hydrodynamic volumes. Hyperbranched polymers are accessed through self-condensing vinyl polymerization of inimers, resulting in branched topologies with degree of branching around 0.5 and controlled molecular weights up to 50,000 g/mol. Functional polymers are readily prepared by ATRP due to the retention of chain-end groups, which can be transformed post-polymerization. These ends are modified via or to incorporate dyes, such as fluorescein for imaging, or biomolecules like proteins for targeted delivery, enabling quantitative conjugation efficiencies over 90%. Inorganic-organic hybrids, exemplified by -silica composites, are synthesized by ATRP from silica surfaces, producing core-shell nanoparticles with improved mechanical toughness, where polymer shells up to 50 nm thick enhance dispersibility in organic media. Representative examples highlight ATRP's versatility in responsive materials. Thermo- and pH-responsive star-like block copolymers, such as poly()-b-poly(2-methoxyethyl -co-oligo() ), are synthesized via photoATRP, exhibiting (LCST) increases with pH in and (UCST) behavior in . Post-2020 developments include degradable networks, like poly(n-butyl ) stars crosslinked via photoATRP, which depolymerize thermally at temperatures below 200°C, achieving near-complete recovery for sustainable .

Practical and industrial uses

ATRP-derived polymers have found applications in coatings and adhesives, particularly in the development of low-volatile (VOC) paints. These materials enable the formulation of waterborne systems that reduce environmental impact while maintaining performance, as demonstrated by advancements in controlled architectures that improve and . In biomaterials, ATRP facilitates the of block copolymers such as PEG-b-PLA for systems, where the amphiphilic structure allows encapsulation and controlled release of therapeutics in response to stimuli like or temperature. These polymers also serve as scaffolds for , providing biocompatible matrices that support cell growth and mimic extracellular environments. For , ATRP enables the of conductive polymers, including polyaniline-modified variants, which enhance charge in devices such as organic light-emitting diodes (OLEDs) and nanostructured films for . These materials offer tunable conductivity and processability, contributing to lightweight, bendable components. In industrial settings, ATRP polymers act as thickeners in , providing rheological control in formulations like shampoos and lotions for improved and . Additionally, they support (EOR) by onto nanoparticles or forming amphiphilic structures that reduce interfacial tension and improve sweep efficiency in reservoirs. Post-2020 developments have leveraged ATRP for through the of renewable monomers and the design of recyclable polymethacrylates, enabling back to monomers under mild conditions to close the material loop and reduce waste. As of , miniemulsion photoATRP has been developed for sustainable production of latexes used in coatings and adhesives, reducing energy use and enabling air-tolerant processes. Despite these advances, industrial adoption faces challenges in scale-up, particularly for batches exceeding quantities, due to issues like catalyst removal, reaction control, and cost in large reactors. Emerging applications include resins, where photoinduced ATRP variants allow precise spatiotemporal control for fabricating complex, functional structures with tailored mechanical properties.

References

  1. [1]
    About ATRP - Matyjaszewski Polymer Group
    ATRP allows scientists to easily form polymers by putting together component parts, called monomers, in a controlled, piece-by-piece fashion.
  2. [2]
    Atom Transfer Radical Polymerization | Chemical Reviews
    This review will focus on the fundamentals of transition metal catalyzed atom transfer radical polymerization (ATRP). We will discuss the current mechanistic ...Introduction · II. Mechanistic Understandings... · Group 8: Ruthenium and Iron
  3. [3]
    Atom Transfer Radical Polymerization - Matyjaszewski Polymer Group
    ATRP is mechanistically related to transition metal mediated atom transfer radical addition (ATRA) reactions.
  4. [4]
    Atom Transfer Radical Polymerization: A Mechanistic Perspective
    Jul 26, 2022 · This Perspective highlights fundamental advances in the design of ATRP reactions and catalysts, focusing on the crucial role that mechanistic studies play.Introduction · Activity and Selectivity of ATRP... · Photocatalyzed ATRP · Outlook
  5. [5]
    Toward Green Atom Transfer Radical Polymerization: Current Status ...
    Atom transfer radical polymerization (ATRP) is a powerful technique for the controlled synthesis of functional polymeric materials. Recent advances in the ...
  6. [6]
    [PDF] Terminology for reversible-deactivation radical polymerization ...
    Nov 18, 2009 · Abstract: This document defines terms related to modern methods of radical polymerization, in which certain additives react reversibly with ...
  7. [7]
    Controlled/"living" radical polymerization. atom transfer radical ...
    Article May 1, 1995. Controlled/"living" radical polymerization. atom transfer radical polymerization in the presence of transition-metal complexes. Click to ...
  8. [8]
    "Living"/Controlled Radical Polymerization. Transition-Metal ...
    Transition-metal-catalyzed atom transfer radical polymerization in the presence of a conventional radical initiator.
  9. [9]
    Matyjaszewski Polymer Group - Carnegie Mellon University
    Professor Matyjaszewski founded the CRP Consortium to expedite transfer of knowledge on CRP procedures to industry in 2001. Virtual Lab Tour. Tap to unmute.Krzysztof Matyjaszewski · About ATRP · Current Group Members · AGET ATRPMissing: Solutions Inc 2000<|control11|><|separator|>
  10. [10]
    ATRP Solutions - Crunchbase Company Profile & Funding
    ATRP Solutions develops and supplies polymers used in manufacturing detergents, paints, personal care, and bio-compatible materials.
  11. [11]
    Diminishing catalyst concentration in atom transfer radical ... - PNAS
    Oct 17, 2006 · ICAR ATRP of styrene (St) was first conducted at low ... J-S Wang, K Matyjaszewski J Am Chem Soc 117, 5614–5615 (1995). Go to ...<|control11|><|separator|>
  12. [12]
    Cytochrome C catalyzed oxygen tolerant atom-transfer radical ...
    Apr 4, 2022 · The cytochrome C catalyzed ATRP showed superior oxygen resistance as it could be performed under aerobic conditions with high dissolved oxygen level.
  13. [13]
    Open-air green-light-driven ATRP enabled by dual photoredox ...
    Sep 20, 2022 · We demonstrate a dual photoredox/copper catalysis that allows open-air ATRP under green light irradiation.Missing: 2020s | Show results with:2020s
  14. [14]
    Future Directions for Atom Transfer Radical Polymerizations
    Feb 27, 2024 · (1) ATRP is a synthetic macromolecular procedure employing controlled radical polymerization to produce well-defined tailored-made polymers.<|control11|><|separator|>
  15. [15]
    Copper-Mediated Living Radical Polymerization (Atom Transfer ...
    Nov 4, 2015 · During the initiation of a normal ATRP process, the starting components, composed of the alkyl halide initiator and activator species, rapidly ...<|control11|><|separator|>
  16. [16]
    Iron Catalysts in Atom Transfer Radical Polymerization - MDPI
    Apr 3, 2020 · In this paper, we review the fundamentals and recent advances of iron-catalyzed ATRP focusing on development of ligands, catalyst design, and techniques used ...Iron Catalysts In Atom... · 2. Ligands And Iron... · 2.1. Halide Salts
  17. [17]
    Metal-free atom transfer radical polymerization with ppm catalyst ...
    Jan 18, 2021 · However, conventional ATRPs rely on transition metal catalysts [i.e., Cu(I), Ru(II)], which will result in transition metal contaminations in ...
  18. [18]
    Evolution and Future Directions of Metal-Free Atom Transfer Radical ...
    Sep 25, 2018 · We highlight key discoveries and recent developments in metal-free ATRP, while providing a perspective for future opportunities in this emerging area.<|control11|><|separator|>
  19. [19]
    Progress in ATRP-derived materials for biomedical applications
    This review summarizes the recent advances in ATRP for the design of functional materials and techniques implemented to advance the biomedical field.1. Introduction · 2. Atrp: Fundamentals And... · 3. Atrp In The Design Of...
  20. [20]
    Living/controlled radical polymerization of methyl methacrylate by ...
    Monodentate phosphines used as ligands include: triphenylphosphine, PPh3 [51,55,59–106], tri(n-butyl)phosphine, PnBu3 [77,83,88,101,105–107], tri(t-butyl) ...
  21. [21]
    ATRP in Protic Media - Matyjaszewski Polymer Group
    Finally, the aqueous ARGET ATRP technique was applied to biological systems by synthesizing a well-defined protein-polymer hybrid by the “grafting from” method.
  22. [22]
    Effect of Ligand and Initiator Structures on the Equilibrium Constants
    The ATRP equilibrium constants obtained vary over 7 orders of magnitude and strongly depend on the ligand and initiator structures.
  23. [23]
    AGET and SARA ATRP of styrene and methyl methacrylate ...
    Two pyridyl-imine ligands, N,N-dimethyl-N′-((pyridin-2-yl)methylene)-ethane-1,2-diamine (paenMe2) and N,N-diethyl-N′-((pyridin-2-yl)methylene)-ethane-1 ...
  24. [24]
    Bidentate guanidine ligands with ethylene spacer in copper ...
    ... ATRP of styrene [19][20][21][22][23][24]. In all ... ATRP studies with the hybrid guanidine ligands ... The sulfonated bipyridine derivative disodium 6.<|separator|>
  25. [25]
    Ligands and characterization for effective bio‐atom‐transfer radical ...
    Sep 30, 2019 · Herein, we describe the synthesis of three new morpholine-based ATRP ligands and their application in polymerizations of oligo(ethylene glycol) ...
  26. [26]
    Current status and outlook for ATRP - ScienceDirect.com
    May 15, 2024 · Cu-based ATRP was first reported in 1995. [259] Carnegie Mellon ... Krzysztof Matyjaszewski: Writing – review & editing, Writing – original draft, ...
  27. [27]
    Development of Environmentally Friendly Atom Transfer Radical ...
    Herein, this review aims to provide a comprehensive understanding of these three types of environmentally friendly ATRP technology and primarily focused on the ...
  28. [28]
  29. [29]
  30. [30]
  31. [31]
  32. [32]
    Determination of Equilibrium Constants for Atom Transfer Radical Polymerization
    ### Summary of K_ATRP Values and Rate Constants from https://pubs.acs.org/doi/10.1021/ja0558591
  33. [33]
    Electrochemical approaches to the determination of rate constants ...
    Dec 20, 2016 · The equilibrium constant can be used to relate the rate of polymerization, Rp, to the concentration of the propagating radicals: R p = k p C M ...
  34. [34]
    Diminishing catalyst concentration in atom transfer radical ... - NIH
    Oct 17, 2006 · The concept of initiators for continuous activator regeneration (ICAR) in atom transfer radical polymerization (ATRP) is introduced, ...
  35. [35]
    Controlled Radical Polymerization: State-of-the-Art in 2014
    May 1, 2015 · In the presence of monomer, MA/DMSO 2/1 (v/v), the value was slightly lower ka1 = 200 M–1 s–1 ( 95). In aqueous media, activation of alkyl ...
  36. [36]
  37. [37]
  38. [38]
    Red-Light-Driven Atom Transfer Radical Polymerization for High ...
    Oct 25, 2023 · (104) The developed dual-catalytic ATRP system exhibited excellent oxygen tolerance, allowing open-air polymerizations with high monomer ...
  39. [39]
    Toward Green Atom Transfer Radical Polymerization: Current Status ...
    Feb 17, 2022 · The classical ATRP method has some limitations including the use of a high concentration of an air-sensitive catalyst, and the consequent ...
  40. [40]
    Atom transfer radical polymerization | Springer Nature Experiments
    ATRP is a synthetic method that harnesses radical polymerization to yield well-defined polymers of controlled architecture from a wide range of monomers.Missing: paper | Show results with:paper<|control11|><|separator|>
  41. [41]
    Recent Developments and Future Challenges in Controlled Radical ...
    Jul 9, 2020 · It should be mentioned that the synthesis of UHMW is dominated by RAFT polymerization with significantly less examples reported on ATRP.
  42. [42]
  43. [43]
  44. [44]
  45. [45]
  46. [46]
    AGET ATRP in the Presence of Air in Miniemulsion and in Bulk - 2006
    Apr 10, 2006 · Ascorbic acid and tin(II) 2-ethylhexanoate are used as reducing agents in miniemulsion and bulk, respectively. An excess of reducing agent ...Missing: reducer | Show results with:reducer
  47. [47]
    A novel immobilized cobalt(II)/copper(II) bimetallic catalyst for atom ...
    A novel immobilized cobalt(II)/copper(II) bimetallic catalyst for atom transfer radical polymerization (ATRP) of methyl methacrylate. Author links open ...Missing: tandem | Show results with:tandem
  48. [48]
    Ultrasonication-Induced Aqueous Atom Transfer Radical ...
    Feb 16, 2018 · A new procedure for ultrasonication-induced atom transfer radical polymerization (sono-ATRP) in aqueous media was developed.
  49. [49]
    Repurposing Biocatalysts to Control Radical Polymerizations
    Sep 4, 2018 · We could show that laccase from Trametes versicolor can overcome this limitation by catalyzing the polymerization of NVIm under ARGET ATRP ...
  50. [50]
    Surface-Initiated Controlled Radical Polymerization: State-of-the-Art ...
    Jan 30, 2017 · As in our previous review, the focus of this contribution will be on the four major surface-initiated controlled radical polymerization (SI-CRP ...
  51. [51]
    Recent Progress on Hyperbranched Polymers Synthesized via ... - NIH
    This short review article summarizes recent reports on using controlled radical polymerization (CRP) of inimers (compounds containing initiating group and vinyl ...
  52. [52]
    ATRP in the design of functional materials for biomedical applications
    Although copper is most commonly used metal catalyst in ATRP, iron compounds, which are generally considered to be less toxic, can also be used [26–29].3.1. Pegylation And... · 4. Drug Delivery Systems · 4.1. Amphiphilic Block...
  53. [53]
    Well-Defined Mesostructured Organic−Inorganic Hybrid Materials ...
    Here, we report a procedure to synthesize well-ordered mesoporous organic−inorganic hybrid materials based on ATRP of methacrylate monomers inside SBA-15 pores ...
  54. [54]
    and pH-responsive star-like polymers synthesized by photoATRP
    Feb 12, 2021 · The thermo and pH-responsive star-like block copolymers are synthesized with controlled molecular weight and low dispersity via photoinduced atom transfer ...
  55. [55]
    Thermally Degradable Poly(n-butyl acrylate) Model Networks ... - NIH
    Feb 12, 2022 · Model poly(n-butyl acrylate) (PBA) networks were prepared by photoinduced atom transfer radical polymerization (photoATRP), followed by curing of polymer stars.
  56. [56]
    Recent Advances in the Application of ATRP in the Synthesis ... - MDPI
    For the purpose of this review, DDSs obtained via ATRP are divided into the four most recently studied types: micelles, polymersomes, and polyplexes formed by ...3.1. 2. Drug Release Induced... · Disulfide-Bearing... · 3.3. Polyplexes
  57. [57]
    Soluble and electrically conductive polyaniline-modified polymers ...
    Aug 8, 2025 · Soluble and electrically conductive polyaniline-modified polymers: Incorporation of biocompatible polymeric chains through ATRP technique: ...
  58. [58]
    Synthesis of sustainable, lightweight and electrically conductive ...
    This work reports on the synthesis of multi-layer graphene oxide functionalized with poly(monomethyl itaconate) using the ARGET-ATRP technique (Fig. 2). This ...Missing: OLEDs | Show results with:OLEDs
  59. [59]
    Specialty Polymers Company ATRP Solutions Picks Up Steam With ...
    Jul 22, 2010 · Early customers are testing a new thickening agent by ATRP ... thickening ingredients currently in use in the personal care and cosmetic industry.
  60. [60]
    Polymer grafted graphene via atom transfer radical polymerization ...
    Oil ... (ATRP): A rheology improver in oil-based drilling fluids ... oil production efficiency in oil fields by means of enhanced oil recovery methods.
  61. [61]
    Photoinduced Iron-Catalyzed ATRP of Renewable Monomers in ...
    Jun 22, 2022 · We demonstrate that renewable monomers can be successfully polymerized into sustainable polymers with controlled molecular weights and narrow molar mass ...
  62. [62]
    Solvent‐Free Chemical Recycling of Polymethacrylates made by ...
    Jul 31, 2023 · A low temperature and solvent-free method to depolymerize polymers made by ATRP and RAFT is reported.Missing: packaging | Show results with:packaging
  63. [63]
    The scale-up of electrochemically mediated atom transfer radical ...
    Oct 1, 2022 · To promote the scale-up of eATRP and reduce the challenges associated to ATRP on an industrial scale, we have embedded an electrochemical O2 ...
  64. [64]
    The Emergence of Reversible–Deactivation Radical Polymerization ...
    Sep 18, 2022 · A suite of photocontrolled reversible–deactivation radical polymerization (photoRDRP) techniques is recently implemented in 3D printing.Introduction · PhotoRDRP-Mediated 3D... · Benefits of Using RDRP for 3D...