Atom transfer radical polymerization

Last updated

Atom transfer radical polymerization (ATRP) is an example of a reversible-deactivation radical polymerization. Like its counterpart, ATRA, or atom transfer radical addition, ATRP is a means of forming a carbon-carbon bond with a transition metal catalyst. Polymerization from this method is called atom transfer radical addition polymerization (ATRAP). As the name implies, the atom transfer step is crucial in the reaction responsible for uniform polymer chain growth. ATRP (or transition metal-mediated living radical polymerization) was independently discovered by Mitsuo Sawamoto [1] and by Krzysztof Matyjaszewski and Jin-Shan Wang in 1995. [2] [3]

Contents

The following scheme presents a typical ATRP reaction:
General ATRP reaction. A. Initiation. B. Equilibrium with dormant species. C. Propagation ATRP general.png
General ATRP reaction. A. Initiation. B. Equilibrium with dormant species. C. Propagation
IUPAC definition for ATRP

Controlled reversible-deactivation radical polymerization in which the deactivation
of the radicals involves reversible atom transfer or reversible group transfer catalyzed usually,
though not exclusively, by transition-metal complexes. [4]

Overview of ATRP

ATRP usually employs a transition metal complex as the catalyst with an alkyl halide as the initiator (R-X). Various transition metal complexes, namely those of Cu, Fe, Ru, Ni, and Os, have been employed as catalysts for ATRP. In an ATRP process, the dormant species is activated by the transition metal complex to generate radicals via one electron transfer process. Simultaneously the transition metal is oxidized to higher oxidation state. This reversible process rapidly establishes an equilibrium that is predominately shifted to the side with very low radical concentrations. The number of polymer chains is determined by the number of initiators. Each growing chain has the same probability to propagate with monomers to form living/dormant polymer chains (R-Pn-X). As a result, polymers with similar molecular weights and narrow molecular weight distribution can be prepared.

ATRP reactions are very robust in that they are tolerant of many functional groups like allyl, amino, epoxy, hydroxy, and vinyl groups present in either the monomer or the initiator. [5] ATRP methods are also advantageous due to the ease of preparation, commercially available and inexpensive catalysts (copper complexes), pyridine-based ligands, and initiators (alkyl halides). [6]

The ATRP with styrene. If all the styrene is reacted (the conversion is 100%) the polymer will have 100 units of styrene built into it. PMDETA stands for N,N,N',N'',N''-pentamethyldiethylenetriamine. ATRP on styrene.png
The ATRP with styrene. If all the styrene is reacted (the conversion is 100%) the polymer will have 100 units of styrene built into it. PMDETA stands for N,N,N′,N′′,N′′-pentamethyldiethylenetriamine.

Components of normal ATRP

There are five important variable components of atom transfer radical polymerizations. They are the monomer, initiator, catalyst, ligand, and solvent. The following section breaks down the contributions of each component to the overall polymerization.

Monomer

Monomers typically used in ATRP are molecules with substituents that can stabilize the propagating radicals; for example, styrenes, (meth)acrylates, (meth)acrylamides, and acrylonitrile. [7] ATRP is successful at leading to polymers of high number average molecular weight and low dispersity when the concentration of the propagating radical balances the rate of radical termination. Yet, the propagating rate is unique to each individual monomer. Therefore, it is important that the other components of the polymerization (initiator, catalyst, ligand, and solvent) are optimized in order for the concentration of the dormant species to be greater than that of the propagating radical while being low enough as to prevent slowing down or halting the reaction. [8] [9]

Initiator

The number of growing polymer chains is determined by the initiator. To ensure a low polydispersity and a controlled polymerization, the rate of initiation must be as fast or preferably faster than the rate of propagation [10] Ideally, all chains will be initiated in a very short period of time and will be propagated at the same rate. Initiators are typically chosen to be alkyl halides whose frameworks are similar to that of the propagating radical. [8] Alkyl halides such as alkyl bromides are more reactive than alkyl chlorides. Both offer good molecular weight control. [8] [9] The shape or structure of the initiator influences polymer architecture. For example, initiators with multiple alkyl halide groups on a single core can lead to a star-like polymer shape. [11] Furthermore, α-functionalized ATRP initiators can be used to synthesize hetero-telechelic polymers with a variety of chain-end groups [12]

Illustration of a star initiator for ATRP. Star ATRP initiator.png
Illustration of a star initiator for ATRP.

Catalyst

The catalyst is the most important component of ATRP because it determines the equilibrium constant between the active and dormant species. This equilibrium determines the polymerization rate. An equilibrium constant that is too small may inhibit or slow the polymerization while an equilibrium constant that is too large leads to a wide distribution of chain lengths. [9]

There are several requirements for the metal catalyst:

  1. There needs to be two accessible oxidation states that are differentiated by one electron
  2. The metal center needs to have reasonable affinity for halogens
  3. The coordination sphere of the metal needs to be expandable when it is oxidized as to accommodate the halogen
  4. The transition metal catalyst should not lead to significant side reactions, such as irreversible coupling with the propagating radicals and catalytic radical termination

The most studied catalysts are those that include copper, which has shown the most versatility with successful polymerizations for a wide selection of monomers.

Ligand

One of the most important aspects in an ATRP reaction is the choice of ligand which is used in combination with the traditionally copper halide catalyst to form the catalyst complex. The main function of the ligand is to solubilize the copper halide in whichever solvent is chosen and to adjust the redox potential of the copper. [13] This changes the activity and dynamics of the halogen exchange reaction and subsequent activation and deactivation of the polymer chains during polymerization, therefore greatly affecting the kinetics of the reaction and the degree of control over the polymerization. Different ligands should be chosen based on the activity of the monomer and the choice of metal for the catalyst. As copper halides are primarily used as the catalyst, amine based ligands are most commonly chosen. Ligands with higher activities are being investigated as ways to potentially decrease the concentration of catalyst in the reaction since a more active catalyst complex would lead to a higher concentration of deactivator in the reaction. However, a too active catalyst can lead to a loss of control and increase the polydispersity of the resulting polymer.

Solvents

Toluene, 1,4-dioxane, xylene, anisole, DMF, DMSO, water, methanol, acetonitrile, or even the monomer itself (described as a bulk polymerization) are commonly used.

Kinetics of normal ATRP

ATRP equilibrium constant

The radical concentration in normal ATRP can be calculated via the following equation:

It is important to know the KATRP value to adjust the radical concentration. The KATRP value depends on the homo-cleavage energy of the alkyl halide and the redox potential of the Cu catalyst with different ligands. Given two alkyl halides (R1-X and R2-X) and two ligands (L1 and L2), there will be four combinations between different alkyl halides and ligands. Let KijATRP refer to the KATRP value for Ri-X and Lj. If we know three of these four combinations, the fourth one can be calculated as:

The KATRP values for different alkyl halides and different Cu catalysts can be found in literature. [14]

Solvents have significant effects on the KATRP values. The KATRP value increases dramatically with the polarity of the solvent for the same alkyl halide and the same Cu catalyst. [15] The polymerization must take place in solvent/monomer mixture, which changes to solvent/monomer/polymer mixture gradually. The KATRP values could change 10000 times by switching the reaction medium from pure methyl acrylate to pure dimethyl sulfoxide. [16]

Activation and deactivation rate coefficients

Deactivation rate coefficient, kd, values must be sufficiently large to obtain low dispersity. The direct measurement of kd is difficult though not impossible. In most cases, kd may be calculated from known KATRP and ka. [14] [17] [18] Cu complexes providing very low kd values are not recommended for use in ATRP reactions.

Retention of chain end functionality



Halogen Conservation in Atom Transfer Radical Polymerization

High level retention of chain end functionality is typically desired. However, the determination of the loss of chain end functionality based on 1H NMR and mass spectroscopy methods cannot provide precise values. As a result, it is difficult to identify the contributions of different chain breaking reactions in ATRP. One simple rule in ATRP comprises the principle of halogen conservation. [19] Halogen conservation means the total amount of halogen in the reaction systems must remain as a constant. From this rule, the level of retention of chain end functionality can be precisely determined in many cases. The precise determination of the loss of chain end functionality enabled further investigation of the chain breaking reactions in ATRP. [20]

Advantages and disadvantages of ATRP

Advantages

ATRP enables the polymerization of a wide variety of monomers with different chemical functionalities, proving to be more tolerant of these functionalities than ionic polymerizations. It provides increased control of molecular weight, molecular architecture and polymer composition while maintaining a low polydispersity (1.05-1.2). The halogen remaining at the end of the polymer chain after polymerization allows for facile post-polymerization chain-end modification into different reactive functional groups. The use of multi-functional initiators facilitates the synthesis of lower-arm star polymers and telechelic polymers. External visible light stimulation ATRP has a high responding speed and excellent functional group tolerance. [21]

Disadvantages

The most significant drawback of ATRP is the high concentrations of catalyst required for the reaction. This catalyst standardly consists of a copper halide and an amine-based ligand. The removal of the copper from the polymer after polymerization is often tedious and expensive, limiting ATRP's use in the commercial sector. [22] However, researchers are currently developing methods which would limit the necessity of the catalyst concentration to ppm. ATRP is also a traditionally air-sensitive reaction normally requiring freeze-pump thaw cycles. However, techniques such as Activator Generated by Electron Transfer (AGET) ATRP provide potential alternatives which are not air-sensitive. [23] A final disadvantage is the difficulty of conducting ATRP in aqueous media.

Different ATRP methods

Activator regeneration ATRP methods

In a normal ATRP, the concentration of radicals is determined by the KATRP value, concentration of dormant species, and the [CuI]/[CuII] ratio. In principle, the total amount of Cu catalyst should not influence polymerization kinetics. However, the loss of chain end functionality slowly but irreversibly converts CuI to CuII. Thus initial [CuI]/[I] ratios are typically 0.1 to 1. When very low concentrations of catalysts are used, usually at the ppm level, activator regeneration processes are generally required to compensate the loss of CEF and regenerate a sufficient amount of CuI to continue the polymerization. Several activator regeneration ATRP methods were developed, namely ICAR ATRP, ARGET ATRP, SARA ATRP, eATRP, and photoinduced ATRP. The activator regeneration process is introduced to compensate the loss of chain end functionality, thus the cumulative amount of activator regeneration should roughly equal the total amount of the loss of chain end functionality.

Activator regeneration Atom Transfer Radical Polymerization Activator Regeneration Atom Transfer Radical Polymerization.svg
Activator regeneration Atom Transfer Radical Polymerization

ICAR ATRP

Initiators for continuous activator regeneration (ICAR) is a technique that uses conventional radical initiators to continuously regenerate the activator, lowering its required concentration from thousands of ppm to <100 ppm; making it an industrially relevant technique.

ARGET ATRP

Activators regenerated by electron transfer (ARGET) employs non-radical forming reducing agents for regeneration of CuI. A good reducing agent (e.g. hydrazine, phenols, sugars, ascorbic acid) should only react with CuII and not with radicals or other reagents in the reaction mixture.

SARA ATRP

A typical SARA ATRP employs Cu0 as both supplemental activator and reducing agent (SARA). Cu0 can activate alkyl halide directly but slowly. Cu0 can also reduce CuII to CuI. Both processes help to regenerate CuI activator. Other zerovalent metals, such as Mg, Zn, and Fe, have also been employed for Cu-based SARA ATRP.

eATRP

In eATRP the activator CuI is regenerated via electrochemical process. The development of eATRP enables precise control of the reduction process and external regulation of the polymerization. In an eATRP process, the redox reaction involves two electrodes. The CuII species is reduced to CuI at the cathode. The anode compartment is typically separated from the polymerization environment by a glass frit and a conductive gel. Alternatively, a sacrificial aluminum counter electrode can be used, which is directly immersed in the reaction mixture.

Photoinduced ATRP

The direct photo reduction of transition metal catalysts in ATRP and/or photo assistant activation of alkyl halide is particularly interesting because such a procedure will allow performing of ATRP with ppm level of catalysts without any other additives.

Other ATRP methods

Reverse ATRP

In reverse ATRP, the catalyst is added in its higher oxidation state. Chains are activated by conventional radical initiators (e.g. AIBN) and deactivated by the transition metal. The source of transferable halogen is the copper salt, so this must be present in concentrations comparable to the transition metal.

SR&NI ATRP

A mixture of radical initiator and active (lower oxidation state) catalyst allows for the creation of block copolymers (contaminated with homopolymer) which is impossible using standard reverse ATRP. This is called SR&NI (simultaneous reverse and normal initiation ATRP).

AGET ATRP

Activators generated by electron transfer uses a reducing agent unable to initiate new chains (instead of organic radicals) as regenerator for the low-valent metal. Examples are metallic copper, tin(II), ascorbic acid, or triethylamine. It allows for lower concentrations of transition metals, and may also be possible in aqueous or dispersed media.

Hybrid and bimetallic systems

This technique uses a variety of different metals/oxidation states, possibly on solid supports, to act as activators/deactivators, possibly with reduced toxicity or sensitivity. [24] [25] Iron salts can, for example, efficiently activate alkyl halides but requires an efficient Cu(II) deactivator which can be present in much lower concentrations (3–5 mol%)

Metal-free ATRP

Trace metal catalyst remaining in the final product has limited the application of ATRP in biomedical and electronic fields. In 2014, Craig Hawker and coworkers developed a new catalysis system involving photoredox reaction of 10-phenothiazine. The metal-free ATRP has been demonstrated to be capable of controlled polymerization of methacrylates. [26] This technique was later expanded to polymerization of acrylonitrile by Matyjaszewski et al. [27]

Mechano/sono-ATRP

Mechano/sono-ATRP uses mechanical forces, typically ultrasonic agitation, as an external stimulus to induce the (re)generation of activators in ATRP. Esser-Kahn, et al. demonstrated the first example of mechanoATRP using the piezoelectricity of barium titanate to reduce Cu(II) species. [28] Matyjaszewski, et al. later improved the technique by using nanometer-sized and/or surface-functionalized barium titanate or zinc oxide particles, achieving superior rate and control of polymerization, as well as temporal control, with ppm-level of copper catalysts. [29] [30] In addition to peizoelectric particles, water and carbonates were found to mediate mechano/sono-ATRP. Mechochemically homolyzed water molecules undergoes radical addition to monomers, which in turn reduces Cu(II) species. [31] Mechanically unstable Cu(II)-carbonate complexes formed in the presence to insoluble carbonates, which oxidizes dimethylsulfoxide, the solvent molecules, to generate Cu(I) species and carbon dioxide. [32]

Biocatalytic ATRP

Metalloenzymes have been used for the first time as ATRP catalysts, in parallel and independently, by the research teams of Fabio Di Lena [33] and Nico Bruns. [34] This pioneering work has paved the way to the emerging field of biocatalytic reversible-deactivation radical polymerization. [35] [36]

Polymers synthesized through ATRP

See also

Related Research Articles

In polymer chemistry, living polymerization is a form of chain growth polymerization where the ability of a growing polymer chain to terminate has been removed. This can be accomplished in a variety of ways. Chain termination and chain transfer reactions are absent and the rate of chain initiation is also much larger than the rate of chain propagation. The result is that the polymer chains grow at a more constant rate than seen in traditional chain polymerization and their lengths remain very similar. Living polymerization is a popular method for synthesizing block copolymers since the polymer can be synthesized in stages, each stage containing a different monomer. Additional advantages are predetermined molar mass and control over end-groups.

In polymer chemistry, ring-opening polymerization (ROP) is a form of chain-growth polymerization in which the terminus of a polymer chain attacks cyclic monomers to form a longer polymer. The reactive center can be radical, anionic or cationic. Some cyclic monomers such as norbornene or cyclooctadiene can be polymerized to high molecular weight polymers by using metal catalysts. ROP is a versatile method for the synthesis of biopolymers.

A substitution reaction is a chemical reaction during which one functional group in a chemical compound is replaced by another functional group. Substitution reactions are of prime importance in organic chemistry. Substitution reactions in organic chemistry are classified either as electrophilic or nucleophilic depending upon the reagent involved, whether a reactive intermediate involved in the reaction is a carbocation, a carbanion or a free radical, and whether the substrate is aliphatic or aromatic. Detailed understanding of a reaction type helps to predict the product outcome in a reaction. It also is helpful for optimizing a reaction with regard to variables such as temperature and choice of solvent.

<span class="mw-page-title-main">Triflate</span> Chemical group (–OSO2CF3) or anion (charge –1)

In organic chemistry, triflate, is a functional group with the formula R−OSO2CF3 and structure R−O−S(=O)2−CF3. The triflate group is often represented by −OTf, as opposed to −Tf, which is the triflyl group, R−SO2CF3. For example, n-butyl triflate can be written as CH3CH2CH2CH2OTf.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound (also known as organostannanes). A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

<span class="mw-page-title-main">Radical polymerization</span> Polymerization process involving free radicals as repeating units

In polymer chemistry, free-radical polymerization (FRP) is a method of polymerization by which a polymer forms by the successive addition of free-radical building blocks. Free radicals can be formed by a number of different mechanisms, usually involving separate initiator molecules. Following its generation, the initiating free radical adds (nonradical) monomer units, thereby growing the polymer chain.

<span class="mw-page-title-main">Reversible addition−fragmentation chain-transfer polymerization</span>

Reversible addition−fragmentation chain-transfer or RAFT polymerization is one of several kinds of reversible-deactivation radical polymerization. It makes use of a chain-transfer agent (CTA) in the form of a thiocarbonylthio compound to afford control over the generated molecular weight and polydispersity during a free-radical polymerization. Discovered at the Commonwealth Scientific and Industrial Research Organisation (CSIRO) of Australia in 1998, RAFT polymerization is one of several living or controlled radical polymerization techniques, others being atom transfer radical polymerization (ATRP) and nitroxide-mediated polymerization (NMP), etc. RAFT polymerization uses thiocarbonylthio compounds, such as dithioesters, thiocarbamates, and xanthates, to mediate the polymerization via a reversible chain-transfer process. As with other controlled radical polymerization techniques, RAFT polymerizations can be performed under conditions that favor low dispersity and a pre-chosen molecular weight. RAFT polymerization can be used to design polymers of complex architectures, such as linear block copolymers, comb-like, star, brush polymers, dendrimers and cross-linked networks.

<span class="mw-page-title-main">Photopolymer</span>

A photopolymer or light-activated resin is a polymer that changes its properties when exposed to light, often in the ultraviolet or visible region of the electromagnetic spectrum. These changes are often manifested structurally, for example hardening of the material occurs as a result of cross-linking when exposed to light. An example is shown below depicting a mixture of monomers, oligomers, and photoinitiators that conform into a hardened polymeric material through a process called curing.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

A Grignard reagent or Grignard compound is a chemical compound with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

<span class="mw-page-title-main">Organocopper chemistry</span> Compound with carbon to copper bonds

Organocopper chemistry is the study of the physical properties, reactions, and synthesis of organocopper compounds, which are organometallic compounds containing a carbon to copper chemical bond. They are reagents in organic chemistry.

In organic chemistry, the Kumada coupling is a type of cross coupling reaction, useful for generating carbon–carbon bonds by the reaction of a Grignard reagent and an organic halide. The procedure uses transition metal catalysts, typically nickel or palladium, to couple a combination of two alkyl, aryl or vinyl groups. The groups of Robert Corriu and Makoto Kumada reported the reaction independently in 1972.

Catalytic chain transfer (CCT) is a process that can be incorporated into radical polymerization to obtain greater control over the resulting products.

Cobalt based catalysts, when used in radical polymerization, have several main advantages especially in slowing down the reaction rate, allowing for the synthesis of polymers with peculiar properties. As starting the reaction does need a real radical initiator, the cobalt species is not the only used catalyst, it is a mediator. For this reason this type of reaction is referred to as cobalt mediated radical polymerization.

<span class="mw-page-title-main">Living free-radical polymerization</span>

Living free radical polymerization is a type of living polymerization where the active polymer chain end is a free radical. Several methods exist. IUPAC recommends to use the term "reversible-deactivation radical polymerization" instead of "living free radical polymerization", though the two terms are not synonymous.

In polymer chemistry, cationic polymerization is a type of chain growth polymerization in which a cationic initiator transfers charge to a monomer, which then becomes reactive. This reactive monomer goes on to react similarly with other monomers to form a polymer. The types of monomers necessary for cationic polymerization are limited to alkenes with electron-donating substituents and heterocycles. Similar to anionic polymerization reactions, cationic polymerization reactions are very sensitive to the type of solvent used. Specifically, the ability of a solvent to form free ions will dictate the reactivity of the propagating cationic chain. Cationic polymerization is used in the production of polyisobutylene and poly(N-vinylcarbazole) (PVK).

<span class="mw-page-title-main">Poly(methacrylic acid)</span> Chemical compound

Poly(methacrylic acid) (PMAA) is a polymer made from methacrylic acid (preferred IUPAC name, 2-methylprop-2-enoic acid), which is a carboxylic acid. It is often available as its sodium salt, poly(methacrylic acid) sodium salt. The monomer is a viscous liquid with a pungent odour. The first polymeric form of methacrylic acid was described in 1880 by Engelhorn and Fittig. The use of high purity monomers is required for proper polymerization conditions and therefore it is necessary to remove any inhibitors by extraction (phenolic inhibitors) or via distillation. To prevent inhibition by dissolved oxygen, monomers should be carefully degassed prior to the start of the polymerization.

<span class="mw-page-title-main">Reversible-deactivation radical polymerization</span> Type of chain polymerization

In polymer chemistry, reversible-deactivation radical polymerizations (RDRPs) are members of the class of reversible-deactivation polymerizations which exhibit much of the character of living polymerizations, but cannot be categorized as such as they are not without chain transfer or chain termination reactions. Several different names have been used in literature, which are:

<span class="mw-page-title-main">Graft polymer</span> Polymer with a backbone of one composite and random branches of another composite

In polymer chemistry, graft polymers are segmented copolymers with a linear backbone of one composite and randomly distributed branches of another composite. The picture labeled "graft polymer" shows how grafted chains of species B are covalently bonded to polymer species A. Although the side chains are structurally distinct from the main chain, the individual grafted chains may be homopolymers or copolymers. Graft polymers have been synthesized for many decades and are especially used as impact resistant materials, thermoplastic elastomers, compatibilizers, or emulsifiers for the preparation of stable blends or alloys. One of the better-known examples of a graft polymer is a component used in high impact polystyrene, consisting of a polystyrene backbone with polybutadiene grafted chains.

<span class="mw-page-title-main">Knotted polymers</span>

Single Chain Cyclized/Knotted Polymers are a new class of polymer architecture with a general structure consisting of multiple intramolecular cyclization units within a single polymer chain. Such a structure was synthesized via the controlled polymerization of multivinyl monomers, which was first reported in Dr. Wenxin Wang's research lab. These multiple intramolecular cyclized/knotted units mimic the characteristics of complex knots found in proteins and DNA which provide some elasticity to these structures. Of note, 85% of elasticity in natural rubber is due to knot-like structures within its molecular chain.
An intramolecular cyclization reaction is where the growing polymer chain reacts with a vinyl functional group on its own chain, rather than with another growing chain in the reaction system. In this way the growing polymer chain covalently links to itself in a fashion similar to that of a knot in a piece of string. As such, single chain cyclized/knotted polymers consist of many of these links, as opposed to other polymer architectures including branched and crosslinked polymers that are formed by two or more polymer chains in combination.

Copper-based reversible-deactivation radical polymerization(Cu-based RDRP) is a member of the class of reversible-deactivation radical polymerization. In this system, various copper species are employed as the transition-metal catalyst for reversible activation/deactivation of the propagating chains responsible for uniform polymer chain growth.

References

  1. Kato, M; Kamigaito, M; Sawamoto, M; Higashimura, T (1995). "Polymerization of Methyl Methacrylate with the Carbon Tetrachloride / Dichlorotris-(triphenylphosphine)ruthenium(II) / Methylaluminum Bis(2,6-di-tert-butylphenoxide) Initiating System: Possibility of Living Radical Polymerization". Macromolecules . 28 (5): 1721–1723. Bibcode:1995MaMol..28.1721K. doi:10.1021/ma00109a056.
  2. Wang, J-S; Matyjaszewski, K (1995). "Controlled/"living" radical polymerization. Atom transfer radical polymerization in the presence of transition-metal complexes". J. Am. Chem. Soc. 117 (20): 5614–5615. doi:10.1021/ja00125a035.
  3. "The 2011 Wolf Prize in Chemistry". Wolf Fund. Retrieved 21 February 2011.
  4. Jenkins, Aubrey D.; Jones, Richard G.; Moad, Graeme (2010). "Terminology for reversible-deactivation radical polymerization previously called "controlled" radical or "living" radical polymerization (IUPAC Recommendations 2010)" (PDF). Pure and Applied Chemistry . 82 (2): 483–491. doi:10.1351/PAC-REP-08-04-03.
  5. Cowie, J. M. G.; Arrighi, V. In Polymers: Chemistry and Physics of Modern Materials; CRC Press Taylor and Francis Group: Boca Raton, Fl, 2008; 3rd Ed., pp. 82–84 ISBN   0849398134
  6. Matyjaszewski, K. "Fundamentals of ATRP Research". Archived from the original on February 22, 2009. Retrieved Jan 7, 2009.
  7. Patten, T. E; Matyjaszewski, K (1998). "Atom Transfer Radical Polymerization and the Synthesis of Polymeric Materials". Adv. Mater. 10 (12): 901–915. doi:10.1002/(sici)1521-4095(199808)10:12<901::aid-adma901>3.0.co;2-b.
  8. 1 2 3 Odian, G. In Radical Chain Polymerization; Principles of Polymerization; Wiley-Interscience: Staten Island, New York, 2004; Vol. , pp 316–321.
  9. 1 2 3 Matyjaszewski, Krzysztof; Xia, Jianhui (2001). "Atom Transfer Radical Polymerization". Chem. Rev. 101 (9): 2921–90. doi:10.1021/cr940534g. ISSN   0009-2665. PMID   11749397.
  10. <!- -Not stated- ->. "Initiators". the Matyjaszewski Polymer Group. Carnegie Mellon University. Archived from the original on July 19, 2011. Retrieved November 30, 2018.
  11. Jakubowski, Wojciech. "Complete Tools for the Synthesis of Well-Defined Functionalized Polymers via ATRP". Sigma-Aldrich. Retrieved 21 July 2010.
  12. <!- -Not stated- ->. "Use of Functional ATRP Initiators". the Matyjaszewski Polymer Group. Carnegie Mellon University. Archived from the original on February 13, 2011. Retrieved November 30, 2018.
  13. <!- -Not stated- ->. "Structural Characterization of an ATRP Catalyst Complex". the Matyjaszewski Polymer Group. Carnegie Mellon University. Retrieved November 30, 2018.
  14. 1 2 Tang, W; Kwak, Y; Braunecker, W; Tsarevsky, N V; Coote, M L; Matyjaszewski, K (2008). "Understanding Atom Transfer Radical Polymerization: Effect of Ligand and Initiator Structures on the Equilibrium Constants". J. Am. Chem. Soc. 130 (32): 10702–10713. doi:10.1021/ja802290a. PMID   18642811.
  15. Braunecker, W; Tsarevsky, N V; Gennaro, A; Matyjaszewski, K (2009). "Thermodynamic Components of the Atom Transfer Radical Polymerization Equilibrium: Quantifying Solvent Effects". Macromolecules . 42 (17): 6348–6360. Bibcode:2009MaMol..42.6348B. doi:10.1021/ma901094s.
  16. Wang, Y; Kwak, Y; Buback, J; Buback, M; Matyjaszewski, K (2012). "Determination of ATRP Equilibrium Constants under Polymerization Conditions". ACS Macro Lett. 1 (12): 1367–1370. doi:10.1021/mz3005378.
  17. Tang, W; Matyjaszewski, K (2007). "Effects of Initiator Structure on Activation Rate Constants in ATRP". Macromolecules . 40 (6): 1858–1863. Bibcode:2007MaMol..40.1858T. doi:10.1021/ma062897b.
  18. Tang, W; Matyjaszewski, K (2006). "Effect of Ligand Structure on Activation Rate Constants in ATRP". Macromolecules . 39 (15): 4953–4959. Bibcode:2006MaMol..39.4953T. doi:10.1021/ma0609634.
  19. Wang, Y; Zhong, M; Zhang, Y; Magenau, A J D; Matyjaszewski, K (2012). "Halogen Conservation in Atom Transfer Radical Polymerization". Macromolecules . 45 (21): 8929–8932. Bibcode:2012MaMol..45.8929W. doi:10.1021/ma3018958.
  20. Wang, Y; Soerensen, N; Zhong, M; Schroeder, H; Buback, M; Matyjaszewski, K (2013). "Improving the "Livingness" of ATRP by Reducing Cu Catalyst Concentration". Macromolecules . 46 (3): 689–691. Bibcode:2013MaMol..46..683W. doi:10.1021/ma3024393.
  21. "Atom-Transfer Radical-Polymerization (ATRP) – Artificial Intelligence for Chemistry" . Retrieved 2019-11-19.
  22. Borman, Stu (October 30, 2006). "Polymers with Safe Amounts of Copper". Chemical & Engineering News. 84 (43): 40–41. doi:10.1021/cen-v084n044.p040 . Retrieved November 30, 2018.
  23. Siegwart, Daniel; Kwan Oh, Jung; Matyjaszewski, Krzysztof (January 1, 2012). "ATRP in the design of functional materials for biomedical applications". Progress in Polymer Science. 37 (1): 18–37. doi:10.1016/j.progpolymsci.2011.08.001. PMC   3604987 . PMID   23525884.
  24. Xiong, De'an; He, Zhenping (15 January 2010). "Modulating the catalytic activity of Au/micelles by tunable hydrophilic channels". Journal of Colloid and Interface Science . 341 (2): 273–279. Bibcode:2010JCIS..341..273X. doi:10.1016/j.jcis.2009.09.045. PMID   19854448.
  25. Chen, Xi; He, Zhenping; et al. (5 August 2008). "Core-shell-corona Au-micelle composites with a tunable smart hybrid shell". Langmuir. 24 (15): 8198–8204. doi:10.1021/la800244g. PMID   18576675.
  26. Treat, Nicolas; Sprafke, Hazel; Kramer, John; Clark, Paul; Barton, Bryan; Read de Alaniz, Javier; Fors, Brett; Hawker, Craig (2014). "Metal-Free Atom Transfer Radical Polymerization". Journal of the American Chemical Society. 136 (45): 16096–16101. doi:10.1021/ja510389m. PMID   25360628.
  27. Pan, Xiangcheng; Lamson, Melissa; Yan, Jiajun; Matyjaszewski, Krzysztof (17 February 2015). "Photoinduced Metal-Free Atom Transfer Radical Polymerization of Acrylonitrile". ACS Macro Letters. 4 (2): 192–196. doi:10.1021/mz500834g.
  28. Mohapatra, Hemakesh; Kleiman, Maya; Esser-Kahn, Aaron Palmer (24 October 2016). "Mechanically controlled radical polymerization initiated by ultrasound". Nature Chemistry. 9 (2): 135–139. doi:10.1038/nchem.2633.
  29. Wang, Zhenhua; Pan, Xiangcheng; Yan, Jiajun; Dadashi-Silab, Sajjad; Xie, Guojun; Zhang, Jianan; Wang, Zhanhua; Xia, Hesheng; Matyjaszewski, Krzysztof (28 April 2017). "Temporal Control in Mechanically Controlled Atom Transfer Radical Polymerization Using Low ppm of Cu Catalyst". ACS Macro Letters. 6 (5): 546–549. doi:10.1021/acsmacrolett.7b00152.
  30. Wang, Zhenhua; Pan, Xiangcheng; Li, Lingchun; Fantin, Marco; Yan, Jiajun; Wang, Zongyu; Wang, Zhanhua; Xia, Hesheng; Matyjaszewski, Krzysztof (4 October 2017). "Enhancing Mechanically Induced ATRP by Promoting Interfacial Electron Transfer from Piezoelectric Nanoparticles to Cu Catalysts". Macromolecules. 50 (20): 7940–7948. Bibcode:2017MaMol..50.7940W. doi:10.1021/acs.macromol.7b01597.
  31. Wang, Zhenhua; Wang, Zhanhua; Pan, Xiangcheng; Fu, Liye; Lathwal, Sushil; Olszewski, Mateusz; Yan, Jiajun; Enciso, Alan E.; Wang, Zongyu; Xia, Hesheng; Matyjaszewski, Krzysztof (20 March 2018). "Ultrasonication-Induced Aqueous Atom Transfer Radical Polymerization". ACS Macro Letters. 7 (3): 275–280. doi: 10.1021/acsmacrolett.8b00027 . ISSN   2161-1653.
  32. Wang, Zhenhua; Lorandi, Francesca; Fantin, Marco; Wang, Zongyu; Yan, Jiajun; Wang, Zhanhua; Xia, Hesheng; Matyjaszewski, Krzysztof (22 January 2019). "Atom Transfer Radical Polymerization Enabled by Sonochemically Labile Cu-carbonate Species". ACS Macro Letters. 8 (2): 161–165. doi:10.1021/acsmacrolett.9b00029.
  33. Ng, Yeap-Hung; Lena, Fabio di; Chai, Christina L. L. (2011-05-24). "PolyPEGA with predetermined molecular weights from enzyme-mediated radical polymerization in water". Chemical Communications. 47 (22): 6464–6466. doi:10.1039/C1CC10989H. ISSN   1364-548X.
  34. Sigg, Severin J.; Seidi, Farzad; Renggli, Kasper; Silva, Tilana B.; Kali, Gergely; Bruns, Nico (2011-11-01). "Horseradish Peroxidase as a Catalyst for Atom Transfer Radical Polymerization: Horseradish Peroxidase as a Catalyst for Atom Transfer …". Macromolecular Rapid Communications. 32 (21): 1710–1715. doi:10.1002/marc.201100349.
  35. Enzyme Catalysis for Reversible Deactivation Radical Polymerization. Ruoyu Li, Weina Kong, Zesheng An. Angew. Chem. Int. Ed. 2022, 61, e202202033 https://doi.org/10.1002/ange.202202033
  36. Li, Ruoyu; Kong, Weina; An, Zesheng (2023-02-14). "Controlling Radical Polymerization with Biocatalysts". Macromolecules. 56 (3): 751–761. doi:10.1021/acs.macromol.2c02307. ISSN   0024-9297.