Photoinduced Electron Transfer Reactions for Macromolecular Syntheses

Size: px
Start display at page:

Download "Photoinduced Electron Transfer Reactions for Macromolecular Syntheses"

Transcription

1 This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes. pubs.acs.org/cr Photoinduced Electron Transfer Reactions for Macromolecular Syntheses Sajjad Dadashi-Silab, Sean Doran, and Yusuf Yagci*,, Department of Chemistry, Istanbul Technical University, Maslak, Istanbul, Turkey Center of Excellence for Advanced Materials Research (CEAMR) and Department of Chemistry, King Abdulaziz University, Jeddah, Saudi Arabia ABSTRACT: Photochemical reactions, particularly those involving photoinduced electron transfer processes, establish a substantial contribution to the modern synthetic chemistry, and the polymer community has been increasingly interested in exploiting and developing novel photochemical strategies. These reactions are efficiently utilized in almost every aspect of macromolecular architecture synthesis, involving initiation, control of the reaction kinetics and molecular structures, functionalization, and decoration, etc. Merging with polymerization techniques, photochemistry has opened up new intriguing and powerful avenues for macromolecular synthesis. Construction of various polymers with incredibly complex structures and specific control over the chain topology, as well as providing the opportunity to manipulate the reaction course through spatiotemporal control, are one of the unique abilities of such photochemical reactions. This review paper provides a comprehensive account of the fundamentals and applications of photoinduced electron transfer reactions in polymer synthesis. Besides traditional photopolymerization methods, namely free radical and cationic polymerizations, step-growth polymerizations involving electron transfer processes are included. In addition, controlled radical polymerization and Click Chemistry methods have significantly evolved over the last few decades allowing access to narrow molecular weight distributions, efficient regulation of the molecular weight and the monomer sequence and incredibly complex architectures, and polymer modifications and surface patterning are covered. Potential applications including synthesis of block and graft copolymers, polymer-metal nanocomposites, various hybrid materials and bioconjugates, and sequence defined polymers through photoinduced electron transfer reactions are also investigated in detail. CONTENTS 1. Introduction Photochemistry and Merging with Polymerization Scope of the Photoinitiation of Chain Polymerization by Photoinduced Electron Transfer Reactions Free Radical Polymerization Dye/Co-Initiator System Type II Photoinitiation Semiconducting Organic/Inorganic Compounds Photoinduced Electron Transfer Reaction of Onium Salts Cationic Polymerization Direct Initiation Free Radical Promoted Cationic Polymerization Cationic Polymerization by Electron Transfer with Singlet and Triplet Excited States Cationic Polymerization by Charge Transfer Complexes Cationic Polymerization by Addition Fragmentation Agents Photoinduced Living Cationic Polymerization Step-Growth Polymerization by Photoinduced Electron Transfer Reactions Polymer-Metal Nanocomposites by Photoinduced Electron Transfer Reactions Block and Graft Copolymers by Photoinduced Electron Transfer Reactions Controlled/Living Radical Polymerizations by Photoinduced Electron Transfer Reactions Atom Transfer Radical Polymerization Mechanistic Explanation Monomer Compatibility Ligands Complex Architectures and Sequence- Controlled Polymers Spatiotemporal Control Reversible Addition Fragmentation Chain Transfer (RAFT) Polymerization by Photoinduced Electron Transfer Reactions Click Chemistry by Photoinduced Electron Transfer Reactions Special Issue: Photochemistry in Organic Synthesis Received: October 5, 2015 Published: January 8, American Chemical Society 10212

2 Chemical s 7.1. Cycloaddition Click Reactions Thiol Ene/Yne Click Chemistry Conclusion and Future Perspective Author Information Corresponding Author Notes Biographies References Note Added after ASAP Publication INTRODUCTION 1.1. Photochemistry and Merging with Polymerization Man has always taken inspiration from nature during the course of innovation. Almost without exception, the great biosphere of Earth is powered by the energy from the sun, transmitted through light radiation, and harnessed in plants and other light harvesting organisms to drive forward the chemical transformations that make life possible. And so chemists in the modern age of science have researched to explore and advance the applicability of light as a reactant in synthetic organic chemistry and its multitude of related fields. The efficacy of light in affecting chemical transformations is due to electrons lying in molecular orbitals within molecules entering into excited states upon the absorption of light photons. The distribution of electrons across molecular orbitals is very different in the excited state compared with the ground state. A molecule entering into an electronically excited state can greatly change its chemical properties and reactivity. This can open avenues of reactivity for reactants that would not have been possible without electronic excitation by light. The fundamentals of this behavior have been described according to groundstate and excited-state potential-energy hypersurface topology. 1 Photochemistry enables a powerful toolkit for a wide array of chemical transformations and is indeed unique in its distinctive ability to fulfill the energetic requirements for conducting such transformations that would be otherwise inaccessible with thermal counterparts. Typically, absorption of a mole of photons of UV light provides as 130 times greater energy as compared to ambient conditions. This huge amount of energy input provided by light would overcome energetic barriers for many chemical reactions to take place. And the fact that this energy for driving chemical reactions forward can be readily available from light, and ideally from inexhaustible sun light, holds great promise for sustainable development and a greener future. Thus, the green nature of photochemistry has been the fundamental inspiration for chemists in exploiting advantages of ubiquitous light and developing photochemical approaches. 2 4 Photochemistry can be employed using photoinitiators to initiate chain processes such as radical or cationic polymerization reactions. The utilization of photochemistry, however, is not limited only to the photoinitiation of chain growth reactions. It is used in a broad sense of applications concerning initiation and control of polymerization, functionalization, decoration, and degradation of polymers and various application aspects. Photocuring and surface coatings have become common practice in many commercial and industrial applications. Biological applications of photopolymers have emerged in dental fillings, artificial bone generation, tissue engineering, and drug delivery arrays. Two-photon polymerization and spatially controlled polymerizations initiated by photochemical means allow fabrication of fine three-dimensional (3D) objects and patterned macromolecular structures. Controlled/living polymerizations together with other advanced chemical tools such as click chemistry have been integrated with photochemistry, opening up new avenues in the synthetic polymer community. Tailor-made polymers with precise structural design have been available in a wide range of topological and morphological varieties, including block, graft, star, gradient, and periodic copolymers, molecular brushes, various hybrid materials and bioconjugates, complex structures, and sequence-defined polymers. All these examples are just but a portion of great advancements and opportunities enabled by photoinitiated polymerization techniques. Considerable efforts have been dedicated toward designing and developing photoinitiation systems suitable for efficient photoinitiation of polymerizations. There are many types of chemically and optically different photoinitiators that can promote polymerization by forming initiating species according to their optical properties. The development has gone through using UV-based high-energy photoinitiating systems to sophisticated chemical and technological advancements allowing the use of very low-energy light sources such as light emitting diodes (LEDs). The polymer community has benefited a lot from advances in the related fields of organic synthesis and visible light photoredox as well as from emerging advanced technologies that enable polymer chemists to tailor novel macromolecular architectures Scope of the In this contribution, we intend to review the fundamentals and applications of photoinduced electron transfer reactions for polymer synthesis. Synthesis of polymers by free radical, cationic, step-growth, and increasingly emerging controlled/ living radical polymerizations enabled by photoinduced electron transfer reactions are addressed in detail. A special emphasis will be made in drawing a detailed mechanistic explanation for photochemical events bringing about polymers. Furthermore, specific applications such as block and graft copolymer formation, metal polymer nanocomposites, etc. will be explained as well. In this connection, we will focus our attention to discussing all aspects of photopolymerization reactions concerning electron transfer processes. As such, some systems operating in different modes of photoactivation other than electron transfer will not be involved in this contribution. Readers who are interested in other areas of photopolymerizations are directed to refer to previously published review articles and book chapters in this area PHOTOINITIATION OF CHAIN POLYMERIZATION BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS Photoinduced chain growth polymerization techniques constitute a considerable contribution to the synthesis of various polymer structures. Free radical and cationic polymerizations have been extensively applied in this context, whereas photoinduced anionic polymerization was limited in scope and application and only a few works have been reported in photoinitiated anionic polymerization. Efforts have been directed toward understanding the mechanistic explanations of the photochemical processes and designing novel photoinitiating systems to enhance and improve the efficiency of such reactions and expand their scope into new areas. The photoinduced electron transfer reactions bringing about radical 10213

3 Chemical s and cationic species to initiate corresponding polymerizations can be realized utilizing chromophore groups in the presence of co-initiator compounds. Therefore, understanding the interaction of chromophores and co-initiator compounds is important for a successful photopolymerization reaction, and various systems have been developed in this regard Free Radical Polymerization Dye/Co-Initiator System. Dye molecules have been employed as visible light photosensitizer compounds in photoinitiated polymerizations. They are capable of inducing electron transfer reactions in the photoexcited state in conjunction with suitable co-initiators. 13 These electron transfer reactions convert the added co-initiator to active initiating sites. Dyes can be both reducing and oxidizing in the photoexcited state. And as such, they can be quenched in the photoexcited state by co-initiators through reductive or oxidative processes solely via photoinduced electron transfer reactions. In the reductive quenching, an electron donor is used as co-initiator which transfers an electron to the dye molecule to form electron-donor radical cation species and reduced dye molecules. On the other hand, in the oxidation quenching, an electron transfer takes place from the dye to an electron acceptor molecule oxidizing dye to radical cation while reducing the co-initiator compound (Scheme 1). Free radical Scheme 1. Photoinduced Electron Transfer Reactions of Dyes polymerization can be initiated directly by radical cations (or initiating compounds resulting from their subsequent fragmentation) formed in the reductive quenching and in the presence of electron donor co-initiators. The oxidative quenching, on the other hand, is generally suitable for the initiation of cationic polymerizations where dye sensitizers are used to sensitize onium photoinitiators to form cationic species. This approach is described in detail in the following sections concerning photosensitization of onium salts. Examples of photoreducible dyes include xanthene dyes, such as rose bengal, eosine Y, and erythrosin B; acridinium dyes, such as acriflavine; phenazine dyes, such as methylene blue; and thiazene dyes, such as thionine. Several studies have appeared that used these systems to initiate free radical polymerization under visible light irradiation As co-initiator, amines, for example, are commonly used as electron donor co-initiators. And iodonium salts are good examples of electron acceptor co-initiators. Moreover, halogen-containing compounds have also been used as co-initiator which can provide initiating radicals by electron transfer resulting in the C-halogen bond dissociation in an oxidative quenching. 32 Allonas and co-workers reported linking electron acceptor groups to the structure of dye molecules donor acceptor concept with favorable results as compared with unlinked structures. 33 It is well-known that in the presence of aliphatic amine coinitiators, α-amino radicals are generated through photoinduced electron and proton transfer to initiate free radical polymerization. However, in a recent study it was found that using methylene blue as the dye photosensitizer and a sacrificial sterically hindered amine co-initiator, namely N,N-diisopropylethylamine (DIPEA), no such radical formation was detected. 34 In fact, it was suggested the strong basicity and sterically hindered nature of DIPEA allows it to serve as a twoelectron and one proton (2e /1H + ) donor, as opposed to oneelectron, one proton transfer reactions in known photochemical processes. A 2e /1H + transfer resulted in decomposition of the amine to nonradical products while generating colorless leucomethylene blue. No polymerization was initiated in the presence of methylene blue and DIPEA. However, adding an iodonium salt led to reoxidation of leuco-methylene blue regenerating the initial ground state dye and reduction of the iodonium salt to generate initiating phenyl radicals. Therefore, the radical polymerization of vinyl monomers was initiated. The mechanism of 2e /1H + transfer-initiated polymerization is illustrated in Scheme 2. The process involved rapid photoexcitation of methylene blue within seconds of irradiation and formation of charge-transfer exciplex by the DIPEA co-initiator which resulted in decomposition of the amine to nonradical (closed-shell) products and formation of leuco-methylene blue via 2e /1H + transfer. Thus, the energy of light was suggested to get stored in the leuco-methylene blue photoproduct without spontaneous radical formation and initiation of polymerization. Energy release was found to occur via redox reactions with the iodonium salt which proceeded in the dark and took a much longer time interval to complete. Despite their ability to generate initiating species and optical characteristics at higher wavelengths, the applicability of dyes in sensitization of polymerization is limited due to drawbacks concerning their poor stability and storage problems. As a result, dye-sensitized systems as compared to other photoinitiators have not been extensively applied in photopolymerization systems Type II Photoinitiation Two-Component Bimolecular Photoinitiators. According to their optical behavior in the formation of initiating species, photoinitiators are generally subdivided into two general classes of Type I or those capable of forming initiating radicals directly upon bond cleavage on absorption of light and Type II which generate radicals in conjunction with various co-initiators. Formation of initiating radicals in Type II photoinitiation occurs by absorption of light by a photoinitiator (ketone type photoinitiators, for example) which is followed by interaction with the co-initiator compounds resulting in radical generation. 35,36 Benzophenone, 37,38 thioxanthone, ketocoumarin, 42 camphorquinone, anthraquinone, 47,48 and their derivatives are examples of the most widely used Type II photoinitiators. The interaction of the excited photoinitiator with the co-initiator, depending on the nature of the co-initiator, may happen through different pathways, including hydrogen abstraction or electron transfer or an involvement of both electron transfer and hydrogen abstraction processes. Although in some cases the probability of quenching triplet photoinitiator with monomers should not be ruled out, the main pathway of formation of radicals is through the interaction with coinitiators. For instance, amines due to their good reductant 10214

4 Chemical s Scheme 2. Photoinduced 2e /1H + Transfer between Methylene Blue Dye and DIPEA Co-Initiator a a Evidence of photoredox cycle of methylene blue in the presence of amine and iodonium salt co-initiators leading to changes in the color of the solution. Reprinted from ref 34. Copyright 2014 American Chemical Society. properties 49 are commonly used as co-initiators in Type II photoinitiation. It is generally believed that amines interact with the excited state photoinitiator primarily by an electron transfer process forming an ion pair intermediate (or an exciplex) of photoinitiator radical anion and amine radical cation Afterward, as depicted in Scheme 3, there follows a proton transfer, which results in the formation of an initiating radical derived from the amine co-initiator and a ketyl radical of photoinitiator. The latter is inactive toward adding to double bonds and rather tends to couple or terminate initiating radicals; however, radicals derived from the co-initiator induce the chain polymerization process. Similarly, various types of co-initiators have been proven to undergo photoinduced electron/proton transfer reactions with triplet photoinitiators. For example, thiols react with Type II photoinitiators in a similar manner to amines, which includes primarily, an electron transfer followed by a proton transfer to form sulfur-based free radicals capable of adding across double bonds and initiating free radical polymerization (Scheme 4) Sulfur-centered radicals are highly active to add to both allyl and vinyl double bonds exhibiting insensitivity toward oxygen inhibition. Thiol ene photopolymerizations of multifunctional thiols and double bonds proceed in a step growthlike manner mediated by free radicals for the formation of polymer networks. 58,59 The centrality of using thiols in photochemical processes deals with their substantial role in one of the most powerful tools of modern chemistry, that is the photoinduced thiol ene click chemistry, which has been the center of research interest from many diverse areas. 60,61 Alternatively, phosphorus-containing compounds were used as electron/hydrogen donor co-initiators in photopolymerization. 62 A report from the laboratories of Laleveé and co-workers revealed that a hydrogen atom abstraction by triplet photoinitiators was observed with phosphorus compounds bearing a labile hydrogen phosphorus bond. The process was proven to occur without the involvement of any electron transfer reactions as no ion pair intermediates were detected in spectroscopic investigations. Whereas, photoinduced electron transfer of the triplet photoinitiator with those phosphorus compounds without any abstractable hydrogen was evidenced through the formation of phosphorus radical cation and ketone radical anion intermediates. The phosphorus radical cation intermediate was reported to further fragment to yield phosphorus-centered initiating radicals (Scheme 5). One way to manipulate and enhance the photoactivity of Type II photoinitiators is by introducing various functionalities onto the structure of the photoinitiator by which the absorption maxima, extinction coefficients, quantum yields, as well as adaptability in different media can be tuned on demand. For example, water-solubilizing agents have been introduced into the structure of benzophenones or thioxanthones to make them soluble in aqueous medium so that they easily initiate photopolymerization of water-soluble monomers (Chart 1). 52, One-Component Bimolecular Photoinitiators. Additionally, one-component Type II photoinitiators consisting of both the chromophore core and co-initiator components have been synthesized and used in radical polymerizations. Without the need of additional co-initiators, one-component photoinitiators produce initiating radicals through an intramolecular and/or intermolecular photoinduced electron/ hydrogen transfer process between the chromophore and donator parts. Amines, thiols, and etherlike electron/hydrogen donors of low-molecular weight and polymeric structures involving, generally, benzophenone or thioxanthone chromophore groups have been synthesized as one-component photoinitiators. A derivative of benzophenone known as 10215

5 Chemical s Scheme 3. Schematic representation of Type II photoinitiation with Benzophenone as An Example of Ketone-Based Photoinitiator and a Tertiary Amine Co-Initiator Scheme 4. Thiols as Co-Initiator in Type II Photoinitiation Michler s ketone bearing amine functionality, for example, is capable of generating initiating radicals upon irradiation through photoinduced intramolecular/intermolecular electron and proton transfers in a similar manner to two-component systems. Similarly, thiol derivatives of thioxanthone are capable of generating sulfur-centered radicals without the need of additional co-initiators (Scheme 6). 67 Representative examples of one-component Type II photoinitiators are collected in Table Macrophotoinitiators. With the aim of facilitating solubility and compatibility of photoinitiators and overcoming some disadvantages of low-molecular weight photoinitiators, which may include migration of byproducts of the photolysis of photoinitiators in cured films due to high volatility, polymeric photoinitiators comprising photoinitiator species incorporated into a polymeric support have been utilized as well. Various techniques, including polymerization of suitable monomers, postmodification, and functionalization processes have been taken advantage of in producing such macrophotoinitiators. 105 Postmodification of polystyrene as a polymeric backbone by treating with thiosalicylic acid, which led to the thioxanthonation of polystyrene, has been reported to prepare macrophotoinitiators with pendant thioxanthone groups. 106 Additionally, this process can bring about water-soluble polystyrenebased polymeric photoinitiators by sulfonation during the thioxanthonation process in a one-pot manner. 107 An amine coinitiator was necessary for polymerization to occur

6 Chemical s Scheme 5. Phosphorous Compounds as Co-Initiator in Type II Photoinitiated Polymerization (PI: Photoinitiator) Scheme 6. Thiol-Derivative of Thioxanthone as One- Component Type II Photoinitiator Quite recently, our group reported the synthesis of thioxanthone-based polymeric networks as heterogeneous macrophotoinitiators. 108 Using various cross-coupling processes such as Sonogashira Hagihara or Friedel Crafts alkylation techniques (the latter also referred to as knitting ), conjugated microporous networks of thioxanthone with specific surface areas of up to 750 m 2 g 1 were obtained using thioxanthone with different suitable comonomers. Dibromothioxanthone with triethynylbenzene were subjected to the Sonogashira Hagihara coupling, which resulted in the formation of a microporous network of thioxanthone with the pore size of 1.4 nm and microporosity of 500 m 2 g 1. Using the Friedel Crafts method, thioxanthone and benzene (or triphenylmethane) were knitted together. It was found that these macrophotoinitiators being two- or three-dimensional networks had strong absorption characteristics at visible regions which, it was reasoned, was due to the strong π interactions of the highly conjugated nature of the network. Free radical and cationic photopolymerizations were achieved in the presence of different co-initiators under visible or sun light irradiation. Reusability was found for all three types of microporous thioxanthone networks in both free radical and cationic polymerizations. The synthesis methodologies of one-component macrophotoinitiators mainly include step-growth and addition 115 (co)polymerization of co-initiator/photoinitiator containing monomers, functionalization by click chemistry techniques, 116 dendrimerization, and other functionalization methods. Examples of one-component macrophotoinitiators developed thus far are collected in Table 2. Alternatively, functionalization of supramolecular or hyper-branched polymeric structures containing amine or etheric functional groups as hydrogen donating groups 117 with photoinitiator moieties have been used to form one-component macrophotoinitiators As an example, dendritic or hyperbranched supramolecular amines were reacted with an epoxy functional thioxanthone molecule resulting in the formation of dendritic structures with thioxanthone end functional groups exhibiting Chart 1. Representative Examples of Water-Soluble Type II Thioxanthone-Based Photoinitiators 10217

7 Chemical s Table 1. One-Component Type II Photoinitiators Carrying Both the Chromophore and Co-Initiator Functionalities 10218

8 Chemical s Table 1. continued 10219

9 Chemical s Table 1. continued photoinitiation activity much higher than small molecular weight photoinitiating systems Hyper-branched poly- (ethylene imine) or dendritic poly(propyleneimine) could be used for this purpose Semiconducting Organic/Inorganic Compounds. The use of semiconducting materials of both organic and inorganic origins as photocatalysts has long been established in heterogeneous photocatalysis. 135 In line with the concepts of sustainable chemistry, substantial applications of heterogeneous semiconductor materials have been achieved in a broad range of perspectives. Semiconductors are light sensitive materials that undergo excitation followed by the release of charge carriers. 136,137 Such a manner of providing electron/hole species is taken advantage of leading to the triggering of an enormous array of chemical reactions Metal oxide nanoparticles such as titanium dioxide (TiO 2 ), zinc oxide (ZnO), and cadmium sulfur (CdS) among others, sized in the range of nm, are typical examples of inorganic semiconductors. Early attempts to use these compounds in photopolymerization were carried out by using ZnO nanoparticles to sensitize the free radical polymerization process. Both the photoinduced electron and hole species have been reported to contribute to the formation of initiating sites In a related work from the authors laboratory, it has been found that both good reductant and oxidant coinitiators can be used with ZnO nanoparticles. The initiating radicals can be formed via reduction of the oxidant co-initiator by the photoinduced electrons or oxidation of reductant by the holes released upon irradiation of nanoparticles. 144 A diphenyliodonium salt was used as the oxidant co-initiator that reduced to the diphenyliodonium radical, which further underwent decomposition giving phenyl radicals. On the other hand, triethylamine (TEA) as a reductant co-initiator interacted with the holes through the nonbonding electrons of the nitrogen atom, which was followed by a hydrogen atom abstraction from another amine molecule giving rise to the formation of initiating α-amino radicals. Polymerization of a water-soluble monomer, acrylamide, was reported in aqueous media in the presence of oxygen. Without the need of additional co-initiators, interaction of the electrons and holes 10220

10 Chemical s Table 2. Selected One-Component Type II Macrophotoinitiators 10221

11 Chemical s Table 2. continued 10222

12 Chemical s Table 2. continued with water and oxygen molecules brought about initiating radicals revealing their potential applicability for overcoming oxygen inhibition problems. Scheme 7 depicts the mechanism of the formation of initiating free radicals by semiconducting ZnO nanoparticles in the presence of various co-initiators. Of great importance, their heterogeneity makes semiconductor nanoparticles efficient reusable photoinitiators as after the reaction they are easily recycled, refined, and applied in further reactions while preserving their reactivity after several usages. Scheme 7. Photoinitiated Free Radical Polymerization Using Semiconductor Nanoparticles: Interaction of the Photoinduced Electron and Hole Species with Different Co- Initiators Nanoparticles of different composition have also been reported to initiate free radical polymerization in a quite similar mechanism, involving the interaction of the photoinduced electron or hole species with initiating components, including additional co-initiators, solvent, or monomer molecules The optical characteristics of these nanoparticles are generally governed and controlled by adjusting the shape, size, and surface modification of nanoparticles. 151 It has been recently found that magnetic iron oxide nanoparticles functionalized with carboxylic acid efficiently initiated free radical polymerization. Carboxylic acid groups used for stabilizing nanoparticles interacted with the photogenerated holes, resulting in a decarboxylation process to form initiating free radicals. 152 A similar decarboxylation process was also reported using additional carboxylic acid co-initiators in the presence of nonfunctionalized TiO 2 nanoparticles. 147 Similar to inorganic nanoparticles, semiconducting organic materials have been developed as heterogeneous photoorganocatalysts for a wider range of catalytic reactions. For example, mesoporous graphitic carbon nitride (mpg-c 3 N 4 ) composed of carbon and nitrogen is highly photosensitive with a large specific surface area and has been proposed as a promising potential heterogeneous organocatalyst for a diverse range of redox reactions. 153 The use of mpg-c 3 N 4 in photoinitiated polymerization was based on photoinduced electron/hole release by mpg-c 3 N 4 under visible light irradiation. In the presence of an oxidant such as amine, initiating free radicals were generated. 154 Of great importance, the heterogeneous nature of mpg-c 3 N 4 made it possible to reuse it in further reactions by retaining catalytic activity in each cycle. Similar strategy has been undertaken with conjugated microporous polymers as heterogeneous photocatalysts

13 Chemical s Upconverting nanoparticles (UCNPs) are worth mentioning in this connection. These materials are highly photosensitive at near-infrared (NIR) region and upon excitation emit UV or visible light. Being photosensitive at NIR or IR regions, these nanoparticles are of high potential for bioapplication since they can penetrate in living tissues without causing any harmful effects. Beyazit, Haupt, and co-workers were the first to apply UCNPs in localized photoinduced polymerization to the surface modification of the particles making them more hydrophilic suitable for many bioapplications. 156 Emitting light at UV and visible region, the UCNPs upon excitation under 980 nm LED illumination were able to sensitize conventional photoinitiating systems to initiate the polymerization process. Bezophenone/TEA or eosin Y/TEA combinations were used as UV or visible light photoinitiating components, respectively. Polymerization of a series of watersoluble monomers was successfully initiated this way leading to the formation of a polymeric hydrophilic shell around the nanoparticles, and the use of functional monomers enabled further modification by attaching biocompatible functional groups (Scheme 8). are used so as to oxidize terminating ketyl radicals while generating new initiating radical species Neckers and co-workers studied photopolymerization of bithiophenefunctionalized gold (or silver) nanoparticles, which was based on an intramolecular electron transfer between thiophene and onium salt leading to bithiophene radical cation species. In the absence of a vinyl monomer, a rapid aggregation of nanoparticles was observed as a result of a combination of thiophene radical cation species, whereas in the presence of an acrylate monomer, free radical polymerization was initiated (Scheme 9). 163,164 Scheme 9. Photopolymerization by Gold Nanoparticles: Photoinduced Electron Transfer Reactions of Bithiophene Functionalized Gold Nanoparticles Scheme 8. Photoinduced Polymerization by Upconverting Nanoparticles Photoinduced Electron Transfer Reaction of Onium Salts. Photoexcited state photoinitiators can be good oxidant or reductant species compared with their ground state form and can be quenched through either a reduction or oxidation process, depending on the characteristics of the present co-initiator. In contrary to hydrogen abstraction wherein the triplet photoinitiator is quenched through a reduction process, as observed with amine co-initiators, an onium salt co-initiator can quench the excited photoinitiator in an oxidation manner through photoinduced electron transfer processes. Onium salts are well-known co-initiators in Type II photoinitiated polymerizations having good redox properties; nevertheless, they are primarily used as photoinitiator in cationic systems (vide infra). In the process, a photoinduced electron transfer from the photoexcited photoinitiator to the onium salt is encountered that forms a photoinitiator radical cation and a neutral onium radical with which both cationic and radical photopolymerizations can be initiated, respectively. In addition to their use as co-initiator for free radical photopolymerization, onium salts are also used as additives in Type II photoinitiated polymerization reactions. Ketyl radicals formed in such systems, as discussed earlier, are capable of terminating propagating radicals. To suppress this disadvantage, onium salts 2.2. Cationic Polymerization Direct Initiation. Photoinitiated cationic polymerization of corresponding monomers such as epoxides and vinyl ethers is widely utilized in various UV-curing and other commercial applications. 165,166 Unlike free radical polymerizations, cationic polymerization does not suffer from oxygen inhibition and so polymerizations can be conducted in air. The use of onium salts in cationic photopolymerization has long become a common practice in relevant commercial applications and academic research arenas Iodonium, 171 sulfonium, phosphonium, 174,175 and pyridinium 176,177 salts among others 178,179 are commonly applied onium salt photoinitiators. These salts contain a heteroatom with a cationic center and an inorganic metal complex anion as the counteranion part such as BF 4, PF 6, and SbF 6. Due to their photosensitivity albeit at relatively short wavelengths, onium salts undergo a direct photolysis upon irradiation through a heterolytic or homolytic bond cleavage giving rise to the formation of cation or radical cation species, respectively. These resulting charged species in some cases may not initiate cationic polymerization of target monomers. However, in the presence of a hydrogen donor, they become highly active to initiate cationic polymerization by Brønsted acids formed through a hydrogen abstraction process. The photolysis of onium salts is presented in the example of diphenyliodonium salt in Scheme 10. A variety of iodonium, pyridinium, sulfonium, phosphonium, and many other types of onium salt photoinitiators developed and successfully used for cationic photopolymerization are given in Chart 2. Phenacyl onium salts are a particular class of onium salt photoinitiators. They can be synthesized by the reaction of phenacyl halide compounds with the corresponding heteroatom containing nucleophiles followed by an anion-exchange process with potassium or sodium salts containing nonnucleophlic counteranions to yield the final product (Scheme 10224

14 Chemical s Scheme 10. Photolysis of Diphenyliodonium Hexafluorophosphate as a Typical Example of Onium Salts 11). 180 These compounds are highly photosensitive, and their photolysis mechanism follows the same trend as other onium salts outlined previously. Although being extensively applied for cationic photopolymerization, phenacyl onium salts are also capable of initiating free radical as well as Zwitterionic polymerizations. Many studies have been reported on the use of pyridinium, anilinium, sulfonium, and other phenacyl-based photoinitiation systems An interesting study revealed a tautomerization behavior associated with the phenacyl benzoylpyridinium salts which resulted from the formation of keto enol isomers upon irradiation. 184 This behavior was found to result in a change in the optical characteristics of the photoinitiator extending its absorption maxima to visible light regions (Scheme 12). In a recent study, a polymeric photoinitiator of phenacylpyridinium salt was synthesized using a polystyrene-b-poly(2- vinylpyridine) support, which was obtained by a living anionic polymerization of styrene and 2-vinylpyridine monomers. Phenacyl bromide was reacted with the pyridine moieties present in the block copolymer support to form phenacylpyridinium salt after an anion-exchange process with potassium hexafluoroantimonate, and the resulting polystyrene-b-poly(2- vinyl phenacylpyridinium) salts were used to photoinduce both free radical and cationic polymerization of corresponding monomers (Scheme 13). 192 Moreover, photolysis of phenacyl groups provided a means to photoswitching the behavior of block copolymers from cationic to neutral states Free Radical Promoted Cationic Polymerization. A practical disadvantage most often associated with onium salts is their photoactivity laying mainly at short wavelength of the UV spectrum between 230 and 300 nm. Such photoactivity requires the use of high-energy irradiation sources to accomplish the photolysis process, which would consequently render it unsuitable for certain purposes. To overcome such problems, indirect photolysis through electron transfer processes has been proposed. 193,194 As stated previously, onium salts can easily be reduced by electron-donating radical species. As a result, free radicals with electron-donating properties are oxidized by the onium salts to form carbocation species suitable for the initiation of cationic polymerizations. This process is referred to as free radical promoted cationic polymerization. The radicals can be generated using various photochemical means or other thermal methods 195,196 as well. This technique allows conducting cationic polymerization initiated by onium salts at higher wavelength using even visible light radical photoinitiators. Therefore, it is possible to tune and broaden the spectral sensitivity of those systems by carefully selecting the photoinitiator. For example, Type I photoinitiators which afford easily oxidizable free radicals by unimolecular bond cleavage are excellent promoters of cationic photopolymerization. The general mechanism of the free radical promoted cationic polymerization using a benzoin Type I photoinitiator and iodonium salt is outlined in Scheme 14. There have been numerous studies using benzoin-type photoinitiators for the promotion of cationic polymerization in the presence of onium salts Acylphosphine oxide photoinitiators have been used to oxidize with onium salts to afford phosphonium cations Acylgermane-based 209 photoinitiators which form germyl radicals upon visible light irradiation have been shown to get oxidized by onium salts generating germanium cations for the initiation of the cationic polymerization of a variety of appropriate monomers Phosphites have been utilized as reducing co-initiators with iodonium salts to initiate visible light cationic polymerization. The mechanism involves the formation of phosphonaryl radicals and subsequent redox reactions with iodonium salt to form cationic species. Azo-based compounds such as phenylazoisobutyronitrile are used as photoinitiators that afford phenyl radicals upon decomposition under visible light irradiation. The phenyl radical adds to the phosphite compounds yielding arylphosphoranyl radicals. The interaction of phosphoranyl radicals with the iodonium salt was proposed to proceed via an electron transfer to the iodonium salt resulting in the oxidization of phosphoranyl radicals to give arylphosphonium cationic species. The phosphonium cations initiated cationic polymerization of cyclic monomers, including cyclohexene oxide and tetrahydrofuran by methylation of the monomer and formation of arylphosphonate product. 214,215 Due to their high nucleophilicity, phosphites can also act as inhibitor of polymerization through chain transfer reactions. Therefore, the concentration of phosphites is required to be carefully balanced to ensure efficient reaction with radicals and subsequently with iodonium salts and yet must not be so great to inhibit the polymerization reaction. Later studies revealed that modification of phosphites with halogen-substituted alkoxy groups reducing their nucleophilicity could suppress any inhibitory activity of phosphite compounds to allow an efficient polymerization process (Scheme 15). 216 Promotion of cationic polymerization by free radicals produced by Type II photoinitiators has been investigated. As Type II photoinitiators thioxanthone, benzophenone, and camphorquinone and related structures have been employed in this connection. In these systems, co-initiators interacting with the excited state photoinitiators are necessary to form electron-donor free radicals. 94, In such systems utilizing Type II photoinitiators for cationic polymerization, the use of amines as co-initiator should be carefully considered as aliphatic amines due to the fact that their strong basicity may terminate propagating cationic chains and prevent polymerization. 220 Aromatic amines are suitable co-initiators for this purpose. Additionally, o-phthaldehyde was employed to promote the cationic polymerization process by the formation of biradicals through intramolecular hydrogen abstraction in the excited 10225

15 Chemical s Chart 2. Common Onium Salts Used in Photopolymerization Reactions state (Scheme 16). 221 Subsequent oxidation of the biradicals by pyridinium salt and further decomposition to yield Bronsted acid initiated cationic polymerization. As previously shown, electron transfer reactions from the radicals formed this way to the onium salt oxidize radicals to cationic species that are capable of initiating cationic polymerization of various monomers. It is worth noting that in those systems using Type II photoinitiators in combination with different types of co-initiators in the presence of onium salts, there is also the possibility of the interaction of the photoexcited photoinitiator with both co-initiator and onium salts as well via reduction or oxidation pathways, respectively. However, this possibility is highly dependent on the characteristics of the photoinitiator and onium salts to directly interact in the excited state. Two-photon absorption mechanism was also applied in free radical promoted cationic polymerizations. Benzodioxinone and naphtadioxinone were shown to act as two-photon absorption photoinitiators, whereby a desirable photoinitiator (i.e., benzophenone) could be formed in a stepwise two-photon absorption process. Absorption of the first photon by benzodioxinone or naphtadioxinone and their subsequent photolysis released benzophenone photoinitiator. 222,223 The second photon absorption was by the benzophenone formed in situ to generate radicals through successive electron transfer and proton abstractions. Subsequent interaction with the iodonium salt and electron transfer reactions resulted in the formation of initiating cationic species, as well as regenerating ground-state benzophenone. The mechanism of the process is outlined in Scheme 17. Neckers and co-workers and others used various dye molecules in conjunction with amine co-initiators to induce free radical promoted cationic polymerization by onium salts under visible light. 220,224,225 In this regard, silane compounds were particularly interesting co-initiators for the promotion of cationic polymerization. Silane co-initiators interacting with the excited state bimolecular photoinitiators, produce silyl radicals that can easily be oxidized by onium salts to form active 10226

16 Chemical s Scheme 11. Synthesis of Phenacyl Onium Salt Photoinitiators and Representative Examples Scheme 13. Phenacyl Pyridinium Salts Based on a Block Copolymer Support as Photoinitiator for Free Radical and Cationic Polymerizations cationic species. The silyl radicals are considered insensitive toward oxygen and can act as oxygen scavengers. Those silane compounds with a labile hydrogen atom form silyl radicals through hydrogen abstraction by the photoinitiator and the disilane structures without any abstractable hydrogen, photoinduced electron transfer results in Si Si bond dissociation bringing about active silyl radicals Laleveé, Fouassier, and co-workers have developed elegant approaches for induction of cationic polymerization in a free radical promoted manner using three component initiating systems. Transitional metal complexes such as Ru II or Ir III have been used in combination with silane or other co-initiators to form initiating cation species with the aid of onium salt oxidants. The transitional metal complexes are highly photoactive at visible light regions and undergo photoexcitation under soft irradiation conditions. In the process, the photoexcited metal complex Ru II * species can be quenched in an oxidative manner by the iodonium salt resulting in the oxidation of the metal complex to Ru III and formation of phenyl radicals. These two components failed to efficiently initiate cationic polymerization of epoxides. However, in the presence of silane, the phenyl radicals abstracted the labile Scheme 12. Photoinduced Keto-Enol Tautomerism of the Phenacyl Benzoylpyridinium Salt a a Reprinted from ref 184. Copyright 2006 American Chemical Society

17 Chemical s Scheme 14. Photoinduced Free Radical Promoted Cationic Polymerization by Type I Photoinitiator Scheme 15. Phosphite Co-Initiators in Visible Light Free Radical Promoted Cationic Polymerization Scheme 16. Photoinduced Free Radical Promoted Cationic Polymerization by Using o-phthaldehyde and a Pyridinium Salt hydrogen of the silane bringing about silyl radicals. At this point, the silyl radicals are oxidized to initiating cationic species through two possible pathways. Oxidation can happen by electron transfer either to the onium salt or through interaction with the oxidized Ru III species. The latter process also regenerates the initial ground state photocatalyst as well. Initiation of cationic polymerization was effectively achieved using this three-component system based on transitional metal photocatalyst and silane co-initiators The mechanism of the process is illustrated in Scheme 18. Various other systems benefiting from the photoinduced electron transfer reactions enabled by transitional metal photoredox catalysis including Ir, Ru, Fe, Cu, etc., and silane or similar co-initiators have been reported for the promotion of cationic polymerization. 230, It is worth mentioning the use of substituted vinyl halide compounds as promoters of cationic polymerization. The carbon-halide bond present in these compounds is readily cleavable upon irradiation at corresponding wavelengths giving rise to the formation of vinyl radicals. An electron transfer then is considered to take place, yielding vinyl cationic and bromine anionic species. 244 Moreover, oxidization of these vinyl radicals by sulfonium or pyridinium salts was employed to form carbocations capable of initiating cationic polymerization (Scheme 19) Cationic Polymerization by Electron Transfer with Singlet and Triplet Excited States. Photosensitization is one such efficient indirect approach for the photolysis of onium salts. 246 Photosensitizers being photochemically active at higher wavelength at near UV or visible light regions are expected to interact in the excited state with onium salts through various pathways resulting in the photolysis of the onium salts. As a result of photosensitization of onium salts through photoinduced electron transfer, radical cations of the photosensitizer and neutral onium radical species are generated. The radical cation so-formed may directly initiate cationic polymerization, or it can interact with a hydrogen donor compound to form a Brønsted acid capable of initiating cationic polymerization reactions (Scheme 20). 247 The photolysis of onium salts by photosensitization is thermodynamically feasible only if the free energy changes (ΔG) of the photoinduced electron transfer process are negative. This can be estimated using the Rehm Weller equation: ox et c 1/2 + red 1/2 Δ G = f [ E (D/D ) E (A/A )] E* + ΔE where f c is the Faraday constant, E ox 1/2 (D/D + ) and E red 1/2 (A/A ) are redox potentials of the donor (D) and acceptor (A) compounds, respectively, E* is the excited state energy of the sensitizer (singlet or triplet), and ΔE c is the Coulombic stabilization energy. The free-energy changes of the electron transfer (ΔG et )of some photoinitiators and photosensitizers with frequently used onium salts, calculated by the Rehm Weller equation, are given in Table 3. Different types of photoinitiators, polynuclear aromatic, highly conjugated compounds, and dyes exhibiting light sensitivity at near UV and visible light regions have been reported as favorable photosensitizers in this regard. Earlier attempts in photosensitization of onium salts were carried out by Crivello using dye molecules. 170 Acridinium, benzothiazolium, and hematoporphyrin type dyes with strong absorption characteristics at higher wavelengths of visible light were used for this purpose (Chart 3). Relatively poor solubility of these dyes in monomer solutions was encountered. Photosensitiza- c 10228

18 Chemical s Scheme 17. Free Radical Promoted Cationic Polymerization by Stepwise Two-Photon Absorption Mechanism Scheme 18. Free Radical Promoted Cationic Polymerization by Transitional Metal Complexes and Silanes in the Presence of Onium Salts Scheme 19. Photoinduced Cationic Polymerization by Substituted Vinyl Ethers: Formation of Cationic Species by Electron Transfer between (a) Radical Fragments or (B) Oxidation by Onium Salts Scheme 20. Photosensitization of an Onium Salt for the Initiation of Cationic Polymerization (PS: Photosensitizer) tion of iodonium salts was effectively achieved with these compounds; however, they were inefficient for the sensitization of sulfonium salts. Electron-rich polynuclear aromatics such as perylene, pyrene, anthracene, and their derivatives have been shown to 10229

19 Chemical s Table 3. Redox Potentials of Photosensitizers and the Free Energy Change of Electron Transfer with Some Onium Salts Calculated by the Rehm-Weller Equation a a E t *: triplet excited energy, E s *: singlet excited energy. Chart 3. Dye-Photosensitizers for Onium Salts Chart 4. Polynuclear Aromatic Compounds As Electron- Transfer Photosensitizer photosensitize onium salts (Chart 4) The ΔG values for the electron transfer sensitization of onium salts by polynuclear aromatic compounds have been found to be quite large and negative, indicating their potential for use as photosensitizers. Substitution was proposed to overcome problems concerning poor solubility of these compounds. Phenothiazine and its derivatives are another class of efficient photosensitizers with excellent reducing properties In a similar manner, thioxanthones, carbazoles, curcumin, 276,277 quinoxaline and related structures, and highly conjugated derivatives of thiophene 242, have been used as long wavelength photosensitizers. Moreover, several other studies have been reported on photosensitization of onium salts based on photoinduced electron transfer reactions for the initiation of cationic polymerization. 218, It was recently shown that ferrocenium salts can promote sensitization of onium salts via photoinduced electron transfer. 297 Photoinduced electron transfer between the excited ferrocenium salt and iodonium salt resulted in the photolysis of the iodonium salt producing ferrocenium radical cations and onium radicals. The radical cation could either undergo further photolysis producing the arene ligand and an active Lewis acid or react with the phenyl radicals to form protonic acid species. The Lewis acid or protonic acid species so-formed were responsible for the initiation of cationic polymerization. Fullerene (C 60 or buckyball) is another such efficient photosensitizer for onium salt photoinitiators to initiate cationic polymerization. It has been shown that photoinduced electron transfer to generate 10230

20 Chemical s initiating cationic species could be efficiently promoted by fullerene under visible light and in the presence of onium salts. 298 Upon photoexcitation of fullerene in the presence of onium salts, an exciplex was formed which was followed by electron transfer from the fullerene to the onium salt, generating fullerene radical cation while reducing onium salts. A variety of monomers including cyclohexene oxide, isobutyl vinyl ether, and N-vinylcarbazole were successfully polymerized using this method. Silver hexafluourophosphate (AgPF 6 ) proved as a more efficient oxidant for fullerene than diphenyliodonium salt as the free energy changes for the electron transfer in the presence of AgPF 6 were more negative and higher monomer conversions were achieved. Fullerene could be used in chlorobenzene solvent or it could alternatively be attached to polystyrene to improve its solubility in organic monomers without the need for using solvents. Onium salts are most often used in three-component photoinitiation systems. These systems consist of a photoinitiator (or photosensitizer), a hydrogen donor (an amine, generally), and an onium salt co-initiator With regard to three-component photoinitiators, triplet state photoinitiator simultaneously reacts with both the amine and onium salt coinitiators through hydrogen abstraction and electron transfer reactions, respectively. The aim of these systems is to enhance the efficiency and rate of photopolymerization reactions. The consequence of hydrogen abstraction from the amine coinitiator brings about an amine-based radical capable of initiation of radical polymerization. A ketyl radical is also formed which in the presence of the onium salt is oxidized to form a protonic acid and a neutral onium radical while regenerating the ground state photoinitiator. Considering the interaction of the triplet photoinitiator with the onium salt, a neutral onium radical and photoinitiator radical cation are formed, in which the latter can sometimes initiate cationic polymerization with the former capable of initiating radical polymerization. In the presence of an amine co-initiator, the formed radical cation abstracts a hydrogen atom yielding a protonic acid and amine-derived radical and, again, regenerating ground state photoinitiator. The general mechanism of this three-component photoinitiation is summarized in Scheme 21. The ability to initiate polymerization of both free radical and cationic monomers, converting noninitiating species to active, initiating components (as in the oxidation of terminating ketyl radicals to generate protonic acid and initiating radical), and more importantly, regeneration of the ground state photoinitiator are the advantages of three-component photoinitiation systems making them highly advantageous for many applications Cationic Polymerization by Charge Transfer Complexes. Pyridinium salts have been shown to be capable of forming ground state charge transfer complexes with electron-rich donors such as methyl- and methoxy-substituted benzene derivatives. 307,308 These complexes absorb at relatively high wavelengths, where other components are virtually transparent. For example, the complex formed between N- ethoxy-4-cyano pyridinium hexafluorophosphate and 1,2,4- trimethoxybenzene exhibited absorption maxima at 420 nm. The absorption maxima of the two components individually are 270 and 265 nm for the pyridinium salt and trimethoxybenzene, respectively. It was found that the charge transfer complexes formed between pyridinium salts and methyl- and methoxy-substituted benzene could act as photoinitiators for the cationic polymerization of cyclohexene oxide and 4- Scheme 21. Three-Component Photoinitiation of Free Radical and/or Cationic Polymerizations vinylcyclohexene oxide. The photoinitiation mechanism primarily involved formation of charge transfer complexes between the two components with subsequent electron transfer bringing about a neutral pyridinium-based radical and radical cation of the benzene derivative. The radical cations formed were considered effective initiating species. The overall mechanism is depicted in Scheme 22. Moreover, since a Scheme 22. Cationic Polymerization through Photoinduced Electron Transfer in Charge Transfer Complexes proton scavenger had no influence on the rate of polymerization, the initiation by Bronsted acid which could be possibly formed by an interaction with hydrogen donor components was excluded. Notably, the charge transfer complexes described above were found applicable for the photoinitiation of epoxide monomers but not for the photoinitiation of other monomers such as vinyl ethers and N-vinylcarbazol monomers. Theses monomers were observed to polymerize in the dark upon addition of the complexes Cationic Polymerization by Addition Fragmentation Agents. Another way to achieve free radical promotion 10231

21 Chemical s of cationic polymerization is by an addition fragmentation technique. 309 This technique enables photosensitization of allyl functional onium salts. It involves the addition of a radical species generated by photochemical or thermal means to the double bond of the allyl group of the onium salt. The radical adduct so-formed tends to fragment to yield onium radical cation species. The cationic polymerization can be initiated by the resulting onium radical cation after further fragmentation, or in the presence of a hydrogen donor it may abstract a hydrogen atom and produce Bronsted acids capable of initiating cationic polymerization. The overall process is depicted in Scheme 23. A variety of mono- or biallyl sulfonium, Scheme 24. Controlled Living Cationic Polymerization in the Presence of Lewis Acids Scheme 23. Cationic Photopolymerization by Addition fragmentation Technique phosphonium, and pyridinium salts have been investigated for the initiation of cationic polymerization using the addition fragmentation technique. 175, Photoinduced Living Cationic Polymerization. In order to establish control over the cationic polymerization systems, Lewis acids have been employed. These compounds enable cationic polymerization of vinyl ether monomers in a controlled/living manner by coordinating with the propagating cationic chains. Mechanistically, a photochemically generated Bro nsted acid adds to the vinyl ether monomer to form a halogen-terminated adduct. The Lewis acid used coordinates with the so-formed halogen-containing adduct ensuring the stabilization of the carbocation center of propagating chains. As a result, the cationic polymerization proceeds in a controlled/ living manner with suppressed chain-breaking reactions due to the coordination of Lewis acids, leading to polymers with controlled molecular weight properties (Scheme 24). Various approaches have been reported in this regard which include direct photolysis of onium salts, free radical promoted, 204 photosensitization, 324 and other approaches such as the use of substituted vinyl halides 325,326 to generate the initiating protonic acid species. In a conceptually distinct approach, Perkowski and coworkers utilized a photoredox concept to carry out controlled/ living cationic polymerization. 327 They used 2,4,6-tri(p-tolyl)- pyrylium tetrafluoroborate as a photoinitiator in combination with methanol to initiate the cationic polymerization of 4- methoxystyrene. Oxidation of the monomer by the photoexcited state photoinitiator was suggested to induce the process generating a styrenyl radical cation species. Methanol was utilized as a reversible chain transfer agent to regulate the concentration and propagation of cationic chain ends. A cationic species formed was captured by methanol generating a proton capable of protonating the additional monomer. This process of nucleophilic addition of methanol to cationic species and subsequent protonation of monomers took place until methanol was totally consumed. As a result, methoxy-captured dormant species and active cationic chains were formed. The methoxy group serving as the chain transfer agent was able to form a transient intermediate of oxonium ion by capturing an active cationic chain end. The oxonium intermediate further fragmented to form active cation and dormant species in a reversible manner, thus regulating the propagation of cationic chains and establishing control over the cationic polymerization process. Polymers with controlled molecular weight properties and narrow distributions were obtained in this way. Additionally, the livingness of the resulting polymers was demonstrated by successful chain extension experiments. 3. STEP-GROWTH POLYMERIZATION BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS Photoinduced electron transfer reactions have been exploited in step-growth polymerization of conjugated monomers as well. For example, interacting photoexcited state thiophene species with electron acceptor compounds such as dinitrobenzene or carbon tetrachloride promoted electron transfer reactions between the thiophene and electron donor compounds This resulted in the formation of corresponding thiophene radical cation species which were then coupled to form polythiophene. Potassium dichromate (Cr(VI)), which is a photochemically active catalyst, was also used to photochemically initiate polymerization of thiophenes. The photoexcited Cr(VI) was suggested to form a complex with the monomer which then resulted in a charge-separation reaction, leading to the formation of radical cation species of the thiophene monomer. 331 Other photochemical strategies concerning photoinduced electron transfer reactions of onium salts and thiophene 10232

22 Chemical s derivatives have also been developed by our group. 332,333 It was shown that an electron transfer from the radical cations formed during the photolysis of the iodonium salt to thiophene was responsible for the initiation of the process. Radical cation species of thiophene were formed that could then undergo dimerization by proton release followed by rearrangement of aromaticity. 334 The dimer thus formed participated in further similar reactions with the iodonium radical cation and subsequently with the thiophene radical cation to further polymerization (Scheme 25). Polymerization took place until a Scheme 25. Photoinduced Electron Transfer Reactions for Step-Growth Polymerization of Thiophene dark precipitate was formed on the inner surface of the reaction tube preventing reaction to continue further. Iodonium, sulfonium, and pyridinium salts were able to promote stepgrowth polymerization of thiophene derivatives. Similar strategies could be applied to the efficient photochemical synthesis of a variety of thiophene derivatives using various types of onium salt photoinitiators (Chart 5). 282,335 In the presence of highly conjugated thiophene derivatives, due to their strong light sensitivity at higher wavelengths, they were able to sensitize the photolysis of onium salts in the photoexcited state under visible light irradiation. In this case also, photoinduced electron transfer between the photoexcited thiophene and onium salt induced photolysis of the salt, yielding thiophene radical cation species. Photosensitization of onium salts by conjugated thiophene derivatives is wellestablished for the initiation of free radical and cationic polymerization (vide supra). Notably, proton release in the course of coupling of thiophene radical cation species could well be taken advantage of for initiating cationic polymerization of appropriate monomers. An interesting approach to photoinduced step-growth polymerization has been recently proposed by Ravoo and coworkers who used a photocatalytic system based on TiO 2 nanoparticles to carry out step-growth polymerization. 336 They took advantage of photoinduced electron/hole charge carriers from nanoparticles to induce reduction/oxidation of ethanol amine to initiate its step-growth polymerization. They used an alcohol-functionalized substrate to grow polymer brushes by a microcontact printing method. It was proposed that the photogenerated hole species could oxidize ethanol amine (or the alcohol groups present on the surface of the substrate) to the corresponding aldehyde. This aldehyde was then reacted with another ethanol amine molecule to produce an imine compound, which was finally reduced by the photogenerated electrons to the secondary amine. The amine molecules grown as such can re-enter the catalytic cycle and consequently form polyethylenimine in a step-growth manner mediated by TiO 2 photocatalysis. The overall process is illustrated in Scheme 26. Using this novel method in such photoinduced step-growth polymerization, the authors were able to prepare microcontact printed polymer brushes. A similar approach could also be applied to the synthesis of polypyrrole using TiO 2 -based nanostructures as both photocatalyst and a template for the polymerization of pyrrole using photoinduced charge carriers. 148, POLYMER-METAL NANOCOMPOSITES BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS Photochemical generation of metallic nanoparticles is arguably one of the most important implementations of photoinduced electron transfer reactions, which has been the subject of various studies in recent years. 338 It is argued that for the generation of metal nanoparticles, photochemical means are considered advantageous over other methodologies in that they offer control of the rate of nanoparticle formation and the possibility of spatiotemporal control. In principle, photochemical means are used for the formation of free radicals that are able to reduce metal ions via electron transfer reactions to form metal nanoparticles. Due to their favorable redox potentials, ketyl radicals availed from both Type I and Type II photoinitiators are promising means for the formation of many types of nanoparticles In the presence of free radical or cationic monomers, simultaneous photoinitiation of corresponding free radical or cationic polymerizations and the formation of metal nanoparticles afford a polymeric matrix containing embedded metal nanoparticles. Nanocomposites of homogeneously distributed metal nanoparticles within the polymeric matrix can thus be obtained by photochemical means. In this regard, gold nanoparticles, for example, were Chart 5. Thiophene Derivatives Polymerized by Photoinduced Induced Electron Transfer Reactions 10233

23 Chemical s Scheme 26. Photoinduced Step-Growth Synthesis of Polyethyleneimine Brushes by Using Semiconducting Photocatalysisa a AFM measurements of patterned brushes with profile height. Reprinted with permission from ref 336. Copyright 2015 Royal Society of Chemistry. Scheme 27. In Situ Formation of Nanoparticles Embedded in Polymer Network by Photoinduced Electron Transfer Reactionsa a TEM images of Au nanoparticles embedded in a crosslinked poly(ethylene glycol) diacrylate network with different concentrations of Au3+ solution: (a) 1 wt % and (b) 5 wt %. Reprinted with permission from ref 349. Copyright 2008 Royal Society of Chemistry. synthesized in situ using AuCl4 as the metal ion precursor and a Type I photoinitiator in the presence of a poly(ethylene glycol) diacrylate (PEGDA) monomer. The photoinitiator affords two initiating and reducing ketyl radicals by which the 10234

24 Chemical s initiation of polymerization and formation of gold nanoparticles were achieved, respectively (Scheme 27). Formation of crosslinked networks enables long run stabilization of the resulting nanoparticles. The formation of gold nanoparticles could be detected by the evolution of the typical surface plasmon absorption band for gold nanoparticles centered at about 600 nm. There are many studies reporting on photochemical synthesis of gold, silver, and other metal nanoparticles via photoinduced electron transfer while concomitantly carrying out radical polymerization to afford metal nanoparticles embedded polymer matrices Also a 2,7-diaminofluorene derivative dye was found to simultaneously initiate free radical polymerization and nanoparticles formation via photoinduced electron transfer. 350 A self-quenching mechanism was proposed to form dye radical cation species which then could abstract a hydrogen atom from the poly(ethylene glycol) monomer forming radical species capable of initiating radical polymerization. In the presence of silver ions, photoinduced electron transfer from the excited state dye directly to the silver ions corresponded to the formation of silver nanoparticles. Adding a hydrogen donor compound like an amine significantly increased the rate and efficiency of both polymerization and nanoparticle formation. Initiation of the polymerization was predominantly through the interaction of the dye/amine compounds via consecutive electron and hydrogen transfers forming initiating α-aminoalkyl radicals. Due to the stabilization role of the amine molecules, samples formed in the presence of amines had uniform size distribution, while in the absence of the amine some agglomeration of the nanoparticles was observed. The key to the formation of metal nanoparticles embedded in polymer matrices via such photochemical reactions is the ability to control the kinetics of the process and the morphology of the resulting nanoparticles. The concentration and type of the photoinitiator employed, the initial concentration of metal ion can directly affect the process and thus the morphology of nanoparticles. Cationically polymerizable monomers have also been used in combination with metal ions to simultaneously form nanocomposites of metal nanoparticles embedded in polymeric matrices. For example, using a cleavable photoinitiator such as 2,2-dimethoxy-2-phenyl acetophenone (DMPA) which produces strong electron donor radicals by bond cleavage under irradiation, the reduction of silver ions could be achieved to form silver nanoparticles. In the meantime, upon donation of an electron, the radicals oxidize to produce active cationic species, which are capable of initiating cationic polymerization in a similar manner to free radical promoted cationic photopolymerizations. A variety of multifunctional cationic monomers including epoxy and vinyl ether monomers have been polymerized via this approach An alternative approach involves the use of photosensitizer compounds such as highly conjugated thiophenes. Sensitization through electron transfer processes from the excited photosensitizer to the metal ion salt induces the formation of metallic nanoparticles, while the radical cations of the oxidized photosensitizer are used to initiate cationic polymerization of epoxy monomers. 284 A photoredox complex, namely tris(triphenylphosphine)- ruthenium(ii) dichloride (Ru(PPh 3 ) 3 Cl 2 ), was shown to promote the in situ synthesis of metallic nanoparticles, while initiating cationic or free radical polymerizations to afford nanocomposite structures. 355 The irradiation of the Ru- (PPh 3 ) 3 Cl 2 complex resulted in the P-Ph bond scission, thus generating phenyl radicals. Free radical polymerization was initiated by the phenyl radicals. In the presence of a metal ion, photoinduced electron transfer took place from the excited Ru II * complex to the silver ion, generating silver nanoparticles.. Furthermore, the phosphine-centered radicals generated from the bond scission of the ligand in the Ru II complex underwent oxidation to reduce metal ions to metallic nanoparticles. Upon addition of a hydrogen donor solvent such as tetrahydrofuran, the formation of metal nanoparticles was accelerated due to the fact that the phenyl radicals abstracted a hydrogen atom from the hydrogen donor generating easily oxidizable radicals by the metal salt used. Also cationic polymerization could be successfully initiated by the cations formed during the reduction of the salt. Laleveé and co-workers have developed sophisticated strategies for the synthesis of novel metallic nanoparticles that do not exist in a salt form to incorporate in polymeric networks. A series of zirconium (Zr) and titanium (Ti) nanoparticles were formed using zirconocene dichloride (Cp 2 ZrCl 2 ) or titanocene dichloride (Cp 2 TiCl 2 ) compounds, respectively, in the presence of a Type I photoinitiator such as DMPA. Conducting the experiments in the presence of oxygen, they observed significant reduction of oxygen inhibition mediated by those organometallic compounds. In fact, peroxyl radicals formed by the reaction of oxygen with the photogenerated radicals from the DMPA photoinitiator were proposed to interact with Cp 2 ZrCl 2 (or Cp 2 TiCl 2 ) to convert nonactive peroxyl radicals to new initiating radicals for initiating free radical polymerization. Thus, significant enhancement of the rate of polymerization in the presence of oxygen was achieved by the added organometallic compounds being used as additives to Type I photoinitiators. The interaction of the peroxyl radicals with the organometallic compounds through a bimolecular homolytic substitution reaction led to the formation of metal-based structures, which finally resulted in the generation of metallic nanoparticles embedded in the polymeric matrix In a quite similar approach, they investigated a series of different Ti-based complexes containing Ti O bonds as additives toward reduction of oxygen inhibition for the free radical polymerization initiated by a bisacylphosphine oxide photoinitiator, as well as photochemical generation of metallic nanoparticles. 359,360 The key to these systems is the formation of metal-based radical compounds which act as peroxyl radical scavengers, and multiple addition of peroxyl radicals leads to the formation of metal nanoparticles. Various applications have been encountered for such nanocomposite systems that contain metal nanoparticles in polymeric networks. Of particular interest have been silvercontaining composites owing to the antibacterial activity of silver nanostructures. For instance, acrylamide-based hydrogels with swelling deswelling properties were photochemically synthesized with simultaneous photogeneration of silver nanoparticles, which exhibited antibacterial activity against a series of standard bacterial samples. 361 Surface modification of a biopolyester, namely poly(3-hydroxybutyrate-co-3-hydroxyvalerate), was photochemically conducted to form antibacterial materials containing silver nanoparticles. 362 For this purpose, butan-2-one as a ketone-type photoinitiator was used which was capable of abstracting a hydrogen atom from the backbone of the to induce "grafting from" of the methacrylic acid monomer. The carboxylic functionality of grafted polymers served as stabilizing agents for the immobilization of silver nanoparticles generated via photoinduced electron transfer 10235

25 Chemical s Scheme 28. Polymer-Bound Onium Salts for Photochemically Formation of Block Copolymers using antraquinone as the photoinitiator and an amine coinitiator. Capping the formed silver nanoparticles by the carboxylic functional groups resulted in the stabilization and homogeneity of these nanoparticles. The obtained material containing silver nanoparticles showed successful antibacterial activity. Finally, polymeric nanocomposites containing gold nanoparticles and glucose oxide as a biomolecule enzyme were formed and investigated as efficient platforms for the biosensing system. 363 The so-called ex situ techniques have also been applied for the generation of polymeric networks containing embedded metal nanostructures. The ex situ techniques utilize preprepared metallic nanostructures, which are generally stabilized and functionalized with photoinitiators to initiate polymerization processes. Silver nanorods or spherical nanoparticles functionalized with thioxanthone have been reported in this regard. 364, BLOCK AND GRAFT COPOLYMERS BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS Complex macromolecular architectures such as block and graft copolymers have been obtained using various photochemical approaches. Indeed, such copolymeric structures consisting of various segments each possessing different properties have unique and novel physiochemical properties that make them suitable for many emerging applications. The advent and development of controlled/living polymerizations has paved the way for versatile syntheses of a variety of different copolymers. The advancement in this area is dealt with in the subsequent sections, whereby photoinduced electron transfer reactions have been shown to combine with controlled/living radical polymerizations for the synthesis of complex macromolecular structures. In this section, however, we will review block copolymerization of monomers from chemically different families that are generally homopolymerized through different conventional polymerization mechanisms. This process in general is referred to as mechanistic transformation, which allows copolymerization of chemically different monomers. 366,367 Photochemical approaches have been demonstrated as efficient methods in promoting such transformations This approach is based on polymerization of a monomer by one mechanism, and, if necessary, endfunctionalization, to obtain macroinitiators used for the initiation of the second monomer through another differing mechanism. The functional groups on one or both ends or middle of prepared polymers are activated by photoinduced electron transfer processes generating suitable sites for polymerization of other monomers. One approach deals with the use of thermal initiators functionalized with photoactive groups. This method is particularly appropriate for combination of radical and cationic polymerizations for the synthesis of block copolymers. As such, first, the radical polymerization of a monomer is conducted thermally. Afterward, the formed polymers with photolabile groups are irradiated in the presence of suitable compounds or co-initiators to generate cationic sites that are capable of initiating cationic polymerization. In this connection, an azo initiator functionalized with photoactive benzoin groups at both ends was used to thermally initiate the radical polymerization of styrene. 200 By doing so, a polystyrene macrophotoinitiator was synthesized with benzoin-functionalized end groups. 371 Irradiation of the macrophotoinitiator resulted in the bond scission of the benzoin group and radical formation which in the presence of oxidant onium salts were oxidized to carbocation by electron transfer processes. Polymerization of cyclohexene oxide monomer could then be initiated in a radical promoted manner affording polystyrene-b-poly(cyclohexene oxide) copolymers. The sequence of the blocks could be arranged by changing the sequence of polymerizations, meaning that the azo initiator could be first irradiated in the presence of onium salts to initiate cationic polymerization of cyclohexene oxide containing an azo linkage in the main chain. On heating up, the latter polymer free radicals were formed to polymerize styrene. Macrophotoinitiators containing Type I photoinitiator functionalities are useful for the formation of block copolymers in the presence of onium salts. The role of onium salt is to 10236

26 Chemical s oxidize photogenerated macroradicals to form carbocation centers Use of polymeric co-initiators in conjunction with Type II photoinitiators is also applied for block or graft copolymerization. In this technique, radicals are formed by a successive electron/proton transfer process of the photoinitiator and polymeric co-initiator, yielding suitable sites for the initiation of second polymerization The immobilization of various onium salts on polymeric supports 384 has been reported as an efficient route to macrophotoinitiators for both cationic and free radical photopolymerizations. This can generally be achieved using postmodification processes on certain polymer chains, so they may act as the supporter to incorporate onium salt functionality. In doing so, problems concerning poor solubility of some onium salts in certain media as well as drawbacks of small molecule onium salt photoinitiators can be overcome to some degree. Perhaps the most significant attraction of developing such macrophotoinitiators, however, is their resulting ability to photochemically generate copolymers of different topologies through mechanistic transformation processes 385 enabled by onium salt functionalities. In this regard, for example, pyridinium-functionalized polytetrahydrofuran (PTHF) macrophotoinitiators were prepared by quenching the oxonium groups of the parent PTHF polymer polymerized by living cationic polymerization thus placing the pyridinium salts at chain ends of the polymer (Scheme 28) Using these pyridinium-functionalized PTHF as macrophotoinitiator, which decomposed on irradiation to form polymeric alkoxy radicals, initiation of a monomer polymerizable by a free radical mechanism [e.g., methyl methacrylate (MMA)] was achieved, leading to the formation of block copolymers containing PTHF and PMMA blocks. Similarly, polystyrene prepared by atom transfer radical polymerization was used as a support to immobilize pyridinium salts. 389 The bromine end group functionality of polystyrene was replaced with alkoxy pyridinium salts with a non-nucleophilic counteranion hexafluoroantimonate. The same procedure was applied to synthesize graft copolymers using a macrophotoinitiator with pendant pyridinium salt groups. A copolymer of polystyrene and polychloromethylstyrene was used as the polymeric support to convert chlorine groups into pyridinium functionalities by a substitution reaction. The resultant polystyrene with pendant pyridinium groups was used to form graft copolymers of a grafting-from method. 390 Transitional metal carbonyls such as rhenium carbonyls (Re 2 (CO) 10 ) or, particularly, dimanganese decacarbonyl (Mn 2 (CO) 10 ) are an interesting class of visible light photoinitiators. Its photodecomposition through Mn Mn bond dissociation forms Mn-centered radicals that are able to abstract halogen atoms from alkyl halide compounds. 391 This process forms carbon-centered radicals by electron transfer processes that are capable of initiating a variety of radical polymerizations, or it can be turned to suitable sites for other polymerization modes. Block and graft copolymerization has been successfully carried out utilizing the photochemistry Mn 2 (CO) 10 in the presence of halide-containing polymers Polymers prepared by ATRP containing bromine end groups could be used as a polymeric halide source to generate radicals by means of Mn 2 (CO) 10. Upon formation of Mn(CO) 5 radicals, the bromine present in the end chain of polystyrene was abstracted, forming radical sites. These sites were then used to initiate radical polymerization of the second monomer. 396 Obviously, such well-defined polymers by ATRP could be used as macrophotoinitiators to initiate cationic polymerization as well. 397 The radicals formed upon abstraction of the chain end bromine functionality of the polymers by Mn(CO) 5 radicals could be oxidized by a photoinduced electron transfer in the presence of onium salts to generate cationic sites capable of initiating cationic polymerization. Cationic polymerization of epoxides and vinyl ethers was conducted to afford block copolymers by combination of ATRP and free radical promoted cation polymerization methods. In a recent study, Mn 2 (CO) 10 was used as a visible light photoinitiator to form polyethylene-g-poly(cyclohexene oxide) (PE-g-PCHO) by a combination of ring opening metathesis polymerization (ROMP) and free radical promoted cationic polymerization techniques. 398 Photogenerated radicals by the manganese photoinitiator abstracted bromine functionality of a brominated polyethylene chain forming radical sites on the backbone of the polymer. These radicals were then oxidized to carbocation in the presence of an iodonium salt to initiate the cationic polymerization of cyclohexene oxide. Another similar approach used dichloromethane as the halogen source to form oxidizable carbon-centered radicals by manganese decacarbonyl photoinitiator. 399 Several other approaches taking advantage of the photochemistry of Mn 2 (CO) 10 and halide compounds have been developed for the synthesis of complex macromolecular structures such as block, graft, hyperbranched, star, etc. copolymers These complex copolymeric structures have found use in various biological applications and controlled drug delivery systems. 402,404,406 A combination of anionic and cationic polymerizations was also used to form graft copolymers of polyethers. 408 First, poly(ethylene oxide-co-ethoxyl vinyl glycidyl ether) was synthesized by anionic ring-opening copolymerization of corresponding monomers. The vinyl ether moieties were then photochemically converted to protic acid groups by using diphenyliodonium iodide photoinitiator. Living cationic polymerization of isobutyl vinyl ether was then initiated in the presence of zinc iodide catalyst by a grafting-from method in the dark. 6. CONTROLLED/LIVING RADICAL POLYMERIZATIONS BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS Merging photochemistry with controlled polymerization techniques has revolutionized the synthesis of well-defined macromolecular architectures. Many attempts have been directed toward adaptation of existing controlled/living polymerization methodologies with photochemical means, which in turn result in photochemically induction, mediation, and control of these processes Going one step further, novel-controlled polymerization techniques have emerged that are solely mediated and controlled by light. Copper-mediated controlled polymerization techniques including atom transfer radical polymerization (ATRP) 416,417 and single-electron transfer-living radical polymerization (SET-LRP), 418,419 reversible addition fragmentation chain transfer (RAFT), 420 and nitroxide-mediated polymerization (NMP) 421 are increasingly becoming well-established photochemical approaches for macromolecular architecture building via photochemical methods. A compelling feature of photoinitiated controlled polymerizations is the ability of establishing on-demand control 10237

27 Chemical s of the process both in terms of temporally manipulating the polymerization reaction as well as gaining spatial control. Thus, fabrication of 3D polymeric structures and functional patterned surfaces can be obtained in a controlled manner using photochemistry, which would have been otherwise inconceivable. In the following sections, various approaches for the synthesis of macromolecular structures via controlled polymerization techniques enabled by photoinduced electron transfer reactions are highlighted. Special emphasis is put on explaining and providing a better understanding of the mechanism of the reactions. And applications for the synthesis of various macromolecular architectures using special catalytic systems are discussed. Moreover, a comprehensive list of monomers, initiators, ligands, chain transfer agents, solvents, etc. used in photoinduced ATRP and RAFT processes is given, and the influence of particular components is discussed Atom Transfer Radical Polymerization Mechanistic Explanation Copper-Based Systems. ATRP is a reversible-deactivation radical polymerization method which uses a transition metal complex as a catalyst to establish control over the polymerization process By reaction of the transition-metal catalyst with alkyl halide compounds used as the initiator, active initiating species are formed in a reversible manner. Copper is the most widely used catalyst for ATRP reactions. The general mechanism of the ATRP process is based on the reaction of a copper catalyst in its lowest oxidation state (Cu I ) with the initiator alkyl halide (R X) to form active radical species capable of initiating the radical polymerization process. Deactivation through back halogen transfer brings about dormant species. Thus, as a consequence of activation deactivation processes enabled by the transition-metal catalyst, control is established over the entire polymerization process. In addition to Cu compounds, other transition metal complexes, including Ru, Fe, Mo, etc. have also been efficiently employed to initiate and control radical polymerization. 425,426 The development of ATRP has been focused on designing and developing new catalyst systems that operate more efficiently under mild and environmentally benign conditions. This has been mainly based on the in situ (re)generation of activator catalyst species, which in turn has led to significant developments in broadening the scope of the process as well as reducing its costs. For example, catalyst loading has been drastically reduced to ppm amounts; 427 tolerance to oxygen has been attained; 428 and applications under biologically relevant conditions 429,430 have been realized. To generate the activator catalyst in situ, several methods have been undertaken. Reducing agents, 427,431,432 various radical initiators, 433,434 electrochemically redox processes, 435 coppercontaining nanoparticles, 436 and photochemical approaches are some techniques undertaken for the generation of activator catalyst (Scheme 29). With regard to photochemical techniques, considerable enhancement of the conventional ATRP process was observed when irradiating under light. Guan and co-workers investigated the influence of irradiation on the conventional ATRP of methyl methacrylate (MMA) employing a copper(i) chloride/ bipyridine (CuCl/bipy) catalyst with the ratio of [MMA]/ [RCl]/[CuCl]/[bipy]:100/0.1/0.3/ While conducting the reaction in the dark reached a 41% reaction within 16 h, employing light to the same reaction conditions resulted in complete conversion of monomer with good control over the Scheme 29. In Situ Generation of Activator Catalyst for ATRP by Means of Various Techniques molecular weight properties of the polymers. This enhancing of the rate and efficiency of the process achieved on irradiation was attributed to the acceleration of the activation of the innersphere complex between the catalyst and the alkyl chloride, [CuCl/bipy and R-Cl], or perhaps through the activation of the catalyst alone by light. Photochemically (re)generation of active catalyst species has been realized by using various photochemical approaches. These include the use of additional photoinitiators, photosensitizers, or photocatalysts to interact with Cu II compounds while reducing them to form an activator catalyst. Additionally, many Cu complexes are highly photosensitive and have been proven to undergo excitation on irradiation and further redox reactions with the ligands or other compounds present in the reaction media have been confirmed. This behavior, of course, is highly dependent on the nature of Cu compounds and the type of ligand used to form complexes with Cu. The consequence of these reactions is the formation of free radicals and reduction of deactivator Cu II complexes to form activator Cu I. The reduction process can be through many different photochemical mechanisms. Thus, the process in which the reduction of Cu II complexes takes place is often referred to as either indirect (in the case of using additional photoinitiator or photocatalyst compounds) or direct (as in the case of not using additional photoinitiators) photogeneration of the activator catalyst for ATRP. It was found that the irradiation of a solution of copper(ii) bromide/n,n,n,n,n -pentamethyldiethylenetriamine (CuBr 2 /PMDETA) under UV light resulted in the disappearance of the characteristic absorption band of the CuBr 2 / PMDETA complex centered around 640 nm. 438 This further confirmed the reduction of Cu II complexes to Cu I through the interaction of the photoexcited ligand with the central metal ion and unimolecular Cu-halogen bond cleavage. As a result, the bulk polymerization of MMA was successfully achieved using equivalent amounts of CuBr 2 /PMDETA with respect to the ATRP initiator irradiated under UV light. The process showed the characteristics of a living/controlled polymerization system, including a linear relationship between the monomer consumption and polymerization time, evolution of molecular weight according to the monomer conversion, and excellent chain end fidelity. Polymers with narrow molecular weight distributions were synthesized having good agreement with theoretical molecular weight values. Moreover, another study revealed the enhancement of the same process by adding small amounts of methanol probably due to the improved solubility of Cu facilitated by methanol and its likely part in the reduction of Cu II. 439 The level of Cu II catalyst loading was successfully decreased to ppm levels in 25% anisole as the solvent in a well

28 Chemical s controlled manner. 440,441 However, decreasing the catalyst loading further to 25 ppm resulted in the loss of control. Matyjaszewski and co-workers along with others investigated the photoinduced ATRP under visible or sunlight irradiation with ppm of Cu catalyst without the use of any conventional photoinitiator or reducing agent. 442,443 A series of ligand components including PMDETA, tris(2-pyridylmethyl) amine (TPMA), and its modified counterpart tris((4-methoxy-3,5- dimethylpyridin-2-yl)methyl)amine (TPMA*) was used to study the effect of the ligand. Mechanistically, photoinduced electron transfer of Cu II /L on irradiation resulted in photoreduction to Cu I and a halogen radical. In the presence of an ATRP initiator, the formed Cu I activator initiated the ATRP process, while halogen radicals formed during photoreduction were also able to initiate chain polymerization. Irradiation was conducted using various light sources. Polymerizations at 392 nm with LED provided a faster polymerization rate and wellcontrolled results as compared to the 450 nm. Sunlight irradiation also resulted in a well-controlled process, however, irradiations at 632 nm failed to initiate the polymerization, indicating inefficiency of the photoactivation at low energy ranges. Furthermore, in a recent study investigating the applicability of such photocatalytic system in aqueous media, the authors found that utilization of a low amount of catalyst resulted in a poor-controlled process, though high concentrations of Cu II catalyst gave well-defined polymers. 444 The poor control over the system in low concentrations of catalyst was attributed to the poor stability of the Cu II deactivator in aqueous media, which could be overcome by supplementing additional halide-containing salt to the system. As a result, utilizing the Cu II /TPMA catalyst as low as 22 ppm with halide salt resulted in a well-controlled process yielding water-soluble polymers with narrow molecular weight distributions and high chain end fidelity in aqueous media. Indeed, many different interactions are conceivable in such photochemical systems. Therefore, to elucidate the exact mechanism of the photochemical ATRP would require detailed consideration of every possible pathway. The group of Matyjaszewski and co-workers thus set up a detailed study of the mechanism of the photogeneration of activators. 445 The contribution and influence of every component present in the system that participates in photochemical reactions leading to the generation of the activator catalyst and finally initiation of polymerization was investigated. It was found that the use of equivalent amounts of ligand with respect to Cu failed to reduce Cu II species under irradiation at 390 nm and, subsequently, no polymerization was achieved. However, supplementing the ligand in excess amounts with respect to Cu (with a 6:1 ratio) induced the reduction of Cu II under irradiation and polymerization was then successfully initiated in a well-controlled manner. This further confirmed the necessity of the presence of uncoordinated ligand species for successful Cu II reduction and initiation of polymerization. Therefore, they concluded that the failure of the system to initiate polymerization where no free ligands were available in the reaction media, rules out the possibility of the photolysis of the Cu II complex to Cu I and the halogen radical as the dominant reaction pathway. The involvement of free, uncoordinated amine ligands was responsible for the photoinitiation processes. This was further proved by a control experiment using a copper(ii) triflate (Cu II (OTf) 2 ) catalyst instead of CuBr 2 with the excess amounts of ligand available. The reaction in the presence of the Cu II (OTf) 2 catalyst gave similar results to those obtained utilizing the CuBr 2 system. Since there was no Cuhalogen bond in Cu II (OTf) 2 to cleave, this observation further ruled out the homolytic Cu halogen bond cleavage upon irradiation as the only dominant mechanism. Interestingly, the nature of the excess amounts of the amine ligand was found to be independent of the type of the amine ligand. Rather it was the concentration of the tertiary amine groups that was important to promote photochemical initiation of ATRP. For this purpose, TEA was used instead of excess amounts of Me 6 TREN, and almost the same polymerization results were achieved in both cases. Together with experimental results and simulation studies, they drew various photochemical pathways with different contributions in the photoinitiation mechanism. The dominant reduction mechanism was suggested to be through free tertiary amine species (either uncoordinated tertiary amine ligand or other tertiary amine compounds) by a photoinduced electron transfer mechanism generating the Cu I complex and amine-centered radical cation. This mechanism resembled a photochemically mediated ARGET ATRP process. A proton transfer by the amine radical cation was suggested to form protonated amine and free radical species capable of initiating polymerization. 446 Photochemical generation of radicals by alkyl halide, ligand, or ligand and monomer was also suggested to be involved in the initiation of the reaction though with very low portions. This mechanism was similar to the photochemical ICAR ATRP process. The general proposed mechanism is depicted in Scheme 30, which was suggested to Scheme 30. Proposed Mechanism of Photoinitiated ATRP by Matyjaszewski and Co-Workers predominantly involve photochemically mediated ARGET ATRP plus a small contribution by photochemical ICAR ATRP processes. Control over the reaction was established by activation deactivation reactions as in classical ATRP. With the experimental results and kinetic simulations, the authors were able to evaluate the contribution of these reaction pathways during the photochemical ATRP process. As explained in the text, the dominant reaction between the Cu II and free amine ligands contributes 90% to the radical (re)generation and Cu reduction. The direct formation of radicals involving alkyl halide or ligand was calculated at about 1%. The contribution of reaction between ligand and macromolecular alkyl halide was found to be 8% of the activator regeneration. Figure 1 shows the fraction contributions of the explained mechanisms. Haddleton and co-workers developed an elegant approach for the synthesis of functional polymers via photoinitiated LRP. 447 A CuBr 2 /Me 6 TREN complex was used as the catalyst, which under irradiation at UV or even visible light was found to photochemically initiate and control the polymerization of a 10239

29 Chemical s Figure 1. Fraction contributions of activator regeneration from each reaction considered for the simulated polymerization of methyl acrylate with CuBr2/Me6TREN irradiating with 392 nm LEDs. Reprinted from ref 445. Copyright 2014 American Chemical Society. series of vinyl monomers. Irradiating a solution of CuBr 2 / Me 6 TREN in DMSO (50%) containing the acrylate monomer in the presence of alkyl halide under visible light initiated the polymerization in a well-controlled manner. An induction period of 3 h was observed, after which a linear relationship between monomer consumption and reaction time demonstrated a well-controlled/living process. Remarkably, the rate of the polymerization process was considerably enhanced when irradiating the system under UV light at 360 nm. Polymerizations proceeded within a 15 h period under daylight, whereas using the artificial UV light, the system was reached to completion within 90 min irradiation. No polymerization was achieved when the CuBr 2 /Me 6 TREN catalyst was used in a 1:1 ratio. However, increasing the ratio of the Me 6 TREN specie from 1 to 2, 3, and 6 (with respect to CuBr 2 ) successfully initiated and controlled polymerization of acrylates. This indicated that the presence of free ligand compounds were essential for the reduction and initiation of polymerization. UV vis spectroscopy was used to investigate the behavior of the polymerization system under irradiation. The CuBr 2 / Me 6 TREN (1:6) solution in DMSO exhibited a strong characteristic absorption band at 950 nm as well as 750 nm, which was attributed to the d-d transitions of the d 9 Cu II complex. Irradiation of this solution under UV light for 90 min resulted in significant reduction of the intensity of the absorption bands, indicating the successful reduction of the Cu II complex. Approximately about 75% reduction of Cu II was achieved. However, continuing irradiation resulted in no further reduction of the Cu II absorption band. Even after long irradiation times (up to 72 h), it remained as such. Moreover, changing in the experimental conditions had effects on the optical behavior of the system. For example, in the presence of alkyl halide, no reduction was detected after a 90 min irradiation. This behavior was attributed to two possible mechanisms. It was suggested that in the presence of alkyl halide it could undergo interaction with the reduced Cu I complex, thus oxidizing it back to the initial Cu II state. Moreover, the photoexcited free ligand species was suggested to preferably interact with the alkyl halide instead of inducing the reduction of Cu II ; this was a more reasonable explanation because in the presence of monomer almost the same behavior was detected (there was no significant reduction in the absorption of Cu II ) (Figure 2). So based on UV vis spectroscopic investigations and systematic control experiments, they proposed an initiation mechanism based on the photoexcitation of the free, uncoordinated Me 6 TREN species which was followed by interaction with the alkyl halide. The reaction of the photoexcited Me 6 TREN with the alkyl halide was through an outer-sphere-single-electron-transfer mechanism resulting in the homolysis of the alkyl halide. This further gave rise to the formation of the required initiating radical and Me 6 TREN radical cation species with a Br counterion. While initiating polymerization, the resultant radicals were proposed to interact with Cu II species, reducing them to Cu I and Figure 2. UV vis changes of the CuBr 2 /Me 6 TREN system under various conditions. Reprinted from ref 447. Copyright 2014 American Chemical Society

30 Chemical s affording dormant polymer chains. The reduced Cu I complex was believed to undergo some possible reactions, including reaction with amine radical cations or disproportionation processes. The proposed mechanism is depicted in Scheme 31. The effect of ligand was crucial in that excess amount of the Scheme 31. Proposed Mechanism of Photoinitiated LRP by Haddleton and Co-Workers ligand was required to initiate the polymerization. More importantly, using TREN which was similar to Me 6 TREN in structure gave almost the same result as that of Me 6 TREN. However, a loss of control was observed in the case of PMDETA as the molecular weight distribution of the polymers were quite high with respect to Me 6 TREN (1.27 and 1.05, respectively) with low monomer conversions (48% vs 98%, respectively). And no polymerization was initiated in the presence of bpy, indicating the crucial role of aliphatic tertiary amine groups in the photoinitiation process. The same group has also designed new Cu catalyst precursors, including copper(ii) formate 448 and copper(ii) gluconate, 449 to conduct photoinduced LRP. The mechanism was proposed to follow reduction of Cu II by similar photoinduced electron transfer processes of counteranion species acting as reducing agents with Cu II to activate the catalyst and initiate the polymerization. Recently, Barner-Kowollik, Haddleton, and co-workers attempted to elucidate the exact mechanism of the photoinitiated LRP process using high-resolution mass spectrometry technique. 450 Combined pulsed-laser polymerization (PLP) technique and electrospray-ionization mass spectrometry (ESI-MS) 454 were used to initiate the polymerization and analyze the resultant compounds, respectively. Each of the participating compounds, namely CuBr 2 /Me 6 TREN and ethyl α-bromoisobutyrate (EBiB), was independently irradiated in a solution of MMA in DMSO to draw a clear idea of whether and to what extent they participate in the photoinitiation process. Due to the low monomer conversions obtained, the authors were able to give an insight into the photoinitiation and earlier stages of polymerization processes. No initiation was observed when irradiating MMA alone in DMSO. However, pulse-laser irradiation of EBiB in MMA/DMSO ([EBiB]: [MMA] = 1:50) in the absence of Cu catalyst initiated the polymerization of MMA. A C Br bond cleavage of the ATRP initiator upon irradiation was suggested to the formation of initiating radicals. Analysis of poly(methyl methacrylate) (PMMA) formed in this course revealed that the polymerization was initiated by both alkyl and bromine radicals. All compounds detected in ESI-MS are shown in Scheme 32. Irradiation of Me 6 TREN in a solution of MMA/DMSO and in the absence of Cu and ATRP initiator also initiated polymerization. Mass spectroscopy analysis confirmed polymerization initiated by Me 6 TREN radicals as well as radicals formed on MMA. Photoexcited Me 6 TREN species were able to abstract a hydrogen atom from the α-position of the ligand, forming an initiating free radical. On the other hand, this hydrogen abstraction could well be achieved by MMA, leading to the formation of initiating free radicals on MMA. Moreover, irradiation of a CuBr 2 /Me 6 TREN solution in MMA/DMSO revealed that the polymerization was exclusively initiated by Me 6 TREN radicals, and no initiation by MMA radicals (as in the former case) was detected. The initial green color of the Cu II solution turned transparent after the irradiation, indicating the successful reduction of Cu II complexes to Cu I. Photoexcited Me 6 TREN species instead of abstracting a hydrogen atom from MMA were shown to reduce Cu II to Cu I by electron transfer reactions and form amine radical cations as well as bromine radicals, which could each contribute to the initiation or reduction processes. Irradiation in the presence of EBiB, Me 6 TREN, and MMA in DMSO, but in the absence of CuBr 2, Scheme 32. Chemical Structures Found in ESI-MS, Resulting from Various Experimental Conditions 10241

31 Chemical s initiated the polymerization by alkyl and Me 6 TREN radicals, whereas no polymers were detected polymerized by bromine or MMA radicals. The bromine radicals were suggested to participate in termination or hydrogen abstraction from the amine. Interestingly, upon irradiation of EBiB, CuBr 2, and MMA in DMSO in the absence of Me 6 TREN ligand, no polymerization was observed. The formed alkyl and bromine radicals were believed to rapidly interact with the Cu II reducing it to Cu I. While reducing Cu II to Cu I, EBiB and elemental bromine (Br 2 ) were also formed under these conditions. The initial green color of the solution of CuBr 2 turned to pale yellow, indicative of the reduction of Cu II and formation of Br 2 species. Finally, in the presence of all components (i.e., MMA, EBiB, CuBr 2, and Me 6 TREN in DMSO solution), the polymerization was initiated only by the alkyl radicals and no polymers initiated by bromine, Me 6 TREN, or MMA radicals were detected. A well-controlled polymerization process was established under these conditions with polymers having narrow molecular weight distributions (<1.3). The authors were also able to detect the formation of Cu I species as a result of the reduction of Cu II in UV vis measurements as new peaks arose at nm. However, this detection, as mentioned by the authors themselves, is highly dependent on the experimental and measurement conditions and would not be clearly seen under every system employed. The proposed mechanism is depicted in Scheme 33. It involves the light-induced C Br bond of the initiator Scheme 33. Proposed Mechanism of Photoinitiated LRP by Barner-Kowollik and Co-Workers dissociating to form alkyl radicals and bromine radicals capable of initiating polymerization and reducing Cu II species, respectively. Photoinduced electron transfer reactions between the photoexcited amine and Cu II also induced the reduction of Cu II to the activator Cu I while forming an amine radical cation. It was also suggested that the Cu II complex can undergo photoexcitation and subsequently be quenched by free ligand species, resulting in the formation of Cu I and amine radical cation species. These radical cation species can form initiating radicals by a deprotonating process. It should be noted that these mechanisms are drawn based on the overall behavior of the system on early stages of the process, and it remains to further explore whether they remain in order of activity while the reaction proceeds to high monomer conversions and high molecular weight polymer chains. In photochemical strategies for conducting ATRP, it is generally considered and mechanistically proved that only Cu II complexes participate in photochemical reactions. And no contribution or participation of Cu I species in photochemical reactions was recorded in such systems. Recently, a new strategy has been developed that was based on the photoactivation of the Cu I complex to induce the ATRP system. 455 A catalyst system consisting of copper(i) bis(1,10-phenanthroline) (Cu I (phen) 2 ) was used. Being highly active at the visible region, the Cu I (phen) 2 catalyst underwent photoexcitation under LED irradiation, which was subsequently quenched by the alkyl halide initiator. As a result, the catalyst was oxidized to its higher oxidation state (Cu II (phen) 2 Br) while forming initiating radicals. The (photo)catalytic cycle can then follow as shown in Scheme 34. In the presence of an amine compound Scheme 34. Photoinduced ATRP by Cu I (phen) 2 (Photo)catalyst such as TEA, a fast process was observed, which was enabled due to the participation of TEA in the regeneration of the activator Cu I and even providing the use of 20 ppm amounts of catalyst while still maintaining a good control over the polymerization process. Irradiation seemed necessary to the initiation of ATRP as the process was inefficient in the dark even within a long period of reaction times. Photoactive compounds have also been taken advantage of for photochemical initiation of ATRP reactions. Photoinitiators, photosensitizers, photocatalysts, etc. are examples of additional photoactive compounds used in this context. These systems are aimed to generate the activator Cu I catalyst with the aid of photochemically formed active compounds by using a photoinitiator or other photoactive compounds. Initiator-chain transfer-termination (iniferter) polymerization is a technique which uses a photoactive dithiocarbamate to photochemically initiate radical polymerization. Dithiocarbamates (DTC)s possess a C S bond readily cleavable by light and have been used to initiate and control radical polymerization processes Reversible deactivation of propagating radicals establishes control over the polymerization in iniferter polymerization. It has been found that combination of the iniferter technique with ATRP processes can overcome some drawbacks associated with these techniques and makes it possible to synthesize photolabile polymers Mechanistically, upon UV irradiation, DTC molecules rapidly generate carbon-centered radicals that induce radical propagation and are deactivated by the Cu complex. This deactivation by Cu helps to set further control over the chain growth. Photochemical approaches based on conventional photoinitiators, photosensitizers, or dyes are also able to initiate and gain control over the ATRP process. The use of photosensitive compounds allows for generation of free radicals with high electron-donating properties by means of photochemical 10242

32 Chemical s methods that are able to reduce deactivator Cu II to activator Cu I complexes. For example, Type I photoinitiators that produce free radical species upon irradiation by bond cleavage have been used for the initiation of ATRP reactions In addition, photoinitiated ATRP reactions have also been conducted with Type II photoinitiators which form free radicals by interacting with various co-initiator compounds through photoinduced electron transfer and/or hydrogen abstraction processes. 468 Similarly, dyes can form radical species by photoinduced electron transfer or hydrogen abstraction reactions as in Type II photoinitiation (Scheme 35). In a Scheme 35. Photoinitiated ATRP with Various Photoinitiating Systems nanoparticles exhibited no absorption characteristics. Photoinduced electrons from dye molecules to the conduction band of the nanoparticles sensitized nanoparticles inducing charge separation and subsequently electron releasing. Another similar hybrid photocatalyst system for ATRP was developed based on TiO 2 /reduced graphene oxide nanocomposites. 473 It was aimed to enhance the photocatalytic activity of TiO 2 by suppressing electron hole charge recombination using graphene oxide. It was suggested that due to their specific nature of higher electron mobility of graphene oxide, the recombination of photogenerated electron hole pairs was prevented. ZnO nanoparticles were also used for the photochemical induction of the ATRP reaction. 474 The reduction of Cu II was achieved using these nanoparticles irradiated under UV light as the characteristic absorption band of the Cu II complex at the nm region decreased to a significant extent, indicating a successful photoreduction process (Figure 3). 475 comparative study, it was found that dyes such as eosin Y and erythrosin B can efficiently photoinitiate the ATRP process. 464 However, control in the dye-sensitized ATRP system was relatively poor with respect to Type I photoinitiated ATRP using the bisacylphosphine oxide photoinitiator. Dimanganese decacarbonyl (Mn 2 (CO) 10 ) was also used as a visible light promoter for ATRP. 469 Photochemically formed mettallo radicals are highly reactive with regard to the generation of initiating radicals by halogen abstraction from the ATRP initiator and can participate in Cu reduction as well. Semiconducting photocatalysis can also be applied to photochemically initiate and control the ATRP process. As previously stated, semiconductor materials are capable of releasing electron and hole charge carriers upon photoexcitation. Photoinduced electrons thus released in this manner can be efficiently used to trigger the photoreduction of Cu II complexes in ATRP reactions. Zhou, Liu, and co-workers used TiO 2 nanoparticles as a heterogeneous photocatalyst for photoinduced ATRP. 470 An aqueous solution of CuCl 2 /bpy in water/methanol containing heterogeneous TiO 2 nanoparticles was irradiated under UV light. The reduction of Cu II species by photoinduced electrons from the nanoparticles was clearly detected as the initially light blue color of the solution turned dark brown after irradiation. The dark brown solution of the Cu I complex turned back to the initial light blue color upon exposure to air. Upon exposure to the atmosphere, Cu I was oxidized to Cu II species and this reduction/oxidation process can be repeated at set intervals by exposing it to light and air. The process was utilized for surface-initiated ATRP in an aqueous environment for polymer brush growth in a controlled manner. The holes were considered to undergo oxidation by methanol present in the reaction medium. 471 A similar photosensitization approach utilizing dye species along with TiO 2 nanoparticles was designed so as to extend the optical characteristics of the system up to the visible light regions. 472 In this regard, dye molecules acting as the visible light absorbing chromophores were exploited to sensitize the charge separation of TiO 2 nanoparticles under visible light irradiation where these Figure 3. UV vis spectra recorded during the photoreduction of Cu II species to Cu I using ZnO nanoparticles as heterogeneous photocatalyst. Reprinted with permission from ref 475. Copyright 2015 Wiley-VCH Verlag GmbH & Co. A key feature of these systems using semiconducting photocatalysis is the heterogeneous nature of the nanoparticles, which makes them easily separable and removable from the reaction media without causing any contamination to the main reactions. Also sized in the range of nm in diameter, semiconductor nanoparticles possess a large surface area for light harvesting. As an alternative to inorganic semiconducting nanoparticles, mpg-c 3 N 4 was used as a heterogeneous organocatalyst. 476 In a similar manner to the inorganic TiO 2 or ZnO nanoparticles, the reduction of Cu II complexes was believed to proceed upon receiving photoinduced electrons released from mpg-c 3 N 4. This system was active under both UV light and natural sunlight irradiation, and irradiation under sunlight resulted in a slightly more controlled process. The general mechanism of the photoinitiated ATRP by using semiconducting photocatalysts is depicted in Scheme Non-Copper-Based Photocatalytic Systems. The photocatalytic activity of transition metal complexes 477,478 has led to significant developments in the synthetic polymer community in developing new (photo)catalytic systems for ATRP reactions that are catalyzed and controlled by photoactive transitional metal complexes other than Cu. The ability of Ir- or Ru-based complexes as alternatives for Cu catalysts has been effectively investigated. In this regard, Hawker and co

33 Chemical s Scheme 36. Photoinitiated ATRP Using Semiconducting Nanoparticles as Heterogeneous Photocatalyst workers developed an efficient Ir-based photoredox catalytic system, namely fac-[ir(ppy) 3 ] (ppy =2-pyridylphenyl), for the photochemical initiation and control of the ATRP reaction. 479,480 A well-controlled, living polymerization system was established using fac-[ir(ppy) 3 ] catalyst in the presence of alkyl halide initiators. The fac-[ir(ppy) 3 ] catalyst is highly photoactive at visible regions and undergoes photoexcitation under visible light irradiation. Mechanistically, it was argued the photoexcited fac-[ir(ppy) 3 ]* complex was quenched by the alkyl halide initiator in an oxidative quenching cycle. This resulted in the formation of initiating radical species which in the presence of monomer initiated the polymerization, as well as forming a highly oxidized form of the catalyst (Ir IV ). The catalytic cycle could then follow to completion by back halogen transfer from the oxidized Ir IV to the propagating radicals bringing about the initial ground state fac-[ir(ppy) 3 ] catalyst and dormant species (Scheme 37). The catalyst loading was photoinitiation mechanism concerning the reduction of Fe III species which required the presence of MMA was proposed. Another approach utilized cyclometalated ruthenium(ii) complexes as the photocatalyst to promote ATRP under visible light irradiation A photoinitiation involving formation of the solvento 18-electron species of cis-[ru(o-c 6 H 4-2-py) (phen) (MeCN) (acetone)] + under irradiation through the intermediacy of the 16-electron five-coordinated complex cis- [Ru(o-C 6 H4 2-py) (phen) (MeCN)] + was believed to be a crucial intermediate of the overall ATRP process. Surface-modified niobium hydroxide nanoparticles were employed in a photoinduced ATRP system. 496 Upon irradiation of the nanoparticles, ligand-to-metal electron transfer followed by a hydrogen atom abstraction took place, yielding an anion and an alkoxy-substituted radical. The radical thus generated was able to reduce the alkyl halide initiator to form initiating radicals, as well as bromination of the alkoxy radical. Deactivation of propagating radicals by back bromine transfer forms dormant chains establishing control over the process. The initial ground state nanoparticles were regenerated through deactivation of the propagating radicals with the ability to reinitiate the polymerization process in a controlled fashion. The nanoparticles due to their heterogeneity were recycled and reused for several times (Scheme 38). Scheme 38. Photoinduced ATRP by Niobium Hydroxide Nanoparticles Scheme 37. Photoinitiated ATRP Using the fac-[ir(ppy) 3 ] Photocatalyst System found to directly influence the livingness and controllability of the process. Loading equivalent amounts of catalyst with respect to the initiator resulted in loss of control, forming polymers with high molecular weight distributions and poor control. This was so probably due to the fact that the Ir complex behaved as a photoinitiator rather than operating as the catalyst. However, reducing the catalyst loading to ppm levels led to good control over the process mediated by light. Several other studies have also reported the same strategy using Ir-based photoredox systems for carrying out photoinitiated ATRP reactions. 237, A dinuclear gold(i) complex has been shown to catalyze the ATRP process in a similar manner to the Ir photoredox system under visible light irradiation. 488 Analogous to Ir photoredox, the iron-based photocatalytic systems have been recently developed Matyjaszewski and co-workers obtained a well-controlled photoinduced process utilizing an iron(iii) catalyst under irradiation. 492 A Metal-Free Photoinitiated ATRP. A great deal of effort has recently been dedicated to developing and exploiting organocatalysis in controlled synthesis of polymers. Photoredox organocatalysis 497,498 is seemingly a promising candidate for replacing metallic catalysts for conducting ATRP or even other controlled polymerization techniques. Photoredox organocatalysis is principally based on the ability of small organic compounds to photochemically catalyze an array of chemical reactions. It bears advantages over transitional metal catalysis such as photoredox organocatalysis and is highly stable, relatively nontoxic, and can be availed from renewable resources with low-cost processes. In order to initiate and control the ATRP reaction, the organic photoredox catalyst should be highly reducing in its photoexcited state to be able to 10244

34 Chemical s activate the initiator alkyl halide in a reversible manner to gain control over the process. Despite the synthetic organic community enjoying the advantages of photoredox organocatalysis for decades, only a few works have recently emerged on its use in polymerization processes. In an attempt to take advantage of photoredox organocatalysis in controlled polymerization, Miyake and Theriot reported the use of perylene as a metal-free photocatalyst for ATRP. 499 Perylene being a strong reductant was previously reported as a photosensitizer in conjunction with other photoinitiators. Irradiating a solution of MMA in DMF containing perylene in the presence of methyl α-bromoisobutyrate (MBiB) as the alkyl halide initiator did afford polymers with relatively low molecular weight distribution. However, a considerable difference was found between theoretical and experimental molecular weight values, which were due to a low initiation efficiency of 6.9%. The use of ethyl α-bromophenylacetate (EBPA) resulted in an increase in the initiation efficiency. A photoinduced electron transfer by the photoexcited state perylene was proposed to directly reduce the alkyl halide initiator bringing about initiating radicals as well as the radical cation of perylene. Chain end group fidelity was successfully proved by MALDI-TOF and chain extension experiments for the formation of block copolymers. Hawker and co-workers designed and developed a more efficient approach utilizing phenothiazine compounds as a metal-free photoredox catalyst for ATRP Methylphenothiazine was able to initiate the polymerization of MMA in the presence of EBPA with broad molecular distribution. A phenylsubstituted phenothiazine (PTH) with a more favorable reduction potential (E red = 2.1 V vs SCE) was specifically synthesized. An efficient photoinitiation was obtained with PTH with good control over the process and molecular weight properties of the resulting polymers. The initiation mechanism was proposed to follow the reduction of the alkyl halide initiator by the photoexcited state catalyst through photoinduced electron transfers. This resulted in the activation of the initiator forming an initiating radical and a photocatalyst radical cation species. Deactivation was followed by back electron transfer reactions, regenerating the initial ground state photocatalyst and affording a dormant polymer chain with bromine end group functionality. This activation deactivation could follow a similar behavior to that described for metalphotocatalyst compounds used in ATRP. Apparently, no polymerization was achieved with photoinitiators such as methylene blue or eosin Y, which are oxidizing in the photoexcited state indicative of the necessity of a photocatalyst to be highly reducing in the photoexcited state. Compatibility of this system with other Cu- or Ir-based ATRP systems was successfully examined. A similar strategy utilizing a series of phenothiazine photocatalysts has also been reported by Matyjaszewski and co-workers (Scheme 39 and Table 4). 501 In this regard, it is worth mentioning the iodine-mediated living radical polymerization initiated by photoorganocatalyst compounds. 502 The C I bond was reversibly cleaved upon irradiation in the presence of the photocatalyst to initiate and control the living radical polymerization process. Although the reaction mechanism was not clear, it was speculated that energy transfer from the photocatalyst to C I bond results in its cleavage and radical formation. Using different photocatalysts active at different regions of the electromagnetic spectrum and a dual initiator selective polymerization of chemically different monomers could be achieved in a one-pot manner. The Scheme 39. Photoinduced Metal-Free ATRP Enabled by Photoorganocatalysts (POC) initiator containing hydroxyl and iodine functionalities was able to initiate radical polymerization of MMA in the presence of a photocatalyst at nm, while tuning the irradiation wavelength to nm initiated cationic polymerization of δ-valerolactone (VL) in the presence of a photo acid generator, triarylsulfonium hexafluorophosphate, active at nm. As a result, block copolymers of chemically different monomers were synthesized in a one-pot system using double photocatalysts by simply adjusting the irradiation wavelength, which could selectively initiate polymerization of certain monomers at certain wavelength regions Monomer Compatibility. A variety of different vinyl monomers have been polymerized using photoinduced ATRP reactions (Chart 6). Methacrylate monomers include methyl methacrylate (MMA), iso-butyl methacrylate (ibma), tert-butyl methacrylate (tbma), benzyl methacrylate (BzMA), hydroxyethyl methacrylate (HEMA), fluorinated monomers such as 2,2,2-trifluoroethyl methacrylate (TFEMA), 2-ethylhexyl methacrylate (EHMA), glycidyl methacrylate (GMA), and poly(ethylene glycol) methacrylate (PEGMA). Polymerization of N,N-dimethylaminoethyl methacrylate (DMAEMA), which was challenging in Ir or other systems due to its highly oxidative properties, was achieved in a well-controlled manner in metal-free ATRP. Water-soluble monomers with salt structures such as 3-sulfopropyl methacrylate potassium salt (SPMA) and 2-methacryloyloxyethyl phosphorylcholine (MPC) were also polymerized. Acrylamide monomers such as methyl acrylate (MA), ethyl acrylate (EA), tert-butyl acrylate (tba), lauryl acrylate (LA), octadecyl acrylate (ODA), isooctyl acrylate (ioa), 2-methoxyethyl acrylate (MEA), glycidyl acrylate (GA), solketal acrylate (SA), hydroxyethyl acrylate (HEA), poly(ethylene glycol) acrylate (PEGA), 2,2,3,3,3- pentafluoropropyl acrylate (PFPA), 3-(dimethylamino)propyl acrylate (DAPA), and [2-(acryloyloxy)ethyl]- trimethylammonium chloride (AETMAC) have been polymerized. Finally, acrylamide-based monomers, including N- isopropylacrylamide (NiPAAm) and N-(2-hydroxypropyl)- methacrylamide (HPMA), as well as styrene (St), acrylonitrile (AN), and methacrylic acid (MMA), are examples of various monomers compatible for photoinduced ATRP catalyzed by both transitional metal catalysts or metal-free photoorganocatalyst systems Ligands. Ligands play a crucial role in promoting the electron transfer reactions with the coordinated Cu ions during the photoreduction process. As explained previously, the structural properties and the amount of amines used as ligand molecule directly influence the efficiency of the photoinduced electron transfer reaction, especially in those systems utilizing no additional photocatalyst compounds. It has been shown that the presence of excess amine species is necessary for successful initiation and control of the process. Tertiary amines have been 10245

35 Chemical s Table 4. Results of Photoinduced Metal-Free ATRP Systems a a MBiB: methyl α-bromoisobutyrate. b EBPA: ethyl α-bromophenylacetate. extensively used and have been proven as efficient ligands. Of particular interest is Me 6 TREN utilized in many diverse systems with excellent controllability over the polymerization process. Other amines used as ligands in photoinduced ATRP reactions include PMDETA, TPMA and its modified version namely tris((4-methoxy-3,5-dimethylpyridin-2-yl)-methyl)- amine (TMPA*), among others. Moreover, the use of TiO 2 nanoparticles as photocatalyst enabled utilization of aromatic ligands such as bpy, which was shown to be inefficient as an electron-transfer promoter in other direct systems. Chart 7 presents amine ligands used in photoinitiated ATRP systems Complex Architectures and Sequence-Controlled Polymers. The development of macromolecular engineering has been directed toward reaching a point at which one can precisely control the exact composition of polymer chains and positioning of functional groups at the molecular level. Inspired by nature in producing precise macromolecular structures such as proteins, peptides, and nucleic acids, chemists have long sought ways to mimic the precision exhibited by nature for the formation of synthetic polymers with precisely controlled structures. 503 Cu-mediated controlled/living radical polymerizations are versatile and promising tools for the synthesis of sequence-controlled polymers, 504 and the potential of photochemical routs in that regard has been investigated as well. Anastasaki, Haddleton, and co-workers have expanded the scope of the photoinitiated living radical polymerization for the synthesis of a variety of functional, multiblock copolymers. Utilizing a CuBr 2 /Me 6 TREN catalyst with an excess free ligand, they showed that multiblock acrylate copolymers could be synthesized in a one-pot manner owing to the excellent control and chain-end fidelity brought about by the exploited photochemical protocol. 505,506 A variety of different acrylate monomers were used. The process involved the iterative addition of monomer in one pot, allowing the incorporation of monomer sequences with varied repeating units across polymer chains while having narrow molecular weight distributions. Significantly, no purification was required between monomer additions as quantitative monomer conversions were achieved in each step. Notably, to maintain a good control along the process, the CuBr 2 /Me 6 TREN complex with specific mole ratios needed to be fed during the iterative additions due to the probable consumption of the necessary amines for a successful photoinitiation process. Moreover, it was found that elevated temperatures had a negative effect in maintaining control over the process as conducting reactions at 50 C resulted in higher molecular weight distributions underlying poor control; however, narrow molecular weight distributions and excellent control were obtained at ambient temperatures. In spite of wellcontrolled synthesis of block copolymers comprised of up to 5 blocks with 100 repeating units in each block having molecular weights of about g mol 1 and a molecular weight distribution of 1.21, further chain extensions resulted in loss of control and failed to reach higher molecular weights and more number of blocks. This limitation in reaching multiple blocks and higher molecular weights could be avoided by using bifunctional initiators furnishing α,ω-telechelic multiblock copolymers. In this case, multiblock copolymers of up to 23 blocks with two repeating units in each block, and also high molecular weight undecablock with a molecular weight of g mol 1, were obtained which had molecular weight distributions as narrow as Finally, the versatility of this approach has been examined in different environments and solvents especially in ionic liquids with high efficiency. 508 A similar approach has also been taken by the group of Junkers, who synthesized multiblock acrylate oligomers and copolymers with the potential in self-assembly applications. 509,510 Moreover, the process was adapted in continuous flow processes, whereby homo and copolymers were efficiently synthesized in 20 min of reactor residence time while reaching quantitative monomer conversions and good control over the molecular weight properties Spatiotemporal Control. By far and large, the most important aspect of such photoinitiated controlled polymerizations is the ability to gain control over the polymerization reaction in terms of temporal and spatial control. In conventional polymerization techniques (i.e., free radical and cationic polymerizations), the role of light is to generate initiating species by photochemical means. Therefore, it is safe to consider its role in the early stages of polymerization in the 10246

36 Chemical s Chart 6. Monomer Compatibility in Photoinduced ATRP Reactions Chart 7. Amine Ligands Used in Photoinduced ATRP Reactions formation of initiating sites without any considerable effect on the propagation or termination steps. However, integration of the concepts of photochemistry with controlled/living radical polymerizations enables chemists to gain control of the process by means of light. It is not therefore restricted only to the formation of initiating species, but the ability to activate deactivate the catalyst by photochemical means opens new opportunities for controlling all the initiation, propagation, and termination steps simply by light. The temporal control of the ATRP process enabled by light is based on the activation of the catalyst and, hence, the reaction by light that can be also deactivated when the light source is removed. Almost in every photoinitiated ATRP systems it is possible to temporarily control the propagation of polymer chains. For this purpose, a reaction tube is subject to intermittent exposures to light and dark. After exposing the reaction mixture for a specific period of time, the polymerization initiates and reaches to some conversions. At this point, turning off the light and keeping the reaction in the dark causes a stop in the propagation of chain growth as almost no or a very low amount of molecular weight evolution is observed in the kinetics studies. Interestingly, re-exposing the same reaction mixture to light reinitiates the reaction that can be stopped again by keeping it in the dark. This repeated light on/off experiments can be made to initiate polymerization by exposure to light while stopping the propagation at any point by switching off that light source. Equally important is spatial control. Spatially controlled functionalization of surfaces and fabrication of polymer brushes with complex architecture by using surface-initiated polymerization techniques have found considerable use in a diverse range of applications. 512,513 In particular, photochemistry is a versatile tool for surface functionalization using controlled polymerization methods. For example, a uniform surface functionalized with the ATRP initiator can be used to grow polymer brushes on the surface of the functionalized substrate by irradiating light in the presence of suitable photocatalyst systems. Moreover, by using a photomask one can spatially control functionalization or obtain chemical concentration gradients. Many researchers have successfully developed novel strategies to establish spatial control and synthesize polymer brushes by merging photochemistry and controlled/living polymerization techniques. Being driven and controlled by light, the growth of polymer chains can be temporarily and spatially manipulated by light. For instance, using TiO 2 semiconductor nanoparticles to induce reduction of Cu catalyst, Zhou and co-workers have grown a series of polymeric brushes of water-soluble monomers from the ATRP initiator functionalized surfaces in a controlled manner with the ability to make copolymer brushes. 470,472 Furthermore, they functionalized the surface of TiO 2 nanowire with the ATRP initiator to grow polymer brushes on TiO 2 nanowires which could also act as the 10247

37 Chemical s Scheme 40. Typical Representation of Surface Initiated Polymerization Using fac-[ir(ppy) 3 ] Photoredox System for the Formation of 3D Objects: (a) Optical Micrograph of Nanoscale-Inclined Plane Formed from a 3D Polymer Brush; (b) 3D AFM Image of Nanoscale-Inclined Plane. Reprinted with permission from ref 518. Copyright 2013 Wiley-VCH Verlag GmbH & Co photocatalyst for the catalyst activation. Jordan and co-workers used a system comprised of CuBr 2 and PMDETA catalyst under irradiation with a household lamp to grow polymer brushes. 514 Successful spatiotemporal control was shown in the formation of a variety of brush block copolymers and photopatterning of functionalized surfaces. Also photoinduced LRP in the presence of CuBr 2 /Me 6 TREN was employed to grow polymer brushes of N-(2-hydroxypropyl)-methacrylamide with potential application as nonfouling agents. 515 Similarly, grafting from cellulose surfaces was reported using the photoinduced surface-initiated polymerization. 516 Recently, Junkers and co-workers demonstrated that such photocontrolled polymer brush formation can be photochemically triggered and controlled by very low Cu catalyst loadings down to a few parts per billion (ppb), reaching to unprecedented film thicknesses of up to 1 μm with a wide scope of monomer compatibility and block copolymer formation. 517 Indeed, such high film thickness was achieved for bulky monomers such as 2- ethylhexyl acrylate or t-butyl acrylate because of their higher excluded volume interactions resulting in less accessible radical centers and hence reduced termination rates. The Ir-based photoredox system pioneered by Hawker and co-workers has been shown particularly suitable for a wide range of photopatterning purposes. Formation of the polymer brush was observed from a substrate functionalized with an ATRP initiator in a temporarily and spatially controlled manner using the fac-[ir(ppy) 3 ] photocatalyst Block copolymer brushes and post modification were conducted to demonstrate the ability of this system to tuning the surface properties of 3D features. Of great importance was the fabrication of complex 3D nanostructures by adjusting the intensity of light where, in turn, the height of polymer brushes being grown can be directly determined. A grayscale lithography mask was used to expose the initiator-functionalized substrate to various light intensities, thus formation of brush polymers with different molecular weight being proportional to the intensity of light was observed. As a result, 3D nanostructures including inclined planes, microprisms, gradients, and arrays of microlenses were obtained (Scheme 40) Reversible Addition Fragmentation Chain Transfer (RAFT) Polymerization by Photoinduced Electron Transfer Reactions The RAFT polymerization is a highly versatile technique for the synthesis of well-defined macromolecular architectures, and a good deal of effort has been dedicated toward understanding the chemistry of the RAFT process and its use in a wide range of areas. 521,522 It was first developed in 1998 by Moad, Rizzardo, Thang, and co-workers 523 and since then has been the focus of various developments in macromolecular engineering. 420, Similar to conventional free radical polymerization, it uses a radical initiator to add to a double bond and initiate the polymerization. However, distinctive from conventional free radical polymerizations is the use of a chain transfer agent, which is also referred to as the RAFT agent, to carry on the polymerization in a controlled manner. As the RAFT agent thiocarbonylthio compounds are widely used. Mechanistically, the polymerization process is initiated by a radical source as in conventional radical polymerizations. In the presence of the RAFT agent, the propagating radical adds to the RAFT agent to form a radical intermediate which can subsequently fragment to generate a new initiating radical and a thiocarbonylthiocontaining polymer chain. The radical formed can initiate new polymer chain growth to form a new propagating radical. The main RAFT equilibrium of activation deactivation is established by degenerative chain transfer between propagating and dormant polymer chains. Photochemical approaches have been employed to initiate and control the RAFT polymerization by means of light. The group of Boyer and co-workers has pioneered versatile and robust techniques for carrying out RAFT processes by exploiting photoinduced electron transfer reactions. In order to activate the polymerization and form initiating species, they used various photochemical routs together with RAFT agents instead of exogenous radical sources, which are prone to affect the livingness of the process and form dead polymers. The fac- [Ir(ppy) 3 ] was used as the photoredox catalyst which can interact in the excited state with the thiocarbonylthio compounds via photoinduced electron transfer reaction. 529,

38 Chemical s Scheme 41. Proposed Mechanism for the RAFT Polymerization Enabled by Photoinduced Electron Transfer Processes by Using Photoredox Catalysis (PC) This interaction reduced thiocarbonylthio agents, which acted as both the initiator and the chain transfer agent, generating initiating radicals while oxidizing the Ir III to Ir IV complexes. The initiating radicals formed may initiate the polymerization and operate the RAFT process or they may participate in the catalytic cycle to regenerate the initial state catalyst while being deactivated. Efficient photoinduced electron transfer between the photoexcited Ir photocatalyst and the thiocarbonylthio agents was proved by fluorescence quenching (or Stern Volmer quenching) studies given the redox potential of the Ir photocatalyst being 1.72 V (versus SCE) and that of the studied RAFT agents higher than that value ( 0.4, 0.6, and 0.9 V vs SCE). Irradiating the fac-[ir(ppy) 3 ] photocatalyst in the presence of different RAFT agents under visible light, welldefined polymers were obtained with narrow molecular weight distributions in a highly controlled manner. The versatility of this technique was successfully investigated on a variety of monomer families. Under optimized experimental conditions, conjugated monomers, including methacrylates, acrylates, styrene, (meth)acrylamide, and isoprene monomers, were successfully polymerized in a well-controlled manner. Moreover, the approach showed good tolerance and efficiency in polymerizing unconjugated monomers such as vinyl acetates and others. Homopolymers with ultrahigh molecular weight of up to g mol 1 was successfully synthesized by optimizing experimental condition, and copolymers with molecular weight of up to g mol 1 and molecular weight distribution of 1.30 were achieved by chain extending macrophotoinitiators prepared by this technique. Also, multiblock copolymers with high molecular weight and narrow distributions were synthesized by iterative monomer addition indicative of high chain-end fidelity preserved throughout the process. The process was driven and controlled exclusively by light as evidenced by light on/off experiments that demonstrated the occurrence of the polymerization under visible light while being stopped in the dark. More importantly, the photoinduced RAFT process based on electron transfer reactions of photoredox catalysts and RAFT agents has been demonstrated to effectively operate even in the presence of oxygen considering the strong reduction properties of the photoredox catalysis could reduce oxygen molecules to inactive superoxide species. 531 Of course, the process in the presence of oxygen proceeded after an induction period of 3 24 h of oxygen consumption depending on the concentration of the catalyst and other experimental conditions; however, almost the same kinetic results were observed in the absence and presence of air. The livingness of the process in the presence of the air was also another significant feature of this approach. These results clearly indicated high efficiency of the photoinduced RAFT processes using photoredox catalysis in furnishing wellcontrolled macromolecules of different types and functionalities. The overall mechanism is illustrated in Scheme 41. Additionally, the efficiency of other photoredox catalysts in promotion of the RAFT polymerization via photoinduced electron transfers has also been examined. Ru II complex, for example, was successfully employed for the polymerization of water-soluble monomers in aqueous media. 532 Successful photoinduced electron transfer of the photoexcited Ru II and thiocarbonylthio compounds in water was responsible for the initiation of different water-soluble monomers. The process was well-controlled and furnished polymers with molecular weight values in good agreement with the theoretical values and narrow molecular weight distributions. The living nature of the system was successfully valued for the block copolymer formation. Significantly, the optimization of the reaction conditions allowed for the conduction of polymerizations in biologically relevant media, and also protein polymer bioconjugates were efficiently synthesized in a controlled manner. Boyer and co-workers subsequently utilized photoorganocatalysis in photoinduced RAFT polymerization. In this regard, chlorophyll a, which is the principal biobased component of photosynthesis in green plants, was proved an efficient promoter and mediator of the RAFT process. 533 It was extracted from spinach, a renewable bioresource, and used at ppm levels under both blue and red lights. Other nontransitional-metal photoorganocatalysts include eosin Y, fluorescein, and porphyrin moieties. 534 Amines could be used as sacrificial electron donors to reduce oxygen inhibition. Various monomer families including (meth)acrylates, (meth)acrylamides containing a large variety of functional groups, such as carboxylic acid, amine, alcohol, and glycidyl groups, were polymerized with excellent control over the process and molecular weight properties of the resulting polymers. Likewise, Johnson and co-workers reported the use of 10-phenylphenothiazine as the photoorganocatalyst in the presence of trithiocarbonate iniferter agents to afford well-defined polymers through electron transfer reactions in the RAFT process. 535 Moreover, surface-modified niobium hydroxide nanoparticles promoted the RAFT process by photoinduced electron release upon irradiation of the nanoparticles activating the RAFT agent. 536 Finally, tertiary amines were shown to catalyze the photoinduced RAFT polymerization undergoing electron transfer reactions with RAFT agents under irradiation. 537 As previously suggested in photoinduced ATRP reactions whereby tertiary 10249

39 Chemical s amines (i.e., free ligand species) promote the photoinduced electron transfer reactions to activate the Cu catalyst and initiate ATRP reaction, in this case, also similar photoinduced electron transfer was encountered but to the RAFT agent. Consequently, an intermediate of the amine radical cation and trithiocarbonate radical anion is formed, which subsequently resulted in the formation of initiating radicals and trithiocarbonate anion species. Degenerative chain transfer followed to control the polymerization via the RAFT process, as previously outlined in Scheme 41. Amine molecules including TEA, PMDETA, and Me 6 TREN were used as the photocatalyst compounds. It was also suggested that covalently binding the RAFT agent and photoredox components can significantly enhance the polymerization process through facilitating the electron transfer processes. 538 The RAFT agents as electron acceptors were conjugated with the photoorganocatalyst porphyrin as the electron donor components to yield electron donor acceptor systems. The spacer length was found to directly influence the electron transfer, and high efficiency was achieved in those systems with the donor acceptor components in as close proximity as possible throughout the reaction. Finally, metalloporphyrin photocatalysts such as zinc porphyrin were shown to selectively activate the RAFT polymerization under light. 539 The zinc porphyrin is a highly photoactive compound exhibiting absorption characteristics in a wide range wavelength from blue to green to orange and to red lights. The selective activation of thiocarbonylthio compounds by the zinc porphyrin photocatalyst through electron transfer was observed which indicated that the activation of trithiocarbonate and consequently initiation of the polymerization was much more efficient than dithiobenzoate compounds. This selective activation was attributed to the possible interaction of the photocatalyst zinc porphyrin and trithiocarbonate agents leading to efficient photoinduced electron transfer. Indeed, such an exceptional behavior was proposed as a novel form of molecular recognition, which is unique to natural processes in biological events. In addition to temporal control by switching on/off the light source, the kinetics of the polymerization was found to be different under different wavelengths. This behavior makes the zinc porphyrin photocatalyst unique in controlling the kinetics of controlled/living polymerization by simply tuning the wavelength of the irradiation light using a single catalyst. Another striking feature of the system developed based on zinc porphyrin photocatalyst was the ability to carry out the RAFT polymerization in a vessel fully open to air which has rarely been achieved in other controlled radical polymerizations. Recently, photoinduced RAFT polymerization was conducted by irradiation under high-wavelength, low-energy nearinfrared (NIR) and far-red light. Shanmugam, Boyer, and coworkers reported the use of a biocatalyst, namely bacteriochlorophyll a, with light harvesting capabilities absorbing light in the NIR and far-red regions as an efficient photoredox catalyst for RAFT process. 540 Bacteriochlorophyll a is a photosynthetic bacteria which enables anoxygenic photosynthesis in dark, deep-sea waters by absorbing NIR and far-red light and utilizing sulfide, elemental sulfur, or hydrogen instead of water as an electron donor. The authors took advantage of bacteriochlorophyll a to carry out RAFT polymerization with low-energy NIR LED light sources diminishing undesirable side-reactions as a result of using high-energy UV and visible light to produce well-defined polymers with controlled molecular weight properties and high chain end fidelity. Furthermore, the NIR light being capable of penetrating depth was used to conduct polymerization with the light source screened by a paper that indicates the potential of this process for biomedical applications where deep penetration is required without causing damage to living tissues. A list of photocatalysts including metal- and organic-based photoredox catalysts employed in photoinitiated RAFT polymerization is collected in Chart 8. A variety of chemically Chart 8. Structures of Photoredox Catalysts Used in RAFT Polymerization different monomers have been shown to be compatible with photoinduced RAFT processes (Chart 9). These include methacrylates such as MMA, glycidyl methacrylate GMA, HEMA, DMAEMA, and oligoethylene glycol methyl ether methacrylate (OEGMA), BzMA, pentafluorophenyl methacrylate (PFPMA); acrylates such as MA, tba, and oligoethylene glycol methyl ether acrylate (OEGA); acrylamide-based monomers such as acrylamide (AAm), DMAAm, NiPAAm, and HPMA; and vinyl acetates such as vinyl acetate (VAc), vinyl pivalate (VP), and vinyl benzoate (VBz). Other monomers also involve St, MAA, N-vinylcarbazole (NVK), N-vinylpyrrolidinone (NVP), dimethyl vinylphosphonate 10250

40 Chemical s Chart 9. Different Monomers Polymerized by Photoinduced RAFT Polymerization (DVP), and isoprene (IP). Importantly, controlled/living syntheses of monomers such as vinyl esters, which are challenging in other systems like photoinduced ATRP, have been successfully conducted taking advantage of photoinitiated RAFT reactions. Also Chart 10 shows different thiocarbonylthio compounds used in photoinduced electron transfer RAFT reactions. The scope of the photoinduced RAFT process has been wide and is being increasingly expanded in terms of synthetic aspects and applications areas. Recently, it has been applied in selfassembly applications and the synthesis of polymeric nanoparticles. Photoinduced dispersion RAFT polymerization was employed in the presence of photoredox catalysts to form polymer nanoparticles of various morphologies in a controlled manner via polymerization-induced self-assembly in dispersion polymerization or using miniemulsion 544 polymerization techniques. Moreover, controlled formation of star polymeric nanoparticles 545 or micelles 546 have been reported in a photocatalyst-free RAFT system under visible light irradiation. Having pioneered sophisticated strategies based on photoinduced electron transfer reactions of photoredox catalysts and RAFT agents to photochemically mediate and control the RAFT polymerization, Boyer and co-workers subsequently exploited this approach to synthesize complex macromolecular architectures and gain further control over the composition of polymer chains in terms of stereochemistry. 547 Isotactic and syndiotactic polymers presenting accurate patterns of tacticty as opposed to atactic ones are of great importance whose chemical and physical properties are significantly imposed and controlled by stereochemical properties. They studied the polymerization of DMAAm monomer initiated by Ir III photoredux catalyst and 2-(n-butyltrithiocarbonate)-propionic acid (BTPA) as the thiocarbonylthio compound under blue LED irradiation. A Lewis acid compound, namely Y(OTf) 3, was also used as the tacticity mediator. In the absence of Y(OTf) 3, polymerization was successfully achieved with controlled molecular weight properties but resulted in atactic polymer chains. Tacticity can be determined by 1 H NMR spectroscopy through analyzing, for example, the methylene peaks present in the backbone of polymer around 1.7 to 0.9 ppm. However, in the presence of Y(OTf) 3 control in terms of both molecular weight and tacticity was established. Indeed, such control on stereochemistry of polymers was due to the complexation of the tacticity mediator Y(OTf) 3 with reaction components. The Lewis acid mediator was suggested to coordinate with the last two pendant amide groups of the propagating radical and incoming monomer unit, resulting in meso type addition and an increase in isotacticity of the resulting polymers. However, in the absence of the mediator, the probability of both meso and racemic type addition is equal due the planar configuration which resulted in atactic chains. The kinetics of the polymerization in the presence of Y(OTf) 3 was found to proceed with high polymerization rates and resulted in polymers with slightly higher molecular weight distributions. This increment in the rate of polymerization was attributed to the increased reactivity of the propagating radicals coordinated by Lewis acid which led to a decrease in the rate of exchange between radical and dormant chains resulting in slightly high molecular weight distributions. This technique further allowed the synthesis of polymers with varying stereotacticity given the efficiency of tacticity control by Lewis acids and its dependence on the nature of the solvent used. It is known that solvents such as DMSO form a tight complex with the Lewis acids preventing control of tacticity. Stereoblock polymers were synthesized by adding a solution of Y(OTf) 3 into a reaction mixture already reached to some monomer conversion with chains being atactic due to the absence of the mediator. The initial atactic segment of polymer chains was followed by a second segment being isotactic in the presence of the Y(OTf) 3 mediator. Given the limited ability of Lewis acids in controlling tacticity of polymer chains in the DMSO solution and taking advantage of temporal control offered by the photocatalytic system, under carefully optimized conditions incredibly complex macromolecular architectures were synthesized. Stereogradient polymers 10251

41 Chemical s Chart 10. Thiocarbonylthio Compounds as Both Initiator and Chain Transfer Agent in Photoinduced RAFT Polymerization Scheme 42. Stereoregulation in Photoinduced RAFT Process Utilizing Ir Photocatalyst and Y(OTf) 3 Mediator for Stereocontrol (Co)polymers containing segments with varying degrees of tacticity could be synthesized by the addition of DMSO with different DMSO/ Y(OTf) 3 ratios (Scheme 42). Thus, such a combination of the advantages of the photoinitiated living radical polymerization in establishing remarkable control over the molecular weight properties and offering spatiotemporal control with other chemical stimuli in controlling the stereochemistry of polymer chains appears a powerful and versatile technique in synthesizing polymers with exact accuracy as presented by nature. Given the nature of the C S bond present in the RAFT agent being easily cleavable by light, photoinitiation of the RAFT process has also been investigated in the absence of any additional photocatalyst and radical source. In fact, such a system was reported as the first example of photocontrolled radical polymerization by Otsu and co-worker who employed a dithiocarbamate as photoiniferter to conduct polymerization under UV light irradiation. 457 The irreversible cleavage of the C S bond resulted in the initiating radical formation and allowed for conducting polymerization in a controlled manner. Various studies have been conducted in this regard reporting the use of thiocarbonyl compounds to initiate and control radical polymerization under irradiation without the need of external photoactive compounds Johnson and coworkers, for example, developed versatile approaches to synthesize photoresponsive hydrogels possessing living characters by using trithiocarbonates under UV- or sunlight. Trithiocarbonate was used to reversibly form initiating radicals by C S bond cleavage to polymerize acrylamide-based monomers in a (photo)controlled manner. 492,556 Using the Diel-Alder click chemistry hydrogels were formed which could further grow in size upon monomer addition under irradiation. However, the use of UV light in such systems with high-energy inputs presents several limitations, especially in bioapplications. Visible light photoinitiators have been proposed as a solution. 554,

42 Chemical s 7. CLICK CHEMISTRY BY PHOTOINDUCED ELECTRON TRANSFER REACTIONS The term click chemistry was first coined in 2001 by Sharpless and co-workers to describe chemical transformations that are very high yielding, produce only inoffensive side-products, need no purification by column chromatography, are highly modular in nature and wide in scope, are stereospecific, although not necessarily enantioselective and provide a high degree of atom economy. 562 Reactions that fulfill the click chemistry criteria and have become to be known as click reactions include cycloaddition reactions such as the 1,3-Huisgen dipolar cycloaddition or copper(i)-catalyzed azide alkyne cycloaddition (CuAAC), the Diels Alder reaction, thiol ene/yne reactions, 570,571 nucleophilic substitution reactions especially with small strained rings like epoxy and aziridine compounds, 572 and reactions involving nonaldol carbonyl compounds such as oximes, hydrazones, and ureas The concept of click chemistry since its inception has captured the attention of chemists, biochemists, and biologists from a wide range of fields with over 5000 scientific articles relating to it published since With regard to polymeric and macromolecular architectures, click reactions have been applied in every imaginable respect from bioconjugation and drug discovery to material sciences and nanotechnology. 576 Many of the thermal click reactions mentioned above have photoinduced counterparts that impart the process with the inherent advantages of photochemical transformations. These advantages include but are not limited to mildness of reaction conditions and spatiotemporal control, properties which otherwise under thermal conditions would not be accessible. Photoclick reactions are tunable duo to their photochemical nature, whereby the wavelength and intensity of the light used can be varied. For two-dimensional (2D) surfaces and 3D spaces, the spatial advantage of photoclick reactions can be made evident as shining the light only in defined locations in the space desired can produce the click reaction only in that area. With such accuracy as can be obtained using photoinduced click reactions, highly precise and tailored macromolecular fabrication is possible Cycloaddition Click Reactions Cycloaddition reactions are those in which unsaturated species combine to form cyclical adducts. They can occur under thermal or photo conditions as well as by metal catalysis or by microwave induction. The reactants are often unreactive toward acids, amines, alcohols, and many other functional groups. Also due to the nature of cycloaddition reactions, small molecule side products are not created and so these features lend to their capacity to fulfill the click chemistry criteria. Molecular oxygen does not affect cycloaddition reactions negatively, and they are known to occur readily in both aqueous and organic media where polar solvents and Lewis acids can accelerate the process The most popular cycloaddition click reaction to date is the CuAAC. The reaction is highly modular and wide in scope, regiospecific, and is highly efficient with mild reaction conditions. Several methods are known to generate the active Cu I species from the more economic and air stable Cu II species, necessary for a successful CuAAC click reaction such as the use of reducing agents, 580 photochemical means, 581,582 electrochemical redox processes, 582,583 and the use of coppercontaining nanoparticles. 584,585 Also our group along with the group of Bowman and others have developed another photochemical method for the in situ generation of Cu I species for the CuAAC click reaction This method involves irradiation of the Cu II /ligand complex with UV light to affect an electron transfer from ligand to metal to reduce the Cu II to Cu I complex. In the direct approach, the UV light promotes an electron transfer from the π electron molecular orbitals of the ligand metal complex to the metal ion to produce the Cu I active species. Alternatively the use of visible light and a photoinitiator has been employed to affect the electron transfer and Cu II s reduction to Cu I. In the indirect approach both UV and visible light can be employed depending on the nature of the photoinitiator used. The absorbed photons produce fragmented electron-donor radical species from the photoinitiator moiety which affects the reduction to Cu I active species via electron transfer as can be observed in Scheme 43. Scheme 43. Photochemical Approach to CuAAC: (a) Direct Cu II to Cu I Reduction by UV Radiation; (b) Indirect Reduction of Cu II to Cu I Using Additional Photoinitiator The influence of amine-based ligands in promoting photoinduced electron transfer to reduce Cu II ions is studied. Bowman and co-workers found that aliphatic amines were efficient in promotion of Cu II reduction and stabilization of resulted Cu I species. 591 Particularly, tertiary amines such as TEA, tetramethylethylenediamine (TMDA), PMDETA, and hexamethylenetetramine (HMTETA) were found to be the most effective ones. Photoinduced electron transfer from the ligand compounds to Cu II was proposed to induce the reaction. Photoreduction of copper(ii) acetate (Cu(OAc) 2 ) in the absence of any photoinitiators was applied by Schubert and coworkers to generate Cu I ions capable of catalyzing CuAAC reactions. 590 Multifunctional azide and alkyne components were successfully photoclicked using the Cu(OAc) 2 catalyst for the synthesis of cross-linked networks with high mechanical properties. As regards indirect systems, the use of both Type I and Type II photoinitiators that produce electron-donating radicals capable of reducing Cu II complexes is a common approach in photoinduction of CuAAC reactions. Various visible light photoinitiators have been used to functionalize polymers or synthesize complex macromolecular structures such as block, graft, and star copolymers as well as polymeric networks. 463,587, Bowman and co-workers photopolymerized a series of multifunctional azide and alkyne containing monomers to cross-linked networks by photoinduced CuAAC strategies in a step-growth polymerization manner. 596 Free radical electrontransfer reduction of Cu II complex was achieved using an acylphosphine oxide photoinitiator under visible light irradiation. Polymers possessed good mechanical properties with high cross-link density and high glass transition temperature. In the process, extensive triazole linkages were formed with potential to use as shape memory materials

43 Chemical s An interesting approach being increasingly developed is the incorporation of photoinitiator moieties as counterion components to the Cu II catalyst structure. For example, acylphosphine oxide as Type I photoinitiator was combined with Cu II species to form a one-component photoinitiator-cu catalyst system. 465 In the process, radicals derived from the combined photoinitiator reduced Cu II ions to induce the click reaction. Vincent and co-workers took a similar approach by combining Type II photoinitiator such as benzophenone derivatives with the Cu catalyst. Photoexcited photoinitiator promoted photoinduced electron transfer and hydrogen abstractions with amine ligands and solvent to achieve photoreduction of Cu II to Cu I, which in turn catalyzes the click reaction The presence of amine-based ligands was necessary for the solubility of photoinitiator-cu compound and successful photoreduction of Cu II species. Temporal control was maintained as the reaction could be quickly halted by removing the light source and bubbling air through the mixture as the molecular oxygen present in the air oxidizes the Cu I complex back to Cu II. This process is reversible and can be restarted if oxygen was purged out of the system by flushing N 2 or argon through followed by further irradiation. As an alternative, recently, Cu II thioxanthone was reported as a onecomponent, ligand-free photocatalyst to conduct the CuAAC reaction providing a successful temporal control over the reaction. 601 In a conceptually distinct contribution, Bear and co-workers reported doping Cu I cations into zinc sulfide quantum dot core shell nanostructures. 602 The CuAAC was achieved by the release of Cu I ions into the solution upon photoexcitation of the nanoparticles in a ligand-free mode. Furthermore, our group investigated the potential of semiconducting heterogeneous photocatalysis in CuAAC reactions. Such systems readily provide electron/hole charge transfers upon illumination. The mpg-c 3 N or ZnO 475 nanoparticles were utilized as a heterogeneous photocatalyst. Reduction of Cu II complex was readily achieved by photogenerated electrons, and CuAAC was conducted in a wide scope with the ability of temporal and reaction rate control by tuning the wavelength or intensity of the light source. The heterogeneous nature of the photocatalytic system enabled reusability in further reactions. 604 Photosensitization by polynuclear aromatic compounds such as perylene, anthracene, and pyrene was used to affect the described redox process. 605 Perylene was shown to be the most efficient photosensitizer for the reduction of Cu II by promoting photoinduced electron transfer in the photoexcited state to the Cu II ions. Temporal control over the reaction is achieved as by switching on or off the activating light irradiation. However, it has been shown that Cu I is not rapidly consumed when the reactions proceed in the dark upon initial activation. This has been shown to be due probably to disproportionation of Cu I to Cu II and Cu 0 leading to the regeneration of Cu I species. 606 This implies that once activation is achieved with the Cu catalyst, constant irradiation with the source of light is not necessary. Visible light photocatalyst has also been utilized in CuAAC reactions. Jain and co-workers reported the use of a bimetallic photocatalyst comprised of Ru as a photosensitizer attached to the Mn complex as a catalytic unit by bipyrimidine bridging the ligand for click transformations. 607 Upon visible light irradiation in the presence of TEA as a sacrificial electron donor, photoinduced electron transfer is facilitated from the Ru photosensitizer to the Mn catalyst, which subsequently was responsible for the reduction of Cu II to Cu I species by transferring electron from Mn to Cu II and hence producing active Cu catalyst for CuAAC reaction. The bimetallic nature of the photocatalyst system comprised of photosensitizer and catalyst units linked together facilitated rapid electron transfer processes to reduce Cu II complexes. To exhibit the spatial control of which can be availed while using the photoinduced CuAAC system, photolithographical patterning over azide-functional plasma polymer film with an alkyne-functional PEG hydrogel through a photomask was carried out (Scheme 44). 608,609 The thickness of the hydrogel Scheme 44. Spatial Control over Photoinduced CuAAC by Photolithographical Patterning a a Image of a hydrogel patterned by photo-cuaac reaction using photomask. Reprinted with permission from ref 608. Copyright 2011 Nature Publishing Group. formed could be precisely tuned by varying the intensity of the UV radiation applied and also the molecular weight of the PEG macromonomer. The PEG hydrogel patterns formed were highly functional and wide in scope with broad applicability for cell-adhesion and chemical functionality. In our group, we have exploited electron transfer dependent photoinduced CuAAC and especially its orthogonal nature to bring about complex macromolecular formation in one-pot simultaneous and sequential double reactions. For example, photoinduced CuAAC and photoinduced thiol ene click were availed of in a one-pot sequential process in that order to form poly(methyl methacrylate)-b-poly(ε-caprolactone)-(n-acetyl-lcysteine) (PMMA-b-PCL-NAC). 588 The obtained PMMA-b- PCL-NAC has been used as a matrix for cell-adhesion and was shown to be as effective as the leading commercially available materials (Scheme 45). The amino acid end functionality also facilitated to anchor the short peptide arginylglycylaspartic acid (RGD) and form PMMA-b-PCL-NAC-RGD, which was then used for integrin avb3-mediated cell adhesion and biosensing studies. 610 The photoinduced CuAAC s orthogonality with the ATRP process was also expanded to the synthesis of block and graft copolymers in one-pot simultaneous CuAAC/ATRP processes 10254

44 Chemical s Scheme 45. Photoinduced Sequential CuAAC and Thiol Ene Click Reactions for Macromolecular Synthesis a a Fluorescence microscope images of DAPI-stained Vero cells cultivated on the (a) commercial well plates and (b) copolymers obtained by sequential CuAAC-thil-ene techniques after 24 h. Reprinted from ref 588. Copyright 2014 American Chemical Society. (Scheme 46). 589,611 Cu I is the active species in both CuAAC and ATRP. The photoinduced electron transfer which occurs upon radiation of the Cu II /PMDETA complex at a wavelength around 350 nm leads to the reduction of Cu II to Cu I forming the active catalyst required to drive forward both the CuAAC and ATRP reactions. In this way, polystyrene-b-poly(methyl methacrylate) and polystyrene-g-poly(methyl methacrylate) could be synthesized using the one-pot photoinduced simultaneous CuAAC/ATRP reactions. Additionally, the combination of polymerization and click concepts to produce functional polymers was recently expanded to cationic polymerization and CuAAC processes. Cationically polymer- ized polymers with various clickable moieties were prepared to perform sequential photoinduced CuAAC reaction, allowing introduction of a variety of functional groups or formation of copolymers and polymer networks. 612 Bowman and co-workers developed a sequential two-stage strategy to wrinkle formation employing the thiol acrylate and photoinduced CuAAC click reactions. 613 The initial step concerned an elastomeric network formation through the thiol ene click reaction between acrylate and multifunctional thiol compounds. The sample produced thereof was subject to a second click step, CuAAC, to modify its properties. Indeed, using propargyl acrylate in the first click stage incorporated suitable propargyl sites within the network to carry out CuAAC reactions. The photoinduced CuAAC was performed by immersing the sample into a Cu II solution,which led to an increase in cross-link density of the network. As the penetration of photoreduced Cu I ions into the elastomeric network was diffusion-limited, the CuAAC was carried out in two dimensions on the surface of the sample. This led to the formation of wrinkles in a facile manner. The topologies of wrinkles could be manipulated by regulating the experimental conditions and concentration of the reactants so that the wavelength and amplitude of the wrinkles could be tuned with the ability to offer spatiotemporal control over the wrinkle generation. Unfortunately, for application in biosystems, the concentrations of copper present in the relevant macromolecular structures that result from the CuAAC are often too high leading to toxicity. For this reason, the use of more activated groups for photoinduced cycloaddition reactions that yield efficiency comparable to the CuAAC has been investigated. These activated groups exist in the form of strained alkynes with electron-withdrawing fluorine substituents which undergo cycloaddition reactions without the need for Cu catalyst and have been termed strain-promoted [3 + 2] cycloaddition of azides and cyclooctynes. 614,615 This Cu I free approach was taken advantage to provide biomolecule labeling. The photoinduced equivalent of this approach was also developed and applied. Strain-promoted alkynes were once again taken advantage of; however, in the photochemical process the strained cyclooctyne species was masked with a cyclo- Scheme 46. Simultaneous Photo CuAAC/ATRP Utilized for Block and Graft Copolymer Synthesis 10255

45 Chemical s propenone species. 616,617 The cyclopropenone species upon irradiation by light with wavelength >350 nm undergoes conversion to the unmasked strained cyclooctyne species, which thereafter can react in the above-described Cu-free azide cyclooctyne click reaction (Scheme 47). Scheme 47. Photoinduced Cu-Free Cyclooctyne-Azide Click Reaction applicability of the photoinduced thiol ene click reaction toward the macromolecular synthesis of complex polymeric architectures. Hawker and co-workers have demonstrated the thiol ene click reaction s utility for rapid complex dendrimer formation Du Prez and co-workers have used the thiol ene click reaction to fashion diblock and star-shaped polymers. 571 The photoinduced thiol ene click reaction relies on the use of a photoinitiator to provide radical species upon light irradiation that can then bring about a radical hydrogen abstraction which in turn leads to a thiol radical species which can undergo the coupling process as can be observed in Scheme 49. Scheme 49. Photoinduced Thiol ene Reaction Mechanism 7.2. Thiol Ene/Yne Click Chemistry Thiol ene/yne reactions bear all the characteristics required to be termed click chemistry. They are high yielding under mild conditions producing highly regioselective products exhibiting largely anti-markovinkov selectivity, and their work-ups are hassle-free with no chromatography required to remove small molecule side-products. The reactants involved are tolerant to a variety of functional groups lending the process with orthogonality with respect to other chemical transformations. They are carried out with benign catalysts and solvents and can often be insensitive toward molecular oxygen (Scheme 48). Scheme 48. Addition of Thiol to An Alkene Moiety by Either Radical or Anionic Catalytic Processes Thiol ene/yne chemistry is defined as the addition of a primary thiol radical species to an alkene or alkyne moiety. The reaction can proceed by two separate mechanisms, either by the free radical mediated or Michael addition processes. When the free radical approach is used, the transformation is labeled the thiol ene/yne reaction; however, when the transformation is catalyzed and proceeds through an anionic pathway, it is labeled the thiol Michael addition. The photoinduced thiol ene/yne click reaction has now become well-established. It lends all the expected advantages of a photochemical process to the reaction such as mild reaction conditions and spatiotemporal control. Groups such as Hoyle, 618 Bowman, 619,620 Lowe 621 and others greatly expanded the sphere of knowledge as it applies to the Our group has thoroughly investigated the influence of the type of photoinitiator on the photoinitiation of thiol ene click chemistry for functionalization of polymers. 61 Thiol or allylmodified polystyrenes with controlled molecular weight properties were successfully functionalized by an appropriate alkene or thiol functionality in the presence of either Type I or Type II photoinitiators with excellent yields achieved in both systems. Type I photoinitiation appeared as slightly more efficient than the other one as higher conversions were obtained by Type I photoinitiators. Thiol groups also served as electron/proton transfer co-initiator for Type II photoinitiators The photoinduced thiol/ene has been recently employed by the group of Bowman and co-workers for the synthesis of sequence defined polymers. 634 A series of nucleobase functional monomers were polymerized using thiol click reactions to form polymers in a sequence-specific manner. The monomers were structurally analogous to oligonucleotides in DNA and comprised of a thiol and an allyalmine or a thiol and an acrylamide functional groups suitable for thiol ene click or thiol-michael reactions, respectively. Thiol ene photopolymerization was used to form nucleobase-containing sequencecontrolled homo polymers as well as diblock copolymers. A combination of thiol ene and thiol-michael approaches led to more specific and robust strategies in synthesizing DNA-like macromolecules such as organogels with reversible cross-links. The process allowed incorporating a wide range of functional groups into the polymer structure in a sequence-controlled manner. Recently investigations have taken place with regards to the photoinduced electron transfer thiol ene reaction photocatalyzed by photoredox catalysts For example, when employing Ru(bpz) 3 (PF 6 ) 2 as the photoredox catalyst, it is possible to carry out thiol ene chemistry with visible light 10256

46 Chemical s irradiation and a catalyst loading of typically 0.25 mol % with quantitative yields obtained in certain cases. Direct photooxidation of thiols with the photocatalyst led to thiyl radicals through photoinduced electron and proton transfers promoting the click reaction; however, this direct path was found to be slow in rate. The reduced state Ru(bpz) 3 + could be regenerated in the presence of oxygen to initiate further chain processes. The rate of the reaction could be effectively accelerated by employing aromatic amines such as anilines as electron-transfer mediators between the photocatalyst and thiol component. The amine mediators were oxidized to the corresponding amine radical cation by an excited state photocatalyst. These amine radical cations were capable of interacting with thiols via consecutive electron- and proton-transfer reactions generating thiyl radicals (Scheme 50). Scheme 50. Photoredox Catalyzed Thiol Ene Reaction The authors of the above work highlighted the photoredox catalyzed thiol ene reaction s excellent efficiency, where 99% yield was obtained for many substrates. Also it was shown to be large in scope where the system could be applied to a variety of thiols and alkenes with positive results under mild conditions using a visible wavelength light source. The photoredox thiol ene system has also been applied to macromolecular architectures for polymer postfunctionalization and step-growth addition polymerization (Scheme 51). 638,639 It was reported that polybutadiene and poly(allyl methacrylates) could be postfunctionalized with thiol side groups by photoredox thiol ene reaction with high conversion obtained within minutes, typically times of <20 min were described (Scheme 52). Similarly, as in the nonpolymeric case, long wavelength blue LED irradiation with low catalyst loading was required, underlining the mild conditions and environmentally conscious nature of the approach. The system was also utilized for polymer formation by stepgrowth addition again using visible light irradiation with only 25 ppm concentration of photoredox catalyst required with respect to olefin concentration. With such low concentrations of catalyst being employed, mild conditions as regards the use of low-energy blue light LED irradiation and high efficiency obtained the visible light photoredox catalyzed thiol ene could in the near future rival the more standard radical thiol ene click reaction for utility and scope. 8. CONCLUSION AND FUTURE PERSPECTIVE Learning to mimic the natural phenomena that drive chemical transformations will significantly advance our understanding of and ability in constructing structurally complex building blocks and molecules in all areas of science. Every moment that passes by, an intricate set of chemical reactions are taking place in living organisms using powerful yet simple toolkits provided by nature that are able to assemble the essential complexity of life. Photosynthesis is an intriguing example of such complex chemical transformations. It smoothly occurs in plants and is based on the efficient utilization of the energy of sunlight as it is converted into chemical energy. This process involves a series of photoinduced electron transfer processes that result in products that are the essential components of fueling life on Earth. We are yet to completely master and mimic this process in order to reach its efficiency and complexity. Nevertheless, we have made significant strides, and in recent years, technological advancements have enabled us to establish molecular complexity through artificial molecular tailoring. 640 Photochemical reactions, particularly those involving photoinduced electron transfer reactions, establish a substantial contribution to the modern synthetic chemistry, and the polymer community has been increasingly interested in exploiting and developing novel photochemical strategies. As described in detail throughout this review, these reactions are readily utilized in almost every aspect of macromolecular architecture synthesis involving initiation, control of the reaction kinetics and molecular structures, functionalization, and decoration, etc. Chain growth polymerizations including different modes of free radical and cationic polymerizations have long been established and expanded upon substantially, together with interesting contributions in step-growth polymerization of various systems. Merging with controlled/living polymerizations, photochemistry has opened up new fascinating and powerful avenues for macromolecular synthesis. Construction of various polymers with incredibly complex structures and specific control over the chain topology, as well as providing the opportunity to manipulate the reaction course through spatiotemporal control, is one of the unique abilities of such photochemical reactions. Two-photon polymerization microfabrication is a very exciting development which avails of the spatial control potential of photochemical polymerization used for 3D printing with ultrafine microresolution. 641 This has already found application in such fields as micro- and nanophotonics, microelectromechanical systems, microfluidics, biomedical implants, and microdevices. Photoinitiating systems for controlled polymerizations which focus on low energy, longer wavelength visible light, and even sunlight have grown in prevalence. Previously, these processes have been centered on ATRP and the photoredox catalyst approach using Cu and Ir. More recently for both cationic and free radical promoted cationic photopolymerization, the use of cheap low-energy LEDs (blue, red, green, and purple laser diodes) has been implemented. Scheme 51. Step-Growth Addition Polymerization by Highly Efficient Visible Light Photoredox Catalysis 10257

Photoinitiation, Photopolymerization, and Photocuring

Photoinitiation, Photopolymerization, and Photocuring Jean-Pierre Fouassier Photoinitiation, Photopolymerization, and Photocuring Fundamentals and Applications Hanser Publishers, Munich Vienna New York Hanser/Gardner Publications, Inc., Cincinnati Contents

More information

5. Photochemistry of polymers

5. Photochemistry of polymers 5. Photochemistry of polymers 5.1 Photopolymerization and cross-linking Photopolymerization The fundamental principle of photopolymerization is based on the photoinduced production of a reactive species,

More information

Radical Initiation 2017/2/ ) Thermal Decomposition of Initiators

Radical Initiation 2017/2/ ) Thermal Decomposition of Initiators adical Initiation Production of radicals (from initiator) to initiate chain polymerization. A variety of initiator systems can be used to bring about the radical polymerization. 1) Thermal Decomposition

More information

KOT 222 Organic Chemistry II

KOT 222 Organic Chemistry II KOT 222 Organic Chemistry II Course Objectives: 1) To introduce the chemistry of alcohols and ethers. 2) To study the chemistry of functional groups. 3) To learn the chemistry of aromatic compounds and

More information

12/27/2010. Chapter 15 Reactions of Aromatic Compounds

12/27/2010. Chapter 15 Reactions of Aromatic Compounds Chapter 15 Reactions of Aromatic Compounds Electrophilic Aromatic Substitution Arene (Ar-H) is the generic term for an aromatic hydrocarbon The aryl group (Ar) is derived by removal of a hydrogen atom

More information

Chem 263 Oct. 12, 2010

Chem 263 Oct. 12, 2010 Chem 263 ct. 12, 2010 Alkyl Side Chain xidation Reaction If the carbon directly attached to the aromatic ring has > 1 hydrogen attached to it, it can be oxidized to the corresponding carboxylic acid with

More information

Chapter 15 Reactions of Aromatic Compounds

Chapter 15 Reactions of Aromatic Compounds Chapter 15 1 Chapter 15 Reactions of Aromatic Compounds Electrophilic Aromatic Substitution Arene (Ar-H) is the generic term for an aromatic hydrocarbon The aryl group (Ar) is derived by removal of a hydrogen

More information

Chapter 7: Alcohols, Phenols and Thiols

Chapter 7: Alcohols, Phenols and Thiols Chapter 7: Alcohols, Phenols and Thiols 45 -Alcohols have the general formula R-OH and are characterized by the presence of a hydroxyl group, -OH. -Phenols have a hydroxyl group attached directly to an

More information

ummary Manipulating Radicals

ummary Manipulating Radicals Manipulating Radicals ummary Modern catalysis research tries to address issues such as material scarcity, sustainability or process costs. One solution is to replace expensive and scarce noble metal catalysts

More information

What are radicals? H. Cl. Chapter 10 Radical Reactions. Production of radicals. Reactions of radicals. Electronic structure of methyl radical

What are radicals? H. Cl. Chapter 10 Radical Reactions. Production of radicals. Reactions of radicals. Electronic structure of methyl radical What are radicals? Radicals are intermediates with an unpaired electron Chapter 10 Radical Reactions H. Cl. Hydrogen radical Chlorine radical Methyl radical Often called free radicals Formed by homolytic

More information

PHOTOINITIATOR BASIC CHEMISTRY.

PHOTOINITIATOR BASIC CHEMISTRY. 1 PHOTOINITIATOR BASIC CHEMISTRY. INTRODUCTION TO FORMULATING AND PRODUCTS FOR LED CURE AND LOW MIGRATION. Youyuan Wu IGM Resins USA Inc 1.0 INTRODUCTION What is a photoinitiator? A substance (other than

More information

Extensive Dark Cure from Controlled Polymerization Based on a Method Using Visible-Light Activated Initiator System

Extensive Dark Cure from Controlled Polymerization Based on a Method Using Visible-Light Activated Initiator System Extensive Dark Cure from Controlled Polymerization Based on a Method Using Visible-Light Activated Initiator System Dongkwan Kim, and Jeffrey W. Stansbury University of Colorado-Denver, School of Dental

More information

Introduction to Macromolecular Chemistry

Introduction to Macromolecular Chemistry Introduction to Macromolecular Chemistry aka polymer chemistry Mondays, 8.15-9.45 am except for the following dates: 01.+29.05, 05.+12.06., 03.07. Dr. Christian Merten, Ruhr-Uni Bochum, 2017 www.ruhr-uni-bochum.de/chirality

More information

JEFFERSON COLLEGE COURSE SYLLABUS CHM201 ORGANIC CHEMISTRY II. 5 Credit Hours. Prepared by: Richard A. Pierce

JEFFERSON COLLEGE COURSE SYLLABUS CHM201 ORGANIC CHEMISTRY II. 5 Credit Hours. Prepared by: Richard A. Pierce JEFFERSON COLLEGE COURSE SYLLABUS CHM201 ORGANIC CHEMISTRY II 5 Credit Hours Prepared by: Richard A. Pierce Revised Date: January 2008 by Ryan H. Groeneman Arts & Science Education Dr. Mindy Selsor, Dean

More information

75. A This is a Markovnikov addition reaction. In these reactions, the pielectrons in the alkene act as a nucleophile. The strongest electrophile will

75. A This is a Markovnikov addition reaction. In these reactions, the pielectrons in the alkene act as a nucleophile. The strongest electrophile will 71. B SN2 stands for substitution nucleophilic bimolecular. This means that there is a bimolecular rate-determining step. Therefore, the reaction will follow second-order kinetics based on the collision

More information

ALCOHOLS AND PHENOLS

ALCOHOLS AND PHENOLS ALCOHOLS AND PHENOLS ALCOHOLS AND PHENOLS Alcohols contain an OH group connected to a a saturated C (sp3) They are important solvents and synthesis intermediates Phenols contain an OH group connected to

More information

Alcohols, Ethers, & Epoxides

Alcohols, Ethers, & Epoxides Alcohols, Ethers, & Epoxides Alcohols Structure and Bonding Enols and Phenols Compounds having a hydroxy group on a sp 2 hybridized carbon enols and phenols undergo different reactions than alcohols. Chapter

More information

Chapter 19: Alkenes and Alkynes

Chapter 19: Alkenes and Alkynes Chapter 19: Alkenes and Alkynes The vast majority of chemical compounds that we know anything about and that we synthesize in the lab or the industrial plant are organic compounds. The simplest organic

More information

A. Loupy, B.Tchoubar. Salt Effects in Organic and Organometallic Chemistry

A. Loupy, B.Tchoubar. Salt Effects in Organic and Organometallic Chemistry A. Loupy, B.Tchoubar Salt Effects in Organic and Organometallic Chemistry 1 Introduction - Classification of Specific Salt Effects 1 1.1 Specific Salt Effects Involving the Salt's Lewis Acid or Base Character

More information

Classifying Organic Chemical Reactions

Classifying Organic Chemical Reactions Chemical Reactivity Organic chemistry encompasses a very large number of compounds ( many millions ), and our previous discussion and illustrations have focused on their structural characteristics. Now

More information

Keynotes in Organic Chemistry

Keynotes in Organic Chemistry Keynotes in Organic Chemistry Second Edition ANDREW F. PARSONS Department of Chemistry, University of York, UK Wiley Contents Preface xi 1 Structure and bonding 1 1.1 Ionic versus covalent bonds 1 1.2

More information

Chapter 9 Aldehydes and Ketones Excluded Sections:

Chapter 9 Aldehydes and Ketones Excluded Sections: Chapter 9 Aldehydes and Ketones Excluded Sections: 9.14-9.19 Aldehydes and ketones are found in many fragrant odors of many fruits, fine perfumes, hormones etc. some examples are listed below. Aldehydes

More information

Nuggets of Knowledge for Chapter 17 Dienes and Aromaticity Chem 2320

Nuggets of Knowledge for Chapter 17 Dienes and Aromaticity Chem 2320 Nuggets of Knowledge for Chapter 17 Dienes and Aromaticity Chem 2320 I. Isolated, cumulated, and conjugated dienes A diene is any compound with two or C=C's is a diene. Compounds containing more than two

More information

Chemistry of Benzene: Electrophilic Aromatic Substitution

Chemistry of Benzene: Electrophilic Aromatic Substitution Chemistry of Benzene: Electrophilic Aromatic Substitution Why this Chapter? Continuation of coverage of aromatic compounds in preceding chapter focus shift to understanding reactions Examine relationship

More information

Chapter 10 Radical Reactions"

Chapter 10 Radical Reactions Chapter 10 Radical Reactions Radicals are intermediates with an unpaired electron H. Cl. Hydrogen radical t Often called free radicals What are radicals? Chlorine radical t Formed by homolytic bond cleavage

More information

Chapter 15. Reactions of Aromatic Compounds. Electrophilic Aromatic Substitution on Arenes. The first step is the slow, rate-determining step

Chapter 15. Reactions of Aromatic Compounds. Electrophilic Aromatic Substitution on Arenes. The first step is the slow, rate-determining step Electrophilic Aromatic Substitution on Arenes Chapter 15 Reactions of Aromatic Compounds The characteristic reaction of aromatic rings is substitution initiated by an electrophile halogenation nitration

More information

Enduring Understandings & Essential Knowledge for AP Chemistry

Enduring Understandings & Essential Knowledge for AP Chemistry Enduring Understandings & Essential Knowledge for AP Chemistry Big Idea 1: The chemical elements are fundamental building materials of matter, and all matter can be understood in terms of arrangements

More information

Organic Chemistry. Second Edition. Chapter 19 Aromatic Substitution Reactions. David Klein. Klein, Organic Chemistry 2e

Organic Chemistry. Second Edition. Chapter 19 Aromatic Substitution Reactions. David Klein. Klein, Organic Chemistry 2e Organic Chemistry Second Edition David Klein Chapter 19 Aromatic Substitution Reactions Copyright 2015 John Wiley & Sons, Inc. All rights reserved. Klein, Organic Chemistry 2e 19.1 Introduction to Electrophilic

More information

PHOTOCATALYSIS: FORMATIONS OF RINGS

PHOTOCATALYSIS: FORMATIONS OF RINGS PHOTOCATALYSIS: FORMATIONS OF RINGS Zachery Matesich 15 April 2014 Roadmap 2 Photoredox Catalysis Cyclizations Reductive Oxidative Redox-neutral Electron Transfer Conclusion http://www.meta-synthesis.com/webbook/11_five/five.html

More information

Chapter 8. Substitution reactions of Alkyl Halides

Chapter 8. Substitution reactions of Alkyl Halides Chapter 8. Substitution reactions of Alkyl Halides There are two types of possible reaction in organic compounds in which sp 3 carbon is bonded to an electronegative atom or group (ex, halides) 1. Substitution

More information

Organic Chemistry. Radical Reactions

Organic Chemistry. Radical Reactions For updated version, please click on http://ocw.ump.edu.my Organic Chemistry Radical Reactions by Dr. Seema Zareen & Dr. Izan Izwan Misnon Faculty Industrial Science & Technology seema@ump.edu.my & iezwan@ump.edu.my

More information

CHEM 251 (4 credits): Description

CHEM 251 (4 credits): Description CHEM 251 (4 credits): Intermediate Reactions of Nucleophiles and Electrophiles (Reactivity 2) Description: An understanding of chemical reactivity, initiated in Reactivity 1, is further developed based

More information

This reactivity makes alkenes an important class of organic compounds because they can be used to synthesize a wide variety of other compounds.

This reactivity makes alkenes an important class of organic compounds because they can be used to synthesize a wide variety of other compounds. This reactivity makes alkenes an important class of organic compounds because they can be used to synthesize a wide variety of other compounds. Mechanism for the addition of a hydrogen halide What happens

More information

Nucleophile. Reaction Intermediate. Introduction to Reaction mechanisms. Definitions 2/25/2012

Nucleophile. Reaction Intermediate. Introduction to Reaction mechanisms. Definitions 2/25/2012 Introduction to Reaction mechanisms Definition A reaction mechanism is the step by step sequence of elementary reactions by which overall chemical change occurs. It is also a detailed description of the

More information

Chapter 24. Amines. Based on McMurry s Organic Chemistry, 7 th edition

Chapter 24. Amines. Based on McMurry s Organic Chemistry, 7 th edition Chapter 24. Amines Based on McMurry s Organic Chemistry, 7 th edition Amines Organic Nitrogen Compounds Organic derivatives of ammonia, NH 3, Nitrogen atom with a lone pair of electrons, making amines

More information

Chemical Engineering Seminar Series

Chemical Engineering Seminar Series Effect of Reaction Conditions on Copolymer Properties Loretta Idowu Keywords: copolymer composition distribution; radical polymerization kinetics; semi-batch starved feed; hydroxyl-functionality Non-functional

More information

Chapter 25: The Chemistry of Life: Organic and Biological Chemistry

Chapter 25: The Chemistry of Life: Organic and Biological Chemistry Chemistry: The Central Science Chapter 25: The Chemistry of Life: Organic and Biological Chemistry The study of carbon compounds constitutes a separate branch of chemistry known as organic chemistry The

More information

Overview of Types of Organic Reactions and Basic Concepts of Organic Reaction Mechanisms

Overview of Types of Organic Reactions and Basic Concepts of Organic Reaction Mechanisms Overview of Types of Organic Reactions and Basic Concepts of Organic Reaction Mechanisms Dr. Solomon Derese 1 A chemical reaction is the transformation of one chemical or collection of chemicals into another

More information

Chapter 17. Reactions of Aromatic Compounds

Chapter 17. Reactions of Aromatic Compounds Chapter 17 Reactions of Aromatic Compounds Electrophilic Aromatic Substitution Although benzene s pi electrons are in a stable aromatic system, they are available to attack a strong electrophile to give

More information

1/4/2011. Chapter 18 Aldehydes and Ketones Reaction at the -carbon of carbonyl compounds

1/4/2011. Chapter 18 Aldehydes and Ketones Reaction at the -carbon of carbonyl compounds Chapter 18 Aldehydes and Ketones Reaction at the -carbon of carbonyl compounds The Acidity of the Hydrogens of Carbonyl Compounds: Enolate Anions Hydrogens on carbons to carbonyls are unusually acidic

More information

Contents. List of. 2 Early pioneers of organic

Contents. List of. 2 Early pioneers of organic Contents List of V 1 1 2 Early pioneers of organic 3 2.1 15 3 Photophysics of 19 3.1 the 21 3.2 The 24 3.3 The Theoreticians' Perspective: A Closer 31 3.3.1 Transition 32 3.3.2 36 3.4 43 Flavin 45 4.1

More information

Module No and Title. PAPER No: 5 ; TITLE : Organic Chemistry-II MODULE No: 25 ; TITLE: S E 1 reactions

Module No and Title. PAPER No: 5 ; TITLE : Organic Chemistry-II MODULE No: 25 ; TITLE: S E 1 reactions Subject Chemistry Paper No and Title Module No and Title Module Tag 5; Organic Chemistry-II 25; S E 1 reactions CHE_P5_M25 TABLE OF CONTENTS 1. Learning Outcomes 2. Introduction 3. S E 1 reactions 3.1

More information

MCAT Organic Chemistry Problem Drill 10: Aldehydes and Ketones

MCAT Organic Chemistry Problem Drill 10: Aldehydes and Ketones MCAT rganic Chemistry Problem Drill 10: Aldehydes and Ketones Question No. 1 of 10 Question 1. Which of the following is not a physical property of aldehydes and ketones? Question #01 (A) Hydrogen bonding

More information

Doctor of Philosophy

Doctor of Philosophy STUDIES ON THE CORROSION INHIBITION BEHAVIOUR OF SOME AMINO ACID SURFACTANT ADDITIVES ABSTRACT SUBMITTED FOR THE AWARD OF THE DEGREE OF Doctor of Philosophy IN APPLIED CHEMISTRY By MOSARRAT PARVEEN UNDER

More information

Chapter 6 Ionic Reactions-Nucleophilic Substitution and Elimination Reactions of Alkyl Halides"

Chapter 6 Ionic Reactions-Nucleophilic Substitution and Elimination Reactions of Alkyl Halides Chapter 6 Ionic Reactions-Nucleophilic Substitution and Elimination Reactions of Alkyl Halides" t Introduction" The polarity of a carbon-halogen bond leads to the carbon having a partial positive charge"

More information

II. 1. TRANSFORMATIONS UNDER THE ACTION OF HEAT OR IRRADIATION

II. 1. TRANSFORMATIONS UNDER THE ACTION OF HEAT OR IRRADIATION CHAPTER II HYDROCARBON TRANSFORMATIONS THAT DO NOT INVOLVE METALS OR THEIR COMPOUNDS n this chapter we will briefly survey the main types of hydrocarbon transformations that occur without the participation

More information

Chem 263 Oct. 6, Single bonds, σ. e - donating Activate Activate ortho and para directing ortho and para directing

Chem 263 Oct. 6, Single bonds, σ. e - donating Activate Activate ortho and para directing ortho and para directing Chem 263 ct. 6, 2009 lectrophilic Substitution of Substituted Benzenes Resonance ffect Inductive ffect C=C, π system Single bonds, σ Strong Weak e - donating Activate Activate ortho and para directing

More information

Chapter 16. Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group. Physical Properties of Aldehydes and Ketones. Synthesis of Aldehydes

Chapter 16. Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group. Physical Properties of Aldehydes and Ketones. Synthesis of Aldehydes Nomenclature of Aldehydes and Ketones Chapter 16 Aldehydes and Ketones I. Aldehydes replace the -e of the parent alkane with -al The functional group needs no number Nucleophilic Addition to the Carbonyl

More information

SchemeII Visble light catalysed [2π+2π] cycloaddition using ruthenium based photocatalyst.

SchemeII Visble light catalysed [2π+2π] cycloaddition using ruthenium based photocatalyst. Executive Summary Principal Investigator : Pawar Hari Ragho Abasaheb Garware College Karve road, Pune UGC Reference No.F.47-626/13(WR)Dt.19/03/2014. Period of report: from 17/06/2014 to 17/06/2016 Title

More information

CHE1502. Tutorial letter 203/1/2016. General Chemistry 1B. Semester 1. Department of Chemistry

CHE1502. Tutorial letter 203/1/2016. General Chemistry 1B. Semester 1. Department of Chemistry E1502/203/1/2016 Tutorial letter 203/1/2016 General hemistry 1B E1502 Semester 1 Department of hemistry This tutorial letter contains the answers to the questions in assignment 3. FIRST SEMESTER: KEY T

More information

Chapter 13 Conjugated Unsaturated Systems

Chapter 13 Conjugated Unsaturated Systems Chapter 13 Conjugated Unsaturated Systems Introduction Conjugated unsaturated systems have a p orbital on a carbon adjacent to a double bond The p orbital can come from another double or triple bond The

More information

Theophylline (TH), the structure of which is presented below, is a bronchial-dilator used for the treatment of asthma.

Theophylline (TH), the structure of which is presented below, is a bronchial-dilator used for the treatment of asthma. 1 EXAM SCIETIIC CULTURE CEMISTRY PRBLEM 1: Theophylline (T), the structure of which is presented below, is a bronchial-dilator used for the treatment of asthma. 3 C 7 1 3 9 1.1 Structural study and acid-base

More information

4. Organic photosynthetic reactions

4. Organic photosynthetic reactions 4. rganic photosynthetic reactions 100 4.1 eactions of ethenes and aromatic compounds Photoreactivity of ethenes E Geometrical isomerization In π-π* excited states, there is effectively no π bond and so

More information

Lecture Topics: I. Electrophilic Aromatic Substitution (EAS)

Lecture Topics: I. Electrophilic Aromatic Substitution (EAS) Reactions of Aromatic Compounds Reading: Wade chapter 17, sections 17-1- 17-15 Study Problems: 17-44, 17-46, 17-47, 17-48, 17-51, 17-52, 17-53, 17-59, 17-61 Key Concepts and Skills: Predict and propose

More information

Chapter 20 Carboxylic Acid Derivatives Nucleophilic Acyl Substitution

Chapter 20 Carboxylic Acid Derivatives Nucleophilic Acyl Substitution Chapter 20 Carboxylic Acid Derivatives Nucleophilic Acyl Substitution Nomenclature: In carboxylic acid chlorides, anhydrides, esters and amides, the parent is the carboxylic acid. In each case be sure

More information

Spectroscopic Quantification of Kinetic Rate Constants for Epoxy-Acrylate Hybrid Photopolymerization

Spectroscopic Quantification of Kinetic Rate Constants for Epoxy-Acrylate Hybrid Photopolymerization Spectroscopic Quantification of Kinetic Rate Constants for Epoxy-Acrylate Hybrid Photopolymerization Brian Dillman University of Iowa, Chemical and Biochemical Eng. Dept., Iowa City, USA Julie L. P. Jessop

More information

A Novel Approach of Using NBS as an Effective and Convenient Oxidizing Agent for Various Compounds a Survey

A Novel Approach of Using NBS as an Effective and Convenient Oxidizing Agent for Various Compounds a Survey Journal of Chemistry and Chemical Sciences, Vol.8(1), 59-65, January 2018 (An International Research Journal), www.chemistry-journal.org ISSN 2229-760X (Print) ISSN 2319-7625 (Online) A Novel Approach

More information

Chapter 10 Radical Reactions

Chapter 10 Radical Reactions Chapter 10 Radical Reactions Introduction Homolytic bond cleavage leads to the formation of radicals (also called free radicals) Radicals are highly reactive, short-lived species Single-barbed arrows are

More information

Nucleic Acid Derivatised Pyrrolidone By: Robert B. Login

Nucleic Acid Derivatised Pyrrolidone By: Robert B. Login ucleic Acid Derivatised Pyrrolidone By: Robert B. Login Templated polymer synthesis is a growing technology with many recent references illustrating for example how employing ucleic acid bases attached

More information

BSc. II 3 rd Semester. Submitted By Dr. Sangita Nohria Associate Professor PGGCG-11 Chandigarh 1

BSc. II 3 rd Semester. Submitted By Dr. Sangita Nohria Associate Professor PGGCG-11 Chandigarh 1 BSc. II 3 rd Semester Submitted By Dr. Sangita Nohria Associate Professor PGGCG-11 Chandigarh 1 Introduction to Alkyl Halides Alkyl halides are organic molecules containing a halogen atom bonded to an

More information

Chapter - III THEORETICAL CONCEPTS. AOPs are promising methods for the remediation of wastewaters containing

Chapter - III THEORETICAL CONCEPTS. AOPs are promising methods for the remediation of wastewaters containing Chapter - III THEORETICAL CONCEPTS 3.1 Advanced Oxidation Processes AOPs are promising methods for the remediation of wastewaters containing recalcitrant organic compounds such as pesticides, surfactants,

More information

Chapter 17 Reactions of Aromatic Compounds. Electrophilic Aromatic Substitution

Chapter 17 Reactions of Aromatic Compounds. Electrophilic Aromatic Substitution Chapter 17 Reactions of Aromatic Compounds Electrophilic Aromatic Substitution Electrophile substitutes for a hydrogen on the benzene ring. Chapter 17: Aromatics 2-Reactions Slide 17-2 1 Mechanism Step

More information

Chapter 22: Amines. Organic derivatives of ammonia, NH 3. Nitrogen atom have a lone pair of electrons, making the amine both basic and nucleophilic

Chapter 22: Amines. Organic derivatives of ammonia, NH 3. Nitrogen atom have a lone pair of electrons, making the amine both basic and nucleophilic hapter 22: Amines. rganic derivatives of ammonia, 3. itrogen atom have a lone pair of electrons, making the amine both basic and nucleophilic 22.1: Amines omenclature. (please read) sp 3 Amines are classified

More information

Organic Molecules, Photoredox, and. Catalysis

Organic Molecules, Photoredox, and. Catalysis Organic Molecules, Photoredox, and Catalysis 1 What is Photoredox Catalysis 2 Transition Metal vs Organic Photoredox Transition Metal Catalysts Organic Catalyst Reprinted (2017) with permission from (Wangelin,

More information

Conjugated Systems, Orbital Symmetry and UV Spectroscopy

Conjugated Systems, Orbital Symmetry and UV Spectroscopy Conjugated Systems, Orbital Symmetry and UV Spectroscopy Introduction There are several possible arrangements for a molecule which contains two double bonds (diene): Isolated: (two or more single bonds

More information

Aromatic Compounds II

Aromatic Compounds II 2302272 Org Chem II Part I Lecture 2 Aromatic Compounds II Instructor: Dr. Tanatorn Khotavivattana E-mail: tanatorn.k@chula.ac.th Recommended Textbook: Chapter 17 in Organic Chemistry, 8 th Edition, L.

More information

Chapter 5. Nucleophilic aliphatic substitution mechanism. by G.DEEPA

Chapter 5. Nucleophilic aliphatic substitution mechanism. by G.DEEPA Chapter 5 Nucleophilic aliphatic substitution mechanism by G.DEEPA 1 Introduction The polarity of a carbon halogen bond leads to the carbon having a partial positive charge In alkyl halides this polarity

More information

Mechanistic Kinetic Modeling of Thiol Michael Addition Photopolymerizations via Photocaged Superbase Generators: An Analytical Approach

Mechanistic Kinetic Modeling of Thiol Michael Addition Photopolymerizations via Photocaged Superbase Generators: An Analytical Approach Supporting Information Mechanistic Kinetic Modeling of Thiol Michael Addition Photopolymerizations via Photocaged Superbase Generators: An Analytical Approach Mauro Claudino a, Xinpeng Zhang a, Marvin

More information

1.1 Is the following molecule aromatic or not aromatic? Give reasons for your answer.

1.1 Is the following molecule aromatic or not aromatic? Give reasons for your answer. Page 1 QUESTION ONE 1.1 Is the following molecule aromatic or not aromatic? Give reasons for your answer. 1.2 List four criteria which compounds must meet in order to be considered aromatic. Page 2 QUESTION

More information

Grafting polystyrene on Cellulose (CNC) by surface initiated. Atom Transfer Radical Polymerization (SI ATRP)

Grafting polystyrene on Cellulose (CNC) by surface initiated. Atom Transfer Radical Polymerization (SI ATRP) Grafting polystyrene on Cellulose (CNC) by surface initiated Abstract Atom Transfer Radical Polymerization (SI ATRP) Zhen Zhang, Gilles Sebe, Xiaosong Wang Grafting polymers on the surface of nanoparticles

More information

Aldehydes and Ketones : Aldol Reactions

Aldehydes and Ketones : Aldol Reactions Aldehydes and Ketones : Aldol Reactions The Acidity of the a Hydrogens of Carbonyl Compounds: Enolate Anions Hydrogens on carbons a to carbonyls are unusually acidic The resulting anion is stabilized by

More information

17.1 Classes of Dienes

17.1 Classes of Dienes 17.1 Classes of Dienes There are three categories for dienes: Cumulated: pi bonds are adjacent. Conjugated: pi bonds are separated by exactly ONE single bond. Isolated: pi bonds are separated by any distance

More information

Photoinitiating Systems for LED-Cured Interpenetrating Polymer Networks

Photoinitiating Systems for LED-Cured Interpenetrating Polymer Networks Journal of Photopolymer Science and Technology Volume 28, Number 1 (2015) 31 35 2015SPST Photoinitiating Systems for LED-Cured Interpenetrating Polymer Networks Suqing Shi 1,2, Feyza Karasu 1, Caroline

More information

Acid-Base -Bronsted-Lowry model: -Lewis model: -The more equilibrium lies to the right = More [H 3 O + ] = Higher K a = Lower pk a = Stronger acid

Acid-Base -Bronsted-Lowry model: -Lewis model: -The more equilibrium lies to the right = More [H 3 O + ] = Higher K a = Lower pk a = Stronger acid Revision Hybridisation -The valence electrons of a Carbon atom sit in 1s 2 2s 2 2p 2 orbitals that are different in energy. It has 2 x 2s electrons + 2 x 2p electrons are available to form 4 covalent bonds.

More information

Figure 4.10 HPLC Chromatogram of the Carbazole-Phenoxy Based Methacrylate

Figure 4.10 HPLC Chromatogram of the Carbazole-Phenoxy Based Methacrylate The percent yield of the methacrylation was 85.2 %, with a purity of 98.2 % determined by HPLC (Figure 4.10). Elemental analysis gave excellent agreement to expected elemental ratios (Table 4.2). Disregarding

More information

Organic Chemistry Review: Topic 10 & Topic 20

Organic Chemistry Review: Topic 10 & Topic 20 Organic Structure Alkanes C C σ bond Mechanism Substitution (Incoming atom or group will displace an existing atom or group in a molecule) Examples Occurs with exposure to ultraviolet light or sunlight,

More information

Organic Chemistry of Drug Degradation

Organic Chemistry of Drug Degradation Organic Chemistry of Drug Degradation Li New Jersey Email: minli88@yahoo.com Publishing Contents Chapter 1 Introduction 1 1.1 Drug Impurities, and the Importance of Understanding Drug Chemistry 1 Characteristics

More information

Topic 1: Quantitative chemistry

Topic 1: Quantitative chemistry covered by A-Level Chemistry products Topic 1: Quantitative chemistry 1.1 The mole concept and Avogadro s constant 1.1.1 Apply the mole concept to substances. Moles and Formulae 1.1.2 Determine the number

More information

Technology offer: Environmentally friendly holographic recording material

Technology offer: Environmentally friendly holographic recording material Technology offer: Environmentally friendly holographic recording material Technology offer: Environmentally friendly holographic recording material SUMMARY Our research group has developed a new photopolymer

More information

Chapter 16 Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group

Chapter 16 Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group Chapter 16 Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group Nomenclature of Aldehydes and Ketones Aldehydes are named by replacing the -e of the corresponding parent alkane with -al

More information

Interaction of Photoexcited Photoinitiators with Nitroxyl Radicals. Igor V. Khudyakov. Department of Chemistry, Columbia University, New York, NY

Interaction of Photoexcited Photoinitiators with Nitroxyl Radicals. Igor V. Khudyakov. Department of Chemistry, Columbia University, New York, NY Interaction of Photoexcited Photoinitiators with Nitroxyl Radicals Igor V. Khudyakov Department of Chemistry, Columbia University, New York, NY Introduction. Formulations which undergo photopolymerization

More information

Reactions at α-position

Reactions at α-position Reactions at α-position In preceding chapters on carbonyl chemistry, a common reaction mechanism observed was a nucleophile reacting at the electrophilic carbonyl carbon site NUC NUC Another reaction that

More information

2.2.7 Research Area Oxidative Coupling Reactions Methods and Mechanisms (M. Klußmann)

2.2.7 Research Area Oxidative Coupling Reactions Methods and Mechanisms (M. Klußmann) 2.2.7 Research Area Oxidative Coupling Reactions Methods and Mechanisms (M. Klußmann) Involved: E. Böß, P. Karier, K. M. Jones, T. Hillringhaus, C. Schmitz, J. Demaerel, B. Schweitzer-Chaput, N. Gulzar,

More information

HALOGENOALKANES (HALOALKANES) Structure Contain the functional group C-X where X is a halogen (F, Cl, Br or I)

HALOGENOALKANES (HALOALKANES) Structure Contain the functional group C-X where X is a halogen (F, Cl, Br or I) alogenoalkanes F322 1 ALOGENOALKANES (ALOALKANES) Structure ontain the functional group X where X is a halogen (F, l, or I) Types alogenoalkanes halogen is attached to an aliphatic skeleton alkyl group

More information

Chapter 12 Alcohols from Carbonyl Compounds: Oxidation-Reduction and Organometallic Compounds

Chapter 12 Alcohols from Carbonyl Compounds: Oxidation-Reduction and Organometallic Compounds Chapter 12 Alcohols from Carbonyl Compounds: Oxidation-Reduction and Organometallic Compounds Introduction Several functional groups contain the carbonyl group Carbonyl groups can be converted into alcohols

More information

Organic Chemistry SL IB CHEMISTRY SL

Organic Chemistry SL IB CHEMISTRY SL Organic Chemistry SL IB CHEMISTRY SL 10.1 Fundamentals of organic chemistry Understandings: A homologous series is a series of compounds of the same family, with the same general formula, which differ

More information

ACTIVATION OF C H BONDS BY LOW-VALENT METAL COMPLEXES ( THE ORGANOMETALLIC CHEMISTRY )

ACTIVATION OF C H BONDS BY LOW-VALENT METAL COMPLEXES ( THE ORGANOMETALLIC CHEMISTRY ) CHAPTER IV ACTIVATION OF C H BONDS BY LOW-VALENT METAL COMPLEXES ( THE ORGANOMETALLIC CHEMISTRY ) n the end of the 1960s the leading specialist in homogeneous catalysis Jack Halpern wrote [1]: to develop

More information

Elimination Reactions Heating an alkyl halide with a strong base causes elimination of a. molecule of HX

Elimination Reactions Heating an alkyl halide with a strong base causes elimination of a. molecule of HX Elimination eactions eating an alkyl halide with a strong base causes elimination of a molecule of X 1. Potassium hydroxide dissolved in ethanol and the sodium salts of alcohols (such as sodium ethoxide)

More information

Polymer Chemistry Prof. Dibakar Dhara Department of Chemistry Indian Institute of Technology, Kharagpur

Polymer Chemistry Prof. Dibakar Dhara Department of Chemistry Indian Institute of Technology, Kharagpur Polymer Chemistry Prof. Dibakar Dhara Department of Chemistry Indian Institute of Technology, Kharagpur Lecture - 10 Radical Chain Polymerization (Contd.) (Refer Slide Time: 00:28) Welcome back, and we

More information

Anionic Polymerization - Initiation and Propagation

Anionic Polymerization - Initiation and Propagation Anionic Polymerization Initiation and Propagation As in free radical polymerization, there are initiation and propagation steps. NH 2 NaNH 2 Na + + NH 2 + H 2 N CH: Propagation proceeds in the usual manner,

More information

UNIT 4 REVISION CHECKLIST CHEM 4 AS Chemistry

UNIT 4 REVISION CHECKLIST CHEM 4 AS Chemistry UNIT 4 REVISION CHECKLIST CHEM 4 AS Chemistry Topic 4.1 Kinetics a) Define the terms: rate of a reaction, rate constant, order of reaction and overall order of reaction b) Deduce the orders of reaction

More information

Innovative. Technologies. Chemie des Klebens Chemistry of Adhesives. Dr. Jochen Stock, Laboratory Manager CRL Germany: Neuss, November 27 th, 2013

Innovative. Technologies. Chemie des Klebens Chemistry of Adhesives. Dr. Jochen Stock, Laboratory Manager CRL Germany: Neuss, November 27 th, 2013 Chemie des Klebens Chemistry of Adhesives Dr. Jochen Stock, Laboratory Manager CRL Germany: Neuss, November 27 th, 2013 Innovative Technologies 1 Overview Chemie des Klebens Chemistry of Adhesives Introduction

More information

Spring Term 2012 Dr. Williams (309 Zurn, ex 2386)

Spring Term 2012 Dr. Williams (309 Zurn, ex 2386) Chemistry 242 Organic Chemistry II Spring Term 2012 Dr. Williams (309 Zurn, ex 2386) Web Page: http://math.mercyhurst.edu/~jwilliams/ jwilliams@mercyhurst.edu (or just visit Department web site and look

More information

Conjugated Dienes and Ultraviolet Spectroscopy

Conjugated Dienes and Ultraviolet Spectroscopy Conjugated Dienes and Ultraviolet Spectroscopy Key Words Conjugated Diene Resonance Structures Dienophiles Concerted Reaction Pericyclic Reaction Cycloaddition Reaction Bridged Bicyclic Compound Cyclic

More information

Organic Chemistry I Lesson Objectives, Lesson Problems, Course Outline Spring 2008

Organic Chemistry I Lesson Objectives, Lesson Problems, Course Outline Spring 2008 Organic Chemistry I Lesson Objectives, Lesson Problems, Course Outline Spring 2008 Lesson Date Assignment Lesson Objective Description Lesson Problems 4 14-Jan Chapter 1 Quiz Describe how bond polarity

More information

235 Organic II. Final Exam Review REACTIONS OF CONJUGATED DIENES 1,2 VS 1,4 ADDITION REACTIONS OF CONJUGATED DIENES

235 Organic II. Final Exam Review REACTIONS OF CONJUGATED DIENES 1,2 VS 1,4 ADDITION REACTIONS OF CONJUGATED DIENES b. the ompound 7 i 1 Spectral Data: singlet, 196.5 ppm singlet, 14.1 ppm singlet, 14.4 ppm doublet, 19.1 ppm doublet, 18.5 ppm 1 MR Mass Spectrum Absorbance Intensity Infrared Spectrum 65 91 9. Structure:

More information

Detailed Course Content

Detailed Course Content Detailed Course Content Chapter 1: Carbon Compounds and Chemical Bonds The Structural Theory of Organic Chemistry 4 Chemical Bonds: The Octet Rule 6 Lewis Structures 8 Formal Charge 11 Resonance 14 Quantum

More information

Ethers. Synthesis of Ethers. Chemical Properties of Ethers

Ethers. Synthesis of Ethers. Chemical Properties of Ethers Page 1 of 6 like alcohols are organic derivatives of water, but lack the labile -OH group. As a result, ethers, except for epoxides, are usually not very reactive and are often used as solvents for organic

More information

Preparation of Alkyl Halides, R-X. Reaction of alkanes with Cl 2 & Br 2 (F 2 is too reactive, I 2 is unreactive): R + X X 2.

Preparation of Alkyl Halides, R-X. Reaction of alkanes with Cl 2 & Br 2 (F 2 is too reactive, I 2 is unreactive): R + X X 2. Preparation of Alkyl alides, R-X Reaction of alkanes with Cl 2 & Br 2 (F 2 is too reactive, I 2 is unreactive): UV R + X 2 R X or heat + X This mechanism involves a free radical chain reaction. A chain

More information

Organic Chemistry. M. R. Naimi-Jamal. Faculty of Chemistry Iran University of Science & Technology

Organic Chemistry. M. R. Naimi-Jamal. Faculty of Chemistry Iran University of Science & Technology Organic Chemistry M. R. Naimi-Jamal Faculty of Chemistry Iran University of Science & Technology Chapter 5-2. Chemistry of Benzene: Electrophilic Aromatic Substitution Based on McMurry s Organic Chemistry,

More information