Next Article in Journal
Isolation and Rheological Characterization of Cellulose Nanofibrils (CNFs) from Coir Fibers in Comparison to Wood and Cotton
Next Article in Special Issue
Study of theThermo-/pH-Sensitivity of Stereo-Controlled Poly(N-isopropylacrylamide-co-IAM) Copolymers via RAFT Polymerization
Previous Article in Journal
Novel Dual-Component Microencapsulated Hydrophobic Amine and Microencapsulated Isocyanate Used for Self-Healing Anti-Corrosion Coating
Previous Article in Special Issue
RAFT Polymerization of Tert-Butyldimethylsilyl Methacrylate: Kinetic Study and Determination of Rate Coefficients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in RAFT Polymerization: Novel Initiation Mechanisms and Optoelectronic Applications

1
Key Laboratory for Advanced Materials and Shanghai Key Laboratory of Functional Materials Chemistry, Institute of Applied Chemistry, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China
2
The State Key Laboratory of Metal Matrix Composites & Shanghai Key Laboratory of Electrical Insulation and Thermal Ageing, School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Dongchuan Road 800, Shanghai 200240, China
3
Center for Advancing Electronics Dresden (CFAED) & Department of Chemistry and Food Chemistry, Technische Universität Dresden, 01062 Dresden, Germany
*
Authors to whom correspondence should be addressed.
Polymers 2018, 10(3), 318; https://doi.org/10.3390/polym10030318
Submission received: 24 February 2018 / Revised: 9 March 2018 / Accepted: 12 March 2018 / Published: 14 March 2018
(This article belongs to the Special Issue RAFT Living Radical Polymerization and Self-Assembly)

Abstract

:
Reversible addition-fragmentation chain transfer (RAFT) is considered to be one of most famous reversible deactivation radical polymerization protocols. Benefiting from its living or controlled polymerization process, complex polymeric architectures with controlled molecular weight, low dispersity, as well as various functionality have been constructed, which could be applied in wide fields, including materials, biology, and electrology. Under the continuous research improvement, main achievements have focused on the development of new RAFT techniques, containing fancy initiation methods (e.g., photo, metal, enzyme, redox and acid), sulfur-free RAFT system and their applications in many fields. This review summarizes the current advances in major bright spot of novel RAFT techniques as well as their potential applications in the optoelectronic field, especially in the past a few years.

Graphical Abstract

1. Introduction

Macromolecules with high molecular weight have many advantages in chemical and physical properties over traditional small molecular materials. A general method for preparation of macromolecules is to convert small molecules into high molecular weight polymers by simple organic reactions, such as coupling reaction polymerization [1,2], condensation polymerization [3,4,5], and radical polymerization [6]. Among them, radical polymerization is an internationally recognized polymerization technique, which plays an important role in practical researches and applications [7]. Wide reports have demonstrated a large number of advantages in radical polymerization, like compatibility with a wide range of (meth)acrylates, vinyl acetate, acrylonitrile, styrenes, and N-vinylpyrrolidone monomers, facile reaction conditions in bulk, solution and emulsion, and tolerance of many functionalities, including electron-donating groups and electro-drawing [8,9,10,11]. However, a free radical polymerization is repeatable, but thus not allowing microstructural control of high livingness and chain length. Thus, the development of new polymerization technology for controlled molecular weight with complex architecture and novel functionalities is highly desirable.
In 1956, Szwarc firstly reported the living polymerization and described the controlled-polymerization of diene and vinyl monomers [12]. This discovery promotes the development of controlled cationic [13,14], anionic [15,16], as well as ring-opening polymerization [17]. Unfortunately, these polymerization methods often required harsh polymerization conditions, e.g., anhydrous and anaerobic environment [18,19]. Close to radical polymerization, reversible deactivation radical polymerizations (RDRP), such as nitroxide-mediated polymerization (NMP) [20,21], metal-mediated living radical polymerization (LRP) [6], atom transfer radical polymerization (ATRP) [22], and reversible addition-fragmentation chain transfer (RAFT) polymerization [23,24], have drawn much attention due to their facile and simple operating mode. Especially, RAFT polymerization reported firstly by CSIRO in 1998 [23] has now emerged as the most widely used polymerization technique. Due to its living and controllable in the whole polymerization process, RAFT polymerization has been considered as a precious pearl in the fabrication of multiblock copolymer and complex polymeric architectures. The development of RAFT polymerization has gone through two main stages: the proposition and illumination of the mechanism; then, the fabrication of various complex polymeric architectures and functional polymeric materials through the RAFT process for practical applications in different fields. To date, a large number of new RAFT polymerization systems have been reported to improve the polymerization performance of conventional RAFT technique, receiving much attention to scientists for the construction of functional polymers in wide applications.
To date, over hundreds of reviews have been published involving in RAFT polymerization, covering a tremendous amount of polymer synthesis and applications, in terms of architectures (block, star, hyperbranched, cross-linked) [25,26,27,28,29], surface-grafted (inorganic nanocomposites, 2D graphene) [30,31,32] and functionalities (stimuli-responsive, biodegradable) [33,34,35,36]. In these studies, optoelectronic functional materials by RAFT polymerization have been widely investigated for their wide application prospect. The advantages of RAFT polymerization permit the accurate control of optoelectronic polymers in molecular weight, molecular weight distributions and architecture. It will bring great benefits for the improvement of optoelectronic devices performance. However, few reviews of RAFT polymerization summed up its related application in the field of optoelectronics. This review mainly consists of two parts: the first part particularly highlights the latest research of novel RAFT technology, including various initiation methods, e.g., photo, metal, enzyme, redox and acid, and sulfur-free RAFT systems. The optical and electronic applications of functional polymers by RAFT technology are also summarized in second part.

2. General Mechanism of RAFT Process

The currently approved mechanism of the RAFT process is depicted in Scheme 1. Initiation, chain propagation and chain termination in RAFT polymerization are somewhat alike to those in conventional radical polymerization. However, the most basic and important step for RAFT process is reversible addition-fragmentation equilibrium (step 2 and step 4). In a typical RAFT process, radical (I) from initiator reacts with monomers (M) to give a propagating radical (Pn) (step 1). Subsequently, the propagating radical (Pn) adds to the C=S group originated from RAFT agent (1) (chain transfer agent, CTA) and an intermediate radical (2) will be generated. Interestingly, the appearance of intermediate radical (2) will restrict the irreversible termination reaction between the propagating, and intermediate radical will also give two types of fragmentation benefited from the activity of C=S group. It is possible to return to their initial state to get a propagating radical (Pn) and RAFT agent, while the intermediate radical (2) may fragment into a new radical (R) and a dormant chain (polymeric thiocarbonylthio compound) (3). Then, the new propagating radical (Pm) will be formed after the reinitation monomer polymerization by the new formative radical (R) (step 3). In this case, a rapid equilibrium will be formed between the dormant species (3) and active propagating radicals (Pn and Pm) (step 2 and step 4), which can enable all of the chains to have equal chance to grow, leading to a similar degree of polymerization. The termination reaction often is inevitable in a radical polymerization system (step 5). For RAFT polymerization, the termination reaction could be minimized, due to the existence of thiocarbonylthio end-group in resulting polymeric chains, which also enables participating in RAFT polymerization as RAFT agent. Using as macroinitiator, other monomers can be polymerized to fabricate multiblock polymers.
As described above, the activity of C=S group in RAFT agent plays a pivotal role in controlling and regulating the polymerization process. Here, the typical types of RAFT agent structures are shown in Scheme 2. The realization of function depends greatly on the substitution of Z- and R-group, respectively. The former Z group often regulates the activity of C=S bond and stabilizes the intermediate radical. Owing to the versatility of monomer compatibility of RAFT, different kinds of Z-groups could be selected to polymerize monomers with different reactivity. At the same time, the latter R-group should be an excellent leaving group when the C=S bond of RAFT agent undergoes the addition of propagating radical, and it also can reinitiate polymerization rapidly and efficiently. In other words, the leaving R-group should have same function and efficiency with the initiator and monomer radical. In brief, an appropriate RAFT agent is vital to achieve controlled and living process in ideal RAFT polymerization. Until now, aromatic dithioester, trithiocarbonates, dithiocarbamates and xanthates are four major categories of typical RAFT agents. Aromatic dithioester and trithiocarbonates are often used to polymerize “More activated” monomers (MAMs), including butadiene, isoprene, styrene, vinylpyridine, (meth)acrylates, (meth)acrylamides, maleic anhydride, maleimide, and acrylonitrile. Z-group of aromatic dithioester and trithiocarbonates can stable the intermediate radical when MAMs radical adds to the C=S bond, attributing to the considerably stable radical generated from MAMs. By contrast, the high reactivity of “Less activated” monomers (LAMs) may produce a relatively stable intermediate radical. Xanthates and dithiocarbamates can deactivate the C=S bond through the delocalization of lone pair of electrons from oxygen in xanthates or nitrogen in dithiocarbamates to thiocarbonyl group, resulting in the destabilization of the radical intermediate for accelerated polymerization. More detailed tutorials for selection of the R-group and Z-group of RAFT agent for various polymerizations could be founded in other reviews [37,38].
The radical is another crucial factor in the RAFT process. The number and concentration of initiators as the parent of radical are closely related to the amount of dead chain and polymerization rate. Kinetic researches demonstrate the maximum number of living chains and the high polymerization rate would be achieved in a low concentration of initiators [39,40,41]. Meantime, monomers with a high propagating rate coefficient and initiators with high decomposition are also achieved. Ordinarily, thermal initiator is preferentially used for RAFT polymerization system, such as azodiisobutyronitrile (AIBN) [42,43,44], 2,2′-azobis-(2,4-dimethylvaleronitrile) (ABVN) [45], or benzoyl peroxide (BPO) [46].

3. Novel RAFT Techniques

In general, a conventional RAFT process employs thermal initiator as a radical source and thiocarbonylthio compound as a RAFT agent. The disadvantages of this thermal decomposition initiation system, like side reaction, self-initiation polymerization of monomer, and irreversible chain transfer, greatly limit their wide polymerizations in some special condition. Recently, new strategies of radical generation induced by other means (e.g., photo, metal, enzyme, redox and acid) and some special RAFT process have been reported to optimize the conventional RAFT polymerization.

3.1. Photo-Induced Initiation

In comparison with that of thermal initiation, a simplified RAFT operating process could be developed by photo-induced initiation, in which the polymerization condition is usually at room temperature without the degassing procedure. More importantly, the polymerization process can be facilely controlled through switching the “on” and “off” of light. This unique photo-controlled RAFT polymerization could be grouped into two types: (i) the homolysis of RAFT agent activated directly by light to give the initiated radical R instead of the conventional radical from added initiator; and, (ii) the generation of radical initiated by photoinitiator or photoredox catalyst through the electron transfer process under the light irradiation, known as photoinduced electron transfer-reversible addition-fragmentation chain transfer (PET-RAFT) (Figure 1).
Without the addition initiator or catalyst, the direct homolysis of RAFT agent using light can avoid contamination that is caused by the addition of external radical resource, especially for metal complex. In this process, RAFT agent plays the dual roles in both a photoinitiator and a chain transfer agent, called photoiniferter, which can generate the stable tertiary carbon-centered radical under the light irradiation. As shown in Table 1, Xanthates [48], dithioesters [47,49] and trithiocarbonates [50], as photoiniferter have been presented for the polymerization of MMA, MA, and acrylic acid in the early days. A 4-cyano-4-((dodecylsulfanylthiocarbonyl)sulfanyl)pentanoic acid (CDTPA)-derived poly(oligo(ethylene glycol) methyl ether methacrylate) (POEGMA) macro-CTA can also be photodissociated directly to conduct RAFT polymerization of BzMA without any photoinitiator, and the morphologies of polymeric nanoparticles could be regulated frequently by the change of the wavelength of light [51]. The higher polymerization efficiency could be obtained by photoinitiator process than that of the traditional RAFT process. For example, the decomposition rate of photoiniferter in the presence of UV irradiation is faster than that of thermal dissociation rate of AIBN at 60 °C, and the polymerization rate would be increased with the increasing of ultraviolet light intensity [52]. Nevertheless, the high energy of the UV light may lead to an irreversible RAFT agent decomposition and unexpected side reaction [53,54]. Recently, Green light radiation had been confirmed as the safe light source, which has been conducted to control the miniemulsion polymerization of butyl methacrylate successfully [55]. Boyer et al. described a RAFT strategy for the precise synthesis of a well-defined oligomer with five different vinyl monomer through the difference of relative activity when the individual unit monomer inserted in RAFT agent under green light initiation. It implements accurate controllability at the monomer level [56]. However, these polymers bearing a thiocarbonylthio end-group at the ω-end are still sensitive to light and lose the end-group fidelity [53,54].
For PET-RAFT approach, Boyer’s group reported the first RAFT polymerization that was activated by photoinduced electron transfer process coupled with photoredox catalyst [58]. Based on the proposed mechanism, the generation of radical R is accompanied by a redox process. Under light irradiation, the RAFT agent is reduced to provide a radical R, along with the formation of oxidized photoredox catalyst. This radical R initiates monomer to become a propagating radical, and enter the equilibrium of activation/deactivation chain transfer process. Different from the photodissociation RAFT agent, this PET-RAFT system can minimize the formation of dead chain to restrict the irreversible light degradation reaction by not using the exogenous radical initiators [58,59]. To date, thiocarbonylthio compounds with higher redox potential are perfect RAFT agent for reduction and initiation [58], and plenty of photoredox catalysts, such as fac-[Ir(ppy)3] [58,60], Ru(bpy)3Cl2 [61,62], zinc porphyrins (ZnTPP) [63,64], tertiary amine (NR3) [65], and Esosin Y [66,67], have been successfully applied to the PET-RAFT polymerization. As a special exception, ZnTPP can activate CPADB (−0.4 V versus SCE) more easily than that of BTPA (−0.6 V versus SCE), may result from the specific coordination activation of trithiocarbonate with ZnTPP [63]. Titanium dioxide (TiO2) has also been demonstrated to be a suitable photoredox catalyst to conduct PET-RAFT polymerization of MMA. However, the high energy of UV light would also decompose dithioester after irradiation time for 500 min, leading to the uncontrollable polymerization process [68]. Thus, visible light induced PET-RAFT polymerization systems have attracted more and more attention for researchers [57,69]. The Boyer’s group reported the PET-RAFT polymerization of photosensitive o-nitrobenzyl methacrylate without photolysis by-product of monomer/polymer by using ZnTPP or PheoA as photoredox catalyst, respectively, under visible red or yellow light [70]. Recently, the lower reduction potential of Bacteriochlorophyll a showed a large molar absorbability in the near-infrared and far-red region, which could reduce RAFT agent efficiently with low absorption overlap. Using bacteriochlorophyll a as a photoredox catalysts, the polymerizations of MMA and tert-butyl methacrylate with controlled molecular weight (PDI < 1.25) were reported with near-infrared and far-red irradiation. These light resources with excellent penetration would expand the potential application of PET-RAFT polymerization in biomedicine and industry [71]. On the other hand, the high potential for oxygen tolerance in PET-RAFT technology enables it to conduct the polymerization process under the air environment without the degassing procedure by freeze-pump-thaw cycles [63,65,66]. Qiao’s group reported the PET-RAFT polymerization of MA by using g-C3N4/amine as cocatalyst in the air condition, in which g-C3N4 cannot only advance the polymerization rate, but also accelerate the consumption of oxygen through the electron transfer from the amine to dissolved oxygen, which is mainly caused by more negative conduction band of g-C3N4. Interestingly, without the external deoxidant, the thioether group from (methylthio) ethyl methacrylate (MTEMA) monomer can quench photosensitized singlet oxygen in PET-RAFT process, which give no distinction product between the presence and the absence of air [72]. The details for the PET-RAFT are listed in Table 2.

3.2. Metal-Catalytic Initiation

A large number of researches and applications of the metal-catalyzed living controlled radical polymerization have been published [6]. Among these studies, the supplemental activator and the reducing agent atom transfer radical polymerization (SARA ATRP) and single electron transfer-controlled living radical polymerization (SET-LRP) technology are the major researches because of their different mechanism model. However, when compared to the SARA ATRP, SET-LRP often proceeded in most polar solvents, such as DMSO, alcohols, DMF, H2O, and binary mixtures of these solvents with trace amount of catalysts at room temperature or below [75,76]. Especially, vinyl chloride, which cannot be prepared using current ATRP catalysts, can successfully polymerize through SET-LRP under mild reaction conditions. Many characteristics of living radical polymerization, including ultrafast polymerization rate, ultrahigh molecular weight, excellent retention of chain-end, and few side reactions can be achieved in SET-LRP. On the bright side, the successful combination of SET-LRP and RAFT technology would give a significant boost in living polymerization with controlled reaction rate [77]. In the SET-RAFT process, the radical (R or Pn) is generated through SET process, in which the electron will transfer from Cu(0) to initiator (R-X) or terminated chain with the halogen end group (Pn-X). At the same time, the oxidized metal product Cu(I)X/L will disproportionate in an instant to produce Cu(0) and Cu(II)X2/L. The radical species (R or Pn) can be polymerized with monomer to proceed the RAFT equilibrium in the presence RAFT agent or be deactivated by Cu(II)X2/L to obtain the dormant state (R-X or Pn-X) (Figure 2). The SET-LRP method in SET-RAFT partly realizes the reversible termination in RAFT process. Various monomers including styrene, acrylates and methacrylates have been polymerized with good monomer conversion rate and low dispersity on basis of the SET-RAFT technology. Owing to the catalytic reaction in click chemistry with oxidized Cu(II), well-defined polymer architecture could be constructed by cooperating of SET-RAFT and click chemistry [78]. In brief, Cu(0) that is used in the SET-RAFT polymerization could be oxidized to be Cu2O to loss activity, while the Cu2O/ascorbic acid (Vc) or CuO/Vc redox pair in click reaction could in situ generate Cu(0), which is used directly for SET-RAFT polymerization [79].
To date, many metal-initiated RAFT systems have been published. As an example of iron (Fe), Zhu’s group reported the SET-RAFT polymerization of MMA at ambient temperature using zero-valent iron powder (Fe(0)) as a metal catalyst [80,81]. The proposed mechanism of Fe(0)-initiated RAFT is great different from that of Cu(0) catalytic system (Figure 3). The RAFT process may be in coexistence with the ATRP process, which mainly plays the control process. Using a hydrophilic RAFT agent of 4-(4-cyanopentanoic acid)dithiobenzoate (CPADB), hydrophilic monomers, including poly(ethylene glycol)monomethyl ether methacrylate (PEGMA) and DMAEMA, can go through a controllable polymerization with Fe(0) catalytic initiation. Meanwhile, Co(0), Ni(0), Zn(0) have also been introduced into SET-RAFT system for well polymerization of GMA and MMA by using 2-cyanoprop-2-yl 1-dithionaphthalate (CPDN) and cumyl dithionaphthalenoate (CDN) as RAFT agent [82].

3.3. Redox Initiation

When compared with thermal- and light-initiated polymerization, the redox-initiated polymerization is well developed and has been applied in a wide range of industrial products, due to its lower activation energy and higher radical generation rate under relative lower temperature. Generally, the effective redox initiator possesses a shorter induction period and higher monomer conversion. As Table 3 shows, persulfates and peroxides are two frequent oxidizing agents, which exhibit excellent redox initiation after coupling with a suitable reducing agent, such as benzoyl peroxide (BPO)/N,N-dimethylaniline (DMA) [83], benzoyl peroxide (BPO)/aromatic tertiary amine functionalized ATP (ATP-ATA) [84], Potassium persulfate (KPS)/L-Ascorbic acid sodium salt (NaAs) [85]. An et al. developed an aqueous redox-initiated RAFT polymerization system for fabrication of amphiphilic block copolymers [85] and highly ordered multiblock copolymer [86]. In the redox-initiated RAFT polymerization process, the permission of low reaction temperature would restrict the probability of side reaction, resulting in well living feature with high conversion (>90%) and low dispersity (<1.10) in the polymerization of NIPAM, acrylamide (AM), MMA, and styrene [87,88]. Employing ammonium persulfate (APS) and sodium formaldehyde sulfoxylate (SFS) redox initiation system, water-soluble triblock copolymer with ultra-high molecular weight (Mn > 106) could be obtained from monomers of AA and AM [89]. Meanwhile, a novel redox-initiation polymerization using Feton reaction has been reported to achieve the polymerization of DMA, NAM, and HEA, with shorter reaction time (1 min) and low polydispersity (D < 1.06) [90].

3.4. Enzyme Initiation

As a green and valuable technology, the biocatalyst enzyme system has been widely utilized in the polymerization of degradable materials [91] and functional monomers [92]. Owing to the nature of high efficiency, benign, and eco-friendly, enzymatic catalysis initiation has also attracted many interests for RDRP researchers [93,94]. An et al. reported that horseradish peroxide (HRP) can initiate the RAFT polymerization of most of monomers directly under differential reaction environment [95]. In the HRP/acetylacetone (ACAC)/H2O2 ternary initiator, H2O2 catalyzed by HRP enzyme can oxidize the ACAC to produce ACAC radicals, which involves the initiation in the RAFT polymerization subsequently. In this system, advantages of fast reaction time (30 min), high monomer conversion (>90%), and low molecular weight distribution in RAFT polymerization could be observed [96]. Initially, enzyme catalysis was used to assist to degas prior to RAFT polymerization [97]. For example, the biosystem of glucose oxidase (GOx) and glucose has been employed to remove dissolved oxygen prior to PET-RAFT polymerization [74]. Recently, the smart cascade method successfully conducted redox initiation RAFT polymerization without conventional deoxygenation procedure [98]. As shown in Figure 4a, the level of dissolved oxygen can be consumed to limited level through GOx catalyzed deoxygenation reaction to produce H2O2, which can directly initiate DMA polymerization with ascorbic acid as a reductant [99]. Recently, multiblock polymers (up to 10 block) with ultrahigh molecular weight (2.3 × 106 g/mol) were synthesized by using enzymatic-initiated RAFT polymerization with pyranose oxidase (P2Ox)-HRP cascade, where P2Ox primarily deoxygenates to ensure the catalysis of HRP for the generation of radical and the initiation of polymerization (Figure 4b) [100].

3.5. Acid Initiation

According to earlier reports, the vinyl ester is easier to be polymerized via cationic mechanism rather than radical mechanism, hence immense effort is required to obtain the copolymer containing vinyl ether by using RAFT radical approach. Lewis acids have be employed to adjust the interconversion between the RAFT radical polymerization and living cationic polymerization [101]. With this in mind, MA was homopolymerized through RAFT radical polymerization and gave a dormant species of C-SC(S)-Z. The addition of Lewis acid would activate this dormant species into cationic species for further dominating the living cationic polymerization of vinyl ester. Further research showed that the organic acids also could be used as initiator for the cationic RAFT polymerization. As an example of isobutyl vinyl ether (IBVE), it could be conducted through cationic RAFT polymerization using a handful of triflic acid as the initiator [102]. Trithiocarbonate and dithiocarbamate with electro-donating nitrogen exhibit good stabilization of cationic intermediate, which could control the molecular weights of IBVE homopolymer (PDI < 1.1) effectively in the triflylimide-initiated RAFT system. Owing to its excellent compatibility, block polymers with IBVE and MA comonomer could be synthesized through the transformation between cationic RAFT polymerization and radical RAFT polymerization (Figure 5a) [103]. Recently, the reaction between ketone and hydroperoxide participated in the organic reaction can generate the radical under acid catalysis [104]. Junkers et al. reported the utilization of p-toluenesulfonic acid to catalyze cyclohexanone/tert-butylhydroperoxide for generation ketone radical, which could be initiated RAFT polymerization of NIPAAM, styrene, and butyl acrylate (BA) with well living control, high end-group fidelities, and low dispersity [105]. However, production of ketone radical from tert-butylhydroperoxide will form the di-tert-butylperoxide, leading to the end of propagating chain in the polymerization (Figure 5b).

3.6. Other Initiation

Distinguish from the SET-RAFT polymerization, the metal-free dissociative electron transfer RAFT (DET-RAFT) method has been developed for the polymerization of styrene and MA at low temperature (30 °C) [106]. As a replacement of Cu(0), inorganic sulfites will transfer one electron to the RAFT agent to form radical anions through the DET process on accounts of a higher reduction potential of sulfites than that of relevant RAFT agent. This radical anion will split into a radical and an anion, following by initiation of RAFT polymerization. The polymers with narrow molecular weights distribution (PDI ≤ 1.3) and high end-group fidelity can be achieved in the DET-RAFT polymerization. Ultrasonic irradiation was once employed to initiate RAFT polymerization to graft polymers on the silica gel [107]. Inspired by the initiation system of Feton reaction catalysis [90], subsequent mechanism research illuminated that the hydroxyl radicals from water play an essential role in polymerization process. Given optimized ultrasonic parameter (frequency: 414 Hz; applied power: 40 W), the polymerization of water-soluble 2-hydroxyethylacrylate showed a high monomer conversion (>90%) within 60 min in water [108]. Besides, an excellent electrochemical-induced RAFT polymerization was reported recently [109]. The irreversible redox process of common RAFT agent makes it inapplicable for direct electrochemical activation. Using BPO or 4-bromobenzenediazonium tetrafluoroborate (BrPhN2+) as the radical initiator, the aryl radicals could be generated by electrochemical reduction for RAFT polymerization. Here, BrPhN2+ is more suitable for electrochemical-initiator than BPO, due to the similar reduction potential between BPO and workable RAFT agent. In the electrochemical-initiated RAFT polymerization, current or potential regulates the generation rate of radical as well as the polymerization degree. Until now, the homopolymers and block copolymers of BA and tBA could be fabricated with narrow molecular weight distribution successfully by electrochemical-initiated RAFT polymerization. All of the detailed summary is listed in Table 4.

3.7. Sulfur-Free RAFT Polymerization

Aside from the research on initiation, there are some fancy findings in RAFT agent. Originally, the products from RAFT polymerization often exist in the thiocarbonate group at the ω-end, which would be destabilized with inappropriate odour and color, limiting their wide applications in many fields. The major improvements have been done by removing or functionalizing the thiocarbonate end group of RAFT products [110]. Out of these efforts, some attempts on screening of sulfur-free RAFT agents have been reported for living radical polymerization successfully [111]. When using α-methylstyrene dimer (AMSD) as RAFT agent, RAFT-like process is observed in the polymerization (Figure 6) [112]. However, the tertiary radical generated from AMSD is relatively unstable, resulting in uncontrollable poly(methacrylate)s with large molecular weight distribution. Methanol also has been investigated as RAFT agent (optimal concentration of 6.5 mol %) for the synthesis of poly(4-methoxystyrene) (PDI ≤ 1.2) bearing methoxy end group [113].
Recently, Haddleton and coworkers developed the sulfur-free RAFT emulsion polymerization for the preparation of methacrylic block copolymers [114,115]. Vinyl-terminated poly(methyl methacrylate) (PMMA) oligomer could be fabricated through the method of catalytic chain transfer polymerization (CCTP) in emulsion [116]. These unsaturated oligomers presented appealing chain transfer constants, which could be continue to be used as chain transfer agent for RAFT polymerization of MMAs, following the addition of potassium persulfate as radical resource (Figure 7) [117,118]. Based on this strategy, the heneicosa and tetracosa quasi multiblock poly(butyl methacrylate) could be achieved with narrow molecular weight distribution (PDI ≤ 1.21). The copolymer of benzyl methacrylate, 2-ethyl hexyl methacrylate, and butyl methacrylate could also be prepared with extremely high conversions (>99%). With the increasing of polymerization degree (DPn = 45) in each block, higher molecular weight hepta-block copolymers were synthesized. Furthermore, the product at a large scale (up to 80 g in a 0.5 L double jacket reactor) could be gained with a low polydispersity (PDI ≤ 1.24), as well as high molecular weight (Mn = 41,300 g/mol). These attractive results suggest the sulfur-free RAFT technology for wide prospects in application.

4. Application

Since conducting polyacetylene was firstly reported by Heeger and coworkers in 1977 [119], optoelectronic polymers, as a new type of organic functional materials, have great research significance and application potential because of their excellent optoelectronic properties, easily preparation, and low cost. However, great efforts have been solely devoted to the synthesis of functional polymers or the optoelectronic applications of existing materials, there have been only few other interdisciplinary efforts in the field of optoelectronics. D’hooge et al. have attempted to carry on interdisciplinary to combine the simulation and radical polymerization [120,121,122], which greatly empowers us to understand and control radical polymerization. According to the experiment data, the kinetic models has been established to evaluate two different precursor routes for the synthesis of poly(p-phenylene vinylene) (PPV) derivative that is often as active material for the optoelectronic devices, and the simulation results contribute to our selection of optimal reactants and reaction conditions [123,124].
However, RAFT polymerization with many good merits also satisfies the interdisciplinary research with an optoelectronic field. First, the undesired metal catalyst can be avoided to use for the synthesis of functional polymers. At the same time, narrow molecular weight distribution will greatly reduced the low molecular weight “impurities” which can degrade performance of semiconductors as electron or hole traps. Then, simplified preparation methods of multiblock copolymers and complex polymer architectures allow us to design and prepare multi-functional macromolecular materials. Finally, RAFT agent can also play the role of binder between organic materials and inorganic/organic materials. All of this suggests that it’s greater prospects for the applying RAFT polymerization to the preparation of optoelectronic materials. The use of functional polymer from RAFT polymerization in the electro-optic field has been prospering in recent years and plenty of related researches have been reported [125,126].
Based on the trans-cis isomerization of azobenzene derivatives under light irradiation, Kuo et al. reported the synthesis of functional material, named PVB-DAP, by RAFT polymerization, which could further interact with thymine-functionalized azobenzene to give a supramolecular complex. This supramolecular material featured the “recordable” and “rewritable” photo-responsive properties, which implied its promising application in optical devices (Figure 8) [127]. The Silicon nanocrystals as the photoluminescent material often suffer from oxidation-induced poor emission performance. Rieger et al. demonstrated the surface modification of silicon nanocrystals by using RAFT polymerization of polystyrene to form core-shell hybrid product, which provided the more stable photoluminescence properties, due to the protection of polymer shell against oxidation [128]. The amino-containing aggregation-induced emission (AIE) fluorescent molecule could be introduced into the side chain of hydrophilic polymers (PEGMA-co-UCL) via the combination of RAFT polymerization and mercaptoacetic acid locking imine (MALI) reaction in one pot, the resulting luminescent polymer with AIE characteristic exhibited the ideal candidates for cell imaging based on its intensive emission and excellent biocompatibity [129]. Meanwhile, AIE also could be introduced to couple with RAFT agent, which then was used to synthesize the AIE active fluorescent polymeric nanoparticles. Zhang et al. found that the AIE-actived polymers with enhanced fluorescence intensity and photostability are beneficial for the application in biological imaging (Figure 9) [130]. Similarly, an amphiphilic copolymer has been synthesized from hydrophobic fluorescein methacryloyl monomer and hydrophilic poly(ethylene glycol) monomethacrylate. The fluorescent organic nanoparticles that were obtained from the self-assemble of this copolymer showed the good biocompatibility and excellent florescence, rendering it a promising material for the application in cell imaging [131]. Kong et al. synthesized an amphiphilic poly(N-vinyl carbazole) (PVK)-based block copolymer with the pendent of tris(8-hydroxyquinoline) aluminum (Alq3) using trithiocarbonate terminated PEG as a RAFT agent [132]. The resonance energy transfer from PVK to Alq3 could give raise to the enhancement of emission from Alq3, which made it an excellent material as electron-transporting and emission layer in OLED. Ru(bpy)32+ not only is an efficient catalyst for PET-RAFT polymerization, but also can be used as a electrogenerated chemiluminescent (ECL) reagent. Liu et al. presented the preparation of pyrene-modified poly(sodium p-styrene sulfonate) (PSS) through PET-RAFT approach using Ru(bpy)32+ as catalyst, which could be directly used to fabricate the solid-state ECL sensor through the electrostatic interaction between Ru(bpy)32+ and PSS. This ECL sensor exhibited the high sensitivity and long term stability (Figure 10) [133].
For electric applications, Chen et al. reported the synthesis of butoxy-substituded triphenylene (TP)-based polymers by RAFT strategy. Owing to the existence of rigid triphenylene, homopolymer could self-assemble into various ordered columnar plastic phases with highest hole mobility of 0.1 cm2 V−1 s−1 (Figure 11) [134]. Subsequently, they developed a cross-linked copolymer with high dielectric constant, which has been demonstrated the superior low-voltage operation performance in the application of organic field-effect transistors [135]. The core-shell nanoparticles can also give the excellent dielectric performance. Jiang and coworkers synthesized a series of core-shell polystyrene (PS)@BaTiO3 nanocomposites by grafting PS from the BaTiO3 surface through in situ RAFT polymerization [136]. The different thickness of PS shell could regulate the dielectric constant of nanoparticles, and the dielectric constant of core-shell BaTiO3@PS is 24, which is roughly eight times larger than that of separate PS (2.76). Meanwhile, the extremely low dielectric loss of BaTiO3@PS had been kept. Chen et al. reported the preparation of cross-linked core-shell block copolymer, poly[poly(ethylene glycol)methylether methacrylate]-block-poly(2,5-dibromo-3-vinylthiophene) (poly(PEGMA)m-b-poly(DB3VT)n) on the basis of RAFT technology, which was applied firstly as charge-storage materials for non-volatile OFET memory devices [137]. They demonstrated the size of the cross-linked nanocomposites and the molecular weights of poly(PEGMA) segments have positive correlation, while the mean diameter of poly(DB3VT) segments can effectively adjust the optical and electronic memory properties. The maximum hole-trapping capacity of the OFET memory device could be achieved by using the block copolymer of poly(PEGMA)m-b-poly(DB3VT)n with a monomer molar ratio of 77:23 (m:n). Recently, Chen’s group reported two kinds of RAFT strategy for the preparation of poly(N-vinylcarbazole) (PVK)-modified graphene oxides (GO). One was “Graft to” method, in which PVK was prepared by traditional RAFT approach by using S-1-dodecyl-S’-(α,α’-dimethyl-α-aceticacid) trithiocarbonate (DDAT) as a RAFT agent, and the product of DDAT-PVK can directly bond onto the isocyanate-functionalized GO with covalent amido linkage [138]. Another was “Graft from” method, where covalent DDAT-functionalized GO (DDAT-GO) was firstly fabricated and used as macroRAFT agent for the polymerization of N-vinylcarbazole through RAFT technology [139]. Both two methods could improve the solubility of GO, which realized the solution process of GO-based materials. When PVK-GO was used for the polymer memories, the device presented an excellent bistable electrical switching and nonvolatile memory effect. Emrick et al. elaborated the preparation of tetrathiafulvalene (TTF)-based polymers using RAFT polymerization, which was then used for the non-covalent surface functionalization of MoS2 to decrease its work function and modulate the electronic band structure [140]. In the same year, Chen and coworkers firstly synthesized covalent PVK-modified MoS2 nanosheets through RAFT polymerization, which exhibited an improvement of solubility in organic solvent. As electric devices, the enhancement of charge transfer of MoS2-PVK guaranteed the excellent optical limiting response and nonvolatile rewritable memory performance (Figure 12) [141,142]. Similar results have also been found in covalent polyacrylonitrile (PAN)-functioned MoS2 nanosheets (MoS2-PAN) [143,144]. These findings give the growing evidences of RAFT polymerization for huge potential in the optoelectronic field.

5. Conclusions

From the basic discoveries to practical applications, RAFT polymerization has become one of the indispensable parts in polymer chemistry. Large number of theoretical and empirical researches are relevant to RAFT polymerization during the past 20 years. New RAFT techniques have been proposed to make up the deficiency of conventional RAFT methods, such as different initiation system (e.g., photo, metal, enzyme, redox, and acid) and sulfur-free RAFT polymerization. Although a great deal of progress has already been achieved, several major challenges still exist for these novel RAFT techniques to achieve their broad applications in optoelectronic field, including new RAFT polymerization system, convenient operating condition without anaerobic treatment, introduction of various optoelectronic properties onto backbone of polymer by RAFT polymerization, and investigation of complex architecture and self-assembly of polymers synthesized by RAFT polymerization on their optoelectronic performances. With the continuous advancement of RAFT technology, we believe that RAFT technology will become the most powerful tool available for design and fabrication of complex architecture polymers for optoelectronic applications.

Acknowledgments

The authors are grateful for the financial support of the National Natural Science Foundation of China (51333002, 21404037, 61378072), Research Fund for the Doctoral Program of Higher Education of China (20120074110004), the Fundamental Research Funds for the Central Universities (WJ1514311), the Chenguang Program (15CG28), and the Pujiang Program (16PJ1402400). Xiaodong Zhuang acknowledge the financial support from NSFC for Excellent Young Scholars (51722304), NSFC (5171101862, 21720102002), Shanghai Committee of Science and Technology (15JC1490500, 16JC1400703), and Open Project Program of the State Key Laboratory of Photocatalysis on Energy and Environment (SKLPEE-KF201702, Fuzhou University).

Author Contributions

Xiangyu Tian and Junjie Ding organized the literature; Xiangyu Tian drafted the manuscript; Bin Zhang assembled the charts. The manuscript was polished through contributions of Feng Qiu, Xiaodong Zhuang and Yu Chen. All authors have given approval to the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sugita, H.; Nojima, M.; Ohta, Y.; Yokozawa, T. Unusual cyclic polymerization through suzuki-miyaura coupling of polyphenylene bearing diboronate at both ends with excess dibromophenylene. Chem. Commun. 2017, 53, 396–399. [Google Scholar] [CrossRef] [PubMed]
  2. Raj, M.R.; Kim, M.; Kim, H.I.; Lee, G.-Y.; Park, C.W.; Park, T. A comparative study on the thermal- and microwave-assisted stille coupling polymerization of a benzodithiophene-based donor-acceptor polymer (PTB7). J. Mater. Chem. A 2017, 5, 3330–3335. [Google Scholar] [CrossRef]
  3. Abdul Kudus, M.H.; Zakaria, M.R.; Hafi Othman, M.B.; Md Akil, H. Preparation and characterization of colloidized diamine/oxidized-graphene via condensation polymerization of carboxyl groups epoxy/oxidized-graphene nanocomposite. Polymer 2017, 124, 186–202. [Google Scholar] [CrossRef]
  4. Nojima, M.; Ohta, Y.; Yokozawa, T. Investigation of catalyst-transfer condensation polymerization for synthesis of poly(p-phenylenevinylene). J. Polym. Sci. Part A 2014, 52, 2643–2653. [Google Scholar] [CrossRef]
  5. Yokozawa, T.; Yokoyama, A. Chain-growth condensation polymerization for the synthesis of well-defined condensation polymers and π-conjugated polymers. Chem. Rev. 2009, 109, 5595–5619. [Google Scholar] [CrossRef] [PubMed]
  6. Kamigaito, M.; Ando, T.; Sawamoto, M. Metal-catalyzed living radical polymerization. Chem. Rev. 2001, 101, 3689–3746. [Google Scholar] [CrossRef] [PubMed]
  7. Moad, G.; Solomon, D.H. The Chemistry of Radical Polymerization, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2006. [Google Scholar]
  8. Matyjaszewski, K.; Davis, T.P. Handbook of Radical Polymerization, 1st ed.; John Wiley & Sons: Hoboken, NJ, USA, 2003. [Google Scholar]
  9. Braunecker, W.A.; Matyjaszewski, K. Controlled/living radical polymerization: Features, developments, and perspectives. Prog. Polym. Sci. 2007, 32, 93–146. [Google Scholar] [CrossRef]
  10. Coessens, V.; Pintauer, T.; Matyjaszewski, K. Functional polymers by atom transfer radical polymerization. Prog. Polym. Sci. 2001, 26, 337–377. [Google Scholar] [CrossRef]
  11. Otsu, T.; Yoshida, M.; Tazaki, T. A model for living radical polymerization. Macromol. Chem. Rapid Commun. 1982, 3, 133–140. [Google Scholar] [CrossRef]
  12. Szwarc, M. ‘Living’ polymers. Nature 1956, 178, 1168–1169. [Google Scholar] [CrossRef]
  13. Crivello, J.V.; Lam, J.H.W. Diaryliodonium salts. A new class of photoinitiators for cationic polymerization. Macromolecules 1977, 10, 1307–1315. [Google Scholar] [CrossRef]
  14. Crivello, J.V. Cationic polymerization—Iodonium and sulfonium salt photoinitiators. In Initiators—Poly-Reactions—Optical Activity, 1st ed.; Springer: Berlin/Heidelberg, Germany, 1984; Volume 62, pp. 1–48. [Google Scholar]
  15. Young, R.N.; Quirk, R.P.; Fetters, L.J. Anionic polymerizations of non-polar monomers involving lithium. In Anionic Polymerization, 1st ed.; Springer: Berlin/Heidelberg, Germany, 1984; Volume 56, pp. 1–90. [Google Scholar]
  16. Hadjichristidis, N.; Pitsikalis, M.; Pispas, S.; Iatrou, H. Polymers with complex architecture by living anionic polymerization. Chem. Rev. 2001, 101, 3747–3792. [Google Scholar] [CrossRef] [PubMed]
  17. Bielawski, C.W.; Grubbs, R.H. Living ring-opening metathesis polymerization. Prog. Polym. Sci. 2007, 32, 1–29. [Google Scholar] [CrossRef]
  18. Xu, G.; Chung, T.C. Borane chain transfer agent in metallocene-mediated olefin polymerization. Synthesis of borane-terminated polyethylene and diblock copolymers containing polyethylene and polar polymer. J. Am. Chem. Soc. 1999, 121, 6763–6764. [Google Scholar] [CrossRef]
  19. Xu, G.; Chung, T.C. Synthesis of syndiotactic polystyrene (s-PS) containing a terminal polar group and diblock copolymers containing s-PS and polar polymers. Macromolecules 1999, 32, 8689–8692. [Google Scholar] [CrossRef]
  20. Nicolas, J.; Guillaneuf, Y.; Lefay, C.; Bertin, D.; Gigmes, D.; Charleux, B. Nitroxide-mediated polymerization. Prog. Polym. Sci. 2013, 38, 63–235. [Google Scholar] [CrossRef]
  21. Hawker, C.J.; Bosman, A.W.; Harth, E. New polymer synthesis by nitroxide mediated living radical polymerizations. Chem. Rev. 2001, 101, 3661–3688. [Google Scholar] [CrossRef] [PubMed]
  22. Matyjaszewski, K.; Xia, J. Atom transfer radical polymerization. Chem. Rev. 2001, 101, 2921–2990. [Google Scholar] [CrossRef] [PubMed]
  23. Chiefari, J.; Chong, Y.K.; Ercole, F.; Krstina, J.; Jeffery, J.; Le, T.P.T.; Mayadunne, R.T.A.; Meijs, G.F.; Moad, C.L.; Moad, G.; et al. Living free-radical polymerization by reversible addition-fragmentation chain transfer: The RAFT process. Macromolecules 1998, 31, 5559–5562. [Google Scholar] [CrossRef]
  24. Perrier, S.; Takolpuckdee, P. Macromolecular design via reversible addition-fragmentation chain transfer (RAFT)/xanthates (madix) polymerization. J. Polym. Sci. Part A 2005, 43, 5347–5393. [Google Scholar] [CrossRef]
  25. Nakatani, R.; Takano, H.; Chandra, A.; Yoshimura, Y.; Wang, L.; Suzuki, Y.; Tanaka, Y.; Maeda, R.; Kihara, N.; Minegishi, S.; et al. Perpendicular orientation control without interfacial treatment of RAFT-synthesized high-χ block copolymer thin films with sub-10 nm features prepared via thermal annealing. ACS Appl. Mater. Interfaces 2017, 9, 31266–31278. [Google Scholar] [CrossRef] [PubMed]
  26. Jiang, Z.-Q.; Zhao, S.-Q.; Su, Y.-X.; Liu, N.; Wu, Z.-Q. Combination of RAFT and Pd(II)-initiated isocyanide polymerizations: A versatile method for facile synthesis of helical poly(phenyl isocyanide) block and star copolymers. Macromolecules 2018, 51, 737–745. [Google Scholar] [CrossRef]
  27. Döhler, D.; Kaiser, J.; Binder, W.H. Supramolecular H-bonded three-arm star polymers by efficient combination of RAFT polymerization and thio-bromo “click” reaction. Polymer 2017, 122, 148–158. [Google Scholar] [CrossRef]
  28. Tochwin, A.; El-Betany, A.; Tai, H.; Chan, K.; Blackburn, C.; Wang, W. Thermoresponsive and reducible hyperbranched polymers synthesized by RAFT polymerisation. Polymers 2017, 9, 443. [Google Scholar] [CrossRef]
  29. Chen, L.; Li, Y.; Yue, S.; Ling, J.; Ni, X.; Shen, Z. Chemoselective RAFT polymerization of a trivinyl monomer derived from carbon dioxide and 1,3-butadiene: From linear to hyperbranched. Macromolecules 2017, 50, 9598–9606. [Google Scholar] [CrossRef]
  30. Sutrisno, J.; Pramudya, I.; Aerken, X.; Fuchs, A. Surface grafting of poly(pentafluorostyrene) on the iron and iron oxide particles via reversible addition fragmentation chain transfer (RAFT) polymerization. J. Appl. Polym. Sci. 2017, 134, 44898. [Google Scholar] [CrossRef]
  31. Kochameshki, M.G.; Marjani, A.; Mahmoudian, M.; Farhadi, K. Grafting of diallyldimethylammonium chloride on graphene oxide by RAFT polymerization for modification of nanocomposite polysulfone membranes using in water treatment. Chem. Eng. J. 2017, 309, 206–221. [Google Scholar] [CrossRef]
  32. Enrique-Jimenez, P.; Quiles-Díaz, S.; Salavagione, H.J.; Wesner, D.; Schönherr, H.; González-Casablanca, J.; García-Quismondo, R.; Martínez, G.; Gómez-Fatou, M.A.; Ania, F.; et al. Control of the structure and properties of SEBS nanocomposites via chemical modification of graphene with polymer brushes. Eur. Polym. J. 2017, 97, 1–13. [Google Scholar] [CrossRef]
  33. Matsukawa, K.; Masuda, T.; Kim, Y.S.; Akimoto, A.M.; Yoshida, R. Thermoresponsive surface-grafted gels: Controlling the bulk volume change properties by surface-localized polymer grafting with various densities. Langmuir 2017, 33, 13828–13833. [Google Scholar] [CrossRef] [PubMed]
  34. Moad, G. Raft polymerization to form stimuli-responsive polymers. Polym. Chem. 2017, 8, 177–219. [Google Scholar] [CrossRef]
  35. Pearce, A.K.; Travanut, A.; Couturaud, B.; Taresco, V.; Howdle, S.M.; Alexander, M.R.; Alexander, C. Versatile routes to functional raft chain transfer agents through the passerini multicomponent reaction. ACS Macro Lett. 2017, 6, 781–785. [Google Scholar] [CrossRef]
  36. Tang, M.; Zhou, M.; Huang, Y.; Zhong, J.; Zhou, Z.; Luo, K. Dual-sensitive and biodegradable core-crosslinked HPMA copolymer-doxorubicin conjugate-based nanoparticles for cancer therapy. Polym. Chem. 2017, 8, 2370–2380. [Google Scholar] [CrossRef]
  37. Moad, G.; Rizzardo, E.; Thang, S.H. Living radical polymerization by the RAFT process? A third update. Aust. J. Chem. 2012, 65, 985–1076. [Google Scholar] [CrossRef]
  38. Perrier, S. 50th anniversary perspective: RAFT polymerization—A user guide. Macromolecules 2017, 50, 7433–7447. [Google Scholar] [CrossRef]
  39. Vana, P.; Davis, T.P.; Barner-Kowollik, C. Kinetic analysis of reversible addition fragmentation chain transfer (RAFT) polymerizations: Conditions for inhibition, retardation, and optimum living polymerization. Macromol. Theory Simul. 2002, 11, 823–835. [Google Scholar] [CrossRef]
  40. Rybel, N.D.; Steenberge, P.H.M.V.; Reyniers, M.F.; Barner-Kowollik, C.; D’Hooge, D.R.; Marin, G.B. An update on the pivotal role of kinetic modeling for the mechanistic understanding and design of bulk and solution raft polymerization. Macromol. Theory Simul. 2017, 26. [Google Scholar] [CrossRef]
  41. De Rybel, N.; Van Steenberge, P.H.M.; Reyniers, M.-F.; D’Hooge, D.R.; Marin, G.B. How chain length dependencies interfere with the bulk RAFT polymerization rate and microstructural control. Chem. Eng. Sci. 2018, 177, 163–179. [Google Scholar] [CrossRef]
  42. Wan, W.-M.; Pan, C.-Y. Formation of polymeric yolk/shell nanomaterial by polymerization-induced self-assembly and reorganization. Macromolecules 2010, 43, 2672–2675. [Google Scholar] [CrossRef]
  43. Lima, V.; Jiang, X.; Brokken-Zijp, J.; Schoenmakers, P.J.; Klumperman, B.; Van Der Linde, R. Synthesis and characterization of telechelic polymethacrylates via RAFT polymerization. J. Polym. Sci. Part A 2005, 43, 959–973. [Google Scholar] [CrossRef]
  44. Yuan, K.; Li, Z.-F.; Lu, L.-L.; Shi, X.-N. Synthesis and characterization of well-defined polymer brushes grafted from silicon surface via surface reversible addition–fragmentation chain transfer (RAFT) polymerization. Mater. Lett. 2007, 61, 2033–2036. [Google Scholar] [CrossRef]
  45. Fu, C.; Yang, B.; Zhu, C.; Wang, S.; Zhang, Y.; Wei, Y.; Tao, L. Synthesis of gradient copolymers by concurrent enzymatic monomer transformation and raft polymerization. Polym. Chem. 2013, 4, 5720–5725. [Google Scholar] [CrossRef]
  46. Lai, J.T.; Filla, D.; Shea, R. Functional polymers from novel carboxyl-terminated trithiocarbonates as highly efficient raft agents. Macromolecules 2002, 35, 6754–6756. [Google Scholar] [CrossRef]
  47. Xu, J.; Shanmugam, S.; Corrigan, N.A.; Boyer, C. Catalyst-free visible light-induced RAFT photopolymerization. In Controlled Radical Polymerization: Mechanisms, 1st ed.; Matyjaszewski, K., Sumerlin, B.S., Tsarevsky, N.V., Chiefari, J., Eds.; American Chemical Society: Washington, DC, USA, 2015; Volume 1187, pp. 247–267. [Google Scholar]
  48. Ajayaghosh, A.; Francis, R. A xanthate-derived photoinitiator that recognizes and controls the free radical polymerization pathways of methyl methacrylate and styrene. J. Am. Chem. Soc. 1999, 121, 6599–6606. [Google Scholar] [CrossRef]
  49. Quinn, J.F.; Barner, L.; Barner-Kowollik, C.; Rizzardo, E.; Davis, T.P. Reversible addition-fragmentation chain transfer polymerization initiated with ultraviolet radiation. Macromolecules 2002, 35, 7620–7627. [Google Scholar] [CrossRef]
  50. Muthukrishnan, S.; Pan, E.H.; Stenzel, M.H.; Barner-Kowollik, C.; Davis, T.P.; Lewis, D.; Barner, L. Ambient temperature raft polymerization of acrylic acid initiated with ultraviolet radiation in aqueous solution. Macromolecules 2007, 40, 2978–2980. [Google Scholar] [CrossRef]
  51. Yeow, J.; Sugita, O.R.; Boyer, C. Visible light-mediated polymerization-induced self-assembly in the absence of external catalyst or initiator. ACS Macro Lett. 2016, 5, 558–564. [Google Scholar] [CrossRef]
  52. Khan, M.Y.; Cho, M.-S.; Kwark, Y.-J. Dual roles of a xanthate as a radical source and chain transfer agent in the photoinitiated RAFT polymerization of vinyl acetate. Macromolecules 2014, 47, 1929–1934. [Google Scholar] [CrossRef]
  53. Zhang, H.; Deng, J.; Lu, L.; Cai, Y. Ambient-temperature RAFT polymerization of styrene and its functional derivatives under mild Long-Wave UV–Vis radiation. Macromolecules 2007, 40, 9252–9261. [Google Scholar] [CrossRef]
  54. Wang, H.; Li, Q.; Dai, J.; Du, F.; Zheng, H.; Bai, R. Real-time and in situ investigation of “living”/controlled photopolymerization in the presence of a trithiocarbonate. Macromolecules 2013, 46, 2576–2582. [Google Scholar] [CrossRef]
  55. Jung, K.; Boyer, C.; Zetterlund, P.B. RAFT iniferter polymerization in miniemulsion using visible light. Polym. Chem. 2017, 8, 3965–3970. [Google Scholar] [CrossRef]
  56. Fu, C.; Huang, Z.; Hawker, C.J.; Moad, G.; Xu, J.; Boyer, C. RAFT-mediated, visible light-initiated single unit monomer insertion and its application in the synthesis of sequence-defined polymers. Polym. Chem. 2017, 8, 4637–4643. [Google Scholar] [CrossRef]
  57. Fu, Q.; Xie, K.; McKenzie, T.G.; Qiao, G.G. Trithiocarbonates as intrinsic photoredox catalysts and RAFT agents for oxygen tolerant controlled radical polymerization. Polym. Chem. 2017, 8, 1519–1526. [Google Scholar] [CrossRef]
  58. Xu, J.; Jung, K.; Atme, A.; Shanmugam, S.; Boyer, C. A robust and versatile photoinduced living polymerization of conjugated and unconjugated monomers and its oxygen tolerance. J. Am. Chem. Soc. 2014, 136, 5508–5519. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, M.; MacLeod, M.J.; Johnson, J.A. Visible-light-controlled living radical polymerization from a trithiocarbonate iniferter mediated by an organic photoredox catalyst. ACS Macro Lett. 2015, 4, 566–569. [Google Scholar] [CrossRef]
  60. Jung, K.; Xu, J.; Zetterlund, P.B.; Boyer, C. Visible-light-regulated controlled/living radical polymerization in miniemulsion. ACS Macro Lett. 2015, 4, 1139–1143. [Google Scholar] [CrossRef]
  61. Xu, J.; Jung, K.; Corrigan, N.A.; Boyer, C. Aqueous photoinduced living/controlled polymerization: Tailoring for bioconjugation. Chem. Sci. 2014, 5, 3568–3575. [Google Scholar] [CrossRef]
  62. Christmann, J.; Ibrahim, A.; Charlot, V.; Croutxé-Barghorn, C.; Ley, C.; Allonas, X. Elucidation of the key role of [ru(bpy)3]2+ in photocatalyzed raft polymerization. ChemPhysChem 2016, 17, 2309–2314. [Google Scholar] [CrossRef] [PubMed]
  63. Shanmugam, S.; Xu, J.; Boyer, C. Exploiting metalloporphyrins for selective living radical polymerization tunable over visible wavelengths. J. Am. Chem. Soc. 2015, 137, 9174–9185. [Google Scholar] [CrossRef] [PubMed]
  64. Xu, J.; Shanmugam, S.; Fu, C.; Aguey-Zinsou, K.-F.; Boyer, C. Selective photoactivation: From a single unit monomer insertion reaction to controlled polymer architectures. J. Am. Chem. Soc. 2016, 138, 3094–3106. [Google Scholar] [CrossRef] [PubMed]
  65. Fu, Q.; McKenzie, T.G.; Tan, S.; Nam, E.; Qiao, G.G. Tertiary amine catalyzed photo-induced controlled radical polymerization of methacrylates. Polym. Chem. 2015, 6, 5362–5368. [Google Scholar] [CrossRef]
  66. Xu, J.; Shanmugam, S.; Duong, H.T.; Boyer, C. Organo-photocatalysts for photoinduced electron transfer-reversible addition-fragmentation chain transfer (PET-RAFT) polymerization. Polym. Chem. 2015, 6, 5615–5624. [Google Scholar] [CrossRef]
  67. Niu, J.; Lunn, D.J.; Pusuluri, A.; Yoo, J.I.; O’Malley, M.A.; Mitragotri, S.; Soh, H.T.; Hawker, C.J. Engineering live cell surfaces with functional polymers via cytocompatible controlled radical polymerization. Nat. Chem. 2017, 9, 537–545. [Google Scholar] [CrossRef] [PubMed]
  68. Cheng, B.-F.; Wang, L.-H.; You, Y.-Z. Photoinduced electron transfer-reversible addition-fragmentation chain transfer (PET-RAFT) polymerization using titanium dioxide. Macromol. Res. 2016, 24, 811–815. [Google Scholar] [CrossRef]
  69. Tucker, B.S.; Coughlin, M.L.; Figg, C.A.; Sumerlin, B.S. Grafting-from proteins using metal-free PET-RAFT polymerizations under mild visible-light irradiation. ACS Macro Lett. 2017, 6, 452–457. [Google Scholar] [CrossRef]
  70. Bagheri, A.; Yeow, J.; Arandiyan, H.; Xu, J.; Boyer, C.; Lim, M. Polymerization of a photocleavable monomer using visible light. Macromol. Rapid Commun. 2016, 37, 905–910. [Google Scholar] [CrossRef] [PubMed]
  71. Shanmugam, S.; Xu, J.; Boyer, C. Light-regulated polymerization under Near-infrared/Far-red irradiation catalyzed by bacteriochlorophyll a. Angew. Chem. Int. Ed. 2016, 55, 1036–1040. [Google Scholar] [CrossRef] [PubMed]
  72. Xu, S.; Ng, G.; Xu, J.; Kuchel, R.P.; Yeow, J.; Boyer, C. 2-(methylthio)ethyl methacrylate: A versatile monomer for stimuli responsiveness and polymerization-induced self-assembly in the presence of air. ACS Macro Lett. 2017, 6, 1237–1244. [Google Scholar] [CrossRef]
  73. Fu, Q.; Ruan, Q.; McKenzie, T.G.; Reyhani, A.; Tang, J.; Qiao, G.G. Development of a robust PET-RAFT polymerization using graphitic carbon nitride (g-C3N4). Macromolecules 2017, 50, 7509–7516. [Google Scholar] [CrossRef]
  74. Tan, J.; Liu, D.; Bai, Y.; Huang, C.; Li, X.; He, J.; Xu, Q.; Zhang, L. Enzyme-assisted photoinitiated polymerization-induced self-assembly: An oxygen-tolerant method for preparing block copolymer nano-objects in open vessels and multiwell plates. Macromolecules 2017, 50, 5798–5806. [Google Scholar] [CrossRef]
  75. Rosen, B.M.; Percec, V. Single-electron transfer and single-electron transfer degenerative chain transfer living radical polymerization. Chem. Rev. 2009, 109, 5069–5119. [Google Scholar] [CrossRef] [PubMed]
  76. Percec, V.; Guliashvili, T.; Ladislaw, J.S.; Wistrand, A.; Stjerndahl, A.; Sienkowska, M.J.; Monteiro, M.J.; Sahoo, S. Ultrafast synthesis of ultrahigh molar mass polymers by metal-catalyzed living radical polymerization of acrylates, methacrylates, and vinyl chloride mediated by SET at 25 °C. J. Am. Chem. Soc. 2006, 128, 14156–14165. [Google Scholar] [CrossRef] [PubMed]
  77. Harihara Subramanian, S.; Prakash Babu, R.; Dhamodharan, R. Ambient temperature polymerization of styrene by single electron transfer initiation, followed by reversible addition fragmentation chain transfer control. Macromolecules 2008, 41, 262–265. [Google Scholar] [CrossRef]
  78. Zhang, W.; Zhang, W.; Zhang, Z.; Zhu, J.; Zhu, X. SET-RAFT polymerization of progargyl methacrylate and a one-pot/one-step preparation of side-chain functionalized polymers via combination of SET-RAFT and click chemistry. Macromol. Rapid Commun. 2010, 31, 1354–1358. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, Q.; Zhang, Z.; Wang, W.; Zhu, J.; Cheng, Z.; Zhou, N.; Zhang, W.; Zhu, X. SET-RAFT of MMA mediated by ascorbic acid-activated copper oxide. J. Polym. Sci. Part A 2012, 50, 1424–1433. [Google Scholar] [CrossRef]
  80. Zhang, Z.; Wang, W.; Cheng, Z.; Zhu, J.; Zhou, N.; Yang, Y.; Tu, Y.; Zhu, X. Zero-valent iron/RAFT agent-mediated polymerization of methyl methacrylate at ambient temperature. Macromolecules 2010, 43, 7979–7984. [Google Scholar] [CrossRef]
  81. Gu, Y.; Zhao, J.; Liu, Q.; Zhou, N.; Zhang, Z.; Zhu, X. Zero-valent iron (Fe(0)) mediated RAFT miniemulsion polymerization: A facile approach for the fabrication of Fe(0)-encapsulated polymeric nanoparticles. Polym. Chem. 2014, 5, 4215–4218. [Google Scholar] [CrossRef]
  82. Gu, Y.; Zhao, J.; Liu, Q.; Pan, X.; Zhang, W.; Zhang, Z.; Zhu, X. Zero valent metal/RAFT agent mediated CRP of functional monomers at room temperature: A promising catalyst system for CRP. Polym. Chem. 2015, 6, 359–363. [Google Scholar] [CrossRef]
  83. Li, G.; Zheng, H.; Bai, R. A facile strategy for the preparation of azide polymers via room temperature RAFT polymerization by redox initiation. Macromol. Rapid Commun. 2009, 30, 442–447. [Google Scholar] [CrossRef] [PubMed]
  84. Yang, H.; Xue, S.; Pan, J.; Gong, F.; Pu, H. Preparation of core-shell attapulgite particles by redox-initiated surface reversible addition-fragmentation chain transfer polymerization via a “graft from” approach. RSC Adv. 2016, 6, 14120–14127. [Google Scholar] [CrossRef]
  85. Liu, G.; Qiu, Q.; Shen, W.; An, Z. Aqueous dispersion polymerization of 2-methoxyethyl acrylate for the synthesis of biocompatible nanoparticles using a hydrophilic RAFT polymer and a redox initiator. Macromolecules 2011, 44, 5237–5245. [Google Scholar] [CrossRef]
  86. Martin, L.; Gody, G.; Perrier, S. Preparation of complex multiblock copolymers via aqueous RAFT polymerization at room temperature. Polym. Chem. 2015, 6, 4875–4886. [Google Scholar] [CrossRef]
  87. Bai, W.; Zhang, L.; Bai, R.; Zhang, G. A very useful redox initiator for aqueous RAFT polymerization of N-isopropylacrylamide and acrylamide at room temperature. Macromol. Rapid Commun. 2008, 29, 562–566. [Google Scholar] [CrossRef]
  88. Zheng, H.; Bai, W.; Hu, K.; Bai, R.; Pan, C. Facile room temperature raft polymerization via redox initiation. J. Polym. Sci. Part A 2008, 46, 2575–2580. [Google Scholar] [CrossRef]
  89. Dao, V.H.; Cameron, N.R.; Saito, K. Synthesis of ultra-high molecular weight aba triblock copolymers via aqueous RAFT-mediated gel polymerisation, end group modification and chain coupling. Polym. Chem. 2017, 8, 6834–6843. [Google Scholar] [CrossRef]
  90. Reyhani, A.; McKenzie, T.G.; Ranji-Burachaloo, H.; Fu, Q.; Qiao, G.G. Fenton-RAFT polymerization: An “on-demand” chain-growth method. Chem. Eur. J. 2017, 23, 7221–7226. [Google Scholar] [CrossRef] [PubMed]
  91. Varma, I.K.; Albertsson, A.-C.; Rajkhowa, R.; Srivastava, R.K. Enzyme catalyzed synthesis of polyesters. Prog. Polym. Sci. 2005, 30, 949–981. [Google Scholar] [CrossRef]
  92. Huang, Z.; Zhang, X.; Zhang, X.; Fu, C.; Wang, K.; Yuan, J.; Tao, L.; Wei, Y. Amphiphilic fluorescent copolymers via one-pot combination of chemoenzymatic transesterification and RAFT polymerization: Synthesis, self-assembly and cell imaging. Polym. Chem. 2015, 6, 607–612. [Google Scholar] [CrossRef]
  93. Ng, Y.H.; di Lena, F.; Chai, C.L. PolyPEGA with predetermined molecular weights from enzyme-mediated radical polymerization in water. Chem. Commun. 2011, 47, 6464–6466. [Google Scholar] [CrossRef] [PubMed]
  94. Silva, T.B.; Spulber, M.; Kocik, M.K.; Seidi, F.; Charan, H.; Rother, M.; Sigg, S.J.; Renggli, K.; Kali, G.; Bruns, N. Hemoglobin and red blood cells catalyze atom transfer radical polymerization. Biomacromolecules 2013, 14, 2703–2712. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, B.; Wang, X.; Zhu, A.; Ma, K.; Lv, Y.; Wang, X.; An, Z. Enzyme-initiated reversible addition-fragmentation chain transfer polymerization. Macromolecules 2015, 48, 7792–7802. [Google Scholar] [CrossRef]
  96. Danielson, A.P.; Bailey-Van Kuren, D.; Lucius, M.E.; Makaroff, K.; Williams, C.; Page, R.C.; Berberich, J.A.; Konkolewicz, D. Well-defined macromolecules using horseradish peroxidase as a RAFT initiase. Macromol. Rapid Commun. 2016, 37, 362–367. [Google Scholar] [CrossRef] [PubMed]
  97. Chapman, R.; Gormley, A.J.; Herpoldt, K.-L.; Stevens, M.M. Highly controlled open vessel raft polymerizations by enzyme degassing. Macromolecules 2014, 47, 8541–8547. [Google Scholar] [CrossRef]
  98. Shenoy, R.; Tibbitt, M.W.; Anseth, K.S.; Bowman, C.N. Formation of core-shell particles by interfacial radical polymerization initiated by a glucose oxidase-mediated redox system. Chem. Mater. 2013, 25, 761–767. [Google Scholar] [CrossRef] [PubMed]
  99. Lv, Y.; Liu, Z.; Zhu, A.; An, Z. Glucose oxidase deoxygenation-redox initiation for RAFT polymerization in air. J. Polym. Sci. Part A 2017, 55, 164–174. [Google Scholar] [CrossRef]
  100. Liu, Z.; Lv, Y.; An, Z. Enzymatic cascade catalysis for the synthesis of multiblock and ultrahigh-molecular-weight polymers with oxygen tolerance. Angew. Chem. Int. Ed. 2017, 56, 13852–13856. [Google Scholar] [CrossRef] [PubMed]
  101. Kumagai, S.; Nagai, K.; Satoh, K.; Kamigaito, M. In-situ direct mechanistic transformation from raft to living cationic polymerization for (meth)acrylate-vinyl ether block copolymers. Macromolecules 2010, 43, 7523–7531. [Google Scholar] [CrossRef]
  102. Uchiyama, M.; Satoh, K.; Kamigaito, M. Cationic raft polymerization using ppm concentrations of organic acid. Angew. Chem. Int. Ed. 2015, 54, 1924–1928. [Google Scholar] [CrossRef] [PubMed]
  103. Aoshima, H.; Uchiyama, M.; Satoh, K.; Kamigaito, M. Interconvertible living radical and cationic polymerization through reversible activation of dormant species with dual activity. Angew. Chem. Int. Ed. 2014, 53, 10932–10936. [Google Scholar] [CrossRef] [PubMed]
  104. Schweitzer-Chaput, B.; Boess, E.; Klussmann, M. Acid-catalyzed activation of peroxyketals: Tunable radical initiation at ambient temperature and below. Org. Lett. 2016, 18, 4944–4947. [Google Scholar] [CrossRef] [PubMed]
  105. Vandenbergh, J.; Schweitzer-Chaput, B.; Klussmann, M.; Junkers, T. Acid-induced room temperature raft polymerization: Synthesis and mechanistic insights. Macromolecules 2016, 49, 4124–4135. [Google Scholar] [CrossRef]
  106. Maximiano, P.; Mendonça, P.V.; Costa, J.R.C.; Haworth, N.L.; Serra, A.C.; Guliashvili, T.; Coote, M.L.; Coelho, J.F.J. Ambient temperature transition-metal-free dissociative electron transfer reversible addition-fragmentation chain transfer polymerization (DET-RAFT) of methacrylates, acrylates, and styrene. Macromolecules 2016, 49, 1597–1604. [Google Scholar] [CrossRef]
  107. Hua, D.; Tang, J.; Jiang, J.; Gu, Z.; Dai, L.; Zhu, X. Controlled grafting modification of silica gel via RAFT polymerization under ultrasonic irradiation. Mater. Chem. Phys. 2009, 114, 402–406. [Google Scholar] [CrossRef]
  108. McKenzie, T.G.; Colombo, E.; Fu, Q.; Ashokkumar, M.; Qiao, G.G. Sono-RAFT polymerization in aqueous medium. Angew. Chem. Int. Ed. 2017, 56, 12302–12306. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, Y.; Fantin, M.; Park, S.; Gottlieb, E.; Fu, L.; Matyjaszewski, K. Electrochemically mediated reversible addition-fragmentation chain-transfer polymerization. Macromolecules 2017, 50, 7872–7879. [Google Scholar] [CrossRef]
  110. Willcock, H.; O’Reilly, R.K. End group removal and modification of RAFT polymers. Polym. Chem. 2010, 1, 149–157. [Google Scholar] [CrossRef]
  111. Zetterlund, P.B.; Yamazoe, H.; Yamada, B. Addition-fragmentation chain transfer involving dimers of α-methylvinyl monomers studied by ESR spectroscopy: Competition between fragmentation and bimolecular termination. Macromol. Rapid Commun. 2003, 24, 197–201. [Google Scholar] [CrossRef]
  112. Liu, J.; Stansbury, J.W. RAFT-mediated control of nanogel structure and reactivity: Chemical, physical and mechanical properties of monomer-dispersed nanogel compositions. Dent. Mater. 2014, 30, 1252–1262. [Google Scholar] [CrossRef] [PubMed]
  113. Perkowski, A.J.; You, W.; Nicewicz, D.A. Visible light photoinitiated metal-free living cationic polymerization of 4-methoxystyrene. J. Am. Chem. Soc. 2015, 137, 7580–7583. [Google Scholar] [CrossRef] [PubMed]
  114. Engelis, N.G.; Anastasaki, A.; Nurumbetov, G.; Truong, N.P.; Nikolaou, V.; Shegiwal, A.; Whittaker, M.R.; Davis, T.P.; Haddleton, D.M. Sequence-controlled methacrylic multiblock copolymers via sulfur-free RAFT emulsion polymerization. Nat. Chem. 2017, 9, 171–178. [Google Scholar] [CrossRef] [PubMed]
  115. Nurumbetov, G.; Engelis, N.; Godfrey, J.; Hand, R.; Anastasaki, A.; Simula, A.; Nikolaou, V.; Haddleton, D.M. Methacrylic block copolymers by sulfur free RAFT (SF RAFT) free radical emulsion polymerisation. Polym. Chem. 2017, 8, 1084–1094. [Google Scholar] [CrossRef]
  116. Haddleton, D.M.; Maloney, D.R.; Suddaby Adam Clarke, K.G.; Richards, S.N. Radical-addition-fragmentation and co-polymerization of methyl methacrylate macromonomers from catalytic chain transfer polymerization (CCTP). Polymer 1997, 38, 6207–6217. [Google Scholar] [CrossRef]
  117. Krstina, J.; Moad, G.; Rizzardo, E.; Winzor, C.L.; Berge, C.T.; Fryd, M. Narrow polydispersity block copolymers by free-radical polymerization in the presence of macromonomers. Macromolecules 1995, 28, 5381–5385. [Google Scholar] [CrossRef]
  118. Moad, C.L.; Moad, G.; Rizzardo, E.; Thang, S.H. Chain transfer activity of ω-unsaturated methyl methacrylate oligomers. Macromolecules 1996, 29, 7717–7726. [Google Scholar] [CrossRef]
  119. Shirakawa, H.; Louis, E.J.; MacDiarmid, A.G.; Chiang, C.K.; Heeger, A.J. Synthesis of electrically conducting organic polymers: Halogen derivatives of polyacetylene, (CH)x. J. Chem. Soc. Chem. Commun. 1977, 578–580. [Google Scholar] [CrossRef]
  120. D’Hooge, D.R.; Van Steenberge, P.H.M.; Derboven, P.; Reyniers, M.-F.; Marin, G.B. Model-based design of the polymer microstructure: Bridging the gap between polymer chemistry and engineering. Polym. Chem. 2015, 6, 7081–7096. [Google Scholar] [CrossRef]
  121. D’hooge, D.R.; Van Steenberge, P.H.M.; Reyniers, M.-F.; Marin, G.B. The strength of multi-scale modeling to unveil the complexity of radical polymerization. Prog. Polym. Sci. 2016, 58, 59–89. [Google Scholar] [CrossRef]
  122. Van Steenberge, P.H.M.; Vandenbergh, J.; D’hooge, D.R.; Reyniers, M.-F.; Adriaensens, P.J.; Lutsen, L.; Vanderzande, D.J.M.; Marin, G.B. Kinetic monte carlo modeling of the sulfinyl precursor route for poly(p-phenylene vinylene) synthesis. Macromolecules 2011, 44, 8716–8726. [Google Scholar] [CrossRef]
  123. Van Steenberge, P.H.M.; D’Hooge, D.R.; Vandenbergh, J.; Reyniers, M.-F.; Adriaensens, P.J.; Vanderzande, D.J.M.; Marin, G.B. Comparative kinetic monte carlo study of the sulfinyl and dithiocarbamate precursor route toward highly regioregular MDMO-PPV. Macromol. Theory Simul. 2013, 22, 246–255. [Google Scholar] [CrossRef]
  124. Zaquen, N.; Van Steenberge, P.H.M.; D’hooge, D.R.; Reyniers, M.-F.; Marin, G.B.; Vandenbergh, J.; Lutsen, L.; Vanderzande, D.J.M.; Junkers, T. Improved mechanistic insights into radical sulfinyl precursor MDMO-PPV synthesis by combining microflow technology and computer simulations. Macromolecules 2015, 48, 8294–8306. [Google Scholar] [CrossRef]
  125. Moad, G.; Chen, M.; Häussler, M.; Postma, A.; Rizzardo, E.; Thang, S.H. Functional polymers for optoelectronic applications by RAFT polymerization. Polym. Chem. 2011, 2, 492–519. [Google Scholar] [CrossRef]
  126. Yen, H.-J.; Liou, G.-S. Solution-processable triarylamine-based high-performance polymers for resistive switching memory devices. Polym. J. 2016, 48, 117–138. [Google Scholar] [CrossRef]
  127. Huang, C.-W.; Wu, P.-W.; Su, W.-H.; Zhu, C.-Y.; Kuo, S.-W. Stimuli-responsive supramolecular materials: Photo-tunable properties and molecular recognition behavior. Polym. Chem. 2016, 7, 795–806. [Google Scholar] [CrossRef]
  128. Hohlein, I.M.; Werz, P.D.; Veinot, J.G.; Rieger, B. Photoluminescent silicon nanocrystal-polymer hybrid materials via surface initiated reversible addition-fragmentation chain transfer (RAFT) polymerization. Nanoscale 2015, 7, 7811–7818. [Google Scholar] [CrossRef] [PubMed]
  129. Wan, Q.; Jiang, R.; Mao, L.; Xu, D.; Zeng, G.; Shi, Y.; Deng, F.; Liu, M.; Zhang, X.; Wei, Y. A powerful “one-pot” tool for fabrication of AIE-active luminescent organic nanoparticles through the combination of RAFT polymerization and multicomponent reactions. Mater. Chem. Front. 2017, 1, 1051–1058. [Google Scholar] [CrossRef]
  130. Zeng, G.; Liu, M.; Jiang, R.; Huang, Q.; Huang, L.; Wan, Q.; Dai, Y.; Wen, Y.; Zhang, X.; Wei, Y. Self-catalyzed photo-initiated raft polymerization for fabrication of fluorescent polymeric nanoparticles with aggregation-induced emission feature. Mater. Sci. Eng. C 2018, 83, 154–159. [Google Scholar] [CrossRef] [PubMed]
  131. Huang, Z.; Zhang, X.; Zhang, X.; Wang, S.; Yang, B.; Wang, K.; Yuan, J.; Tao, L.; Wei, Y. Synthesis of amphiphilic fluorescent copolymers with smart pH sensitivity via RAFT polymerization and their application in cell imaging. Polym. Bull. 2017, 74, 4525–4536. [Google Scholar] [CrossRef]
  132. Fang, F.; Kong, F.; Zhang, X.; Lin, M.; Zhang, Y.; Ding, S. Enhanced emission from Alq3 in amphiphilic block copolymers PEG-b-PVK-co-Alq3 synthesized via RAFT polymerization. J. Appl. Polym. Sci. 2017, 134, 44573. [Google Scholar] [CrossRef]
  133. Chen, T.; Xu, Y.; Peng, Z.; Li, A.; Liu, J. Simultaneous utilization of a bifunctional ruthenium complex as an efficient catalyst for raft controlled photopolymerization and a sensing probe for the facile fabrication of an ECL platform. Polym. Chem. 2016, 7, 5880–5887. [Google Scholar] [CrossRef]
  134. Mu, B.; Hao, X.; Chen, J.; Li, Q.; Zhang, C.; Chen, D. Discotic columnar liquid-crystalline polymer semiconducting materials with high charge-carrier mobility via rational macromolecular engineering. Polym. Chem. 2017, 8, 3286–3293. [Google Scholar] [CrossRef]
  135. Wang, Y.; Huang, X.; Li, T.; Wang, Z.; Li, L.; Guo, X.; Jiang, P. Novel crosslinkable high-k copolymer dielectrics for high-energy-density capacitors and organic field-effect transistor applications. J. Mater. Chem. A 2017, 5, 20737–20746. [Google Scholar] [CrossRef]
  136. Yang, K.; Huang, X.; Xie, L.; Wu, C.; Jiang, P.; Tanaka, T. Core-shell structured polystyrene/BaTiO3 hybrid nanodielectrics prepared by in situ raft polymerization: A route to high dielectric constant and low loss materials with weak frequency dependence. Macromol. Rapid Commun. 2012, 33, 1921–1926. [Google Scholar] [CrossRef] [PubMed]
  137. Lo, C.T.; Watanabe, Y.; Oya, H.; Nakabayashi, K.; Mori, H.; Chen, W.C. Non-volatile transistor memory devices using charge storage cross-linked core-shell nanoparticles. Chem. Commun. 2016, 52, 7269–7272. [Google Scholar] [CrossRef] [PubMed]
  138. Zhang, B.; Chen, Y.; Zhuang, X.; Liu, G.; Yu, B.; Kang, E.-T.; Zhu, J.; Li, Y. Poly(N-vinylcarbazole) chemically modified graphene oxide. J. Polym. Sci. Part A 2010, 48, 2642–2649. [Google Scholar] [CrossRef]
  139. Zhang, B.; Chen, Y.; Xu, L.; Zeng, L.; He, Y.; Kang, E.-T.; Zhang, J. Growing poly(N-vinylcarbazole) from the surface of graphene oxide via raft polymerization. J. Polym. Sci. Part A 2011, 49, 2043–2050. [Google Scholar] [CrossRef]
  140. Selhorst, R.C.; Puodziukynaite, E.; Dewey, J.A.; Wang, P.; Barnes, M.D.; Ramasubramaniam, A.; Emrick, T. Tetrathiafulvalene-containing polymers for simultaneous non-covalent modification and electronic modulation of MoS2 nanomaterials. Chem. Sci. 2016, 7, 4698–4705. [Google Scholar] [CrossRef]
  141. Cheng, H.; Dong, N.; Bai, T.; Song, Y.; Wang, J.; Qin, Y.; Zhang, B.; Chen, Y. Covalent modification of MoS2 with poly(N-vinylcarbazole) for solid-state broadband optical limiters. Chem. Eur. J. 2016, 22, 4500–4507. [Google Scholar] [CrossRef] [PubMed]
  142. Fan, F.; Zhang, B.; Cao, Y.; Chen, Y. Solution-processable poly(N-vinylcarbazole)-covalently grafted MoS2 nanosheets for nonvolatile rewritable memory devices. Nanoscale 2017, 9, 2449–2456. [Google Scholar] [CrossRef] [PubMed]
  143. Shi, M.; Dong, N.; He, N.; Wan, Y.; Cheng, H.; Han, M.; Wang, J.; Chen, Y. MoS2 nanosheets covalently functionalized with polyacrylonitrile: Synthesis and broadband laser protection performance. J. Mater. Chem. C 2017, 5, 11920–11926. [Google Scholar] [CrossRef]
  144. Fan, F.; Zhang, B.; Song, S.; Liu, B.; Cao, Y.; Chen, Y. Pyrolytically modified polyacrylonitrile-covalently grafted MoS2 nanosheets for a nonvolatile rewritable memory device. Adv. Electron. Mater. 2018, 4, 1700397. [Google Scholar] [CrossRef]
Scheme 1. Mechanism of reversible addition-fragmentation chain transfer (RAFT) polymerization.
Scheme 1. Mechanism of reversible addition-fragmentation chain transfer (RAFT) polymerization.
Polymers 10 00318 sch001
Scheme 2. Typical RAFT agent structures and different types of RAFT agents.
Scheme 2. Typical RAFT agent structures and different types of RAFT agents.
Polymers 10 00318 sch002
Figure 1. Proposed mechanism for photo-induced initiation RAFT. (a) Photodissociation RAFT process; (b) photoinduced electron (PET)-RAFT process. Note: PC, photoredox catalyst. Reprinted with permission from [47]. Copyright 2015 American Chemical Society.
Figure 1. Proposed mechanism for photo-induced initiation RAFT. (a) Photodissociation RAFT process; (b) photoinduced electron (PET)-RAFT process. Note: PC, photoredox catalyst. Reprinted with permission from [47]. Copyright 2015 American Chemical Society.
Polymers 10 00318 g001
Figure 2. Proposed mechanism (a) and RAFT agents (b) for single electron transfer (SET)-RAFT. Reprinted with permission from [77,78,79]. Copyright 2008 American Chemical Society.
Figure 2. Proposed mechanism (a) and RAFT agents (b) for single electron transfer (SET)-RAFT. Reprinted with permission from [77,78,79]. Copyright 2008 American Chemical Society.
Polymers 10 00318 g002
Figure 3. Proposed hybrid mechanism for RAFT polymerization of glycidyl methacrylate through Fe(0) catalysis initiation. Reprinted with permission from [82]. Copyright 2014 Royal Society of Chemistry.
Figure 3. Proposed hybrid mechanism for RAFT polymerization of glycidyl methacrylate through Fe(0) catalysis initiation. Reprinted with permission from [82]. Copyright 2014 Royal Society of Chemistry.
Polymers 10 00318 g003
Figure 4. (a) Glucose oxidase (GOx) deoxygenation redox initiation RAFT polymerization cascade technique; (b) polymerization with pyranose oxidase (P2Ox)-horseradish peroxide (HRP) cascade catalysis initiation RAFT polymerization. Adapted from [99], Copyright 2016, with permission from John Wiley & Sons Inc.; and [100], Copyright 2017, with permission from John Wiley & Sons Inc.
Figure 4. (a) Glucose oxidase (GOx) deoxygenation redox initiation RAFT polymerization cascade technique; (b) polymerization with pyranose oxidase (P2Ox)-horseradish peroxide (HRP) cascade catalysis initiation RAFT polymerization. Adapted from [99], Copyright 2016, with permission from John Wiley & Sons Inc.; and [100], Copyright 2017, with permission from John Wiley & Sons Inc.
Polymers 10 00318 g004
Figure 5. (a) Reversibly switching mechanism between cationic and radical polymerization of vinyl ether and acrylate monomers with the proper regulation of radical species (RC) and Lewis acids (MtXn); (b) Mechanism of RAFT polymerization with acid-induced cyclohexanone/tert-butylhydroperoxide initiation system at room temperature. Adapted from [103], Copyright 2014, with permission from John Wiley & Sons Inc.; and [105], Copyright 2016, with permission from American Chemical Society.
Figure 5. (a) Reversibly switching mechanism between cationic and radical polymerization of vinyl ether and acrylate monomers with the proper regulation of radical species (RC) and Lewis acids (MtXn); (b) Mechanism of RAFT polymerization with acid-induced cyclohexanone/tert-butylhydroperoxide initiation system at room temperature. Adapted from [103], Copyright 2014, with permission from John Wiley & Sons Inc.; and [105], Copyright 2016, with permission from American Chemical Society.
Polymers 10 00318 g005
Figure 6. Proposed RAFT mechanism using α-methylstyrene dimer (AMSD) as RAFT agent. Reprinted with permission from [112]. Copyright 2014 Elsevier.
Figure 6. Proposed RAFT mechanism using α-methylstyrene dimer (AMSD) as RAFT agent. Reprinted with permission from [112]. Copyright 2014 Elsevier.
Polymers 10 00318 g006
Figure 7. Proposed mechanism of sulfur-free RAFT process. Reprinted with permission from [115]. Copyright 2016 Royal Society of Chemistry.
Figure 7. Proposed mechanism of sulfur-free RAFT process. Reprinted with permission from [115]. Copyright 2016 Royal Society of Chemistry.
Polymers 10 00318 g007
Figure 8. Synthesis of (a) PVB-DAP; (b) Azo-T; and (c) hydrogen-bond based supramolecular complex; (d) The variations of water contact angles of PVB-DAP/Azo-T hydrogen-bond based supramolecular complexes after several cycles between UV irradiation and being placed in the dark. Reprinted with permission from [127]. Copyright 2015 Royal Society of Chemistry.
Figure 8. Synthesis of (a) PVB-DAP; (b) Azo-T; and (c) hydrogen-bond based supramolecular complex; (d) The variations of water contact angles of PVB-DAP/Azo-T hydrogen-bond based supramolecular complexes after several cycles between UV irradiation and being placed in the dark. Reprinted with permission from [127]. Copyright 2015 Royal Society of Chemistry.
Polymers 10 00318 g008
Figure 9. The preparation of TPE-poly(St-PEGMA) fluorescent polymeric nanoparticles (FPNs) and their cell imaging applications. Reprinted with permission from [130]. Copyright 2018 Elsevier.
Figure 9. The preparation of TPE-poly(St-PEGMA) fluorescent polymeric nanoparticles (FPNs) and their cell imaging applications. Reprinted with permission from [130]. Copyright 2018 Elsevier.
Polymers 10 00318 g009
Figure 10. (a) The fabrication of pyrene-terminated PSS via PET-RAFT polymerization and production of Ru(bpy)32+ adsorbed PSS concurrently; (b) Diagrammatic sketch for the fabrication of Ru/PSS based solid-sate ECL sensor. Reprinted with permission from [133]. Copyright 2016 Royal Society of Chemistry.
Figure 10. (a) The fabrication of pyrene-terminated PSS via PET-RAFT polymerization and production of Ru(bpy)32+ adsorbed PSS concurrently; (b) Diagrammatic sketch for the fabrication of Ru/PSS based solid-sate ECL sensor. Reprinted with permission from [133]. Copyright 2016 Royal Society of Chemistry.
Polymers 10 00318 g010
Figure 11. Well-defined side-chain discotic liquid crystal polymers 4Pm-n synthesized by RAFT polymerization and the spacer length-dependent self-organized superstructures, including columnar plastic phases via butoxy-substituted triphenylene (TP) in pairs stacking modes and columnar liquid crystal mesophases from single TP discotic units. Reprinted with permission from [134]. Copyright 2017 Royal Society of Chemistry.
Figure 11. Well-defined side-chain discotic liquid crystal polymers 4Pm-n synthesized by RAFT polymerization and the spacer length-dependent self-organized superstructures, including columnar plastic phases via butoxy-substituted triphenylene (TP) in pairs stacking modes and columnar liquid crystal mesophases from single TP discotic units. Reprinted with permission from [134]. Copyright 2017 Royal Society of Chemistry.
Polymers 10 00318 g011
Figure 12. (a) Covalent PVK-modified MoS2 nanosheets and (b) Current-voltage (I–V) characteristics of ITO/MoS2-PVK/Au memory device. Reprinted with permission from [144].
Figure 12. (a) Covalent PVK-modified MoS2 nanosheets and (b) Current-voltage (I–V) characteristics of ITO/MoS2-PVK/Au memory device. Reprinted with permission from [144].
Polymers 10 00318 g012
Table 1. RAFT agents and monomers for direct photodissociation RAFT.
Table 1. RAFT agents and monomers for direct photodissociation RAFT.
PhotoiniferterPolymerization
Polymers 10 00318 i001Methyl methacrylate (MMA);
Methyl acrylate (MA);
Styrene [47]
Polymers 10 00318 i002Styrene [48]
Polymers 10 00318 i003Acrylic acid (AA) [50]
Polymers 10 00318 i004Vinyl acetate (VAc) [52]
Polymers 10 00318 i005MA [54]
Polymers 10 00318 i006Butyl methacrylate (BMA) [55,57];
MMA [57];
MA [57]
Polymers 10 00318 i007MA [57]
Polymers 10 00318 i008Benzyl methacrylate (BzMA) [51]
Table 2. RAFT agents, photoredox catalysts and monomers for PET-RAFT.
Table 2. RAFT agents, photoredox catalysts and monomers for PET-RAFT.
RAFT agentPhotoredox CatalystMonomer
Polymers 10 00318 i009 Polymers 10 00318 i010
ZnTPP [63]; Eosin Y [66] Polymers 10 00318 i011
Fluorescein [66]; Pheo A [70];
fac-[Ir(ppy)3] [58]; Ru(bpy)3Cl2 [61]; TiO2 [68];
Bacteriochlorophyll a [71]
MMA [58,63,66,68,71];
N-(2-hydroxypropyl) methacrylamide (HPMA) [58];
Oligo(ethylene glycol) methyl ether acrylate (OEGA) [61];
Oligo(ethyleneglycol) methyl ether methacrylate (OEGMA) [61];
Hydroxylethyl methacrylate (HEMA) [66];
N,N-dimethyl aminoethyl methactylate (DMAEMA) [66];
Pentafluorophenyl methacrylate (PFDMA) [66];
Methacrylic acid (MAA) [66];
Hydroxylpropyl methacrylamide (HPMA) [66];
Glycidyl methacrylate (GMA) [66];
Oligo-ethylene glycol methyl ether methacrylate (OEGMA) [66];
O-nitrobenzyl methacrylate(NBMA) [70]
Polymers 10 00318 i012fac-[Ir(ppy)3] [58];
Ru(bpy)3Cl2 [61];
Polymers 10 00318 i013
ZnTPP [63]; Eosin Y [67]
MA [58,63];
Styrene [58,63];
Oligo(ethyleneglycol) methyl ether acrylate (OEGA) [58,61];
Tert-butyl acrylate (tBA) [58];
N-isopropylacrylamide (NIPAAM) [58,61,63];
Isoprene [58];
N,N-dimethyl acrylamide (DMA) [58,61,63];
N,N’-diethylacrylamide (DEA) [61];
2-(dimethylamino)ethyl acrylate (DMAEA) [63];
Polyethylene glycol (PEG)-based acrylamides [67]
Polymers 10 00318 i014 Polymers 10 00318 i015
ZnTPP [63,70,72]
MMA [63];
N-(2-hydroxylpropyl) methacrylamide (HPMA) [63];
O-nitrobenzyl methacrylate (NBMA) [70];
2-(methylthio) ethyl methacrylate; (MTEMA) [72]
Polymers 10 00318 i016 Polymers 10 00318 i017
ZnTPP [63]
MMA [63]
Polymers 10 00318 i018 Polymers 10 00318 i019
ZnTPP [63];
fac-[Ir(ppy)3] [58]
Styrene [63];
MA [58]
Polymers 10 00318 i02010-phenylphenothiazine (PTH) [59]NIPAAM [59];
DMA [59];
tBA [59];
Ethylene glycol methyl ether acrylate (EGMEA) [59]
Polymers 10 00318 i021fac-[Ir(ppy)3] [58]VAc [58];
Vinyl pivalate (VP) [58];
Dimethyl vinylphosphonate (DVP) [58];
N-vinyl pyrolidinone (NVP) [58]
Polymers 10 00318 i022 Polymers 10 00318 i023
TEA [65]; PMDETA [65];
Polymers 10 00318 i024
g-C3N4 [73];
Me6TREN [65]
MMA [65];
BuMA [65];
Ethyl methacrylate (EMA) [65];
MA [73];
DMA [73]
Polymers 10 00318 i025 Polymers 10 00318 i026
TEA [65]; PMDETA [65]
Polymers 10 00318 i027
g-C3N4 [73];
Me6TREN [65]
MA [65];
Styrene [65];
NIPAAM [65];
MA [73];
DMA [73]
Polymers 10 00318 i028 Polymers 10 00318 i029
TEA [65]; PMDETA [65];
Polymers 10 00318 i030
Me6TREN [65]
OEGMA [65]
Polymers 10 00318 i031 Polymers 10 00318 i032
Eosin Y [69]
DMA [69];
2-hydroxyethyl acrylate (HEA) [69];
Acid functional 2-acrylamido-2-methyl-1-propane-sulfonic acid (AMPS) [69]
NIPAAM [69];
Sodium styrenesulfonate (NaSS) [69];
Poly(ethylene glycol) acrylate (PEGA) [69]
Polymers 10 00318 i033g-C3N4 [73]MA [73];
DMA [73]
Polymers 10 00318 i034g-C3N4 [73]MMA [73]
R: Polymers 10 00318 i035Sodium phenyl-2,4,6-trimethyl-benzoylphosphinate (SPTP) [74]HPMA [74]
Table 3. RAFT agents, redox initiator and monomers for RAFT via redox initiation.
Table 3. RAFT agents, redox initiator and monomers for RAFT via redox initiation.
RAFT AgentRedox InitiatorMonomer
Polymers 10 00318 i036BPO/DMA [83,88]4-azidophenyl methacrylate (APM) [83];
MA [83,88];
MMA [83,88];
Styrene [83,88]
Polymers 10 00318 i037BPO/DMA [83,88];
BPO/ATP-ATA [84]
4-azidophenyl methacrylate (APM) [83];
MA [83,88];
MMA [83,84,88];
Styrene [83,88]
Polymers 10 00318 i038KPS/NaAs [85]2-methoxyethyl acrylate (MEA) [85];
Poly(ethylene glycol) methyl ether methacrylate (PEGDA) [85]
Polymers 10 00318 i039Tert-butyl hydroperoxide (TBHP)/ascorbic acid (AsAc) [86];
TBHP/Bruggolite® FF7
(Brug 7) [86]
4-acryloylmorpholine;
HEA;
DEA;
EGMEA;
AA;
MA;
NIPAAM;
PEGA;
DMA;
2-carboxyethylacrylate (CEA) [86]
Polymers 10 00318 i040K2S2O8/Na2S2O3 [87];
H2O2/Fe2+ [90]
NIPAAM [87];
AM [87];
DMA [90];
NAM [90];
HEA [90]
Polymers 10 00318 i041APS/SFS [89]AA [89];
AM [89]
Table 4. RAFT agents, initiator and monomers for RAFT polymerization.
Table 4. RAFT agents, initiator and monomers for RAFT polymerization.
RAFT AgentInitiatorMonomer
Polymers 10 00318 i042Na2S2O4 [106];
ultrasonic irradiation [107];
electronreduction of BPO or BrPhN2+ [109]
Styrene [106];
MA [106,107,109];
MMA [109];
BA [109]
Polymers 10 00318 i043Na2S2O4 [106];
electronreduction of BPO or BrPhN2+ [109]
MMA [106];
GMA [106];
2-(diisopropylamino) ethyl methacrylate (DPA) [106];
DMAEMA [106];
MMA [109];
BA [109]
Polymers 10 00318 i044ultrasonic irradiation [108]HEA;
OEGA;
DMA;
N-acryloylmorpholine (NAM);
OEGMA [108]

Share and Cite

MDPI and ACS Style

Tian, X.; Ding, J.; Zhang, B.; Qiu, F.; Zhuang, X.; Chen, Y. Recent Advances in RAFT Polymerization: Novel Initiation Mechanisms and Optoelectronic Applications. Polymers 2018, 10, 318. https://doi.org/10.3390/polym10030318

AMA Style

Tian X, Ding J, Zhang B, Qiu F, Zhuang X, Chen Y. Recent Advances in RAFT Polymerization: Novel Initiation Mechanisms and Optoelectronic Applications. Polymers. 2018; 10(3):318. https://doi.org/10.3390/polym10030318

Chicago/Turabian Style

Tian, Xiangyu, Junjie Ding, Bin Zhang, Feng Qiu, Xiaodong Zhuang, and Yu Chen. 2018. "Recent Advances in RAFT Polymerization: Novel Initiation Mechanisms and Optoelectronic Applications" Polymers 10, no. 3: 318. https://doi.org/10.3390/polym10030318

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop