Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Mar 23.
Published in final edited form as: Chem Rev. 2022 Jan 4;122(6):5476–5518. doi: 10.1021/acs.chemrev.1c00409

Rational Design of Photocatalysts for Controlled Polymerization: Effect of Structures on Photocatalytic Activities

Chenyu Wu 1, Nathaniel Corrigan 2, Chern-Hooi Lim 3, Wenjian Liu 4, Garret Miyake 5, Cyrille Boyer 6
PMCID: PMC9815102  NIHMSID: NIHMS1858245  PMID: 34982536

Abstract

Over the past decade, the use of photocatalysts (PCs) in controlled polymerization has brought new opportunities in sophisticated macromolecular synthesis. However, the selection of PCs in these systems has been typically based on laborious trial-and-error strategies. To tackle this limitation, computer-guided rational design of PCs based on knowledge of structure-property-performance relationships has emerged. These rational strategies provide rapid and economic methodologies for tuning the performance and functionality of a polymerization system, thus providing further opportunities for polymer science. This review provides an overview of PCs employed in photocontrolled polymerization systems and summarizes their progression from early systems to the current state-of-the-art. Background theories on electronic transitions are also introduced to establish the structure-property-performance relationships from a perspective of quantum chemistry. Typical examples for each type of structure-property relationships are then presented to enlighten future design of PCs for photocontrolled polymerization.

Graphical Abstract

graphic file with name nihms-1858245-f0025.jpg

1. INTRODUCTION

1.1. Photocatalysts for Chemical Transformations

Catalysts are effective in flattening the energy profile of a chemical reaction. That is, by virtue of an appropriate catalyst, a reaction with an insurmountable energy barrier can take place with appreciable yield under milder conditions. For example, biological systems rely on enzymes as catalysts to catalyze complex biosynthetic reactions that produce the chemicals needed to sustain life.13 By mimicking this superior functionality, artificial catalysts have been progressively developed that have enabled a myriad of industrial reactions.47 This catalytic strategy has become commonplace and has broken new ground for chemical synthesis and chemical engineering, facilitating the production of commodity chemicals, pharmaceuticals, and finding use in minerals and petroleum refining, and energy generation, among many others.810

In particular, photocatalysts (PCs) are a special class of catalyst which harness light and convert solar energy into chemical energy.11 A PC operates by first absorbing light to generate an electronically excited state (PC*), which can then undergo an electron or energy transfer reaction. While many photosensitized chemical reactions exist, PCs provide greater specificity for chemical reaction activation and thus limit unwanted side reactions.12 With PCs, the use of light as an external stimulus provides temporal/spatial control over the reaction and allows chemical transformations over a wider temperature range, including at ambient temperatures. These merits have led to a revolution in organic11,13,14 and macromolecular7,1517 synthesis.

Macmillan, Yoon, Stephenson, and others led pioneering discoveries of new light-driven organic transformations catalyzed by PCs.11,14,1820 Though highly efficient, many of the earlier PCs only work with ultraviolet (UV) or violet/blue light. Another limitation is their reliance on precious transition metal centers such as iridium and ruthenium.11,14 To circumvent the toxicity and natural scarcity associated with these rare transition metals,21 fully organic PCs were carefully developed with the aid of a highly sophisticated organic synthesis tool-kit,13 allowing for organic PCs with diverse functional groups and readily customizable properties. As such, by rational design, specific and targeted functionalities can be judiciously incorporated into organic PCs, which greatly increases the versatility of these PCs for applications in chemical and material syntheses.22

Nevertheless, a rational strategy for constructing PCs is still needed. Due to sophistication and versatility of organic chromophore cores and substituents, specification of an “ideal” PC has been usually preceded by synthesis of a library of premodified compounds coupled with laborious screening. Without a priori knowledge, this lab-based strategy can be rather costly and poorly transferable between different areas and applications. In this regard, further efforts are needed to unravel the structure-property-performance relationships between the chemical structure of PCs and performance of photocontrolled polymerization. Indeed, by understanding the underlying mechanisms, the catalytic performance of PCs can be correlated with a range of properties including optical, photophysical, and electrochemical properties that govern the consecutive light absorption, excited state evolution, and electron/energy transfer processes in photocatalytic cycles.13,23 These predictive relationships are the key to enabling the rational design of PCs.24 Notably, PC design aided by theoretical calculations can reduce labor and time spent in discovering new functional PCs. In this vein, and because of a concomitant surge in the number of new PC structures and their applications, exceptional advances have been seen in the rational design of PCs over the past decade.

1.2. Photocatalysts for Photocontrolled Polymerization

Extending beyond small molecule transformations, photochemical transformations have endowed macromolecular synthesis with exceptional reaction control under mild and ambient conditions for the fabrication of functional materials, finding applications, for example, in electronics, coatings, and dentistry.15 Indeed, using photochemical reactions for regulating polymerization can allow for many external control approaches in materials processing, with a typical example being photoresist technology for the preparation of microprocessors.2527 If photochemical reactions undergo a catalytic cycle to reversibly activate and deactivate the polymerization, such a system is termed photocontrolled polymerization. Coupled with various modern techniques for polymer synthesis, which allow fine control over chain growth processes, photocontrolled polymerization catalyzed by PCs has provided a versatile toolbox for manufacturing functional polymer materials. The merits of photocontrolled polymerization systems include temporal and spatial control imparted by light,15 reaction orthogonality arising from light absorptions at discrete wavelengths,2834 and precise molecular weight control aided by modern polymerization techniques.17 Over the past decade, photocontrolled polymerizations have found broad applications, such as producing antimicrobial polymers,35,36 decorating biocompounds or organisms with functional polymers,37,38 synthesizing polymeric micelles with complex architectures,3943 surface patterning,4447 and 3D/4D additive manufacturing.4850 In addition, within a photocontrolled polymerization system, PCs uniquely regulate the activation/deactivation cycle; propagating species are generated under irradiation to allow chain growth and reversibly recombine with desired chain-end groups to become dormant upon cessation of irradiation. This is in contrast with photoinitiated polymerization systems, which may also effectively generate propagating species, but where removal of the irradiation stimulus does not lead to rapid and reversible recombination of the propagating species.51

Given the commanding role of PCs in controlling the activation/deactivation cycles in photocontrolled polymerizations, the performance of photocontrolled polymerizations, i.e., the efficiency,5254 robustness/tolerance to reaction conditions,5558 and orthogonality to different systems,31,59 are highly dependent on the properties of PCs. However, early PCs for photocontrolled polymerization systems were usually selected by screening transition-metal-based PCs6063 and organic PCs6467 that were reported successful in organic synthesis.13,14,68 Although some of the early PCs for polymerization were selected based on the similarity of chemical structures with respect to established examples,6972 such approaches are still essentially trial-and-error and do not properly take into account structure-property relationships. It was not until recently that the rational design of PCs was achieved by utilizing the established structure-property relationships to selectively tailor properties and functionalities of chromophores by the aid of computational analysis.52,7375

In this regard, the rational design of PCs with application-specific features has become imperative for addressing the central challenges confronted in various photocontrolled polymerization systems and expanding their application scope, such as decreasing the catalyst loading,52,54 ultrafast polymerization for rapid materials manufacturing,49 polymerization in ultralow volume for high-throughput systems,35 and oxygen-tolerant polymerization for open-air production of polymers.55 Apart from these, one of the most appealing advantages of rational PC design lies in its potential for developing orthogonal polymerization systems,28,29 where carefully considered combinations of PCs and photosystems consequently result in numerous possibilities for selectivity and orthogonality to enable precise control over complex hybrid polymerization systems.

1.3. Scope of the Review

This review aims to summarize the guiding principles (structure-property-performance relationships), theories, and methodologies which can be applied for the design of PCs in photocontrolled polymerizations. Particular focus will be given to photoredox catalysts as a major and most useful type of PC for catalyzing photocontrolled polymerization. Photoredox catalysts operate via a specific mechanism where the excited state PC undergoes electron transfer to activate the initiator and subsequent back electron transfer to deactivate the initiator. In the following context, “PC” exclusively refers to photoredox catalysts. This review serves as a guide to researchers who seek to rapidly design a PC for a photocontrolled polymerization system with specific functionalities and demand for certain application scenarios. Accordingly, this review will not make an exhaustive list of existing PCs but will instead provide methodologies and tools allowing rational design of PCs. We will begin with a brief outline of existing photocontrolled polymerization techniques and commonly employed PC series. Background theories and knowledge required for rational PC design are addressed in detail, including those for light absorption and excitation of PCs and rate constants of electronic transitions and electron transfer reactions. Correlation between performance of photocontrolled polymerization systems and key PC properties of concern are summarized. As the main body of the review, we will then successively present the knowledge established on PC structure-property-performance relationships and their use in the control of polymerizations.

2. PHOTOCONTROLLED POLYMERIZATION

Photocontrolled polymerization encompasses a broad range of polymerization techniques, including radical, cationic, ring opening, and ring-opening metathesis processes. In this section, these polymerizations are analyzed and classified according to their photocatalytic mechanisms.

2.1. Photocatalysis in Photocontrolled Polymerization

Photocontrolled polymerization is a particular kind of polymerization controlled by photocatalysis in which the PC reversibly activates and deactivates the polymerization. As shown in Scheme 1, photocatalysis is governed by photophysical processes (1), activation reactions (2), and deactivation reactions (6), whereas polymerization is performed via initiation (3), chain growth (4), and chain transfer reactions (5). The first step (1) of photocatalysis involves several photophysical processes such as photon absorption, fluorescence, phosphorescence, internal conversion (IC), and intersystem crossing (ISC; 1PC* → 3PC*). In the second step (2, activation step), the excited PC* reacts with an initiator through electron or energy transfer reactions. As a consequence, the initiator reaches a new electronic state, which decomposes in step (3) to form a propagating species and an end group species. The former can react with monomers, thereby initiating the polymerization. Upon sequential addition of monomers, the propagating species becomes polymeric and increases its chain length. During this chain growth (4), the propagating polymer chain can in some cases be deactivated by chain transfer (5) or by reversible recombination (6) to form a dormant chain, thereby completing the catalytic cycle. Since the photophysical (1), activation (2), and deactivation (6) steps are the key steps of photocontrolled polymerization, they are the focus of this review.

Scheme 1.

Scheme 1.

Key Steps in Photo-Controlled Polymerization

To illustrate the various photophysical processes in step (1), we can invoke a Jablonski diagram (Scheme 2), where the electronic states and associated vibrational levels of a molecule are shown. Upon absorbing a photon of appropriate frequency, the molecule initially at the vibrational ground state (ν = 0) of the electronic ground state S0 (i.e., S0,0) is vertically excited to some vibrational excited state (ν > 0) of S1 (denoted as S1,ν) or higher states in view of the Franck–Condon principle.76,77 Thereafter, the molecule undergoes rapid vibrational relaxation to S1,0. From S1,0, the excited molecule can vertically return to S0,ν via fluorescence by emitting a photon (Scheme 2, green arrow), which is also governed by the Franck–Condon principle. Alternatively, the excited molecule can undergo a nonradiative internal conversion (IC S1,0-S0,ν), with the extra electronic energy being converted into vibrational energy to be released through vibrational relaxation. When the spin–orbit coupling is appreciable, the molecule at S1,0 can also access a triplet manifold (e.g., ISC S1,0-T2,ν) and then rapidly release the vibrational energy. Similarly, the T1,0 molecule can reach S0,ν through phosphorescence by emitting a photon, or via a nonradiative ISC T1,0-S0,ν process, followed by the release of vibrational energy. Generally, vibrational relaxation and IC between Sn,0 and S1,ν and between Tn,0 and T1,ν are fast events with a time scale at the picosecond level, whereas fluorescence, phosphorescence, IC S1,0-S0,ν, and ISCs T1,0-S0,ν and S1,0-Tn,ν are slower events characteristic of a time scale between nanoseconds and seconds.

Scheme 2. Jablonski Diagram of Excited Statesa.

Scheme 2.

aNote: A, absorption; F, fluorescence; P, phosphorescence; IC, internal conversion; ISC, intersystem crossing; and VR, vibrational relaxation.

In principle, both 1PC* and 3PC* generated from step (1) can serve as the reactant in the activation step (2). However, in photocontrolled polymerization systems, the PC is present at low concentration (typically 10–1000 ppm relative to the initiator). Therefore, PC* must be sufficiently long-lived to diffuse and interact with the initiator. Due to the spin-forbidden feature of the ISC and phosphorescence, the lifetime of 3PC* (microseconds to milliseconds) is usually much longer than that of 1PC* (nanoseconds to microseconds). Therefore, 3PC* rather than 1PC* is usually the major active species in step (2) of photocontrolled polymerization, unless the concentration of PC or the initiator is high or PC and the initiator form a complex before photoexcitation. For this reason, this review mainly focuses on activation via the 3PC* species.

The collision of 3PC* with an initiator R-X forms an exciplex 3(PC/R-X)*, which can then provide different forms which are symbolized by PC•+/(R-X)•−, PC•−/(R-X)•+, or PC/3(R-X)*. The first two forms (PC•+/(R-X)•−, PC•−/(R-X)•+) are obtained by electron transfer (ET), whereas the last one is obtained by triplet energy transfer (TET). The three forms correspond to three different catalytic cycles as shown in Scheme 3. In the oxidative quenching of 3PC* (Scheme 3A), an electron is transferred from 3PC* to the initiator R-X, resulting in a cationic P•+ and an anionic (R-X)•−. The latter is further decomposed to an end group anion X and a propagating radical R for radical polymerization. PC•+ and X may form a PC•+/X complex in the solution, thereby facilitating the deactivation process where PC•+/X and another propagating species R′ recombine into PC and R′-X to complete the catalytic cycle. In contrast, in the reductive quenching of 3PC* (Scheme 3B), an electron is transferred from R-X to 3PC*, leading to an anionic PC•−, an end group radical X, and a propagating cation R+ which can be employed for cationic polymerization. Energy transfer from 3PC* to R-X via TET mechanism (Scheme 3C) results in the formation of homolytic cleavage of 3(R-X)*, with an end group radical X and a propagating radical R for radical polymerization.

Scheme 3. Catalytic Cycles of (a) ET via Oxidative Quenching Pathway, (b) ET via Reductive Quenching Pathway, and (c) TET in Photocontrolled Polymerizationa.

Scheme 3.

aNote: R-X, initiator; Pn-X, polymer chain capped with an end group X; X or X, end group in (a)/(b,c); R and Pn, propagating radicals for radical polymerization in (a,c); R+ and Pn+, propagating cations for cationic polymerization in (b); and M, monomer.

2.2. Classification of Photocontrolled Polymerization Systems

The distinct catalytic cycles depicted in Scheme 3 allow for an easy classification of photocontrolled polymerization systems. Both the oxidative quenching (Scheme 3A) and TET (Scheme 3C) pathways produce propagating radicals R, which are suitable for photoreversible deactivation radical polymerization (photo-RDRP, Scheme 4). Hence, photo-RDRP is compatible with a broad range of common vinyl monomers including acrylates, methacrylates, acrylamides, methacrylamides, styrenics, and others. In contrast, the reductive quenching pathway (Scheme 3B) produces propagating cationic species (R+), which are suitable for photoliving cationic polymerization (photo-LCP) and photoring opening metathesis polymerization (photo-ROMP).78,79 As shown in Scheme 4, photo-LCP is compatible with vinyl ethers, styrenics, and heterocyclic monomers (lactone and others), whereas photo-ROMP can polymerize many cyclic olefins.

Scheme 4. Classification of Photocontrolled Polymerizations with Their Corresponding Monomers (Left) and Compatible Catalytic Cycles (Right)a.

Scheme 4.

aNote: ET, electron transfer; OQP, oxidative quenching pathway; RQP, reductive quenching pathway; and TET, triplet energy transfer.

Major types of photo-RDRP techniques include photocontrolled atom transfer radical polymerization (photo-ATRP)51,60,64,72,8085 and photoinduced electron transfer-reversible addition-fragmentation chain transfer (PET-RAFT) polymerization,55,56,61,73,75,86 while other types include photocontrolled-nitroxide-mediated radical polymerization (photo-NMP),87,88 photocontrolled-cobalt-mediated radical polymerization (photo-CMRP),89,90 photocontrolled-iodine transfer polymerization (photo-ITP),91 photocontrolled-reversible complexation mediated living radical polymerization (photo-RCMP),92 and photocontrolled-organotellurium-mediated radical polymerization (photo-TERP),93 as listed in Scheme 4. Major types of photo-LCP techniques include photocationic RAFT polymerization31,65,94 and photocationic NMP (Scheme 4).95 In the absence of additional cocatalysts, photo-RDRP systems follow an oxidative quenching pathway. On the other hand, photo-LCP systems and photo-ROMP follow a reductive quenching pathway. Interestingly, PET-RAFT polymerization and photo-ATRP can also proceed through a reductive quenching pathway in the presence of additional electron-donating cocatalysts like tertiary amines.66,67,96,97 While there is evidence for TET in some mechanistic pathways for photocontrolled polymerization (such as photo-NMP,88 photo-ITP,91 and isolated cases of PET-RAFT polymerization98100 and photo-TERP93), ET reactions are mostly accepted as the dominant mechanism. As such, this review will only focus on ET pathways and indicate any possible involvement of TET where needed. It should also be noted that many of these polymerization techniques can also proceed through different catalytic mechanisms in the presence of additives,66,67,96 which is beyond the scope of this review.

2.2.1. Photocontrolled Atom Transfer Radical Polymerization (ATRP).

Photo-ATRP typically proceeds via an oxidative quenching mechanism, as shown in Scheme 5A. Reductive quenching mechanisms are also possible, but the addition of sacrificial reductants results in polymers with higher dispersity.101,102 By absorption of a photon, the ground state PC is first excited to the Sn state subsequently relaxing to S1 (1PC*); it can then undergo ISC to reach a long-lived T1 state (3PC*). Either 1PC* or 3PC* can reduce the ATRP initiator R-X (activation of photo-ATRP) to yield a propagating carbon-centered radical capable of initiating radical polymerization (monomer propagation), a halide ion X (a bromide ion Br, or a chloride ion Cl), and the radical cation PC•+. After a certain number of monomer additions to the carbon-centered radical, the PC•+/X pair encounters the propagating radical R to oxidize it and afford a dormant halogen capped polymer chain (Pn-X) and the regenerated ground state PC (deactivation of photo-ATRP). Because of the presence of the end group in Pn-X, the polymer chain can be repetitively reactivated by 3PC* for continued monomer additions followed by recapping of the chain carrier with the X. The greatest advantage of photo-ATRP systems over the traditional thermally mediated copper-catalyzed ATRP systems is the ability to externally and reversibly activate and deactivate the ATRP process with visible light to control chain growth. As presented in Scheme 5B, common initiators proven suitable for photo-ATRP include ethyl-α-bromophenylacetate (EBPA), ethyl-α-chlorophenylacetate (EClPA), ethyl α-bromoisobutyrate (EBiB), and benzyl α-bromoisobutyrate (BnBiB).

Scheme 5. (A) Proposed Mechanism of Photo-ATRP via an Oxidative Quenching Pathway and (B) Chemical Structures of ATRP Initiators Commonly Used in Photo-ATRPa.

Scheme 5.

aNote: ISC, intersystem crossing; M, monomer; X, bromine (Br) and chlorine (Cl) are the most commonly employed end groups (in some specific cases, iodine (I) was also employed); EBPA, ethyl-α-bromophenylacetate; EClPA, ethyl-α-chlorophenylacetate; EBiB, ethyl α-bromoisobutyrate; and BnBiB, benzyl α-bromoisobutyrate.

An important factor in the activation and deactivation of photo-ATRP is the ET reaction, specifically whether ET occurs via inner-sphere (ISET) or outer-sphere (OSET) electron transfer. For the activation process, the most accepted mechanism is OSET.72,103 The corresponding products are PC•+ and a negatively charged initiator (R-X)•− which decomposes into a propagating radical species (R) and a catalyst-halide complex, i.e., PC•+/X (OSET-2-body in Scheme 6). While this complex can reversibly dissociate to form free ions (i.e., PC•+ and X) in solution, this is an equilibrium process and is highly dependent on the nature of the PC, the halide anion X, and the solvent. Indeed, the solvent polarity will have a significant effect, influencing the complex binding energy and lifetime of these species, and correspondingly, the ability for efficient deactivation of the propagating radical species to occur.103

Scheme 6. Activation Pathways Proposed in Photo-RDRPa.

Scheme 6.

aNote: OSET-2-body, outer sphere electron transfer (OSET) that generates 2 bodies including the PC•+/X ion pair and R; OSET-3-body, OSET followed by decomposition of (R-X)•− into X and R.

Another activation pathway (OSET-3-body in Scheme 6) is also possible but results in three separate species (R, PC•+, and X).72 In this case, deactivation of photo-ATRP requires the recombination of all three species to provide a dormant polymeric species and return the catalyst to the ground state (i.e., 3-body-OSET in Scheme 7). Since such a three-body collision is entropically improbable, especially given the low concentrations of these species in solution, a two-body collision is a more likely pathway for deactivation in photo-ATRP (i.e., 2-body-OSET in Scheme 7).72,103 This two-body collision thus relies on the persistence of the PC•+/X complex in solution (produced upon the activation process via OSET-2-body in Scheme 6). Miyake and co-workers calculated the standard complexation Gibbs free energy for the formation ΔG0complex) of complexes featuring oxidized N,N-diaryl dihydrophenazines coupled with bromide anions, i.e., PC•+/Br.103 For the PC, 5,10-di(2-naphthyl)-5,10-dihydrophenazine, the ΔG0complex in DMF was found to be −4.2 kcal/mol, while the value in THF was −11.1 kcal/mol, suggesting that the formation of these complexes is exergonic in both these solvents. Notably, the more exergonic ΔG0complex in THF was ascribed to the lower solvent polarity, which favored complex association. Interestingly, the deactivation behavior of O-ATRP using this and similar PCs with charge transfer (CT) character was found to be tunable by adjusting the solvent polarity.103

Scheme 7. Deactivation Pathways Proposed in Photo-RDRPa.

Scheme 7.

aNote: 2-body-OSET, two bodies (the PC•+/X ion pair and R) undergo OSET to recover PC and R-X; 3-body-OSET, three bodies (PC•+, X,and R) undergo OSET to recover PC and R-X; and 3-body-ISET, three bodies (PC•+, X and R) undergo ISET to recover PC and R-X.

Matyjaszewski and co-workers have also studied O-ATRP mechanisms, finding that the addition of Br(using tetra-n-butylammonium bromide salt) provided more efficient deactivation during the polymerization of MMA catalyzed by N-phenyl phenothiazine under 365 nm irradiation.72 In this case, the addition of a large excess of free Br pushed the complex-free-ion equilibrium (i.e., PC•+/X ← → PC•+ + X) toward the formation of the complex (PC•+/X), and thus provided a higher concentration of effective deactivator, i.e., the catalyst complex, rather than the free ions. This data is consistent with Miyake and co-workers’ previous works and supports the deactivation occurring via a two-body collision event between the propagating radical species and the associated PC•+/X complex. These observations clearly supported the OSET-2-body (Scheme 6) as the dominant activation pathway and 2-body-OSET (Scheme 7) as the dominant deactivation pathway.

Further questions regarding the activation and deactivation mechanisms can be answered by studying reaction paths of these systems along with their energy barriers via quantum chemical calculations. The transition state theory104 (TST) or Rice-Ramsperger-Kassel-Marcus theory105 (RRKM) can be used to evaluate kinetic aspects of the minimum energy reaction paths. Alternatively, nonadiabatic molecular dynamics106 may be performed to simulate the activation/deactivation processes.

Initial photo-ATRP systems were catalyzed by transition-metal-based PCs (Scheme 8A). The first example of a photo-ATRP was reported by Fors and Hawker,60 who utilized fac-[Ir(ppy)3] (Scheme 8A, 1) as a PC to reversibly activate and deactivate the ATRP process. Following this initial work, more transition-metal-based PCs were exploited. For example, Le Lagadec, Alexandrova, and co-workers explored the use of Ru-based complexes (Scheme 8A, 2,3) to effectively catalyze photo-ATRP of methyl methacrylate (MMA), n-butyl acrylate (BA), and styrene (St) under visible light in the presence of EBiB as the initiator.62 Lalevée, Fensterbank, Goddard, Ollivier, and co-workers discovered a novel Au-based complex (Scheme 8A, 4) as PC for photo-ATRP of various methacrylates with EBPA as the initiator, under sunlight or visible light.107 Similarly, Lalevée and co-workers further demonstrated a class of Fe-based PCs (Scheme 8A, 57) for photo-ATRP of MMA and BA initiated by EBPA.108 Poly, Lalevée, and co-workers reported a Cu-based PC109 (Scheme 8A, 8) that also involves an oxidative quenching pathway in photo-ATRP.85

Scheme 8.

Scheme 8.

Chemical Structures of Representative PCs in Photo-ATRP Systems, Including (A) Transition-Metal-Based PCs, (B) Perylene, (C) Phenothiazine Derivatives, (D) Dihydrophenazine Derivatives, and (E) Phenoxazine Derivatives

In an effort to eliminate transition metal residues for more widespread applications, researchers started to seek metal-free alternatives to catalyze photo-ATRP. In this regard, Fors, Hawker, and co-workers demonstrated the use of an organic PC N-phenyl phenothiazine (Scheme 8C, 10) to effectively catalyze organocatalyzed photo-ATRP (O-ATRP) of methacrylates with EBPA as the initiator via an oxidative quenching pathway.64 Matyjaszewski and co-workers further expanded the monomer scope of O-ATRP to acrylonitrile.81 At the same time, independently, Miyake and Theriot demonstrated the successful implementation of perylene (Scheme 8B, 9) to catalyze O-ATRP of MMA with EBPA as the initiator in a controlled manner.110 Matyjaszewski and co-workers revisited the proposed mechanism behind O-ATRP and expanded the scope of organic PCs to more phenothiazine-based PCs including π-extended derivatives (Scheme 8C).72

Since 2016, aided by computationally directed discovery, Miyake and co-workers consecutively developed and explored broad libraries of dihydrophenazine (Scheme 8D),80 phenoxazine (Scheme 8E),74,111 and dimethyl-dihydroacridine derivatives (Scheme 9A)112 as efficient organic PCs for photo-ATRP. Remarkably, reduced catalyst loadings were achieved by pursuing core-extended organic PCs with CT excited states, which broadened absorption peaks to facilitate visible light absorption at longer wavelengths.54,111 The importance of the N-aryl substituent was highlighted with phenoxazines, where changing the N-aryl group from phenyl to 1-naphthalene greatly enhanced the ISC process to increase the triplet quantum yield (ΦT) to long-lived T1 states.113 Kim et al. reported a computer-aided molecular design/screening strategy to directly construct organic PCs with donor–acceptor (D–A) scaffolds from a large library of common organic donor and acceptor moieties.52 In addition, highly conjugated thienothiophene derivatives were investigated by Yagci and co-workers as PCs for application in O-ATRP (Scheme 9B).114 In photo-ATRP (or O-ATRP), the initiator efficiency is defined as the number of polymer chains with the “living” end group divided by the total number of initiators employed (i.e., the portion of initiators that are activated). High initiator efficiency can be achieved by using PCs with the following critical characteristics: (i) sufficiently reducing excited states, (ii) maximized population of long-lived T1 states, (iii) highly oxidizing PC•+ that can effectively deactivate the propagating radical, and (iv) PCs that are chemically inert to radicals.

Scheme 9.

Scheme 9.

Chemical Structures of Representative PCs in O-ATRP Systems, Including (A) Dimethyl-Dihydroacridine Derivatives and (B) Organic Donor–Acceptor Scaffolds

2.2.2. Photo-Induced Reversible Addition–Fragmentation Chain Transfer (RAFT) Polymerization.

In the case of additive-free PET-RAFT polymerization, an oxidative quenching cycle is followed for the generation of carbon-centered propagating radicals, i.e., via an ET reaction61,115 (Scheme 10A). Alternatively, a TET (Scheme 10B) mechanism can also be followed in certain cases. The ET mechanism is characterized by an electron transfer from the excited state PC* to the RAFT agent, which subsequently fragments to form a propagating radical species, an anionic thiocarbonylthio species, and a P•+ species. Recombination closes the catalytic cycles and generates a dormant thiocarbonylthio-capped polymer chain.

Scheme 10. Proposed Mechanisms for PET-RAFT Polymerization Activated by (A) ET via the Oxidative Quenching Pathway or (B) TET and (C) Chemical Structures of Thiocarbonylthio RAFT Agents Commonly Used in PET-RAFT Polymerizationsa.

Scheme 10.

aTrithiocarbonates: BTPA, BSTP, CDTPA, and CPDTC; dithiobenzoates: CPADB and CDB; xanthate. Note: M, monomer; Z, Z group of the RAFT agent.

For the activation step in PET-RAFT polymerization, an outer-sphere mechanism (e.g., OSET) should proceed regardless of an electron or energy transfer mechanism. However, the products of these reactions do differ considerably; for PET-RAFT polymerization via an ET mechanism, the products are a radical-cation PC species (P•+), an anionic radical thiocarbonylthio species (X•−), and the standard propagating radical (R). Hence, the ET mechanism will require recombination of three species, which will be entropically improbable, similar to the photo-ATRP deactivation. Thus, it is possible that some PC•+X•− species in PET-RAFT polymerization display exergonic ΔG0complex values and thus avoid the three-body collision requirement. It must be noted that there is limited information regarding the deactivation in PET-RAFT polymerization; further studies are required to fully elucidate this behavior.

As in any chemical process, both the thermodynamics and kinetics of the system will determine the resulting reaction behavior. For both PET-RAFT and photo-ATRP polymerization systems, removal of the light source results in rapid cessation of monomer propagation.116 This behavior indicates a rapid deactivation process, which is consistent with retention of the associated P•+X•− complex, allowing bimolecular recombination to close the catalytic cycles. Again, these systems are highly dependent on the reaction conditions as well as the structure of the chemical species involved, and investigation of specific systems should be performed to uncover the reaction behavior more clearly.

Alternatively, for a TET mechanism, the propagating radical species is formed via cleavage of the RAFT agent weak C–S bond, in a similar fashion to the photoiniferter process proposed by Otsu.117 This results in the formation of a radical thiocarbonylthio species (X) and a propagating radical species (R). In this case, a TET mechanism will provide only two products that require recombination to close the catalytic cycle. The only TET-activated systems discovered so far are PET-RAFT catalyzed by Ir- and Ru-based catalysts, reported by Allonas, Boyer, and co-workers using transient absorption spectroscopy and computational studies.98,100 As recently reported by Falvey and co-workers,118 it should be noted that higher triplet states such as T2 of the RAFT agent may be responsible for the TET-activated systems as T1 does not possess sufficient excitation energy to cleave the C–S bond but T2 presented the capability in some cases. Specifically, they found that some dithioester RAFT agents exhibited low-energy T1 states with adiabatic values calculated at ~33 kcal mol−1, leading to no (or negligible) decomposition in laser flash photolysis observations. On the other hand, the trithiocarbonate RAFT agents were found with ~44 kcal/mol T1 excitation energies, which are higher but still could not lead to effective decomposition. On the contrary, anionic RAFT•− resulting from ET reactions were found to undergo rapid decomposition to generate radicals. On the basis of these results, Falvey and co-workers drew a conclusion that PET-RAFT polymerization catalyzed by dyes with low-energy T1 cannot activate polymerization by TET but should by ET reactions instead.118 Further evidence for an ET mechanism in longer wavelength absorbing dyes characteristic of low-energy T1 has also been provided. For example, Smith, Seal, and co-workers combined experimental and computational evidence to conclude that the commonly used ZnTPP catalyzes PET-RAFT via an ET process.115 Hence, at least for PCs with longer-wavelength absorption,56 an ET mechanism is most likely.

Similar to that mentioned in section 2.2.1, accurate identification of an ET/TET mechanism in activation/deactivation of PET-RAFT polymerization should be conducted by calculating the minimum energy reaction paths, whereupon TST or RRKM methods based on the energy barrier can solve the kinetic problem.

Interestingly, the robustness of PET-RAFT polymerization has been demonstrated by Konkolewicz et al.,119 who showed that changes to reaction conditions such as light intensities and reactor geometries result in only marginal changes to polymerization control. While factors such as the light intensity and chemical concentrations will change the polymerization kinetics, polymer materials still display low dispersities under a range of conditions.

Inspired by photo-ATRP, Boyer and co-workers introduced Ir-based PC fac-[Ir(ppy)3] and Ru-based PC Ru(bpy)32+ (Scheme 11A, 57) to RAFT polymerization.61 These PCs were shown to be suitable for dithiobenzoate, trithiocarbonate, and xanthate RAFT agents. With dependence on the selection of the RAFT agent, a range of monomer families were polymerizable through PET-RAFT polymerization with fac-[Ir(ppy)3] as the PC, including (meth)acrylates, (meth)acrylamides, styrenics, isoprene, vinyl pivalate (VP), N-vinyl pyrrolidinone (NVP), and dimethyl vinylphosphonate (DVP). Boyer and co-workers explored the use of ZnTPP as a PC for PET-RAFT polymerization of acrylates, acrylamides, and styrene regulated by trithiocarbonates, as well as methacrylates regulated by dithiobenzoates.86 Notably, although ZnTPP can be regarded as a transition-metal-based complex, it does not exhibit any metal-to-ligand charge transfer (MLCT) excited states because of the inert nature of Zn; instead, ZnTPP mainly exhibits local excitation upon exposure to light.115 The oxygen tolerance featured in PET-RAFT polymerization catalyzed by ZnTPP was discovered based on the capability of ZnTPP to photosensitize oxygen (O2) via triplet–triplet annihilation (TTA) to yield singlet oxygen (1O2), which can be scavenged by monomers or the solvent.6,55 Boyer and co-workers further reported water-soluble derivatives for aqueous PET-RAFT polymerization.43,120

Scheme 11.

Scheme 11.

Chemical Structures of Representative PCs Employed in PET-RAFT Polymerization Systems, Including (A) Transition-Metal-Based PCs fac-[Ir(ppy)3] (1) and Ru(bpy)32+ (57), (B) Porphyrin Derivatives, (C) Naturally Derived Chlorophyll a (62), Bacteriochlorophyll a (63), and Pheophorbide a (64) and (D) Xanthene Derivatives

The naturally derived chlorophyll a (62),70 bacteriochlorophyll a (63),69 and pheophorbide a,59 (64) shown in Scheme 11C have also been explored for mediating PET-RAFT polymerization. While the former two natural PCs are generally successful, pheophorbide a (64) was found to only be compatible with dithiobenzoates for the polymerization of methacrylates and unsuccessful for the polymerization of acrylates in the presence of trithiocarbonates. The metal-free tetraphenyl porphyrin (TPP, Scheme 11B, 60) as a red-light-absorbing organic PC displayed similar performance to pheophorbide a and is active for dithiobenzoates but inactive with trithiocarbonates.86 Zhang and co-workers discovered a reduced TPP with a bacteriochlorin core (RTPP, Scheme 11B, 61) as an effective organic PC for PET-RAFT polymerization of both acrylates and methacrylates in the presence of CDTPA or CDB as RAFT agents.121 Most encouragingly, the bacteriochlorin-based RTPP has strong absorption in the far-red to near-infrared (NIR) range of the absorption spectrum, which makes it an unprecedented NIR-absorbing organic PC for efficient PET-RAFT polymerization. To explore more candidates as organic PCs for PET-RAFT polymerization, the Boyer group screened other common organic dyes and identified a Br-substituted fluorescein derivative Eosin Y (Scheme 11D, 65) as an effective PC for PET-RAFT polymerization of MMA in the presence of CPADB as the RAFT agent.67

In collaboration with the Miyake group, the Boyer group demonstrated the use of other halogenated derivatives of this dye class (Scheme 11D) as efficient PET-RAFT PCs and compared their performance in PET-RAFT polymerization in relation to their structure and properties.73,75 In particular, a newly synthesized Br-substituted xanthene dye tetra-bromofluorescein (Scheme 11D, 67) was discovered as a highly sensitive and efficient pH/light dual-responsive organic PC for PET-RAFT polymerization of acylates, acrylamides, methacrylates, and methacrylamides in the presence of BTPA, CPADB, and CPDTC (Scheme 10C).75 The merits of halogenated xanthene dyes are their exceptionally strong green-to-orange light absorption and efficient ISC to yield considerable long-lived T1 states. These properties make them ideal organic PCs for PET-RAFT polymerization, requiring only ~20 ppm catalyst loadings to yield an excellent polymerization efficiency.73 Good oxygen tolerance was also achieved in these xanthene dye-catalyzed systems due to their high ΦT, which enables scavenging of O2 species via TTA. To maximize the oxygen tolerance performance in organic PET-RAFT polymerization, another xanthene dye substituted with four heavy bromine atoms and four iodine atoms (Scheme 11D, 70) was judiciously designed and synthesized, exhibiting the best oxygen tolerance among organic PCs for PET-RAFT polymerization.73 Additionally, a few effective candidates developed for O-ATRP were also implemented for PET-RAFT polymerization, which were demonstrably successful.53,122,123

2.2.3. Photocontrolled Living Cationic Polymerization (LCP).

A typical photo-LCP system is the photocationic RAFT polymerization developed by Fors et al., which is characteristic of a reductive quenching pathway.65 Photocationic RAFT polymerization is activated by a photoexcited PC abstracting an electron from the RAFT agent IBDTC or IBTTC, forming a propagating cation and a leaving thiocarbonylthio anion (Scheme 12). The propagating cation can then undergo a cationic RAFT process to polymerize monomers in a controlled manner. By occasional diffusive collision, the negatively charged PC•− can reduce the propagating cation to close the catalytic cycle and restore the ground state PC and provide a polymer chain capped with a RAFT agent end group (deactivation). Photocationic RAFT polymerization is capable of effectively polymerizing vinyl ether monomers because of the electron-donating alkyloxy ligands attached to the vinyl group, which facilitates the propagation of cations. Common RAFT agents that can regulate photocationic RAFT polymerization include IBDTC and IBTTC (Scheme 12B), whose R group is an alkyloxy-substituted secondary carbon that mimics the poly-(vinyl ether) chain to optimize initiation of photocationic RAFT polymerization. Although 3PC* is denoted as the active species for the subsequent reductive quenching PET (Scheme 12), it still remains unclear whether 3PC* or 1PC* or both should be the dominant active excited state for photocationic RAFT polymerization,124 despite the fact that 3PC* appears to be more likely considering its much longer lifetime.

Scheme 12. (A) Proposed Mechanism for Photocationic RAFT Polymerization and (B) Common RAFT Agents and (C) Proposed Mechanism for Photocationic NMP Polymerization, (D) Common Monomers, and (E) Common Alkoxyamine Initiatorsa.

Scheme 12.

aNote: IBDTC, S-1-isobutoxyethyl N,N-diethyl dithiocarbamate; IBTTC, S-1-isobutoxylethyl S′-ethyl trithiocarbonate. M, monomer; Z, RAFT agent Z group.

In photocationic RAFT polymerization, the activation/deactivation mechanisms have been much less studied. Commonly, a cationic 1PC+ is used to ensure sufficient oxidizing capability in excited states, and hence 3PC+* is populated. Accordingly, in an OSET scheme (reductive quenching pathway), PC and (R-X)•+ are formed and the latter decomposes to X and R+ for cationic polymerization. Although current reports assumed that PC and X were formed, existing evidence may not rule out the possibility of forming 1PC+X ion pairs from the activation process, especially considering the tendency of the RAFT end group to be anionic (X, as with those in PET-RAFT) and the tendency of the PC being restored to the original cationic 1PC+ ground state. Indeed, the formation of the 1PC+X ion pair may facilitate a two-body collision with R+ which may effectively deactivate the polymerization to recover 1PC+ and R-X. Overall, the exact activation/deactivation mechanisms of photocationic RAFT polymerization are still under debate and further investigations are needed, including computational study of the ISET mechanism, to uncover the reaction paths.

The first class of PCs that were proven to be effective for photocationic RAFT polymerization were triphenylpyrylium derivatives (Scheme 13A).124 However, these organic dye-catalyzed systems still showed some dark polymerization, i.e., continued polymerization upon cessation of irradiation.65,124 To circumvent this problem and obtain ideal temporal control for photocationic RAFT polymerization, Fors and co-workers further introduced a series of modified Ir-based complexes (Scheme 13C), which has resulted in greatly improved temporal control for photocationic RAFT polymerization, despite some reduction in polymerization rates.94 More recently, Liao and co-workers have introduced a new metal-free cationic polymerization of vinyl ethers utilizing bisphosphonium salt doped anthanthrenes as organic PCs (Scheme 13D). These PCs present very strong visible light absorption and tunable redox potentials. The authors demonstrated temporal control as well as control of the molecular weights of a broad range of vinyl ether polymers.125 In particular, compared with other existing organic PCs for photocationic RAFT polymerization, these bisphosphonium salts display better temporal control with no dark polymerization after ceasing light irradiation.

Scheme 13.

Scheme 13.

Chemical Structures of Representative PCs in Photo-LCP and Photo-ROMP Systems, Including (A) Triphenylpyrylium Derivatives, (B) Triphenylthiopyrylium Derivatives, (C) Substituted Ir-Based Complexes, and (D) Bisphosphonium Salt Derivatives

Another photo-LCP system is photocationic NMP developed by Kamigaito and co-workers,95 whose photocatalytic mechanism (Scheme 12C) is similar to its RAFT counterpart (Scheme 12A), albeit with replacement of RAFT agents with alkoxyamines activated by mesolytic cleavage (Scheme 12E). With regard to the polymerization mechanism, it should be noted that alkoxyamines do not undergo chain transfer like the RAFT equilibrium and thus rely on the redox cycle to retain controllable chain growth. Apart from the vinyl ether monomers applied in cationic RAFT, photocationic NMP is also compatible with styrene derivatives (Scheme 12D).

2.2.4. Photocontrolled Ring-Opening Metathesis Polymerization (ROMP).

Photo-ROMP, also known as metal-free ROMP, was established by Boydston and co-workers.79,126 Photo-ROMP is mechanistically distinct from traditional ROMP systems that are typically initiated using organometallic Ru, Mo, and W alkylidene complexes.127,128 As shown in Scheme 14A, photo-ROMP operates via a reductive quenching pathway in which an excited state PC oxidizes an enol ether initiator to produce an enol ether radical cation chain carrier. This chain carrier can react with strained cyclic olefins to provide a cyclobutane radical cation intermediate species, which then undergoes rapid ring opening to form the propagating species. Subsequent reduction of the propagating species by the reduced PC•− closes the catalytic cycle and provides an enol ether capped polymer chain. Under this mechanism, the ring opening of the cyclobutane intermediate must be sufficiently fast to overcome reduction of the radical cation intermediate to form stable cyclobutane side-products.126

Scheme 14. (A) Proposed Mechanism for Photo-ROMP via a Reductive Quenching Pathway and (B–C) Chemical Structures of the (B) Initiators and (C) Monomers Used in Photo-ROMPa.

Scheme 14.

aNote: ISC, intersystem crossing; DCPD, dicyclopentadiene.

In much the same way as early PC selection for photo-ATRP and PET-RAFT was based on PCs from organic synthetic transformations, the first PCs for photo-ROMP were selected based on prior work on the photoredox-mediated synthesis of cyclobutanes under visible light by Nicewicz et al.,129 namely triphenyl pyrylium derivatives. Furthermore, because of its reductive quenching pathway, the PC scope for photo-ROMP is similar to that for photo-LCP and includes other blue light absorbing pyrylium and thiopyrylium derivatives.78 By systematically studying these varied pyrylium (Scheme 13A) and thiopyrylium (Scheme 13B) PCs, Boydston et al. observed higher efficiency of thiopyrylium derivatives over their corresponding pyrylium counterparts.78 Critically, these PCs must have sufficient reduction potential in the excited state to oxidize the enol ether initiators, which typically have reduction potentials in the range of 1.4 V versus SCE.126 Notably, however, it was shown that using PCs with higher excited state reduction potential led to slower polymerizations under analogous conditions.78 This effect has been suggested to be a result of overoxidation of the enol ether initiators and demonstrates the need for carefully selected PC-initiator pairs. Regardless, a range of enol ether initiators have been successfully employed in photo-ROMP, as shown in Scheme 14B.

Photo-ROMP has expanded its monomer scope to a range of functional norbornene and dicyclopentadiene (DCPD) monomers (Scheme 14C) for the preparation of block (co)-polymers.130135 Interestingly, Boydston et al. showed that endo-DCPD, exo-DCPD monomers can be used to prepare linear polyDCPD polymers; these monomers readily form cross-linked thermoset materials when polymerized via traditional ROMP, and the linear forms have been previously inaccessible. These linear polymers could be readily cross-linked via a secondary thiol–ene reaction if required.130 In addition, ether, silyl, and chloro-functionalized norbornenes, among others, were all shown to be suitable monomers or comonomers for photo-ROMP, leading to a broad range of functional polymers.131,132,134 The high functional group tolerance shows the versatility of this technique and potential applicability for orthogonal chemistries.133,134 Furthermore, photo-ROMP has been successfully performed under ambient conditions, which increases its appeal from an industrial perspective.126,131

More recent work in photo-ROMP has shown that the counterion for the pyrylium PC can have a marked influence on the photo-ROMP.132,136 In particular, it was shown that strongly coordinating counterions chloride, bromide, and nitrate precluded the photo-ROMP of norbornene. Conversely, noncoordinating counterions such as tetrafluoroborate (BF4), hexafluorophosphate (PF6), and tetrakis(3,5-bis-(trifluoromethyl)phenyl)borate (BArF4) all provided effective photo-ROMP of norbornene. More interestingly, as the size of the counterion increased from BF4 to PF6 to BArF4, the trans alkene content of the resulting polynorbornene decreased from 79% to 74% to 58%.136 Moreover, solvent and temperature were also shown to affect the stereochemistry of polynorbornene produced through photo-ROMP, with >98% trans alkene content achievable using dichloromethane at low temperatures (−76 °C). Quantum chemical calculations showed that the smaller counterions stabilized a pro-trans isomer of the radical cation species formed after ring opening, leading to a higher trans-content in the final polynorbornene.136

3. GUIDING THE DESIGN OF PHOTOCATALYSTS FOR PHOTOCONTROLLED POLYMERIZATION

3.1. Qualitative Structure-Property-Performance Relationships

Irrespective of other functionalities, by assuming the same polymerization formulation (monomer type, initiator, and solvent, etc.), key performance parameters of a photocontrolled polymerization include the initiator efficiency, dispersity of polymers, and the apparent polymerization (propagation) rate kpapp. As mentioned before, the photocatalytic process initiating a photocontrolled polymerization is divided into four main stages: (i) light absorption, (ii) populating the excited state species responsible for catalysis, (iii) electron/energy transfer to generate propagating species, and (iv) the deactivation reaction to complete the catalytic cycle. Tuning key properties of the PC with respect to each of these stages (Scheme 15) will result in variations in the performance of photocontrolled polymerization systems.

Scheme 15. Roadmap of the Photocatalytic Pathway of a Typical Photocontrolled Polymerizationa.

Scheme 15.

aNote: λmax, maximum absorption wavelength of a given absorption peak; εmax, maximum molar extinction coefficient of a given absorption peak; Eex,S0–S1, excitation energy of the S1 state; Eex,S0-T1, excitation energy of the T1 state; f, oscillator strength of excitation to an excited state; τS1, lifetime of the S1 state; τT1, lifetime of the T1 state; kF, fluorescence rate constant; kIC, internal conversion rate constant; kISC, intersystem crossing rate constant; ϕT, triplet quantum yield; IP, ionization potential; EA, electron affinity; E*ox, excited state oxidation potential; E*red, excited state reduction potential; EUSOMO, energy level of the upper singly occupied molecular orbital; and ELSOMO, energy level of the lower singly occupied molecular orbital.

The properties of PCs depend on their electronic structures and are thus strongly correlated with their chemical structures. As a PC can be seen as a chromophore core with functional substituents attached, the properties of PCs can be manipulated by systematically changing their substituents. These functional substituents can include but are not limited to electron-donating groups, neutral groups, electron-withdrawing groups, heavy atom groups, and a myriad of conjugated groups with different electronic structures (as shown in Scheme 16A). Commonly used organic chromophore cores for a PC (which this review mainly focuses on) vary from highly reducing/oxidizing variants to large conjugated variants and water-soluble variants (Scheme 16B). By the aid of existing and new chromophore cores and substituents, PCs can be rationally designed according to the required performance of photocontrolled polymerization for a given application.

Scheme 16.

Scheme 16.

Various (A) Functional Substituents and (B) Organic Chromophore Cores Commonly Used in Designing a PC in Photocontrolled Polymerization

To this end, it is important to determine the guiding principles for rational design of a PC to attain better performance of photocontrolled polymerization (e.g., higher polymerization rates, higher initiator efficiency, and lower dispersity polymers) as well as tunable absorption wavelengths. Such guiding principles can be summarized as structure-property-performance relationships. Scheme 17 provides an overview of key relationships known to date with regard to variations of chemical groups listed in Scheme 16. These structure-property-performance relationships are discussed in section 3.1.1 for tuning light absorption, section 3.1.2 for increasing ΦT, and section 3.1.3 for enhancing activation/deactivation efficiencies, respectively. On the other hand, existing examples which reflected or made use of these structure-property-performance relationships are demonstrated in detail in section 3.2.

Scheme 17.

Scheme 17.

Overview of Key Structure-Property-Performance Relationships Known to Date in Designing a PC

3.1.1. Tuning Light Absorption.

3.1.1.1. Tuning Absorption Wavelengths.

The absorbed photon must possess energy hv close to the energy gap between the final state Sn and the initial state S0,137,138 i.e., the corresponding excitation energy. Accordingly, substituents that can tune the excitation energy by moving up/down relevant orbitals should in turn tune the absorption wavelength λ=cv (vide infra).

For a typical ππ* excitation, which is useful and commonly seen in organic PCs, increasing the dimensions of the π-conjugation can stabilize the lowest π* orbital or destabilize the highest π orbital, thereby extending λ of the maximum of the absorption peak (λmax). By contrast, directly attaching atoms with nonbonded (n) electrons (i.e., alkyloxy, alkylthio, or (di)alkylamino substituents) on the chromophore core will mix the high-lying n orbital into the highest π orbital of the chromophore core and hence extend λmax.139,140 Similarly, heavy halogen substituents can also mix their n orbitals into the highest π orbital, while a heavier halogen substituent with a higher-lying n orbital leads to longer λmax. On the other hand, more electron-withdrawing substituents extend λmax mostly by stabilizing the lowest π* orbital.

3.1.1.2. Maximizing the Molar Extinction Coefficient.

With respect to a specific absorption peak associated with an electronic transition if, the molar extinction coefficient ε has the maximum value εmax at λmax of an absorption peak and has a relation:

εmaxfifΔν¯FWHM (1)
fif=4meπc3e2λ|Φf|μ^|Φi|2 (2)
Φf|μ^|Φi=Φf(r)μ^Φi(r)dr (3)

where fif is the probability of the electronic transition defining the relative strength of the electronic transition, Δν¯fwhm is the full width at half-maximum of the absorption peak, me is the weight of an electron and e is the charge of an electron, and r is the position vector representing any point in the space. Φf|μ^|Φi is the electronic transition dipole moment, which is an integral in the whole space. Usually, for a typical ππ* excitation, larger π conjugation of the chromophore scales up the distribution of the transition density ρif(r)=Φf(r)Φi(r), rendering the amplitude of its dipole moment |Φf|μ^|Φi| larger and hence enhancing εmax.

3.1.1.3. Broadening the Absorption Peak.

Under the quantum-mechanical Franck–Condon principle76,77 and the displaced harmonic oscillator (DHO) model,137,138 higher Huang–Rhys factor S corresponds to a larger displacement between the potential energy surfaces of the two states in the electronic transition, which results in a broadening effect in the absorption peak. For example, flexible molecules usually exhibit a larger difference between the equilibrium geometries of Sn and S0, which leads to higher S and broader absorption peaks.56 This phenomenon is also commonly seen in the rather broad emission peaks of charge transfer states, where large deviation in equilibrium geometries is attributed to the vastly different electronic wave function of the charge transfer Sn as compared to S0.80,141

3.1.2. Increasing the Triplet Quantum Yield.

On the basis of the rate constants (kF for fluorescence, kIC for IC, and kISC for ISC), the quantum yield ΦT1 of 3PC* can be calculated according to eq 4 (Scheme 18).75 Accordingly, approaches to enhance ΦT can be practically categorized as (i) increasing kISC, (ii) reducing kF, or (iii) reducing kIC.75

ΦT1=[kISC,S1T2(1kISC,T2S1kISC,T2S1+kIC,T2T1)+kISC,S1T1kISC,T1S1]/[kF+kIC,S1S0+kISC,S1T2(1kISC,T2S1kISC,T2S1+kIC,T2T1)+kISC,S1T1kISC,T1S1] (4)
Scheme 18. Jablonski Diagram for Rate Constants of Typical Excited State Decay Pathwaysa.

Scheme 18.

aNote: kF, fluorescence rate constant; kP, phosphorescence rate constant; kIC, internal conversion rate constant between two electronic states of the same multiplicity; kISC, intersystem crossing rate constant from a singlet state to a triplet state; and kISC, “reversed” intersystem crossing rate constant from a triplet state to a singlet state.

3.1.2.1. Increasing kISC.

In accordance with Fermi’s golden rule expression regarding ISC,142144 kISC is proportional to the squared amplitude of spin–orbit coupling matrix elements (SOCME), i.e., |Φf|H^SO|Φi|2, whereas SOCME scales as the fourth power of the atomic charge ZN manifesting the heavy-atom effect and also becomes appreciable when the initial and final states differ in an orbital of different angular momentum (e.g., 1 (π,π*) → 3(n,π*)) manifesting the El-Sayed rule.145 Also, an energy gap law exists such that narrower energy gap between the initial and final states is responsible for exponentially higher kISC.146

Briefly, the heavy atom effect can be introduced by heavy substituents such as heavy halogens,73 whereas the El-Sayed rule can be implemented by including 1 (π,π*) → 3(n,π*) transition characters or ligand-to-ligand charge transfer characters in the ISC process. By minimizing the energy gap between S1 and a final state in the triplet manifold whose energy is close to, but lower than S1, the energy gap law takes effect.52

In addition, by considering the second-order perturbation theory and the first- and second-order dependence of SOCME on nuclear motions, there is a spin-vibronic coupling effect142 commonly seen in chromophore core-twisted molecules, whose mechanism is beyond the present scope for qualitative discussions. For a detailed outline of spin-vibronic coupling, the interested reader is directed to a recent review.142

3.1.2.2. Decreasing kF and kIC.

According to Fermi’s golden rule expression for fluorescence,143,147 kF is proportional to the squared transition dipole moment |Φf|μ^|Φi|2, which is greatly diminished with poor overlap between Φf and Φi. Hence, S1 with charge transfer characters should largely retard fluorescence due to poor overlap between wave functions of S1 and S0. On the other hand, the IC process results from the coupling of nuclear motions into the electronic transition. Accordingly, kIC can be effectively reduced by using more rigid and planar molecules with less nuclear mobility.

3.1.3. Completing the Catalytic Cycle by Efficient Electron Transfer Reactions.

As addressed in section 2.1, photocontrolled polymerization needs a complete photocatalytic cycle in both oxidative (Scheme 19A) and reductive (Scheme 19B) quenching pathways. Given the same monomer and reaction formulation, efficient polymerization and good control over the polymerization process is attained by the combination of effective activation and effective reversible deactivation. Hence, the redox properties of the PC corresponding to activation (3PC* → PC•+ for oxidative quenching and 3PC* → PC•− for reductive quenching) and deactivation (PC•+ → PC for oxidative quenching and PC•− → PC for reductive quenching) are the key to effective control over the polymerization process.

Scheme 19. (A–B) Illustration of the Photocatalytic Cycles via (A) the Oxidative Quenching Pathway and (B) the Reductive Quenching Pathway and (C) Derivation of Redox Potentials from Gibbs Free Energiesa.

Scheme 19.

aNote: E0(B/A): the standard potential for the A → B half reaction of the A + C → B + D electron transfer reaction. E0(triplet): the triplet excitation energy in eV.

In accordance with the Marcus formula for the electron/energy transfer rate constant kET (i.e., Marcus theory),148 it was concluded that the electron transfer rate kET scales exponentially with more negative ΔG0 of the electron transfer reaction (3PC* + R-X → PC•+ + (R-X)•− for oxidative quenching and 3PC* + R-X → PC•− + (R-X)•+ for reductive quenching). Within the scope of this review, we focus on the half reaction of PC in the electron transfer reaction and only deal with the redox property ΔG0PC of PC in ΔG0=ΔG0PC+ΔG0initiator for both the activation and deactivation reaction. On the other hand, the experimental convention for quantifying the redox properties of PCs is most commonly the redox potentials obtained using cyclic voltammetry. In this context, ΔG0PC is expressed as ΔG0PC=F[EPC], where F is the Faraday constant (F = 23.06 kcal V−1 mol−1) and EPC is the redox potential for PC. By implementing the standard redox potential E0 with reference to the saturated calomel electrode (SCE), the relation between E0 and ΔG0PC is given in Scheme 19C. Consequently, we can employ the E0 metric convention to reinterpret the exponential dependence of kET on ΔG0.

Scheme 20 illustrates the property-performance relationships which associate the energy levels of frontier orbitals to redox potentials and the efficiencies of activation/deactivation processes in photocontrolled polymerization. For the oxidative quenching pathway (Scheme 20A), a higher-lying upper-SOMO of 3PC* leads to smaller ionization potential (IP) and hence more negative E0(PC•+/3PC*) dictating more efficient activation, whereas a lower-lying HOMO of PC leads to larger electron affinity (EA) and thereby more positive E0(PC•+/PC) for more efficient deactivation. For the reductive quenching pathway (Scheme 20B), a lower-lying lower-SOMO of 3PC* and correspondingly larger EA result in more positive E0(3PC*/PC•−) for efficient activation while a higher-lying LUMO of PC and a smaller IP give rise to more negative E0(PC/PC•−) for efficient deactivation. In the following context, we discuss the cases for activation and deactivation, respectively, regarding the oxidative quenching pathway (Scheme 20A) and the reductive quenching pathway (Scheme 20B).

Scheme 20.

Scheme 20.

(A–B) Illustration of the Relationships among Activation/Deactivation Efficiencies, Redox Potentials, and Energy Levels of Frontier Molecular Orbitals for (A) the Oxidative Quenching Pathway and (B) the Reductive Quenching Pathway

3.1.3.1. Effective Activation.

With respect to oxidative quenching, more negative E0(PC•+/3PC*) signifies a less negative ΔG0PC (Scheme 19C) which indicates a higher-lying upper SOMO (Scheme 20A). Because in a typical organic PC, the upper SOMO mostly consists of contributions from the π* orbital of the chromophore core, methods that can increase the π* energy level contribute to more negative E0(PC•+/3PC*) and enhance the reducing ability of 3PC*, thus facilitating more effective electron transfer via oxidative quenching. These methods include introducing less electron-withdrawing (i.e., more electron-donating) substituents or smaller dimensions of the π conjugation to destabilize the π* orbital. Alternatively, higher-lying upper SOMO can also be achieved by mixing the high-lying n orbitals (where the n electrons occupy) into the π* orbital; substituents with n electrons (e.g., alkyloxy, alkylthio, or (di)alkylamino groups) directly connected to the π-conjugated chromophore core usually work well.80

By contrast, more efficient reductive quenching of 3PC* is represented by more positive E0(3PC*/ PC•−) and lower-lying lower SOMO, where the vacancy (or the “hole”) lies after electron excitation (Scheme 20B). In organic PCs, the lower SOMO of 3PC* is usually π orbitals of the chromophore where more electron-withdrawing substituents decreases the π orbital energy level and leads to a more positive E0(3PC*/PC•−) and thus a higher oxidizing ability of 3PC*.

3.1.3.2. Reversible Deactivation.

In all photocontrolled polymerization mechanisms, as shown in Scheme 19, reversible deactivation is attained through the last step of a catalytic cycle, where the propagating species is oxidized by PC•+ (in an oxidative quenching catalytic cycle) or reduced by PC•− (in a reductive quenching catalytic cycle). Effective and reversible deactivation is highly important in three aspects: (i) Upon activation of an initiator, propagation of the disassociated active species (radical/cation) through additions of monomers is continued until the propagating radical is deactivated by PC•+ or PC•− leading to a dormant chain. Efficient deactivation thus minimizes dark polymerization and imparts precise temporal/spatial control by switching the light on/off. (ii) In photocontrolled polymerization, successful deactivation plays a major role in controlling the average polymer chain length. After activation of the initiator, the propagating species must be rapidly deactivated so as to allow only a few additions of monomers per activation cycle. This increases the chance of uniform growth of different chains to attain narrower molecular weight distributions. Although this may not be obvious in RAFT-based techniques where chain transfer exists, it can be a major factor in photo-ATRP techniques. (iii) Effective deactivation also keeps radical concentration low to prevent radical–radical quenching leading to nonunity initiator efficiency or nonuniform molecular weight distribution.

Consequently, the deactivation process usually needs to be highly exothermic for photocontrolled polymerization, so that the dormant polymer chain can be generated for the next photoactivation event. In this case with regard to oxidative quenching, the PC•+ species should be highly oxidizing with E0(PC•+/PC) excessively more positive than E0(initiator/initiator•−), contributed by lower-lying HOMO of PC (Scheme 20A); with regard to reductive quenching, the PC•− species should be highly reducing with E0(PC/PC•−) excessively more negative than E0(initiator•+/initiator), contributed by higher-lying LUMO of PC (Scheme 20B).

3.2. Examples for Catalyst Design in Photocontrolled Polymerization

3.2.1. Tuning Absorption: Bathochromic and Hyperchromic Effect.

3.2.1.1. Heavier Halogen Substitution.

Perhaps the most well-known strategy to tune the λmax and εmax of a chromophore is the introduction of halogen substituents. While it is generally accepted that heavier atom substitution results in longer λmax and slightly higher εmax,149 heavy halogen substitution is the most effective and readily applicable method. For example, recent reports on modifying iridium complexes with heavier halogens attributed the observed bathochromic shift and hyperchromic shift to the heavy halogen substituents, which stabilizes the lowest unoccupied molecular orbital (LUMO) while leaving the highest occupied molecular orbital (HOMO) mostly unperturbed.150,151

Similarly, halogenated xanthene dyes (XAN-1, XAN-2, XAN-3, and XAN-4 as examples in Figure 1) were implemented in PET-RAFT polymerization to explore how the variation of halogen substitution tunes λmax and εmax of the chromophore.73 Theoretical calculations showed that the HOMO → LUMO transition has more than 98% contribution to the S0 → S1 vertical excitation of these halogenated xanthene dyes and indicated that the evolution of λmax by halogen substitution can be exclusively described by the nature and energy levels of the HOMO and LUMO. Specifically, because the nonbonded (n) electrons of the halogen substituents have a significant contribution to the population of the HOMO, heavier halogens that naturally exhibit higher energy n electrons (i.e., I > Br) lead to higher HOMO energy levels. This phenomenon was shown by comparing the xanthene dyes XAN-1 with iodine substituents (Figure 1A) and XAN-2 with bromine substituents (Figure 1B). On the other hand, these xanthene core halogens do not contribute appreciably to the LUMO, and thus LUMO energy levels are less effected when comparing XAN-1 (Figure 1A) and XAN-2 (Figure 1B). Consequently, the heavier I-substituted XAN-1 has narrower HOMO/LUMO energy gap and longer λmax = 548 nm compared to Br-substituted XAN-2 with λmax = 540 nm. A similar trend was observed for I-substituted XAN-3 with longer λmax = 563 nm compared to Br-substituted XAN-4 with λmax = 555 nm.73 It was also found that heavier halogen substitution also leads to an increase in εmax (Figure 1), which had contributions from both narrower S1/S0 energy difference and higher transition electric dipole moment by the heavy atom effect.73 Because of the notable bathochromic shift brought by introduction of heavier halogen, XAN-3 (Rose Bengal, λmax = 563 nm) was reported to enable PET-RAFT polymerization mediated by yellow (560 nm) LED light, in comparison to the more commonly known XAN-2 (Eosin Y, λmax = 540 nm) which favors PET-RAFT polymerization under green (530 nm) LED light.73

Figure 1.

Figure 1.

Chemical structures, frontier molecular orbitals, and their energy levels of the ground states of (A) XAN-1 (Erythrosine B), (B) XAN-2 (Eosin Y), (C) XAN-3 (Rose Bengal), and (D) XAN-4 (Phloxine B). q: percentage of the MO localized on the specified element. Reprinted from ref 73. Copyright 2019 American Chemical Society.

3.2.1.2. Chromophore Core-Twisting.

By introducing bulky substituents to the chromophore that produce significant steric hindrance, an originally planar chromophore core can be twisted, which leads to large variation in its molecular properties. Researchers have found that core-twisting of some common organic chromophores can lead to a bathochromic shift and broadening of the absorption peaks. These core-twisted dyes include nanocarbon (e.g., acene) derivatives,152 perylene diimide (PDI) derivatives,153 and porphyrin derivatives.154,155

Gidron and co-workers recently reviewed the helically locked twisted acene derivatives, focusing on the relationship between their twist angles and the resultant effect on their optical and electronic properties.152 Overall, with the increase in the twist angle, the twisted acene derivatives exhibited slight destabilization of the HOMO as well as slight stabilization of the LUMO (Figure 2).152,156 This is caused by orbital rehybridization where greater σ character is mixed into the π orbitals as a result of core-twisting.157,158 Consequently, the narrowed HOMO/LUMO energy gap led to a slight bathochromic shift as observed in the absorption spectra.152,156 This general trend is applicable to most twisting nanocarbons.152

Figure 2.

Figure 2.

(A) Planar, bent, and twisted conformations of two adjacent p-orbitals. (B) Calculated HOMO and LUMO energy levels of anthracene (as a simplified example of twisted acene derivatives) with different twist angles. Reprinted from ref 152. Copyright 2019 American Chemical Society.

Similarly, Holten and co-workers investigated conformationally distorted porphyrin derivatives and discovered that the nonplanarity of the macrocycle induced by the steric hindrance of adjacent bulky peripheral substituents largely affects the S1 state.159 Specifically, the saddle deformations of twisted porphyrins tend to destabilize their HOMOs and stabilize their LUMOs, thereby yielding smaller HOMO/LUMO energy gaps and a bathochromic shift in the absorption spectra.159 Senge more generally reviewed a large library of core-twisted porphyrins and concluded that macrocycle distortion of porphyrin derivatives can destabilize the π system.154 For example, zinc(II) 2,3,7,8,12,13,17,18-octaethyl-5,10,15,20-tetraphenylporphyrin (POR-2, Figure 3) is an excellent core-twisting porphyrin dye whose planar porphyrin counterpart is zinc(II) 5,10,15,20-tetraphenylporphyrin (POR-1, Figure 3). While POR-2 only has eight additional peripheral ethyl groups in comparison to 9, these ethyl groups mostly impose geometrical impact on the chromophore without any notable electron-donating/withdrawing effect or any conjugated effect. As reported in the literature, the experimentally resolved structure of POR-2 indeed exhibits a saddle-distorted nonplanar macrocycle because of the steric hindrance formed between ethyl groups and phenyl groups.154,155 This distortion led to a significant ~85 nm bathochromic shift in λmax from 600 nm of ZnTPP86 to 685 nm of ZnOETPP.160 While POR-1 was reported in PET-RAFT polymerization to mediate one of the first orange (~600 nm) light-mediated polymerizations,55,86 the large bathochromic shift of λmax from POR-1 to POR-2 induced by a core-twisting effect has further enabled the implementation of deep-red (~700 nm) light in PET-RAFT polymerization.

Figure 3.

Figure 3.

Chemical structures of POR-1 and POR-2.

3.2.1.3. Introduction of Atoms with Nonbonded Electrons Attached to the Chromophore.

Recently, Zhao and co-workers synthesized a series of PDI derivatives and reported the strong bathochromic shift in absorption caused by alkylthio substitution on the bay positions of PDI derivatives.161 As shown in Figure 4, because of the bay substituents, PDI-2, PDI-3, PDI-4, PDI-5, and PDI-6 all exhibit slight core-twisting characters as compared to the planar PDI-1. However, this nonplanarity is not significant and apparently the core-twisting halogen-substituted PDI-4 and PDI-5 exhibited comparable λmax to PDI-1 (λmax = 523 nm). Therefore, the alkyl/aryl sulfide substituents of PDI-2 (λmax = 560 nm) and PDI-3 (λmax = 552 nm) must have played an important role in the observed ~30 nm bathochromic shift in λmax as compared to PDI-1. The authors observed that the presence of higher energy n electrons in the S atoms increased the HOMO energy level while having a reduced effect on the LUMO energy level, leading to lower HOMO/LUMO energy gaps (Figure 4). Consequently, these n → π* characters in S0 → S1 excitation (~99% contribution from HOMO → LUMO transition) result in the notable bathochromic shift of λmax in PDI-2 and PDI-3 with respect to PDI-1. On the other hand, the bathochromic shift in PDI-6 is smaller, which is due to the strong electron-withdrawing effect of F substituents that greatly withdraw the n electrons on S and reduce their impact on the chromophore.161

Figure 4.

Figure 4.

Chemical structures, molecular geometries, and visualized frontier molecular orbitals (isovalue = 0.03) of PDI derivatives PDI-1, PDI-2, PDI-3, PDI-4, PDI-5, and PDI-6. Reprinted with permission from ref 161. Copyright 2019 The Royal Society of Chemistry.

Würthner and co-workers observed similar bathochromic shifts for PDI derivatives substituted with aryl ethers at the bay positions, which led to a 56 nm bathochromic shift in λmax.139 More strikingly, the authors observed 181 nm bathochromic shift in λmax by introducing pyrrolidinyl substituents on the bay positions,139 which is likely also due to the high energy n electrons in the N atoms contributing to n → π* characters of the S0 → S1 excitation.

Recently, Martynov and co-workers employed (TD)DFT methods and spectroscopic methods to theoretically interpret the spectral properties across a library of unsubstituted and alkoxy-substituted free-base or zinc phthalocyanine derivatives PHC-1, PHC-2, PHC-3, PHC-4, PHC-5, and PHC-6 (Figure 5).140 Two types of alkoxy substitutions are possible for phthalocyanine derivatives, i.e., the peripheral β-alkoxy-substitution (e.g., PHC-2 and PHC-5, Figure 5B,E) and the nonperipheral α-alkoxy-substitution (e.g., PHC-3 and PHC-6, Figure 5C,F). Interestingly, the two types perform very differently in affecting the spectral properties of the phthalocyanine chromophores. In the context of the peripheral β-alkoxy-substituted PHC-2 and PHC-5, λmax of their Q-bands are located at 703 and 679 nm, respectively, almost the same as their unsubstituted counterparts PHC-1 and PHC-4 (Figure 5A,D). In contrast, the nonperipheral α-alkoxy-substituted PHC-3 (λmax = 772 nm) and PHC-6 (λmax = 751 nm) exhibit pronounced bathochromic shift in λmax as large as ~70 nm (Figure 5C,F). The authors concluded that the nonperipheral α-alkoxy-substitution can strongly destabilize HOMO because of the larger HOMO population on α-carbons resulting in stronger mesomeric interactions with n electrons of the O atoms, which consequently leads to a narrowed HOMO/LUMO energy gap and the observed large Q-band bathochromic shift.140 Another report discovered that the Q-band bathochromic shift is enhanced with an increasing number of nonperipheral α-alkoxy-substituents.162 Similarly, Martin and co-workers observed the same nonperipheral α-alkoxy-substitution effect in generating a large Q-band bathochromic for naphthalocyanines.163

Figure 5.

Figure 5.

Chemical structures and molecular geometries of (A) free-base phthalocyanine PHC-1, (B) free-base β-alkyloxy-substituted phthalocyanine PHC-2, (C) free-base α-alkyloxy-substituted phthalocyanine PHC-3, (D) zinc(II) phthalocyanine PHC-4, (E) zinc(II) β-alkyloxy-substituted phthalocyanine PHC-5, and (F) zinc(II) α-alkyloxy-substituted phthalocyanine PHC-6. Reprinted from ref 140. Copyright 2019 American Chemical Society.

Boyer, Liu, and co-workers implemented PHC-4 and PHC-6 in PET-RAFT polymerization via an oxygen-mediated reductive quenching pathway and performed (TD)DFT calculations to gain a deeper insight into the structure-property relationships (Figure 6). UV–vis spectroscopy was performed in dimethyl sulfoxide for PHC-4 (Figure 6C, λmax = 661 nm) and PHC-6 (Figure 6G, λmax = 723 nm). For both dyes, the S0 → S1 vertical excitation has over 95% contribution from the HOMO → LUMO transitions. The installation of highly electron-donating alkoxy groups80 in PHC-6 significantly increases the HOMO and LUMO energy levels compared to PHC-4 but increases the HOMO to a larger extent.73 This effect was coupled with excitation from n orbitals of the oxygen atom to π* orbitals of the tetraaza-prophyrin core, in addition to the excitation from the partial phenyl-π orbital to the π* orbital of the tetraaza-prophyrin (Figure 6F), thus narrowing the energy gap (Figure 6E) and leading to a 62 nm bathochromic shift with respect to PHC-4 (Figure 6A,B). Consequently, while PHC-4 enabled PET-RAFT polymerization under red (660 nm) LED light (Figure 6D), the implementation of PHC-6 in PET-RAFT polymerization not only enabled the use of deep-red (~700 nm) LED light but also enabled the use of near-infrared (NIR, 780 nm) light for efficient PET-RAFT polymerization via the oxygen-mediated reductive quenching pathway (Figure 6H).

Figure 6.

Figure 6.

(A,E) Chemical structures and energy level diagrams, (B,F) visualization of HOMO, LUMO/LUMO+1 (isovalue = 0.3), (C,G) UV–vis spectra with λmax denoted, and (D,H) plots of ln([M]0/[M]t) versus time revealing kpapp and temporal control for model PET-RAFT polymerization via the oxygen-mediated reductive quenching pathway energy, corresponding to (A–D) zinc(II) phthalocyanine PHC-4 and (E–H) the zinc(II) β-alkyloxy-substituted phthalocyanine PHC-6, respectively. Percentage contribution of the dominant molecular pairs contributing to the S0 → S1 electronic transition is denoted in red. Reprinted with permission from ref 56. Copyright 2021 Nature Publishing Group.

3.2.1.4. Extended π-Conjugation.

Extension of π-conjugated systems is one of the most applied guiding principles for designing long-wavelength absorbing dyes.164 This approach is based on the conventional recognition that larger π-conjugation contributes to bathochromic shifts. Although there are exceptions,164,165 this statement is still useful in most cases given sufficient understanding of the core chromophore before modification.164 The most commonly seen π-expanded dyes include polyaromatic dyes164 and porphyrin derivatives.166 It should be noted the statement that larger π-conjugation chromophores can be excited by longer-wavelength irradiation is only valid for the dimension of π-conjugation in the chromophore rather than the fused rings in the substituents, which was also mentioned by Ruiz-Morales.167

Adachi and co-workers reported the effect of π-conjugation extension on polynaphthalenetetracarboxylic dianhydride diimide dyes PDI-1, PDI-7, and PDI-8 (Figure 7).164 From PDI-1 (λmax = 526 nm, εmax = 81500 M−1 cm−1) to PDI-7 (λmax = 650 nm, εmax = 93300 M−1 cm−1) and PDI-8 (λmax = 764 nm, εmax = 162000 M−1 cm−1), with the π-conjugation extension along the long molecular axis, significant bathochromic shifts in λmax and notable hyperchromic shifts in εmax were observed. As shown in Figure 6, the extended π-conjugation leads to destabilization of HOMO and increases the HOMO energy level while scarcely affecting the LUMO energy level. This overall effect contributes to a narrowed HOMO/LUMO energy gap and therefore contributes to notable bathochromic shifts. The authors further explained that bridges connecting each naphthalene fragment are characterized as antibonding connections in HOMOs of these derivatives. Due to the increasing content of these antibonding connections from PDI-1 to PDI-7 and PDI-8, the HOMO is more destabilized and exhibits a higher energy level, while the LUMO always has bonding connections in all bridges and is thus not affected (Figure 7).164 Meanwhile, the larger π-conjugation leads to an increase in the transition dipole moment and consequently yields enhanced εmax.164

Figure 7.

Figure 7.

Chemical structures and HOMOs/LUMOs (and corresponding energy levels) of polynaphthalenetetracarboxylic dianhydride diimide dyes PDI-1, PDI-7, and PDI-8. Reprinted from ref 164. Copyright 2001 American Chemical Society.

On the other hand, Sessler and co-workers reviewed a large library of π-extended boron dipyrromethenes (BODIPYs), π-extended sapphyrins, π-extended porphycenes, π-extended corroles, π-extended rosarins, π-extended cyclo[n]pyrroles, and other types of expanded porphyrins with π-extension.166 All of these π-extended dye classes exhibit larger bathochromic shifts in λmax with larger π-extension.166 Accordingly, Boyer, Liu, and co-workers investigated a series of tetraazaporphyrin/phthalocyanine derivatives and revealed λmax = 582 nm for PHC-7, λmax = 661 nm for PHC-4, and λmax = 749 nm for PHC-8, which covers wavelengths from orange to near-infrared light (Figure 8).56 For all three dyes, HOMO → LUMO transitions exhibit dominant (88–99%) contribution to the S1 vertical excitation. As the tetraaza-porphyrin core is very electron-withdrawing,168 overall, the perphenyl rings in PHC-4 exhibit a more notable electron-donating conjugated effect rather than electron-withdrawing inductive effect. Hence, the addition of four phenyl rings moves both the HOMO and LUMO energy levels of PHC-4 up but moves up the LUMO to a much lesser extent (because of larger conjugation stabilizing the π* orbital), compared with PHC-7, thus narrowing the energy gap and leading to a 79 nm bathochromic shift (Figure 8A,B). Indeed, as seen in Figure 8B, bottom, the HOMO → LUMO transition exhibits partial phenyl-π to tetraazaporphyrin-π* excitation. Similarly, larger aryl rings in PHC-8 yield an even more notable electron-donating conjugation effect, increasing both the HOMO and LUMO but moving up the latter to a lesser extent (larger conjugation stabilizing the π* orbital), thus leading to a much narrower energy gap (Figure 8C) and 88 nm bathochromic shift as compared to PHC-4 (Figure 8B,C). These dyes were implemented in PET-RAFT polymerization via the oxygen-mediated reductive quenching pathway, which achieved efficient polymerization under orange (~600 nm, PHC-7) and red (660 nm, PHC-4) lights.56

Figure 8.

Figure 8.

(A–C) Chemical structures, UV–vis spectra with λmax denoted, energy level diagrams, and visualization of HOMO and LUMO/LUMO+1 (isovalue = 0.3) of (A) zinc tetramethyl tetraazaporphyrin PHC-7, (B) zinc phthalocyanine PHC-4, and (C) zinc naphthalocyanine PHC-8. Percentage contribution of the dominant molecular pairs contributing to the S0 → S1 electronic transition is denoted in red. Reprinted with permission from ref 56. Copyright 2021 Nature Publishing Group.

3.2.2. Tuning Excited State Evolutions: Enhancing the Triplet Quantum Yield.

3.2.2.1. Constructing Charge Transfer States.

As discussed earlier, it is essential to populate T1 states in photocontrolled polymerization systems as the sufficiently long-lived T1 states are more likely to encounter the initiators for electron transfer reactions, while simultaneously suppressing undesirable back electron transfer.13,23,52 However, most purely organic chromophores lack efficient spin–orbit coupling and hence exhibit rather slow ISC and low ΦT.52 In organic PCs, the use of donor–acceptor scaffolds with charge transfer excited states can in some cases result in high ΦT. Typically, these charge-transfer PCs are vertically photoexcited to a Franck–Condon singlet state (SFC, a locally excited higher singlet state Sn above S1) and rapidly relax to the charge transfer S1 state (SCT). From here, the ISC process occurs to reach the corresponding charge transfer triplet state (TCT′) which stays in the long-lived triplet manifold awaiting electron transfer with appropriate substrates. Upon reaching the SCT state, the fluorescence process is greatly suppressed due to the minimal transition dipole moment induced by the spatially segregated molecular orbitals of the “excited electron” and the “hole” (see section 3.1.2.2). Similarly, the IC process is also greatly suppressed due to poor overlap of orbitals (see section 3.1.2.2). Consequently, a reasonable kISC can more likely compete with the suppressed kF and kIC, rendering high ΦT of the PC. On the other hand, by rationally constructing multiple charge transfer states where ISC occurs between a SCT and a TCT′, characteristic of different molecular orbital types, this SCT → TCT′ transition should bear a significant SOCME according to the El-Sayed rule and hence an enhanced kISC.52,169 While the TCT′ state is constructed in a way that displays an energy level slightly lower than that of SCT, the resultant narrow singlet–triplet energy gap ΔEST should also contribute to promoting spin–orbit coupling and kISC (the energy gap law).22,52 By combination of these considerations, a high ΦT can be obtained by simultaneous suppressing kF + kIC and promoting kISC through rationally constructing charge transfer states.

By attaching aryl groups with low-lying π* (either by extended conjugation or electron-withdrawing substitution), Miyake and co-workers constructed a library of dihydrophenazine derivatives,80 phenoxazine derivatives,111 and dimethyl-dihydroacridine derivatives,112 bearing donor–acceptor motifs and characteristically displaying charge transfer excited states as effective O-ATRP PCs with high ΦT and highly reducing T1 states. As an example, Damrauer, Miyake, and co-workers used two comparable π-extended phenoxazine derivatives to demonstrate the effect of charge transfer characters of excited states on photophysical properties of a PC.113 They discovered that the N-phenyl-substituted phenoxazine derivative PXZ-1 (Figure 9A) possesses a ΦT = 0.11. As shown in Figure 9C, after local excitation in the phenoxazine core of PXZ-1 reaching the Franck–Condon state SFC, the electron density is rapidly shifted to one of the biphenyl substituents reaching the charge transfer state SCT-Biph within 10 ps after solvent reorganization and conformational relaxation. However, the subsequent ISC process occurs as a SCT-Biph→ TCT-Biph transition does not have a narrow ΔEst nor does it exhibit orthogonality in its orbital angular momentum modes. Consequently, the kisc of SCT-Biph → TCT-Biph transition is limited and cannot effectively compete with kf + kic of the SCT-Biph → S0 transitions, resulting in a low ΦT of 0.11 (Figure 9A). In stark contrast, by replacing the N-phenyl group of PXZ-1 with the N-1-naphthyl group, the π-extended derivative PXZ-2 (Figure 9B) exhibits a remarkably high ΦT of 0.91. As presented in the energy diagram (Figure 9C), upon local excitation in the phenoxazine core of PXZ-2 reaching SFC and further SCT-Biph, the electron density is further shifted to the N-naphthyl group, reaching SCT-Naph within ~20 ps. Subsequently, two ISC pathways can occur (respectively denoted as A and B in Figure 9C). The ISC pathway A has an SCT-Naph → TCT-Naph transition as the rate-limiting step, which bears a very small ΔEST; however, although possible, the pathway A was not observed likely because of a lack of change in molecular orbital types.113 On the other hand, the ISC pathway B directly couples SCT-Naph with TCT-Biph. The kisc of SCT-Naph → TCT-Biph transition is greatly enhanced because of the notable change in orbital angular momentum which greatly facilitates spin–orbit coupling. Consequently, PXZ-2 exhibits a high ΦT of 0.91 arising from the excited state charge transfer characters.113 The high ΦT = 0.91 of PXZ-2 has endowed the PXZ-2-catalyzed O-ATRP polymerization of MMA with an excellent initiator efficiency of close to unity, as well as good control of the polymerization, exhibiting a resultant dispersity of 1.1–1.2. This is lower than its other derivatives in the phenoxazine dye class.

Figure 9.

Figure 9.

(A,B) Chemical structures and molecular geometries of (A) PXZ-1 and (B) PXZ-2, with ΦT denoted below. (C) Diagram of excited and ground state energy levels with alternative ISC pathways presented for PXZ-1 and PXZ-2. Less likely pathways are indicated in dashed lines.113 Reprinted from ref 113. Copyright 2018 American Chemical Society.

Following this study, Damrauer, Miyake, and co-workers investigated a class of N-aryl 3,7-diphenyl phenoxazine derivatives PXZ-3, PXZ-4, and PXZ-5 (Figure 10AC) with varied N-aryl substitution to establish more detailed relationships between ΦT and the charge transfer character of excited states.169 Unlike the previous report, the 3,7-diphenyl group chosen to replace the 3,7-dibiphenyl group isolated the impact of N-aryl substitution on the charge transfer character and ΦT. As shown in Figure 10C for PXZ-3, upon photoexcitation to the Franck–Condon excited state SFC, it rapidly relaxed to an emitting state Sdeloc characteristic of an excited electron delocalized to the whole chromophore. Subsequently, Sdeloc can either relax to S0 by fluorescence/IC or TCT-Phen(core) which exhibits partial charge transfer to a molecular orbital spanning over the phenoxazine core and its one-side adjoining phenyl group. Because of the difference in molecular orbital types between Sdeloc and TCT-Phen(core), the spin–orbit coupling of the Sdeloc → TCT-Phen(core) transition is enhanced, leading to a reasonable kISC capable of competing over kF and kIC and resulting in a ΦT = 0.30 (Figure 10A). On the other hand, by replacing the N-phenyl group with a N-1-naphthyl group, the resultant PXZ-4 exhibited an additional intermediate singlet state SCT-Naph with charge transfer character to the N-1-naphthyl group, which is relaxed from Sdeloc (Figure 10C). Due to the charge transfer character of the emitting state SCT-Naph, the irradiative SCT-Naph → S0 transition is greatly impeded, leading to a low kF = 3.0 × 106 s−1 of PXZ-4 as compared to PXZ-3 (kF = 2.1 × 108 s−1). Meanwhile, the more pronounced change in molecular orbital types of the SCT-Naph → TCT-Phen(core) transition led to enhanced spin–orbit coupling according to the El-Sayed rule. Correspondingly, PXZ-4 bears a higher kISC = 1.5 × 108 s−1 compared with PXZ-3 (kisc = 9.3 × 107 s−1). The overall effect of retarded fluorescence and enhanced ISC, contributed to the high ΦT = 0.95 of PXZ-4 as compared to PXZ-3T = 0.30). On the other hand, the N-2-naphthyl counterpart PXZ-5 exhibits a naphthyl-to-phenoxazine center–center distance of 0.533 nm larger than that of PXZ-4 (0.437 nm). Consequently, PXZ-5 exhibits less Coulombic stabilization than PXZ-4 and thus PXZ-5 possesses a slightly higher energy SCT-Naph compared to PXZ-4, resulting in a larger ΔEST of PXZ-5 than that of PXZ-4. Therefore, PXZ-5 exhibits a slightly lower ΦT = 0.92 as compared to PXZ-4T = 0.95), arising from the variation in the N-naphthyl connectivity.169

Figure 10.

Figure 10.

(A–C) Chemical structures of (A) PXZ-3, (B) PXZ-4, and (C) PXZ-5 with kF, kIC, kISC, and ΦT denoted below. (D) Energy level diagram of PXZ-3, PXZ-4, and PXZ-5 showing dominant decay pathways of excited states. Reprinted from ref 169. Copyright 2019 American Chemical Society.

3.2.2.2. Heavy Atom Effect.

On the basis of the heavy atom effect,170 ΦT of fluorescein dyes can be simply tuned by the halogenation strategy, where the spin–orbit coupling between S1 and Tn is enhanced by halogen substituents with larger atomic numbers Z (see section 3.1.2.1), leading to higher kISC and ΦT. Recently, Boyer and Liu reported a synthesized PC XAN-6 (Figure 11B), which possesses ΦT = 0.58 and features four Br substituents; the ΦT was significantly increased compared to its nonsubstituted counterpart fluorescein XAN-5T = 0.03, Figure 11A).75 As denoted in Figure 11A,B, XAN-5 and XAN-6 bear highly comparable kF and both exhibit negligible kIC; however, heavier substituents of XAN-6 result in remarkably higher kISC by enhanced spin–orbit coupling compared to XAN-5.75 Overall, this heavy atom effect leads to much higher ΦT = 0.58 of XAN-6. Consequently, although XAN-5 has better redox properties favoring PET-RAFT polymerization, it was inefficient (kpapp ≈ 0 min−1) due to its low ΦT.67 By contrast, because of the considerable ΦT = 0.58, the XAN-6-catalyzed model PET-RAFT polymerization exhibited a kpapp of 0.016 min−1, which is considered rather efficient.75

Figure 11.

Figure 11.

(A,B) Chemical structures of (A) XAN-5 and (B) XAN-6, with kF, kIC, kISC, and ΦT denoted below. (C) Computationally derived Jablonski diagram of XAN-6 and its derivation of ΦT from different excited state decay rate constants. Reprinted from ref 75. Copyright 2019 American Chemical Society.

Following this trend, heavier halogen substitution appears to result in higher ΦT for the halogenated xanthene dyes XAN-1, XAN-2, XAN-3, and XAN-4 mentioned above (Figure 1).73 XAN-1 substituted with heavier I atoms bears ΦT = 0.62–0.69, which almost doubles that of the Br-substituted XAN-2T = 0.28–0.32). Although XAN-1 and XAN-2 possess similar absorption and redox properties, the high ΦT = 0.62–0.69 has enabled much faster XAN-1-catalyzed model PET-RAFT polymerization with a kpapp = 0.023 min−1, as compared to that of the XAN-2-catalyzed system (kpapp = 0.014 min−1) with low T1 population (ΦT = 0.28–0.32), under the same polymerization conditions.73

3.2.2.3. Chromophore Core-Twisting.

By state-of-the-art theoretical and experimental studies, researchers clarified the important role of core-twisting/nonplanarity in promoting ISC via enhanced spin–orbit coupling.171,172 Hariharan and co-workers initially investigated a class of core-twisted PDI derivatives with multiple Br substitutions in the bay positions and discovered a significant enhancement in triplet generation.173 However, to explore the isolated effect of core-twisting on ISC, the impact from heavy atoms must be excluded. With this consideration, Hariharan and co-workers further imparted light-atom-based twisting into the core chromophore of PDI derivatives forming PDI-10 and PDI-11 with the planar PDI-9 as a comparison (Figure 12). Indeed, the planar analogue PDI-9 possesses negligible ΦT ~ 0 because of the weak spin–orbit coupling which diminishes ISC. By imparting core-twisting induced by steric hindrance at one side of the bay regions of the PDI core (PDI-10), it yields an enhanced ΦT ~ 0.10 coupled with a kisc = 1 × 109 s−1 (Figure 12B). By imparting this core-twisting at both sides of the bay regions, the resultant PDI-11 possesses a more twisted structure and exhibits an even higher ΦT ~ 0.30 with a higher kisc = 4 × 1010 s−1 (Figure 12C). Quantum chemical calculations revealed that the C–H and C═C bending vibration modes in the twisted regions contribute to efficient ISC between 1(ππ*) and 3(ππ*) driven by the Herzberg–Teller vibronic coupling.153,174

Figure 12.

Figure 12.

Chemical structures and representation of molecular geometries of PDI-9, PDI-10, and PDI-11, with ΦT and kISC denoted below the dye labels. Reproduced with permission from ref 153. Copyright 2017 The Royal Society of Chemistry.

Very recently, Behrends and co-workers constructed a class of twisted acene derivatives Ant-Cn (n = 3–6) tethered with alkyl tethers (n = propyl-tether to hexyl-tether) to investigate the effect of core-twisting in ΦT of chromophores (the strategy illustrated as Figure 13, top).175 With shortened alkyl bridges, the twist angle is increased from 23° with a propyl bridge (Ant-C3), 30° with a butyl-tether (Ant-C4), 32° with a pentyl-tether (Ant-C5), to 38° with a hexyl-tether (Ant-C6). As a comparison, a planar open reference compound without a tether (open) featuring 0° twist angle is also studied. Time-resolved electron paramagnetic resonance (EPR) spectroscopy was performed to obtain the transient EPR spectra of all these compounds (Figure 13, bottom). The highly similar spectral shapes for open, Ant-C6, Ant-C5, and Ant-C4 allow direct comparison and clearly reveal increased ΦT with shortened tethers. Despite a different spectral shape of Ant-C3 (because of significant stretching-induced variation in spin polarization175), Ant-C3 bears the highest ΦT as expected (Figure 13, bottom). Overall, a clear trend was observed by the authors toward increased ΦT for more twisted compounds tethered with shorter alkyl tethers.175

Figure 13.

Figure 13.

(Top) Chemical structures of twisted acene derivatives Ant-Cn (n = 3–6) tethered with alkyl bridges (n = propyl bridge to hexyl bridge). (Bottom) Transient EPR spectra for Ant-Cn and the open (no alkyl bridges) reference compound. Reproduced with permission from ref 175. Copyright 2019 The Royal Society of Chemistry.

Gidron and co-workers broadly reviewed the consequences of core-twisting nanocarbon molecules on their properties (including ΦT) and summarized the proposed mechanisms.152 For example, Brédas and co-workers demonstrated a direct correlation between the magnitude of spin–orbit coupling and the degree of nonplanarity in nonplanar aromatic heterocyclic compounds.176 Combined with the above-mentioned findings, chromophore core-twisting can generally enhance spin–orbit coupling and accelerate ISC, thereby enhancing ΦT of the chromophore.

Similarly, Holten and co-workers investigated a variety of core-twisted nonplanar metal-free porphyrin derivatives substituted with bulky alkyl groups that induced steric hindrance.159 They observed an increase in kISC primarily from enhanced spin–orbit coupling in these nonplanar porphyrin complexes. To interpret this observation, they mentioned that ISC in porphyrins is commonly deemed to rely on the wave function overlap involving the central N atoms in the macrocycle and arising from out-of-plane distortion in S1 and/or Tn.159 Indeed, previous theoretical studies177 and experimental studies178 on chlorophylls argue that ISC can be modulated by the N-centered perturbations. Furthermore, Shchupak, Ivashin, and Sagun reported that common planar porphyrin derivatives exhibit small spin–orbit coupling matrix elements; however, the nonplanar porphyrin derivatives with more distortion of the macrocycle in T1 exhibits a considerable increase in SOCME, which is induced by a mixture of s character to π-orbitals in T1 compared to S1 and hence enhances spin–orbit coupling.179

3.2.3. Tuning Redox Properties.

3.2.3.1. Photo-ATRP via the Oxidative Quenching Pathway.

As shown in Figure 14A, the first photo-ATRP PC fac-Ir(ppy)3 reported by Hawker and Fors 60,84 exhibits a T1 state that is strongly reducing, E0(PC•+/3PC*) = E0[Ir(IV)/Ir(III)*] = −1.73 V, versus SCE (redox potential are all presented vs SCE herein). In light of the reduction potential of a typical ATRP initiator EBP (E0(EBP/EBP•−) = −0.74 V), the reduction of the alkyl bromide initiator by fac-Ir(ppy)3 is highly exothermic, supportive of efficient initiation.72,80 On the other hand, the PC•+ species of fac-Ir(ppy)3 is a strong oxidant exhibiting E0(PC•+/PC) = E0[Ir(IV)/Ir(III)] = 0.77 V, which is much higher than E0(EBP/EBP•−) = −0.74 V and is thus capable of oxidizing the propagating radical and enabling reversible deactivation (Scheme 2A).72 A metal-free version of photo-ATRP was also presented by Hawker et al.64 and Matyjaszewski et al.72 with PTZ-1 (Figure 14B) as the PC, which possesses highly reducing 1PC* with E0(PC•+/1PC*) = −2.10 V (thus indicating highly reducing 3PC*) and highly oxidizing PC•+ with E0(PC•+/PC) = 0.68 V to complete the oxidative quenching cycle of photo-ATRP.72

Figure 14.

Figure 14.

Chemical structures of (A) fac-Ir(ppy)3 and (B) PTZ-1 with E0(PC•+/3PC*) and E0(PC•+/PC) denoted below the dye label. Reproduced from ref 72. Copyright 2016 American Chemical Society.

The advantage of the organic scaffolds over the transition-metal-based compounds in photocatalysis have motivated the study of organic chromophores. As the chromophore core predetermines the benchmark redox properties, substituents attached to the chromophore provide the ability to rationally tune the redox properties based on the electronic nature of the substituting fragment. Thus, Miyake and co-workers compared the redox potentials of N,N-diphenyl dihydrophenazine (DPZ-1, Figure 15A), N-phenyl phenoxazine (PXZ-6, Figure 15B), and N-phenyl phenothiazine (PTZ-1, Figure 15C).111 The three chromophore cores all exhibit highly reducing T1 states with DFT-computed E0(PC•+/3PC*) < −2.0 V, as computed under the same level of theory.111 Specifically, among the three, DPZ-1 has the most negative E0(PC•+/3PC*) of −2.25 V, followed by PXZ-6 and PTZ-1. On the other hand, E0(PC•+/PC) of all the three chromophore cores are all much higher than E0(initiator/initiator•−) of the EPA initiator (−0.74 V, Scheme 2B), and are sufficient for deactivation of O-ATRP.111

Figure 15.

Figure 15.

Chemical structures of the original chromophore core structures (A) DPZ-1 as a dihydrophenazine derivative, (B) PXZ-6 as a phenoxazine derivative, and (C) PTZ-1 as a phenothiazine derivative, with DFT-computed E0(PC•+/3PC*) and E0(PC•+/PC) denoted below the dye label. Reproduced from ref 111. Copyright 2016 American Chemical Society.

Miyake and co-workers reported dihydrophenazine derivatives DPZ-1, DPZ-2, DPZ-3, and DPZ-4 (Figure 15) as PCs for O-ATRP.80 DFT predictions and experimental measurements displayed similar trends, where DPZ-1, DPZ-2, DPZ-3, and DPZ-4 all exhibit strong E0(PC•+/3PC*) < −2.0 V (computed). The ATRP initiator ethyl α-bromophenylacetate (EBP, Scheme 2B) exhibits a computed E0(initiator/initiator•−) = −0.74 V, thus all these PCs with computed E0(PC•+/3PC*) < −2.0 V are capable of efficiently activating EBP via PET.80 In particular, they demonstrated that by installing electron-donating methoxy groups to DPZ-1, the resultant DPZ-2 possesses stronger E0(PC•+/3PC*) = −2.36 V compared to DPZ-1 (−2.34 V). On the other hand, installation of more electron-withdrawing substituents leads to a lower E0(PC•+/3PC*), i.e., −2.24 V for DPZ-3 and −2.06 V for DPZ-4 compared with −2.34 V for the reference DPZ-1 (Figure 16). Moreover, the relatively stable PC•+ of these PCs formed from PET with computed E0(PC•+/PC) = ~0.1 to 0.2 V are sufficiently oxidizing to deactivate the propagating radical. Indeed, the anion radical of the EPA initiator has E0(initiator/initiator•−) of −0.74, −0.86, and −0.71 V after zero, one, and two monomer (methyl methacrylate) additions, which are much more negative than E0(PC•+/PC) of PCs DPZ-1, DPZ-2, DPZ-3, and DPZ-4, allowing the O-ATRP oxidative quenching cycle to be completed.80

Figure 16.

Figure 16.

Chemical structures of dihydrophenazine derivatives (A) DPZ-1, (B) DPZ-2, (C) DPZ-3, and (D) DPZ-4, with DFT-computed E0(PC•+/3PC*) denoted below the label. Reproduced with permission from ref 80. Copyright 2016 American Association for the Advancement of Science.

The authors further designed and synthesized DPZ-5 and DPZ-6 (Figure 17A,B). While exhibiting notable charge transfer characters in T1 (Figure 17C,D), DPZ-5 and DPZ-6 still possess strong computed E0(PC•+/3PC*) of −2.20 V and −2.12 V, respectively, which were comparable to the original DPZ-1 (Figure 16A). Indeed, in addition to the 3PC* being sufficiently reducing to activate O-ATRP, those with charge transfer T1 states (i.e., DPZ-3, DPZ-5, and DPZ-6) yielded much more efficient O-ATRP polymerization with better control and high initiator efficiency.80

Figure 17.

Figure 17.

(A,B) Chemical structures of dihydrophenazine derivatives (A) DPZ-5 and (B) DPZ-6, with computed E0(PC•+/3PC*) denoted below the label. (C,D) T1 frontier molecular orbitals of (C) DPZ-5 and (D) DPZ-6 which visualizes the higher-lying SOMO (top) and the lower lying SOMO (bottom). Reproduced with permission from ref 80. Copyright 2016 American Association for the Advancement of Science.

Miyake and co-workers further discovered phenoxazine derivatives as O-ATRP PCs and sought to tune E0(PC•+/3PC*) by modification of the chromophore core with substituents. As shown in Figure 18, installation of the electron-withdrawing trifluoromethyl substituent on the para position of the N-phenyl group (PXZ-7) leads to slightly less negative E0(PC•+/3PC*) = 2.03 V, as compared to the original benchmark PXZ-6 exhibiting E0(PC•+/3PC*) = 2.11 V. By replacing the N-phenyl group with a N-1-naphthyl group (PXZ-8), the resulting expanded conjugative effect leads to weaker E0(PC•+/3PC*) = −1.84 V, apparently due to the electron-withdrawing conjugated effect of the N-1-naphthyl group with excited electron residing on its low-lying π* orbital (Figure 18G). Consequently, the higher SOMO, where the excited electron resides, is localized in the low-lying π* orbital of the N-1-naphthyl group, leading to a higher IP of the T1 state and thus weakened E0(PC•+/3PC*). The charge transfer character was incorporated into the T1 state of 29 by this modification (Figure 18B), which is believed to improve the O-ATRP performance by enhancing ΦT. 30 substituted with a N-2-naphthyl group is highly similar to 29, exhibiting computed E0(PC•+/3PC*) = −1.90 V, versus SCE and a T1 with the charge transfer character.111

Figure 18.

Figure 18.

(A–D) Chemical structures of phenoxazine derivatives (A) PXZ-6, (B) PXZ-7, (C) PXZ-8, and (D) PXZ-9 with computed E0(PC•+/3PC*) denoted below the label. (E–H) T1 frontier molecular orbitals of (E) PXZ-6, (F) PXZ-7, (G) PXZ-8, and (H) PXZ-9 which visualizes the higher-lying SOMO (top) and the lower lying SOMO (bottom). Reproduced from ref 111. Copyright 2016 American Chemical Society.

Miyake and co-workers also investigated the possible range of E0(PC•+/3PC*) that the phenoxazine PC scaffold can access, with substituted derivatives possessing electron-withdrawing, electron-donating, or conjugation-extended substituents, which yielded a series of derivatives with E0(PC•+/3PC*) ranging between −1.40 V and −2.20 V.74 As shown in Figure 19A, more electron-donating substituents (e.g., the methoxy group) enhances the 3PC* reducing ability, rendering E0(PC•+/3PC*) more negative, whereas more electron-withdrawing substituents (e.g., the trifluoromethyl group and the nitrile group) decreases the 3PC* reducing ability while rendering E0(PC•+/3PC*) less negative. On the other hand, larger dimensions of the conjugation (or cross-conjugation) systems also diminishes the reducing capability of 3PC*. As shown in Figure 19B, PXZ-6 with the smallest conjugation/cross-conjugation bears the most negative E0(PC•+/3PC*) and PXZ-16 with the largest conjugation/cross-conjugation bears the least negative E0(PC•+/3PC*) among the four derivatives.

Figure 19.

Figure 19.

Chemical structures of (A) PXZ-10, PXZ-11, PXZ-12, PXZ-13, PXZ-5, PXZ-14, and PXZ-15 for demonstration of the electron-donating/withdrawing effect on computed E0(PC•+/3PC*) and (B) PXZ-6, PXZ-9, PXZ-5 and PXZ-16 for demonstration of the effect of extended conjugation on computed E0(PC•+/3PC*), with E0(PC•+/3PC*), E0(PC•+/PC), and ET denoted below the label. Reproduced from ref 74. Copyright 2018 American Chemical Society.

3.2.3.2. PET-RAFT Polymerization via the Oxidative Quenching Pathway.

One particular feature of PET-RAFT polymerization that is advantageous relative to photo-ATRP is that the initiator efficiency is easier to control because in the RAFT equilibrium the propagating species can react with a dormant RAFT agent or a dormant polymer chain with a RAFT end-group, thus inducing efficient chain transfer (Scheme 5A). Furthermore, PCs with less reducing T1 states are also suitable for PET-RAFT polymerization because RAFT agents are thermodynamically more facile to reduce than ATRP initiators, under the proviso that they are still thermodynamically capable of activating some portion of the RAFT agents; the other RAFT agents can participate in chain growth by the chain transfer mechanism. Accordingly, for a PC capable of catalyzing PET-RAFT polymerization, its E0(PC•+/3PC*) has to be stronger (i.e., more negative) than E0(RAFT/RAFT•−) of the RAFT agents but need not be overwhelmingly exothermic.

Following the initial work where fac-Ir(ppy)3 was implemented in PET-RAFT polymerization,61 Boyer and co-workers further reported the use of POR-1 and POR-3 (Figure 20) to catalyze PET-RAFT polymerization.86 POR-1 bears a computed E0(PC•+/3PC*) of −1.24 V and is thus capable of activating BTPA (Scheme 6C) to polymerize acrylates (e.g., MA, Figure 20C) as well as activating CPADB (Scheme 6C) to polymerize methacrylates.86 By contrast, the replacement of the highly electron-donating Zn2+ cation with two H+ cations, renders POR-3 with a less reducing T1, characterized by a computed E0(PC•+/3PC*) of −1.00 V, which makes it unable to catalyze polymerization of acrylates (e.g., MA, Figure 20D) and only effective in activating CPADB to polymerize methacrylates. Although the weaker reducing capability of the T1 state POR-3 is inefficient for the oxidative quenching pathway, the reductive quenching induced by addition of triethylamines can be enabled by POR-3 to polymerize acrylates. A recent study by Boyer, Liu, and co-workers more explicitly revisited this selectivity using thermodynamic studies (Figure 20E,F) where POR-1 was found to be thermodynamically favorable for completing the catalytic cycle.56 Comparatively, POR-3 failed in the activation of the RAFT agent (i.e., a derivative of BTPA) and thus cannot catalyze polymerization of acrylates via oxidative quenching.

Figure 20.

Figure 20.

(A,B) Chemical structures of (A) POR-1 and (B) POR-2. (C,D) Plot of ln([M]0/[M]t) versus time revealing kpapp and temporal control for model PET-RAFT polymerization via oxidative quenching catalyzed by (C) POR-1 and (D) POR-2. (E,F) Proposed photocatalytic cycles with the Gibbs free energy change (ΔG) of key steps denoted in comparison to corresponding thresholds for PET-RAFT polymerization via the oxidative quenching pathway, catalyzed by (E) POR-1 and (F) POR-2, respectively. Green stands for favorable and red for inert. Colored values are ΔG of the corresponding process in kcal/mol, in conjunction with the derived thresholds. Reproduced with permission from ref 56. Copyright 2021 Nature Publishing Group.

Boyer, Miyake, Liu, and co-workers investigated the effect of varied halogen substituents (H, Cl, Br, and I) on E0(PC•+/3PC*) of halogenated xanthene dyes XAN-1, XAN-6, XAN-2, XAN-7, XAN-3, and XAN-4 (Figure 20AF) as PET-RAFT PCs.73,75 As shown in Figure 21G, the series H, I, Br, and Cl are substituents with increasing electronegativity and electron-withdrawing character in halogenated xanthene dyes. With more halogen substituents of higher electronegativity, the overall more notable electron-withdrawing effect leads to less negative E0(PC•+/3PC*) values (Figure 21H). Consequently, XAN-1 with four I substituents exhibits the strongest E0(PC•+/3PC*) of −1.35 V, while XAN-4 possessing eight strongly electron-withdrawing halogens exhibits the weakest E0(PC•+/3PC*) of −0.91 V. The polymerization rate kpapp was reported to be highly dependent on E0(PC•+/3PC*) of these halogenated xanthene dyes, where more negative E0(PC•+/3PC*) in combination with higher ΦT leads to higher kpapp.73,75

Figure 21.

Figure 21.

(A–F) Chemical structures of (A) XAN-1, (B) XAN-6, (C) XAN-2, (D) XAN-7, (E) XAN-3, and (F) XAN-4. (G) Electronegativities of the halogen substituents versus H to represent their electron-withdrawing capability. (H) Computed E0(PC•+/3PC*) of the PCs in comparison. Reproduced from ref 73 and ref 75. Copyright 2019 American Chemical Society.

3.2.3.3. Photocationic RAFT Polymerization via the Reductive Quenching Pathway.

With respect to photo-LCP systems as well as photo-ROMP, current studies mostly assume 1PC* as the catalytic species without consideration of 3PC*, despite the lack of evidence. However, because the energy levels and electronic structures of 1PC* are very similar to the corresponding 3PC* state, except that 3PC* is lower in energy, herein we follow existing studies and use properties of 1PC* for the discussion in the following context.

A library of triphenylpyrylium and triphenylthiopyrylium derivatives (Figure 22AE) was compared by Fors and co-workers as PCs for photocationic RAFT polymerization.124 The pyrylium core bears a positive charge and is thus highly electron-withdrawing, which by contrast renders the methyl group in PY-2 and the methoxy group in PY-3 highly electron-donating. As a consequence, these electron-donating substituents increase the energies of both the lower SOMO of 1PC* and LUMO of PC, leading to less positive E0(1PC*/PC•−) and more negative E0(PC/PC•−) of PY-2 and PY-3 as compared to the unsubstituted PY-1. Since the methoxy group is a stronger electron-donating group, PY-3 displayed much less positive E0(1PC*/PC•−) and hence contributed to much slower polymerization of isobutyl vinyl ethers (6 monomer additions per photon, Figure 22C). On the other hand, PY-2 and PY-1 contributed to similar polymerization rates which were measured to be around 35 monomer additions per photon,94 in spite of less positive E0(1PC*/PC•−) of PY-2 (Figure 22B) compared to PY-1 (Figure 22A).124 Unexpectedly, using PY-4 (Figure 22D) as PC led to slow polymerization which the authors attributed to poor solubility, while TPY-1 (Figure 22F) also led to slow polymerization but was found to have a high ΦT = 0.94 (underlying mechanism yet unclear).124 Meanwhile, the authors discovered that the more negative E0(PC/PC•−) of PY-3 resulted in rapid chain-end recapping and hence much better short-time temporal control where polymerization is effectively halted in the dark compared with PY-1.124 However, further studies showed that, after dark periods lasting over several hours, notable monomer conversions still occurred with PY-3.94 To tackle this problem, Fors and co-workers implemented a series of Ir-based complexes (Figure 22FH) bearing more negative E0(PC/PC•−).94 Indeed, IR-1 (Figure 21F) and IR-2 (Figure 22G) bearing E0(PC/PC•−) = −0.69 V and −0.79 V, respectively, displayed excellent temporal control where no monomer conversion was observed in >20 h dark periods after ceasing light irradiation. Interestingly, the authors observed ultraslow polymerization with IR-3 (Figure 22H) which was proposed to be due to a much higher rate of polymer recapping relative to propagation as a result of highly negative E0(PC/PC•−) = −1.16 V; this led to sluggish polymerization with <1 monomer addition per photon. In general, IR-1 and IR-2 also contributed to relatively slow photocationic RAFT polymerization because of the less oxidizing 1PC*, indicated by less positive E0(1PC*/PC•−) combined with more negative E0(PC/PC•−), which results in faster polymer recapping compared with triphenylpyrylium and triphenylthiopyrylium derivatives.94

Figure 22.

Figure 22.

Chemical structures, redox potentials, and polymerization efficiency for PCs used in photocationic polymerization. Triphenylpyrylium derivatives124 (A) PY-1, (B) PY-2, (C) PY-3, (D) PY-4, triphenylthiopyrylium derivative124 (E) TPY-1, Ir-based complexes94 (F) IR-1, (G) IR-2, (H) IR-3, and bisphosphonium salt derivatives125 (I) BPP-1, (J) BPP-2, and (K) BPP-3.

More recently, Liao and co-workers demonstrated the use of bisphosphonium salt derivatives (Figure 22IK) to achieve efficient photocationic RAFT polymerization at low catalyst loading of 10 ppm relative to the monomer isobutyl vinyl ether.125 Significantly, this allowed achieved strict temporal control to be achieved with organic PCs. While BBP-1 (Figure 22I) suffered from poor temporal control at high monomer conversions, by installation of relatively electron-donating methyl groups, the resultant BBP-2 (Figure 22J) and BBP-3 (Figure 22K) exhibited more negative E0(PC/PC•−) and strict temporal control with no monomer conversion observed over long dark periods lasting ~3 h.125

3.2.3.4. Photocationic NMP via the Reductive Quenching Pathway.

Kamigaito et al. compared Ir-based complexes (Figure 23AC) in photocationic NMP polymerization and discovered that IR-1 (Figure 23A) bearing the most positive E0(1PC*/PC•−) = 1.70 V exhibited the highest efficiency for photocationic NMP of isobutyl vinyl ether, whereas IR-4 (Figure 23B) with less positive E0(1PC*/PC•−) = 1.23 V yielded much slower polymerization (Figure 23D).95 On the other hand, IR-5 (Figure 23B, a commonly used oxidative quenching PC) has a much less positive E0(1PC*/PC•−) = 0.33 V and exhibited almost no activity for photocationic NMP via the reductive quenching pathway (Figure 23D). By examining the chemical structures for the three Ir-based PCs, it is evident that the much more positive E0(1PC*/PC•−) of IR-1 and IR-4 comes from the positively charged and highly electron-withdrawing iridium core structure, compared to the neutral IR-5. On the other hand, the electron-donating (relative to the positively charged iridium core) tertbutyl groups renders IR-4 less positive E0(1PC*/PC•−) than IR-1.

Figure 23.

Figure 23.

Chemical structures of Ir-based complexes (A) IR-1, (B) IR-4, and (C) IR-5 with E0(PC*/PC•−) vs SCE denoted below the label. (D) Polymerization of isobutyl vinyl ether via photocationic NMP catalyzed by IR-1 (red), IR-4 (blue), and IR-5 (orange). Reproduced with permission from ref 95. Copyright 2019 Wiley-VCH.

3.2.3.5. Photo-ROMP via the Reductive Quenching Pathway.

Boydston and co-workers made comparisons between the pyrylium and thiopyrylium PCs in photo-ROMP. Apart from PY-1 (Figure 24A), PY-2 (Figure 24B), PY-3 (Figure 24D), and TPY-1 (Figure 24E) discussed in section 3.2.3.3, the additional phenyl groups in PY-5 (Figure 24C) are also electron-donating (due to conjugated effect) relative to the positively charged pyrylium core, leading to less positive E0(1PC*/PC•−).78 It was found that norbornene monomers are easily oxidized and thus the use of PCs bearing more positive E0(1PC*/PC•−) tend to result in a larger portion of oxidized norbornene and correspondingly reduced polymerization after 150 min irradiation, i.e., after reaching the maximum conversions.78 Indeed, as shown in Figure 24A,B, PY-1 with the most positive E0(1PC*/PC•−) yielded the least polymerization, with only 9% monomer conversion and also the most oxidation (91%); by contrast, PY-3 with much less positive E0(1PC*/PC•−) led to 75% monomer conversion and only 25% oxidization of norbornene monomers. Interestingly, the thiopyrylium derivatives generally possessed improved ratio of polymerization given each type of substitution compared with their pyrylium counterparts (Figure 24). This could be due to higher ΦT of thiopyrylium compounds relative to pyrylium compounds (because of the heavy atom effect see section 3.1.2.1), which led to a population of less energized 3PC*. As aforementioned, the populated 3PC* is characteristic of less positive E0(1PC*/PC•−) compared to 1PC* with similar electronic structure. For example, PY-1 exhibits ΦT = 0.42 and TPY-1 exhibits ΦT = 0.94.124

Figure 24.

Figure 24.

Chemical structures of PCs used in photo-ROMP, their redox potentials, and portions of oxidized norbornene vs polymerization. Triphenylpyrylium derivatives (A) PY-1, (B) PY-2, (C) PY-5, and (D) PY-3 and triphenylthiopyrylium derivatives (D) TPY-1, (E) TPY-2, (F) TPY-3, and (G) TPY-4. A portion of the norbornene monomers being polymerized and being oxidized after full conversion was achieved within 150 min in photo-ROMP catalyzed by each PC, respectively.

4. CONCLUDING REMARKS AND OUTLOOK

Considering the ever-growing versatility of functional polymers in various applications, discovery of PCs via a trial-and-error strategy represents a rate-determining step in finding polymerization systems that meet the demand for sophisticated macromolecular syntheses. To tackle this limitation, rational computer-guided strategies for PC design have emerged, providing rapid and economic methodologies for discovering new PCs. This rational strategy is based on the acknowledgment of structure-property-performance relationships. As shown in Scheme 21, the performance of the photocontrolled polymerization can be correlated with the properties of the PC, which can be tuned by rational design of their chemical structure.

Scheme 21.

Scheme 21.

Rational Design of a PC to Control the Performance of Photocontrolled Polymerization

The underlying logic of the structure-property-performance relationships has been introduced through theoretical derivations and augmented by physical and chemical observations. Accordingly, there are many examples of applications of these guiding principles promoting polymerization efficiency by enhancing absorption, triplet yield, redox capabilities, as well as choice of activation wavelengths through structural variation of the PC chromophores. To date, a number of useful PCs have been discovered for different photocontrolled polymerization, which have resulted from implementation of the structure-property-performance relationships outlined in this review. Scheme 22 lists some typical examples52,54,73,86,125,180,181 that outcompete their counterparts with respect to certain aspects of performance in photocontrolled polymerization of acrylates, methacrylates, acrylamides, methacrylamides, and vinyl ethers. Nevertheless, a rational approach in PC design for photocontrolled polymerization has tremendous potential in unrealized applications.

Scheme 22.

Scheme 22.

A Few Examples of the PC-Initiator Pairs Designed with Better Performance in Photocontrolled Polymerization Systems to Date

To provoke future exploration on this topic, we propose the following future directions. (i) Developing NIR-light-regulated polymerization systems. Although great effort has been made to red-shift the maximum absorption wavelengths of PCs, the majority of current PCs do not absorb light beyond 800 nm. There are, however, great benefits to developing PCs with longer wavelength absorption. Indeed, far-red light with wavelengths above 800 nm and infrared light show increased depth penetration through human tissues and other materials.182 The ability to harness these longer wavelengths will open the possibility to perform photocontrolled polymerization and other organic transformations within tissues or vessels and in a spatiotemporally controlled manner. This goal can be achieved by collaboratively customizing PCs,183 modifying initiators and designing new quenching pathways mediated by additives, in a rational manner. However, using lower energy light also introduces a thermodynamic boundary on the redox properties of photoexcited PCs. To overcome this boundary defined by the energetics of a single photon, synthetic strategies are emerging that combine the energetics of two photons into a single redox event,184 coupling photochemistry with electrochemistry,185 and direct excitation to T1.186 (ii) Tackling aqueous solubility issues of large π-conjugation PCs.187 As mentioned earlier, the enhancement of photophysical properties in many chromophores is achieved through extending the π conjugation. This extended conjugation is usually accompanied by a large increase in hydrophobicity for large fused ring structures commonly used as PCs. As such, to enable broadened applications in 3D printing and biomedical scenarios, additional design parameters should be included in future PC design. The balance between photophysical properties and solubility of the new PCs should be considered depending on the intended application. (iii) Using a computer-guided approach to interactively tune properties of both the PC and the initiator. In order to design new highly selective reactions between PC-initiator pairs, the properties of both the PC and the initiator should be carefully designed. Advancements in computational capabilities such as machine learning will accelerate the discovery and development of new systems and enable the development of new cutting-edge orthogonal polymerization systems. Indeed, the selective reactivity of excited state PCs to the initiators can enable a much broader range of chemistries extending beyond polymerization. Complex multifunctional materials should be realized as a result. For example, by judiciously designing PCs to selectively match specific initiators, single-unit monomer-insertion or atom transfer radical addition can be achieved for the precise placement of functional groups within polymer chains.59,188 (iv) Designing PCs with more isolated absorption wavelengths for orthogonal polymerization. Combining point (i) and point (iii), increasing the wavelength selectivity of PCs will enable multiple polymerization systems to be independently activated. PCs should be designed that show limited absorption over predesigned wavelengths ranges, while still showing appreciable absorption in other wavelength ranges.28 Combinations of PCs with precisely tunable and nonoverlapping absorptions will enable new orthogonal photosystems in chemical and material syntheses.28,29 As shown in Scheme 23A, orthogonal photosystems have the potential to synthesize polymers with complex block sequences or architectures by switching between different wavelengths. This can greatly alleviate the burden in polymer purification and could facilitate a facile approach in rapid synthesis of complex polymeric materials. On the other hand, orthogonal systems can enable surface patterning4547,108,189,190 with different materials orthogonally controlled by different wavelengths, which may much increase the information stored per area.28,29,33,34,43,191 This direction is challenging and can only be achieved by delicate rational design of two or more PCs along with the initiator by probing into both the optical and chemical aspects, and with the help of computational studies. In addition, the design of these PCs should also consider the other design principles mentioned previously to ensure not only strong light absorption but also efficient photochemical processes thereafter. (v) Designing PCs for additive manufacturing. Photoinduced additive manufacturing uses a photosensitive resin where multifunctional monomers undergo rapid polymerization to form 3D polymer networks in the area of irradiation (Scheme 23B).50 3D objects can be printed in a layer-by-layer manner using a repetitive irradiation program. The use of light in additive manufacturing can offer higher resolution and faster build speeds to produce geometrically complex objects compared with their thermally mediated 3D printing counterparts.50,192 By using photocontrolled polymerization, many additional merits can be imparted into additive manufacturing.49,193195 For example, the “reversible deactivation” feature in photo-RDRP and the “living” feature in photo-LCP result in well-defined polymeric frameworks and enable chain extension from an existing layer, forming covalent bonds and, hence, strengthening binding between layers. Indeed, the dormant polymer end-group at the surface of printed objects can facilitate continued printing by chain extension and surface processing by chemical modifications. Importantly, implementation of orthogonal photocontrolled polymerization systems could gain facile control over nanostructuration of 3D networkers,48 hence enabling tunable mechanical properties.196198 Moreover’ the extensive range of PCs and corresponding photochemistries available to perform photocontrolled polymerization can provide versatile control over the polymerization process and contribute to materials with diverse physical and chemical properties. However, PCs used in photoinduced additive manufacturing must exhibit high ΦT upon irradiation to populate 3PC* to deactivate oxygen during polymerization by photosensitization. More importantly, the PC should be very efficient so as to catalyze ultrafast polymerization for additive manufacturing. In addition, PCs absorbing longer wavelength light may be sought to increase light penetration into layers of resin. In this regard, it is both complicated and challenging to develop a PC for a particular photoinduced additive manufacturing system, and a computer-aided rational approach is certainly needed. (vi) Developing Biocompatible PCs for photocontrolled polymerization in bioapplications. With regard to the construction or modification of biorelated polymeric materials, thermally initiated polymerization systems are usually inaccessible as biomaterials such as proteins, cells, and living tissues are sensitive to high temperatures. On the other hand, visible light is benign to these biomaterials and controlled polymerization can facilitate production of polymers with well-defined molecular weights. Hence, photocontrolled polymerization has been used in a range of biorelated areas where maintaining bioactivities is important including fabricating protein–polymer bioconjugates63,199 and engineering cell surfaces with polymers37 (Scheme 23C). In these directions, the PCs should be carefully chosen/designed to be nontoxic, biocompatible, and water-soluble. Due to the exposure to oxygen in biosystems, high ΦT of PC may also be needed as mentioned in point (v). With regard to in vivo applications, infrared (IR) absorbing PCs may be desirable for better light penetration. Additionally, biodegradability200202 or photo-bleaching96 is another key feature desired in bioapplications, in order to more easily eliminate the impact of residual PCs after synthesis.

Scheme 23.

Scheme 23.

Some Examples of Future Directions for Photocontrolled Polymerization

We anticipate these future directions will further expand the current toolbox for photocontrolled polymerization systems and potentially other photocatalytic chemical transformations. On the other hand, we expect that the knowledge of structure-property-performance relationships presented in this review will inspire rational PC design for photocontrolled polymerization to facilitate chemical, polymer, and material syntheses. It may also promote future studies on these guiding principles and the theoretical backgrounds behind them.

ACKNOWLEDGMENTS

C.B., N.C., and W.L. acknowledge the support by the Australian Research Council (ARC) via Discovery Research program (DP190100067 and DP210100094). G.M. acknowledges the Colorado Office of Economic Development & International Trade and the National Institutes of Health (Award R35GM119702). W.L. acknowledges the support by the National Natural Science Foundation of China (Grant nos. 21833001 and 21973054), Mountain Tai Climb Program of Shandong Province, and Key-Area Research and Development Program of Guangdong Province (Grant 2020B0101350001). C.W. acknowledges the support by the National Natural Science Foundation of China (Grant no. 22101155) and the China Postdoctoral Science Foundation (Grant nos. 2021M691918 and 2021TQ0191).

GLOSSARY

1PC*

the lowest singlet excited state photocatalyst

3PC*

the lowest triplet excited state photocatalyst

ATRP

atom transfer radical polymerization

BA

n-butyl acrylate

BnBiB

benzyl α-bromoisobutyrate

BSTP

3-benzylsulfanylthiocarbonylthiosulfanyl propionic acid

BTPA

2-(butylthiocarbonothioyl) propionic acid

CDB

cumyl dithiobenzoate

CDTPA

4-cyano-4-[(dodecylsulfanylthiocarbonyl)-sulfanyl]pentanoic acid

CPADB

4-cyanopentanoic acid dithiobenzoate

CPDTC

2-cyano-2-propyl dodecyl trithiocarbonate

CT

charge transfer

D–A

donor–acceptor

DCPD

dicyclopentadiene

DFT

density functional theory

DHO

displaced harmonic oscillator

DVP

dimethyl vinylphosphonate

E 0

standard redox potential

EA

electron affinity

EBiB

ethyl α-bromoisobutyrate

EBPA

ethyl-α-bromophenylacetate

EClPA

ethyl-α-chlorophenylacetate

EPR

electron paramagnetic resonance

ES

excited state

f

oscillator strength

GS

ground state

HOMO

highest occupied molecular orbital

IBDTC

S-1-isobutoxyethyl N,N-diethyl dithiocarbamate

IBTTC

S-1-isobutoxylethyl S′-ethyl trithiocarbonate

IC

internal conversion

IP

ionization potential

ISC

intersystem crossing

k ET

electron transfer rate constant

k F

fluorescence rate constant

k IC

internal conversion rate constant

k ISC

intersystem crossing rate constant

kp app

apparent propagation rate

LED

light emitting diode

LUMO

lowest unoccupied molecular orbital

LSOMO

lower singly occupied molecular orbital

M

monomer

MLCT

metal-to-ligand charge transfer

MMA

methyl methacrylate

ms

millisecond

μs

microsecond

NIR

near-infrared

NMP

nitroxide-mediated radical polymerization

ns

nanosecond

NVP

N-vinyl pyrrolidinone

O-ATRP

organocatalyzed atom transfer radical polymerization

OQP

oxidative quenching pathway

P3T

photoinduced Dexter/triplet energy transfer

PC

photocatalyst

PC•+

radical cation of photocatalyst

PC•−

radical anion of photocatalyst

PDI

perylene diimide

PET

photoinduced electron transfer

PET-RAFT

photoinduced electron/energy transfer-reversible addition–fragmentation chain transfer

photo-ATRP

photocontrolled-atom transfer radical polymerization

photo-CMRP

photocontrolled-cobalt-mediated radical polymerization

photo-ITP

photocontrolled-iodine transfer polymerization

photo-LCP

photocontrolled-living cationic polymerization

photo-NMP

photocontrolled-nitroxide-mediated radical polymerization

photo-RCMP

photocontrolled-reversible complexation mediated living radical polymerization

photo-RDRP

photocontrolled-reversible deactivation radical polymerization

photo-ROMP

photocontrolled-ring-opening metathesis polymerization

photo-TERP

photocontrolled-organotellurium-mediated radical polymerization

ps

picosecond

RAFT

reversible addition–fragmentation chain transfer

ROHF

restricted open-shell Hartree–Fock

ROMP

ring-opening metathesis polymerization

RQP

reductive quenching pathway

RTPP

reduced tetraphenyl porphyrin

s

second

SCE

saturated calomel electrode

SOCME

spin–orbit coupling matrix element

SOMO

singly occupied molecular orbital

S0

singlet ground state

S1

lowest singlet excited state

S2

Second lowest singlet excited state

SCT

charge transfer singlet excited state

SFC

Franck–Condon singlet excited state

Sn

nth singlet excited state

St

styrene

T1

lowest triplet excited state

T2

second triplet singlet excited state

TCT

charge transfer triplet excited state

Tn

nth triplet excited state

TDDFT

time-dependent density functional theory

TPP

tetraphenyl porphyrin

TS

transition state

TTA

triplet–triplet annihilation

USOMO

upper singly occupied molecular orbital

UV

ultraviolet

VP

vinyl pivalate

VR

vibrational relaxation

ZnTPP

Zn(II) tetraphenyl porphyrin

ε max

maximum molar extinction coefficient

λ max

maximum absorption wavelength

τ

excited state lifetime

ΦT

triplet quantum yield

Biographies

Chenyu Wu received his Ph.D. in Chemical Engineering under the supervision of Prof Cyrille Boyer at University of New South Wales, Sydney, in 2020. He is currently a post doctorate at Shandong University where his research focuses on computer-aided rational design of photocatalysts for polymerization, mechanistic understanding of photocontrolled polymerization and development of algorithms for molecular property calculations.

Nathaniel Corrigan received his Ph.D. in Chemical Engineering under the supervision of Prof Cyrille Boyer at University of New South Wales, Sydney, Australia in 2019. His Ph.D. thesis focused on the combination of visible light mediated reversible deactivation radical polymerization (RDRP) and flow chemistry for advanced macromolecular synthesis. He is currently a research associate at UNSW Sydney where his research focuses on exploiting visible light for controlled radical polymerization, flow chemistry, and materials synthesis via 3D printing approaches.

Chern-Hooi Lim received his Ph.D. in Chemical Engineering (2015) from the University of Colorado Boulder under the supervision of Prof. Charles Musgrave. He later joined Prof. Garret Miyake as a postdoctoral researcher and was awarded NIH F32 Postdoctoral Fellowship to advance the development of photocatalysis and organic photocatalysts for efficient small molecule and macromolecule syntheses. He is now the cofounder and CEO of New Iridium Inc, developing photocatalytic solutions that focus on utilizing biobased and CO2 feedstocks for significantly greener and more economical chemical manufacturing. Dr. Lim was recently named to the C&EN Talented 12 (Class of 2021) for his contribution in advancing the science of photocatalysis.

Wenjian Liu obtained his Ph.D. in 1995 from Peking University and was a Cheung Kong professor at Peking University between 2001 and 2018. He is now a chair professor at Shandong University and the inaugural director of Qingdao Institute for Theoretical and Computational Sciences. He is an elected member of the International Academy of Quantum Molecular Science, fellow of the Royal Society of Chemistry, and fellow of the Asia-Pacific Association of Theoretical and Computational Chemists. He was honored to receive the Annual Medal of the International Academy of Quantum Molecular Science, the Bessel Award of the Humboldt Foundation, and the People and Fukui Medals of the Asia-Pacific Association of Theoretical and Computational Chemists. His research focuses on relativistic theories and methods for molecular electronic structure and magnetic properties, as well as wave function-based methods for strongly correlated electrons.

Garret Miyake is an Associate Professor and the Dr. Robert Williams Professor of Organic Chemistry at Colorado State University. He earned a B.S. at Pacific University. He performed Ph.D. studies with Eugene Chen at Colorado State University before conducting postdoctoral research with Robert Grubbs at the California Institute of Technology. The Miyake group has research interests in the fields of photoredox catalysis, organocatalyzed atom-transfer radical polymerization, sustainable plastics, and the synthesis of block copolymers that self-assemble to photonic crystals. He has been recognized by the 2021 Journal of Polymer Science Innovation Award, the 2017 ACS Division of Polymer Chemistry Herman F. Mark Young Scholar Award, a Camille Dreyfus Teacher-Scholar Award, a Cottrell Scholar Award, and a Sloan Research Fellowship. He is also a cofounder of New Iridium and Cypris Materials.

Cyrille Boyer obtained his Ph.D. in 2006 from the University of Montpellier. He is a full professor and deputy head of school in the School of Chemical Engineering, at the University of New South Wales, Sydney, and codirector of Australian Centre for Nanomedicine. He is an elected member of European Academy of Sciences and has published over 340 research articles, which have generated over 25,000 citations, resulting in an H-index of 90. He was selected as a “Highly Cited Researcher in Chemistry” by Web of Science in 2018, 2019, and 2020, and that for “exceptional cross-field performance” in 2021. His research group works on the development of new polymerization techniques using photocatalysts, resulting in the discovery of PET-RAFT polymerization in 2014 and the development of photosingle unit monomer insertion (photo-SUMI). He uses these polymerization tools for the preparation of functional polymers which find applications in nanomedicine, energy, and materials science. He was the recipient of the 6th Polymer International-IUPAC Award for Creativity in Applied Polymer Science or Polymer Technology, 2018 Award of Excellence in Chemical Engineering awarded by ICHEM, 2018 Polymer Chemistry Lectureship awarded by Royal Society of Chemistry, 2015 Prime Minister Prize for Physical Science in Australia, 2016 ACS Biomacromolecules/Macromolecules Award, 2016 Le Fevre Memorial Award (Australia Academia of Science), and 2014 Scopus Young Research Award.

Footnotes

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.chemrev.1c00409

The authors declare no competing financial interest.

Contributor Information

Chenyu Wu, Qingdao Institute for Theoretical and Computational Sciences, Institute of Frontier and Interdisciplinary Science, Shandong University, Qingdao 266237, China; Centre for Advanced Macromolecular Design (CAMD), School of Chemical Engineering and Australian Centre for NanoMedicine, School of Chemical Engineering, UNSW Sydney, Sydney, NSW 2052, Australia.

Nathaniel Corrigan, Centre for Advanced Macromolecular Design (CAMD), School of Chemical Engineering and Australian Centre for NanoMedicine, School of Chemical Engineering, UNSW Sydney, Sydney, NSW 2052, Australia.

Chern-Hooi Lim, Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523, United States; New Iridium Incorporated, Boulder, Colorado 80303, United States.

Wenjian Liu, Qingdao Institute for Theoretical and Computational Sciences, Institute of Frontier and Interdisciplinary Science, Shandong University Qingdao 266237, China.

Garret Miyake, Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523, United States.

Cyrille Boyer, Centre for Advanced Macromolecular Design (CAMD), School of Chemical Engineering and Australian Centre for NanoMedicine, School of Chemical Engineering, UNSW Sydney, Sydney, NSW 2052, Australia.

REFERENCES

  • (1).Key HM; Dydio P; Clark DS; Hartwig JF Abiological Catalysis by Artificial Haem Proteins Containing Noble Metals in Place Of Iron. Nature 2016, 534, 534–537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Cao Y; Li X; Ge J Enzyme Catalyst Engineering toward the Integration of Biocatalysis and Chemocatalysis. Trends Biotechnol. 2021, in press, DOI: 10.1016/j.tibtech.2021.01.002. [DOI] [PubMed] [Google Scholar]
  • (3).Toogood HS; Scrutton NS Enzyme Engineering Toolbox - a ’Catalyst’ for Change. Catal. Sci. Technol 2013, 3, 2182–2194. [Google Scholar]
  • (4).MacMillan DW The Advent and Development of Organocatalysis. Nature 2008, 455, 304–308. [DOI] [PubMed] [Google Scholar]
  • (5).Blanco V; Leigh DA; Marcos V Artificial Switchable Catalysts. Chem. Soc. Rev 2015, 44, 5341–5370. [DOI] [PubMed] [Google Scholar]
  • (6).Xu SH; Ng G; Xu JT; Kuchel RP; Yeow J; Boyer C 2-(Methylthio)ethyl Methacrylate: a Versatile Monomer for Stimuli Responsiveness and Polymerization-Induced Self-Assembly in the Presence of Air. ACS Macro Lett. 2017, 6, 1237–1244. [DOI] [PubMed] [Google Scholar]
  • (7).Corrigan N; Shanmugam S; Xu J; Boyer C Photocatalysis in Organic and Polymer Synthesis. Chem. Soc. Rev 2016, 45, 6165–6212. [DOI] [PubMed] [Google Scholar]
  • (8).Joshi-Pangu A; Levesque F; Roth HG; Oliver SF; Campeau LC; Nicewicz D; DiRocco DA Acridinium-Based Photocatalysts: a Sustainable Option in Photoredox Catalysis. J. Org. Chem 2016, 81, 7244–7249. [DOI] [PubMed] [Google Scholar]
  • (9).Busacca CA; Fandrick DR; Song JJ; Senanayake CH The Growing Impact of Catalysis in the Pharmaceutical Industry. Adv. Synth. Catal 2011, 353, 1825–1864. [Google Scholar]
  • (10).Hagen J Industrial Catalysis: A Practical Approach. Wiley: 2015, 17–78. [Google Scholar]
  • (11).Shaw MH; Twilton J; MacMillan DW Photoredox Catalysis in Organic Chemistry. J. Org. Chem 2016, 81, 6898–6926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).König B Chemical Photocatalysis; De Gruyter, 2020, pp 1–44. [Google Scholar]
  • (13).Romero NA; Nicewicz DA Organic Photoredox Catalysis. Chem. Rev 2016, 116, 10075–10166. [DOI] [PubMed] [Google Scholar]
  • (14).Prier CK; Rankic DA; MacMillan DW Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis. Chem. Rev 2013, 113, 5322–5363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Chen M; Zhong M; Johnson JA Light-Controlled Radical Polymerization: Mechanisms, Methods, and Applications. Chem. Rev 2016, 116, 10167–10211. [DOI] [PubMed] [Google Scholar]
  • (16).Zivic N; Bouzrati-Zerelli M; Kermagoret A; Dumur F; Fouassier JP; Gigmes D; Lalevée J Photocatalysts in Polymerization Reactions. ChemCatChem 2016, 8, 1617–1631. [Google Scholar]
  • (17).Dadashi-Silab S; Doran S; Yagci Y Photoinduced Electron Transfer Reactions for Macromolecular Syntheses. Chem. Rev 2016, 116, 10212–10275. [DOI] [PubMed] [Google Scholar]
  • (18).Schultz DM; Yoon TP Solar Synthesis: Prospects in Visible Light Photocatalysis. Science 2014, 343, 1239176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Yoon TP; Ischay MA; Du J Visible Light Photocatalysis as a Greener Approach to Photochemical Synthesis. Nat. Chem 2010, 2, 527–532. [DOI] [PubMed] [Google Scholar]
  • (20).Narayanam JM; Stephenson CR Visible Light Photoredox Catalysis: Applications in Organic Synthesis. Chem. Soc. Rev 2011, 40, 102–113. [DOI] [PubMed] [Google Scholar]
  • (21).Ravelli D; Fagnoni M; Albini A Photoorganocatalysis. What for? Chem. Soc. Rev 2013, 42, 97–113. [DOI] [PubMed] [Google Scholar]
  • (22).Speckmeier E; Fischer TG; Zeitler K A Toolbox Approach To Construct Broadly Applicable Metal-Free Catalysts for Photoredox Chemistry: Deliberate Tuning of Redox Potentials and Importance of Halogens in Donor-Acceptor Cyanoarenes. J. Am. Chem. Soc 2018, 140, 15353–15365. [DOI] [PubMed] [Google Scholar]
  • (23).Orr-Ewing AJ Perspective: How Can Ultrafast Laser Spectroscopy Inform the Design of New Organic Photoredox Catalysts for Chemical and Materials Synthesis? Struct. Dyn 2019, 6, 010901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (24).Lu PT; Chung KY; Stafford A; Kiker M; Kafle K; Page ZA Boron Dipyrromethene (BODIPY) in Polymer Chemistry. Polym. Chem 2021, 12, 327–348. [Google Scholar]
  • (25).Zhang J; Tan KL; Hong GD; Yang LJ; Gong HQ Polymerization Optimization of SU-8 Photoresist and its Applications in Microfluidic Systems and MEMS. J. Micromech. Microeng 2001, 11, 20–26. [Google Scholar]
  • (26).Lin Y; Gao C; Gritsenko D; Zhou R; Xu J Soft Lithography Based on Photolithography and Two-Photon Polymerization. Microfluid. Nanofluid 2018, 22, 97. [Google Scholar]
  • (27).O’Brien J; Hughes PJ; Brunet M; O’Neill B; Alderman J; Lane B; O’Riordan A; O’Driscoll C Advanced Photoresist Technologies for Microsystems. J. Micromech. Microeng 2001, 11, 353–358. [Google Scholar]
  • (28).Corrigan N; Ciftci M; Jung K; Boyer C Mediating Reaction Orthogonality in Polymer and Materials Science. Angew. Chem., Int. Ed 2021, 60, 1748–1781. [DOI] [PubMed] [Google Scholar]
  • (29).Corrigan N; Boyer C 100th Anniversary of Macromolecular Science Viewpoint: Photochemical Reaction Orthogonality in Modern Macromolecular Science. ACS Macro Lett. 2019, 8, 812–818. [DOI] [PubMed] [Google Scholar]
  • (30).Fu C; Xu J; Boyer C Photoacid-Mediated Ring Opening Polymerization Driven by Visible Light. Chem. Commun 2016, 52, 7126–7129. [DOI] [PubMed] [Google Scholar]
  • (31).Kottisch V; Michaudel Q; Fors BP Photocontrolled Interconversion of Cationic and Radical Polymerizations. J. Am. Chem. Soc 2017, 139, 10665–10668. [DOI] [PubMed] [Google Scholar]
  • (32).Xiong XH; Xue LL; Cui JX Phototriggered Growth and Detachment of Polymer Brushes with Wavelength Selectivity. ACS Macro Lett. 2018, 7, 239–243. [DOI] [PubMed] [Google Scholar]
  • (33).Hiltebrandt K; Pauloehrl T; Blinco JP; Linkert K; Borner HG; Barner-Kowollik C Lambda-Orthogonal Pericyclic Macromolecular Photoligation. Angew. Chem., Int. Ed 2015, 54, 2838–2843. [DOI] [PubMed] [Google Scholar]
  • (34).Hiltebrandt K; Kaupp M; Molle E; Menzel JP; Blinco JP; Barner-Kowollik C Star Polymer Synthesis Vialambda-Orthogonal Photochemistry. Chem. Commun 2016, 52, 9426–9429. [DOI] [PubMed] [Google Scholar]
  • (35).Judzewitsch PR; Zhao L; Wong EHH; Boyer C High-Throughput Synthesis of Antimicrobial Copolymers and Rapid Evaluation of their Bioactivity. Macromolecules 2019, 52, 3975–3986. [Google Scholar]
  • (36).Judzewitsch PR; Nguyen TK; Shanmugam S; Wong EHH; Boyer C Towards Sequence-Controlled Antimicrobial Polymers: Effect of Polymer Block Order on Antimicrobial Activity. Angew. Chem., Int. Ed 2018, 57, 4559–4564. [DOI] [PubMed] [Google Scholar]
  • (37).Niu J; Lunn DJ; Pusuluri A; Yoo JI; O’Malley MA; Mitragotri S; Soh HT; Hawker CJ Engineering Live Cell Surfaces with Functional Polymers via Cytocompatible Controlled Radical Polymerization. Nat. Chem 2017, 9, 537–545. [DOI] [PubMed] [Google Scholar]
  • (38).Tucker BS; Coughlin ML; Figg CA; Sumerlin BS Grafting-From Proteins Using Metal-Free PET-RAFT Polymerizations under Mild Visible-Light Irradiation. ACS Macro Lett. 2017, 6, 452–457. [DOI] [PubMed] [Google Scholar]
  • (39).Yeow J; Shanmugam S; Corrigan N; Kuchel RP; Xu JT; Boyer C A Polymerization-Induced Self-Assembly Approach to Nanoparticles Loaded with Singlet Oxygen Generators. Macromolecules 2016, 49, 7277–7285. [Google Scholar]
  • (40).Szymanski JK; Perez-Mercader J Direct Optical Observations of Vesicular Self-Assembly in Large-Scale Polymeric Structures During Photocontrolled Biphasic Polymerization. Polym. Chem 2016, 7, 7211–7215. [Google Scholar]
  • (41).Albertsen AN; Szymanski JK; Perez-Mercader J Emergent Properties of Giant Vesicles Formed by a Polymerization-Induced Self-Assembly (PISA) Reaction. Sci. Rep 2017, 7, 41534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (42).Xu S; Zhang T; Kuchel RP; Yeow J; Boyer C Gradient Polymerization-Induced Self-Assembly: a One-Step Approach. Macro-mol Rapid Commun 2020, 41, 1900493. [DOI] [PubMed] [Google Scholar]
  • (43).Xu S; Yeow J; Boyer C Exploiting Wavelength Orthogonality for Successive Photoinduced Polymerization-Induced Self-Assembly and Photo-Crosslinking. ACS Macro Lett. 2018, 7, 1376–1382. [DOI] [PubMed] [Google Scholar]
  • (44).Poelma JE; Fors BP; Meyers GF; Kramer JW; Hawker CJ Fabrication of Complex Three-Dimensional Polymer Brush Nanostructures through Light-Mediated Living Radical Polymerization. Angew. Chem., Int. Ed 2013, 52, 6844–6848. [DOI] [PubMed] [Google Scholar]
  • (45).Fors BP; Poelma JE; Menyo MS; Robb MJ; Spokoyny DM; Kramer JW; Waite JH; Hawker CJ Fabrication of Unique Chemical Patterns and Concentration Gradients with Visible Light. J. Am. Chem. Soc 2013, 135, 14106–14109. [DOI] [PubMed] [Google Scholar]
  • (46).Pester CW; Narupai B; Mattson KM; Bothman DP; Klinger D; Lee KW; Discekici EH; Hawker CJ Engineering Surfaces through Sequential Stop-Flow Photopatterning. Adv. Mater 2016, 28, 9292–9300. [DOI] [PubMed] [Google Scholar]
  • (47).Discekici EH; Pester CW; Treat NJ; Lawrence I; Mattson KM; Narupai B; Toumayan EP; Luo YD; McGrath AJ; Clark PG; et al. Simple Benchtop Approach to Polymer Brush Nanostructures Using Visible-Light-Mediated Metal-Free Atom Transfer Radical Polymerization. ACS Macro Lett. 2016, 5, 258–262. [DOI] [PubMed] [Google Scholar]
  • (48).Chen M; Gu Y; Singh A; Zhong M; Jordan AM; Biswas S; Korley LT; Balazs AC; Johnson JA Living Additive Manufacturing: Transformation of Parent Gels into Diversely Functionalized Daughter Gels Made Possible by Visible Light Photoredox Catalysis. ACS Cent. Sci 2017, 3, 124–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Zhang Z; Corrigan N; Bagheri A; Jin J; Boyer C A Versatile 3D and 4D Printing System through Photocontrolled RAFT Polymerization. Angew. Chem. Int. Ed 2019, 58, 17954–17963. [DOI] [PubMed] [Google Scholar]
  • (50).Jung K; Corrigan N; Ciftci M; Xu J; Seo SE; Hawker CJ; Boyer C Designing with Light: Advanced 2D, 3D, and 4D Materials. Adv. Mater 2020, 32, 1903850. [DOI] [PubMed] [Google Scholar]
  • (51).Pan XC; Tasdelen MA; Laun J; Junkers T; Yagci Y; Matyjaszewski K Photomediated Controlled Radical Polymerization. Prog. Polym. Sci 2016, 62, 73–125. [Google Scholar]
  • (52).Singh VK; Yu C; Badgujar S; Kim Y; Kwon Y; Kim D; Lee J; Akhter T; Thangavel G; Park LS; et al. Highly Efficient Organic Photocatalysts Discovered via a Computer-Aided-Design Strategy for Visible-Light-Driven Atom Transfer Radical Polymerization. Nat. Catal 2018, 1, 794–804. [Google Scholar]
  • (53).Song Y; Kim Y; Noh Y; Singh VK; Behera SK; Abudulimu A; Chung K; Wannemacher R; Gierschner J; Luer L; Kwon MS Organic Photocatalyst for ppm-Level Visible-Light-Driven Reversible Addition-Fragmentation Chain-Transfer (RAFT) Polymerization with Excellent Oxygen Tolerance. Macromolecules 2019, 52, 5538–5545. [Google Scholar]
  • (54).Cole JP; Federico CR; Lim CH; Miyake GM Photoinduced Organocatalyzed Atom Transfer Radical Polymerization Using Low ppm Catalyst Loading. Macromolecules 2019, 52, 747–754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Corrigan N; Rosli D; Jones JWJ; Xu JT; Boyer C Oxygen Tolerance in Living Radical Polymerization: Investigation of Mechanism and Implementation in Continuous Flow Polymerization. Macromolecules 2016, 49, 6779–6789. [Google Scholar]
  • (56).Wu C; Jung K; Ma Y; Liu W; Boyer C Unravelling an Oxygen-Mediated Reductive Quenching Pathway for Photopolymerisation under Long Wavelengths. Nat. Commun 2021, 12, 478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).McCarthy B; Miyake GM Organocatalyzed Atom Transfer Radical Polymerization Catalyzed by Core Modified N-Aryl Phenoxazines Performed under Air. ACS Macro Lett. 2018, 7, 1016–1021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Liarou E; Whitfield R; Anastasaki A; Engelis NG; Jones GR; Velonia K; Haddleton DM Copper-Mediated Polymerization without External Deoxygenation or Oxygen Scavengers. Angew. Chem., Int. Ed 2018, 57, 8998–9002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Xu J; Shanmugam S; Fu C; Aguey-Zinsou KF; Boyer C Selective Photoactivation: From a Single Unit Monomer Insertion Reaction to Controlled Polymer Architectures. J. Am. Chem. Soc 2016, 138, 3094–3106. [DOI] [PubMed] [Google Scholar]
  • (60).Fors BP; Hawker CJ Control of a Living Radical Polymerization of Methacrylates by Light. Angew. Chem., Int. Ed 2012, 51, 8850–8853. [DOI] [PubMed] [Google Scholar]
  • (61).Xu J; Jung K; Atme A; Shanmugam S; Boyer C A Robust and Versatile Photoinduced Living Polymerization of Conjugated and Unconjugated Monomers and its Oxygen Tolerance. J. Am. Chem. Soc 2014, 136, 5508–5519. [DOI] [PubMed] [Google Scholar]
  • (62).Alfredo NV; Jalapa NE; Morales SL; Ryabov AD; Le Lagadec R; Alexandrova L Light-Driven Living/Controlled Radical Polymerization of Hydrophobic Monomers Catalyzed by Ruthenium-(II) Metalacycles. Macromolecules 2012, 45, 8135–8146. [Google Scholar]
  • (63).Xu J; Jung K; Corrigan NA; Boyer C Aqueous Photoinduced Living/Controlled Polymerization: Tailoring for Bioconjugation. Chem. Sci 2014, 5, 3568–3575. [Google Scholar]
  • (64).Treat NJ; Sprafke H; Kramer JW; Clark PG; Barton BE; Read de Alaniz J; Fors BP; Hawker CJ Metal-Free Atom Transfer Radical Polymerization. J. Am. Chem. Soc 2014, 136, 16096–16101. [DOI] [PubMed] [Google Scholar]
  • (65).Kottisch V; Michaudel Q; Fors BP Cationic Polymerization of Vinyl Ethers Controlled by Visible Light. J. Am. Chem. Soc 2016, 138, 15535–15538. [DOI] [PubMed] [Google Scholar]
  • (66).Liu XD; Zhang LF; Cheng ZP; Zhu XL Metal-Free Photoinduced Electron Transfer-Atom Transfer Radical Polymerization (PET-ATRP) via a Visible Light Organic Photocatalyst. Polym. Chem 2016, 7, 689–700. [Google Scholar]
  • (67).Xu JT; Shanmugam S; Duong HT; Boyer C Organo-Photocatalysts for Photoinduced Electron Transfer-Reversible Addition-Fragmentation Chain Transfer (PET-RAFT) Polymerization. Polym. Chem 2015, 6, 5615–5624. [Google Scholar]
  • (68).Srivastava V; Singh PP Eosin Y Catalysed Photoredox Synthesis: a Review. RSC Adv. 2017, 7, 31377–31392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (69).Shanmugam S; Xu J; Boyer C Light-Regulated Polymerization under Near-Infrared/Far-Red Irradiation Catalyzed by Bacteriochlorophyll a. Angew. Chem., Int. Ed 2016, 55, 1036–1040. [DOI] [PubMed] [Google Scholar]
  • (70).Shanmugam S; Xu J; Boyer C Utilizing the Electron Transfer Mechanism of Chlorophyll a under Light for Controlled Radical Polymerization. Chem. Sci 2015, 6, 1341–1349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (71).Corrigan N; Xu JT; Boyer C A Photoinitiation System for Conventional and Controlled Radical Polymerization at Visible and NIR Wavelengths. Macromolecules 2016, 49, 3274–3285. [Google Scholar]
  • (72).Pan X; Fang C; Fantin M; Malhotra N; So WY; Peteanu LA; Isse AA; Gennaro A; Liu P; Matyjaszewski K Mechanism of Photoinduced Metal-Free Atom Transfer Radical Polymerization: Experimental and Computational Studies. J. Am. Chem. Soc 2016, 138, 2411–2425. [DOI] [PubMed] [Google Scholar]
  • (73).Wu C; Corrigan N; Lim CH; Jung K; Zhu J; Miyake G; Xu J; Boyer C Guiding the Design of Organic Photocatalyst for PET-RAFT Polymerization: Halogenated Xanthene Dyes. Macromolecules 2019, 52, 236–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (74).McCarthy BG; Pearson RM; Lim CH; Sartor SM; Damrauer NH; Miyake GM Structure-Property Relationships for Tailoring Phenoxazines as Reducing Photoredox Catalysts. J. Am. Chem. Soc 2018, 140, 5088–5101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (75).Wu C; Chen H; Corrigan N; Jung K; Kan X; Li Z; Liu W; Xu J; Boyer C Computer-Guided Discovery of a pH-Responsive Organic Photocatalyst and Application for pH and Light Dual-Gated Polymerization. J. Am. Chem. Soc 2019, 141, 8207–8220. [DOI] [PubMed] [Google Scholar]
  • (76).Franck J; Dymond EG Elementary Processes of Photochemical Reactions. Trans. Faraday Soc 1926, 21, 536–542. [Google Scholar]
  • (77).Condon EU Nuclear Motions Associated with Electron Transitions in Diatomic Molecules. Phys. Rev 1928, 32, 858–872. [Google Scholar]
  • (78).Pascual LMM; Dunford DG; Goetz AE; Ogawa KA; Boydston AJ Comparison of Pyrylium and Thiopyrylium Photooxidants in Metal-Free Ring-Opening Metathesis Polymerization. Synlett 2016, 27, 759–762. [Google Scholar]
  • (79).Ogawa KA; Goetz AE; Boydston AJ Metal-Free Ring-Opening Metathesis Polymerization. J. Am. Chem. Soc 2015, 137, 1400–1403. [DOI] [PubMed] [Google Scholar]
  • (80).Theriot JC; Lim CH; Yang H; Ryan MD; Musgrave CB; Miyake GM Organocatalyzed Atom Transfer Radical Polymerization Driven by Visible Light. Science 2016, 352, 1082–1086. [DOI] [PubMed] [Google Scholar]
  • (81).Pan X; Lamson M; Yan JJ; Matyjaszewski K Photoinduced Metal-Free Atom Transfer Radical Polymerization of Acrylonitrile. ACS Macro Lett. 2015, 4, 192–196. [DOI] [PubMed] [Google Scholar]
  • (82).Pan X; Malhotra N; Simakova A; Wang Z; Konkolewicz D; Matyjaszewski K Photoinduced Atom Transfer Radical Polymerization with ppm-Level Cu Catalyst by Visible Light in Aqueous Media. J. Am. Chem. Soc 2015, 137, 15430–15433. [DOI] [PubMed] [Google Scholar]
  • (83).Tasdelen MA; Ciftci M; Yagci Y Visible Light-Induced Atom Transfer Radical Polymerization. Macromol. Chem. Phys 2012, 213, 1391–1396. [Google Scholar]
  • (84).Treat NJ; Fors BP; Kramer JW; Christianson M; Chiu C-Y; Read de Alaniz J; Hawker CJ Controlled Radical Polymerization of Acrylates Regulated by Visible Light. ACS Macro Lett. 2014, 3, 580–584. [DOI] [PubMed] [Google Scholar]
  • (85).Yang QZ; Dumur F; Morlet-Savary F; Poly J; Lalevée J Photocatalyzed Cu-Based ATRP Involving an Oxidative Quenching Mechanism under Visible Light. Macromolecules 2015, 48, 1972–1980. [Google Scholar]
  • (86).Shanmugam S; Xu J; Boyer C Exploiting Metalloporphyrins for Selective Living Radical Polymerization Tunable over Visible Wavelengths. J. Am. Chem. Soc 2015, 137, 9174–9185. [DOI] [PubMed] [Google Scholar]
  • (87).Su JH; Liu XX; Hu JQ; You Q; Cui YY; Chen YY Photo-Induced Controlled Radical Polymerization of Methyl Methacrylate Mediated by Photosensitive Nitroxides. Polym. Int 2015, 64, 867–874. [Google Scholar]
  • (88).Guillaneuf Y; Bertin D; Gigmes D; Versace D-L; Lalevee J; Fouassier J-P Toward Nitroxide-Mediated Photopolymerization. Macromolecules 2010, 43, 2204–2212. [Google Scholar]
  • (89).Zhao Y; Yu M; Fu X Photo-Cleavage Of The Cobalt-Carbon Bond: Visible Light-Induced Living Radical Polymerization Mediated by Organo-Cobalt Porphyrins. Chem. Commun 2013, 49, 5186–5188. [DOI] [PubMed] [Google Scholar]
  • (90).Kermagoret A; Wenn B; Debuigne A; Jerome C; Junkers T; Detrembleur C Improved Photo-Induced Cobalt-Mediated Radical Polymerization in Continuous FlowPhotoreactors. Polym. Chem 2015, 6, 3847–3857. [Google Scholar]
  • (91).Koumura K; Satoh K; Kamigaito M Manganese-Based Controlled/Living Radical Polymerization of Vinyl Acetate, Methyl Acrylate, and Styrene: Highly Active, Versatile, and Photoresponsive Systems. Macromolecules 2008, 41, 7359–7367. [Google Scholar]
  • (92).Ohtsuki A; Goto A; Kaji H Visible-Light-Induced Reversible Complexation Mediated Living Radical Polymerization of Methacrylates with Organic Catalysts. Macromolecules 2013, 46, 96–102. [Google Scholar]
  • (93).Nakamura Y; Yamago S Organotellurium-Mediated Living Radical Polymerization under Photoirradiation by a Low-Intensity Light-Emitting Diode. Beilstein J. Org. Chem 2013, 9, 1607–1612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (94).Kottisch V; Supej MJ; Fors BP Enhancing Temporal Control and Enabling Chain-End Modification in Photoregulated Cationic Polymerizations by Using Iridium-Based Catalysts. Angew. Chem., Int. Ed 2018, 57, 8260–8264. [DOI] [PubMed] [Google Scholar]
  • (95).Konya M; Uchiyama M; Satoh K; Kamigaito M Cationic Polymerization via Activation of Alkoxyamines Using Photoredox Catalysts. ChemPhotoChem. 2019, 3, 1100–1108. [Google Scholar]
  • (96).Figg CA; Hickman JD; Scheutz GM; Shanmugam S; Carmean RN; Tucker BS; Boyer C; Sumerlin BS Color-Coding Visible Light Polymerizations to Elucidate the Activation of Trithiocarbonates Using Eosin Y. Macromolecules 2018, 51, 1370–1376. [Google Scholar]
  • (97).Nomeir B; Fabre O; Ferji K Effect of Tertiary Amines on the Photoinduced Electron Transfer Reversible Addition Fragmentation Chain Transfer (PET-RAFT) Polymerization. Macromolecules 2019, 52, 6898–6903. [Google Scholar]
  • (98).Corrigan N; Xu JT; Boyer C; Allonas X Exploration of the PET-RAFT Initiation Mechanism for Two Commonly Used Photocatalysts. ChemPhotoChem. 2019, 3, 1193–1199. [Google Scholar]
  • (99).Allegrezza ML; Konkolewicz D PET-RAFT Polymerization: Mechanistic Perspectives for Future Materials. ACS Macro Lett. 2021, 10, 433–446. [DOI] [PubMed] [Google Scholar]
  • (100).Christmann J; Ibrahim A; Charlot V; Croutxe-Barghorn C; Ley C; Allonas X Elucidation of the Key Role of [Ru(bpy)3](2+) in Photocatalyzed RAFT Polymerization. ChemPhysChem 2016, 17, 2309–2314. [DOI] [PubMed] [Google Scholar]
  • (101).Zhang G; Song IY; Ahn KH; Park T; Choi W Free Radical Polymerization Initiated and Controlled by Visible Light Photocatalysis at Ambient Temperature. Macromolecules 2011, 44, 7594–7599. [Google Scholar]
  • (102).Furst L; Matsuura BS; Narayanam JM; Tucker JW; Stephenson CR Visible Light-Mediated Intermolecular C-H Functionalization of Electron-Rich Heterocycles with Malonates. Org. Lett 2010, 12, 3104–3107. [DOI] [PubMed] [Google Scholar]
  • (103).Lim CH; Ryan MD; McCarthy BG; Theriot JC; Sartor SM; Damrauer NH; Musgrave CB; Miyake GM Intramolecular Charge Transfer and Ion Pairing in N,N-Diaryl Dihydrophenazine Photoredox Catalysts for Efficient Organocatalyzed Atom Transfer Radical Polymerization. J. Am. Chem. Soc 2017, 139, 348–355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (104).Eyring H The Activated Complex in Chemical Reactions. J. Chem. Phys 1935, 3, 107–115. [Google Scholar]
  • (105).Marcus RA Unimolecular Dissociations and Free Radical Recombination Reactions. J. Chem. Phys 1952, 20, 359–364. [Google Scholar]
  • (106).Nelson TR; White AJ; Bjorgaard JA; Sifain AE; Zhang Y; Nebgen B; Fernandez-Alberti S; Mozyrsky D; Roitberg AE; Tretiak S Non-adiabatic Excited-State Molecular Dynamics: Theory and Applications for Modeling Photophysics in Extended Molecular Materials. Chem. Rev 2020, 120, 2215–2287. [DOI] [PubMed] [Google Scholar]
  • (107).Nzulu F; Telitel S; Stoffelbach F; Graff B; Morlet-Savary F; Lalevee J; Fensterbank L; Goddard JP; Ollivier C A Dinuclear Gold(I) Complex as a Novel Photoredox Catalyst for Light-Induced Atom Transfer Radical Polymerization. Polym. Chem 2015, 6, 4605–4611. [Google Scholar]
  • (108).Telitel S; Dumur F; Campolo D; Poly J; Gigmes D; Fouassier JP; Lalevee J Iron Complexes as Potential Photocatalysts For Controlled Radical Photopolymerizations: a Tool for Modifications and Patterning of Surfaces. J. Polym. Sci., Part A: Polym. Chem 2016, 54, 702–713. [Google Scholar]
  • (109).Konkolewicz D; Schroder K; Buback J; Bernhard S; Matyjaszewski K Visible Light and Sunlight Photoinduced ATRP with ppm of Cu Catalyst. ACS Macro Lett. 2012, 1, 1219–1223. [DOI] [PubMed] [Google Scholar]
  • (110).Miyake GM; Theriot JC Perylene as an Organic Photocatalyst for the Radical Polymerization of Functionalized Vinyl Monomers through Oxidative Quenching with Alkyl Bromides and Visible Light. Macromolecules 2014, 47, 8255–8261. [Google Scholar]
  • (111).Pearson RM; Lim CH; McCarthy BG; Musgrave CB; Miyake GM Organocatalyzed Atom Transfer Radical Polymerization Using N-Aryl Phenoxazines as Photoredox Catalysts. J. Am. Chem. Soc 2016, 138, 11399–11407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (112).Buss BL; Lim CH; Miyake GM Dimethyl Dihydroacridines as Photocatalysts in Organocatalyzed Atom Transfer Radical Polymerization of Acrylate Monomers. Angew. Chem. Int. Ed 2020, 59, 3209–3217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (113).Sartor SM; McCarthy BG; Pearson RM; Miyake GM; Damrauer NH Exploiting Charge-Transfer States for Maximizing Intersystem Crossing Yields in Organic Photoredox Catalysts. J. Am. Chem. Soc 2018, 140, 4778–4781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (114).Kutahya C; Allushi A; Isci R; Kreutzer J; Ozturk T; Yilmaz G; Yagci Y Photoinduced Metal-Free Atom Transfer Radical Polymerization Using Highly Conjugated Thienothiophene Derivatives. Macromolecules 2017, 50, 6903–6910. [Google Scholar]
  • (115).Seal P; Xu J; Luca S; Boyer C; Smith SC Unraveling Photocatalytic Mechanism and Selectivity in PET-RAFT Polymerization. Adv. Theor. Simul 2019, 2, 1900038. [Google Scholar]
  • (116).Dolinski ND; Page ZA; Discekici EH; Meis D; Lee IH; Jones GR; Whitfield R; Pan X; McCarthy BG; Shanmugam S; et al. What Happens in the Dark? Assessing the Temporal Control of Photo-Mediated Controlled Radical Polymerizations. J. Polym. Sci., Part A: Polym. Chem 2019, 57, 268–273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (117).Otsu T Iniferter Concept and Living Radical Polymerization. J. Polym. Sci., Part A: Polym. Chem 2000, 38, 2121–2136. [Google Scholar]
  • (118).Thum MD; Wolf S; Falvey DE State-Dependent Photochemical and Photophysical Behavior of Dithiolate Ester and Trithiocarbonate Reversible Addition-Fragmentation Chain Transfer Polymerization Agents. J. Phys. Chem. A 2020, 124, 4211–4222. [DOI] [PubMed] [Google Scholar]
  • (119).Kurek PN; Kloster AJ; Weaver KA; Manahan R; Allegrezza ML; De Alwis Watuthanthrige N; Boyer C; Reeves JA; Konkolewicz D How Do Reaction and Reactor Conditions Affect Photoinduced Electron/Energy Transfer Reversible Addition-Fragmentation Transfer Polymerization? Ind. Eng. Chem. Res 2018, 57, 4203–4213. [Google Scholar]
  • (120).Shanmugam S; Xu JT; Boyer C Aqueous RAFT Photopolymerization with Oxygen Tolerance. Macromolecules 2016, 49, 9345–9357. [Google Scholar]
  • (121).Cao HL; Wang GC; Xue YD; Yang GL; Tian J; Liu F; Zhang WA Far-Red Light-Induced Reversible Addition-Fragmentation Chain Transfer Polymerization Using a Man-Made Bacteriochlorin. ACS Macro Lett. 2019, 8, 616–622. [DOI] [PubMed] [Google Scholar]
  • (122).Huang Z; Zhang L; Cheng Z; Zhu X Reversible Addition-Fragmentation Chain Transfer Polymerization of Acrylonitrile under Irradiation of Blue LED Light. Polymers 2017, 9, 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (123).Theriot JC; Miyake GM; Boyer CA, N,N-Diaryl Dihydrophenazines as Photoredox Catalysts for PET-RAFT and Sequential PET-RAFT/O-ATRP. ACS Macro Lett. 2018, 7, 662–666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (124).Michaudel Q; Chauvire T; Kottisch V; Supej MJ; Stawiasz KJ; Shen L; Zipfel WR; Abruna HD; Freed JH; Fors BP Mechanistic Insight into the Photocontrolled Cationic Polymerization of Vinyl Ethers. J. Am. Chem. Soc 2017, 139, 15530–15538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (125).Zhang X; Jiang Y; Ma Q; Hu S; Liao S Metal-Free Cationic Polymerization of Vinyl Ethers with Strict Temporal Control by Employing an Organophotocatalyst. J. Am. Chem. Soc 2021, 143, 6357–6362. [DOI] [PubMed] [Google Scholar]
  • (126).Lu P; Kensy VK; Tritt RL; Seidenkranz DT; Boydston AJ Metal-Free Ring-Opening Metathesis Polymerization: From Concept to Creation. Acc. Chem. Res 2020, 53, 2325–2335. [DOI] [PubMed] [Google Scholar]
  • (127).Grubbs RH; Carr DD; Hoppin C; Burk PL Consideration of the Mechanism of the Metal Catalyzed Olefin Metathesis Reaction. J. Am. Chem. Soc 1976, 98, 3478–3483. [Google Scholar]
  • (128).Trnka TM; Grubbs RH The Development of L2 × 2RuCHR Olefin Metathesis Catalysts: an Organometallic Success Story. Acc. Chem. Res 2001, 34, 18–29. [DOI] [PubMed] [Google Scholar]
  • (129).Riener M; Nicewicz DA Synthesis Of Cyclobutane Lignans via an Organic Single Electron Oxidant-Electron Relay System. Chem. Sci 2013, 4, 2625–2629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (130).Goetz AE; Boydston AJ Metal-Free Preparation of Linear and Cross-Linked Polydicyclopentadiene. J. Am. Chem. Soc 2015, 137, 7572–5. [DOI] [PubMed] [Google Scholar]
  • (131).Goetz AE; Pascual LMM; Dunford DG; Ogawa KA; Knorr DB; Boydston AJ Expanded Functionality of Polymers Prepared Using Metal-Free Ring-Opening Metathesis Polymerization. ACS Macro Lett. 2016, 5, 579–582. [DOI] [PubMed] [Google Scholar]
  • (132).Pascual LMM; Goetz AE; Roehrich AM; Boydston AJ Investigation of Tacticity and Living Characteristics of Photoredox-Mediated Metal-Free Ring-Opening Metathesis Polymerization. Macromol. Rapid Commun 2017, 38, 1600766. [DOI] [PubMed] [Google Scholar]
  • (133).Krappitz T; Jovic K; Feist F; Frisch H; Rigoglioso VP; Blinco JP; Boydston AJ; Barner-Kowollik C Hybrid Photoinduced Copolymerization of Ring-Strained and Vinyl Monomers Utilizing Metal-Free Ring-Opening Metathesis Polymerization Conditions. J. Am. Chem. Soc 2019, 141, 16605–16609. [DOI] [PubMed] [Google Scholar]
  • (134).Lu P; Boydston AJ Integration Of Metal-Free Ring-Opening Metathesis Polymerization and Organocatalyzed Ring-Opening Polymerization through a Bifunctional Initiator. Polym. Chem 2019, 10, 2975–2979. [Google Scholar]
  • (135).Kensy VK; Tritt RL; Haque FM; Murphy LM; Knorr DB Jr; Grayson SM; Boydston AJ Molecular Weight Control via Cross Metathesis in Photo-Redox Mediated Ring-Opening Metathesis Polymerization. Angew. Chem., Int. Ed 2020, 59, 9074–9079. [DOI] [PubMed] [Google Scholar]
  • (136).Yang X; Gitter SR; Roessler AG; Zimmerman PM; Boydston AJ An Ion-Pairing Approach to Stereoselective Metal-Free Ring-Opening Metathesis Polymerization. Angew. Chem., Int. Ed 2021, 60, 13952–13958. [DOI] [PubMed] [Google Scholar]
  • (137).Englman R; Jortner J The Energy Gap Law for Radiationless Transitions in Large Molecules. Mol. Phys 1970, 18, 145–164. [Google Scholar]
  • (138).Lin SH; Chang CH; Liang KK; Chang R; Shiu YJ; Zhang JM; Yang TS; Hayashi M; Hsu FC Ultrafast Dynamics and Spectroscopy of Bacterial Photosynthetic Reaction Centers. Adv. Chem. Phys 2002, 121, 1–88. [Google Scholar]
  • (139).Chen Z; Baumeister U; Tschierske C; Würthner F Effect of Core Twisting on Self-Assembly and Optical Properties of Perylene Bisimide Dyes in Solution and Columnar Liquid Crystalline Phases. Chem. - Eur. J 2007, 13, 450–465. [DOI] [PubMed] [Google Scholar]
  • (140).Martynov AG; Mack J; May AK; Nyokong T; Gorbunova YG; Tsivadze AY Methodological Survey of Simplified TD-DFT Methods for Fast and Accurate Interpretation of UV-Vis-NIR Spectra of Phthalocyanines. ACS Omega 2019, 4, 7265–7284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (141).Theriot JC; McCarthy BG; Lim CH; Miyake GM Organocatalyzed Atom Transfer Radical Polymerization: Perspectives on Catalyst Design and Performance. Macromol. Rapid Commun 2017, 38, 1700040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (142).Penfold TJ; Gindensperger E; Daniel C; Marian CM Spin-Vibronic Mechanism for Intersystem Crossing. Chem. Rev 2018, 118, 6975–7025. [DOI] [PubMed] [Google Scholar]
  • (143).Niu Y; Peng Q; Deng C; Gao X; Shuai Z Theory of Excited State Decays and Optical Spectra: Application to Polyatomic Molecules. J. Phys. Chem. A 2010, 114, 7817–7831. [DOI] [PubMed] [Google Scholar]
  • (144).Peng Q; Niu Y; Shi Q; Gao X; Shuai Z Correlation Function Formalism for Triplet Excited State Decay: Combined Spin-Orbit and Nonadiabatic Couplings. J. Chem. Theory Comput 2013, 9, 1132–1143. [DOI] [PubMed] [Google Scholar]
  • (145).El-Sayed MA Spin—Orbit Coupling and the Radiationless Processes in Nitrogen Heterocyclics. J. Chem. Phys 1963, 38, 2834–2838. [Google Scholar]
  • (146).Ma H; Peng Q; An Z; Huang W; Shuai Z Efficient and Long-Lived Room-Temperature Organic Phosphorescence: Theoretical Descriptors for Molecular Designs. J. Am. Chem. Soc 2019, 141, 1010–1015. [DOI] [PubMed] [Google Scholar]
  • (147).Peng Q; Yi Y; Shuai Z; Shao J Toward Quantitative Prediction of Molecular Fluorescence Quantum Efficiency: Role of Duschinsky Rotation. J. Am. Chem. Soc 2007, 129, 9333–9339. [DOI] [PubMed] [Google Scholar]
  • (148).Marcus RA On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. I. J. Chem. Phys 1956, 24, 966–978. [Google Scholar]
  • (149).Lohmann W Halogen-Substitution Effect on the Optical Absorption Bands of Uracil. Z. Naturforsch., C: J. Biosci 1974, 29, 493. [DOI] [PubMed] [Google Scholar]
  • (150).De Angelis F; Fantacci S; Evans N; Klein C; Zakeeruddin SM; Moser JE; Kalyanasundaram K; Bolink HJ; Gratzel M; Nazeeruddin MK Controlling Phosphorescence Color and Quantum Yields in Cationic Iridium Complexes: a Combined Experimental and Theoretical Study. Inorg. Chem 2007, 46, 5989–6001. [DOI] [PubMed] [Google Scholar]
  • (151).Baranoff E; Curchod BF; Monti F; Steimer F; Accorsi G; Tavernelli I; Rothlisberger U; Scopelliti R; Gratzel M; Nazeeruddin MK Influence of Halogen Atoms on a Homologous Series of Bis-Cyclometalated Iridium(III) Complexes. Inorg. Chem 2012, 51, 799–811. [DOI] [PubMed] [Google Scholar]
  • (152).Bedi A; Gidron O The Consequences of Twisting Nanocarbons: Lessons from Tethered Twisted Acenes. Acc. Chem. Res 2019, 52, 2482–2490. [DOI] [PubMed] [Google Scholar]
  • (153).Nagarajan K; Mallia AR; Muraleedharan K; Hariharan M Enhanced Intersystem Crossing in Core-Twisted Aromatics. Chem. Sci 2017, 8, 1776–1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (154).Senge MO Exercises in Molecular Gymnastics-Bending, Stretching and Twisting Porphyrins. Chem. Commun 2006, 243–256. [DOI] [PubMed] [Google Scholar]
  • (155).Sparks LD; Medforth CJ; Park MS; Chamberlain JR; Ondrias MR; Senge MO; Smith KM; Shelnutt JA Metal Dependence of the Nonplanar Distortion of Octaalkyltetraphenylporphyrins. J. Am. Chem. Soc 1993, 115, 581–592. [Google Scholar]
  • (156).Norton JE; Houk KN Electronic Structures and Properties of Twisted Polyacenes. J. Am. Chem. Soc 2005, 127, 4162–4163. [DOI] [PubMed] [Google Scholar]
  • (157).Haddon RC pi-Electrons in Three Dimensiona. Acc. Chem. Res 1988, 21, 243–249. [Google Scholar]
  • (158).Jug K; Hiberty PC; Shaik S Sigma-Pi Energy Separation In Modern Electronic Theory for Ground States Of Conjugated Systems. Chem. Rev 2001, 101, 1477–500. [DOI] [PubMed] [Google Scholar]
  • (159).Gentemann S; Medforth CJ; Forsyth TP; Nurco DJ; Smith KM; Fajer J; Holten D Photophysical Properties of Conformationally Distorted Metal-Free Porphyrins - Investigation into the Deactivation Mechanisms of the Lowest Excited Singlet-State. J. Am. Chem. Soc 1994, 116, 7363–7368. [Google Scholar]
  • (160).Zhang L; Wu C; Jung K; Ng YH; Boyer C An Oxygen Paradox: Catalytic Use of Oxygen in Radical Photopolymerization. Angew. Chem., Int. Ed 2019, 58, 16811–16814. [DOI] [PubMed] [Google Scholar]
  • (161).Zhou Y; Xue B; Wu C; Chen S; Liu H; Jiu T; Li Z; Zhao Y Sulfur-Substituted Perylene Diimides: Efficient Tuning Of LUMO Levels and Visible-Light Absorption via Sulfur Redox. Chem. Commun 2019, 55, 13570–13573. [DOI] [PubMed] [Google Scholar]
  • (162).Zhong A; Zhang Y; Bian Y Structures and Spectroscopic Properties of Nonperipherally and Peripherally Substituted Metal-Free Phthalocyanines: a Substitution Effect Study Based on Density Functional Theory Calculations. J. Mol. Graphics Modell 2010, 29, 470–480. [DOI] [PubMed] [Google Scholar]
  • (163).Jakubikova E; Campbell IH; Martin RL Effects of Peripheral and Axial Substitutions on Electronic Transitions of Tin Naphthalocyanines. J. Phys. Chem. A 2011, 115, 9265–9272. [DOI] [PubMed] [Google Scholar]
  • (164).Adachi M; Nagao Y Design of Near-Infrared Dyes Based on π-Conjugation System Extension 2. Theoretical Elucidation of Framework Extended Derivatives of Perylene Chromophore. Chem. Mater 2001, 13, 662–669. [Google Scholar]
  • (165).Adachi M; Murata Y Relationship between π-Conjugation Size and Electronic Absorption Spectrum: Novel π-Conjugation Size Dependence of Indoaniline Dyes. J. Phys. Chem. A 1998, 102, 841–845. [Google Scholar]
  • (166).Roznyatovskiy VV; Lee CH; Sessler JL π-Extended Isomeric and Expanded Porphyrins. Chem. Soc. Rev 2013, 42, 1921–1933. [DOI] [PubMed] [Google Scholar]
  • (167).Ruiz-Morales Y HOMO-LUMO Gap As An Index Of Molecular Size and Structure for Polycyclic Aromatic Hydrocarbons (PAHs) and Asphaltenes: A Theoretical Study. I. J. Phys. Chem. A 2002, 106, 11283–11308. [Google Scholar]
  • (168).Banfi S; Montanari F; Quici S; Barkanova SV; Kaliya OL; Kopranenkov VN; Luk’yanets EA Porphyrins and Azaporphines as Catalysts in Alkene Epoxidations with Peracetic Acid. Tetrahedron Lett. 1995, 36, 2317–2320. [Google Scholar]
  • (169).Sartor SM; Lattke YM; McCarthy BG; Miyake GM; Damrauer NH Effects of Naphthyl Connectivity on the Photophysics of Compact Organic Charge-Transfer Photoredox Catalysts. J. Phys. Chem. A 2019, 123, 4727–4736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (170).Koziar JC; Cowan DO Photochemical Heavy-Atom Effects. Acc. Chem. Res 1978, 11, 334–341. [Google Scholar]
  • (171).Lewis FD; Zuo X Activated Decay Pathways for Planar vs Twisted Singlet Phenylalkenes. J. Am. Chem. Soc 2003, 125, 8806–8813. [DOI] [PubMed] [Google Scholar]
  • (172).Huertas-Hernando D; Guinea F; Brataas A Spin-Orbit Coupling in Curved Graphene, Fullerenes, Nanotubes, and Nanotube Caps. Phys. Rev. B: Condens. Matter Mater. Phys 2006, 74, 155426. [Google Scholar]
  • (173).Nagarajan K; Mallia AR; Reddy VS; Hariharan M Access to Triplet Excited State in Core-Twisted Perylenediimide. J. Phys. Chem. C 2016, 120, 8443–8450. [Google Scholar]
  • (174).Henry BR; Siebrand W Spin-Orbit Coupling In Aromatic Hydrocarbons - Analysis of Nonradiative Transitions Between Singlet and Triplet States in Benzene and Naphthalene. J. Chem. Phys 1971, 54, 1072. [Google Scholar]
  • (175).Tait CE; Bedi A; Gidron O; Behrends J Photoexcited Triplet States of Twisted Acenes Investigated by Electron Paramagnetic Resonance. Phys. Chem. Chem. Phys 2019, 21, 21588–21595. [DOI] [PubMed] [Google Scholar]
  • (176).Schmidt K; Brovelli S; Coropceanu V; Beljonne D; Cornil J; Bazzini C; Caronna T; Tubino R; Meinardi F; Shuai Z; Brédas JL Intersystem Crossing Processes in Nonplanar Aromatic Heterocyclic Molecules. J. Phys. Chem. A 2007, 111, 10490–10499. [DOI] [PubMed] [Google Scholar]
  • (177).Bowman MK Intersystem Crossing in Photosynthetic Pigments. Chem. Phys. Lett 1977, 48, 17–21. [Google Scholar]
  • (178).Tait CD; Holten D Effects of Mg-Solvent Coordination State on Some Excited-State Processes of Bacteriochlorophyll-a. Photobioch. Photobiop 1983, 6, 201–209. [Google Scholar]
  • (179).Shchupak EE; Ivashin NV; Sagun EI Conformational Dynamics and Spin-Orbit Interactions in a Series of Sterically Hindered Porphyrins in the Lowest Triplet State. Opt. Spectrosc 2013, 115, 37–47. [Google Scholar]
  • (180).Ryan MD; Theriot JC; Lim CH; Yang H; Lockwood AG; Garrison NG; Lincoln SR; Musgrave CB; Miyake GM Solvent Effects on the Intramolecular Charge Transfer Character of N,N-Diaryl Dihydrophenazine Catalysts for Organocatalyzed Atom Transfer Radical Polymerization. J. Polym. Sci., Part A: Polym. Chem 2017, 55, 3017–3027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (181).Boyer C; Corrigan NA;Jung K; Nguyen D; Nguyen TK; Adnan NN; Oliver S; Shanmugam S; Yeow J Copper-Mediated Living Radical Polymerization (Atom Transfer Radical Polymerization and Copper(0) Mediated Polymerization): from Fundamentals to Bioapplications. Chem. Rev 2016, 116, 1803–949. [DOI] [PubMed] [Google Scholar]
  • (182).Yang J; Zhang X; Zhang X; Wang L; Feng W; Li Q Beyond the Visible: Bioinspired Infrared Adaptive Materials. Adv. Mater 2021, 33, 2004754. [DOI] [PubMed] [Google Scholar]
  • (183).Fabian J; Nakazumi H; Matsuoka M Near-Infrared Absorbing Dyes. Chem. Rev 1992, 92, 1197–1226. [Google Scholar]
  • (184).Glaser F; Kerzig C; Wenger OS Multi-Photon Excitation in Photoredox Catalysis: Concepts, Applications, Methods. Angew. Chem. Int. Ed 2020, 59, 10266–10284. [DOI] [PubMed] [Google Scholar]
  • (185).Yu Y; Guo P; Zhong JS; Yuan YF; Ye KY Merging Photochemistry with Electrochemistry in Organic Synthesis. Org. Chem. Front 2020, 7, 131–135. [Google Scholar]
  • (186).Ravetz BD; Tay NES; Joe CL; Sezen-Edmonds M; Schmidt MA; Tan Y; Janey JM; Eastgate MD; Rovis T Development of a Platform for Near-Infrared Photoredox Catalysis. ACS Cent. Sci 2020, 6, 2053–2059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (187).Allison-Logan S; Fu Q; Sun Y; Liu M; Xie J; Tang J; Qiao GG From UV to NIR: A Full-Spectrum Metal-Free Photocatalyst for Efficient Polymer Synthesis in Aqueous Conditions. Angew. Chem., Int. Ed 2020, 59, 21392–21396. [DOI] [PubMed] [Google Scholar]
  • (188).Huang Z; Noble BB; Corrigan N; Chu Y; Satoh K; Thomas DS; Hawker CJ; Moad G; Kamigaito M; Coote ML; Boyer C; Xu J Discrete and Stereospecific Oligomers Prepared by Sequential and Alternating Single Unit Monomer Insertion. J. Am. Chem. Soc 2018, 140, 13392–13406. [DOI] [PubMed] [Google Scholar]
  • (189).Olivier A; Meyer F; Raquez JM; Damman P; Dubois P Surface-Initiated Controlled Polymerization as a Convenient Method for Designing Functional Polymer Brushes: from Self-Assembled Monolayers to Patterned Surfaces. Prog. Polym. Sci 2012, 37, 157–181. [Google Scholar]
  • (190).Wang CG; Li FF; Goto A Photo-Controlled Organocatalyzed Living Radical Polymerization and its Application to Polymer Brush Synthesis on Surface. J. Photopolym. Sci. Technol 2017, 30, 379–383. [Google Scholar]
  • (191).Peterson BM; Kottisch V; Supej MJ; Fors BP On Demand Switching of Polymerization Mechanism and Monomer Selectivity with Orthogonal Stimuli. ACS Cent. Sci 2018, 4, 1228–1234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (192).Lee K; Corrigan N; Boyer C Rapid High-Resolution 3D Printing and Surface Functionalization via Type I Photoinitiated RAFT Polymerization. Angew. Chem., Int. Ed 2021, 60, 8839–8850. [DOI] [PubMed] [Google Scholar]
  • (193).Bagheri A; Fellows CM; Boyer C Reversible Deactivation Radical Polymerization: from Polymer Network Synthesis to 3D Printing. Adv. Sci 2021, 8, 2003701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (194).Shi XB; Zhang J; Corrigan N; Boyer C PET-RAFT Facilitated 3D Printable Resins with Multifunctional RAFT Agents. Mater. Chem. Front 2021, 5, 2271–2282. [Google Scholar]
  • (195).Zhang ZH; Corrigan N; Boyer C Effect of Thiocarbonylthio Compounds on Visible-Light-Mediated 3D Printing. Macromolecules 2021, 54, 1170–1182. [Google Scholar]
  • (196).Dolinski ND; Page ZA; Callaway EB; Eisenreich F; Garcia RV; Chavez R; Bothman DP; Hecht S; Zok FW; Hawker CJ Solution Mask Liquid Lithography (SMaLL) for One-Step, Multimaterial 3D Printing. Adv. Mater 2018, 30, 1800364. [DOI] [PubMed] [Google Scholar]
  • (197).Dolinski ND; Callaway EB; Sample CS; Gockowski LF; Chavez R; Page ZA; Eisenreich F; Hecht S; Valentine MT; Zok FW; Hawker CJ Tough Multimaterial Interfaces through Wavelength-Selective 3D Printing. ACS Appl. Mater. Interfaces 2021, 13, 22065–22072. [DOI] [PubMed] [Google Scholar]
  • (198).Schwartz JJ; Boydston AJ Multimaterial Actinic Spatial Control 3D and 4D Printing. Nat. Commun 2019, 10, 791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (199).Theodorou A; Liarou E; Haddleton DM; Stavrakaki IG; Skordalidis P; Whitfield R; Anastasaki A; Velonia K Protein-Polymer Bioconjugates via a Versatile Oxygen Tolerant Photoinduced Controlled Radical Polymerization Approach. Nat. Commun 2020, 11, 1486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (200).Hortensteiner S Chlorophyll Degradation during Senescence. Annu. Rev. Plant Biol 2006, 57, 55–77. [DOI] [PubMed] [Google Scholar]
  • (201).Hortensteiner S; Krautler B Chlorophyll Breakdown in Higher Plants. Biochim. Biophys. Acta, Bioenerg 2011, 1807, 977–88. [DOI] [PubMed] [Google Scholar]
  • (202).Wu C; Shanmugam S; Xu J; Zhu J; Boyer C Chlorophyll a Crude Extract: Efficient Photo-Degradable Photocatalyst for PET-RAFT Polymerization. Chem. Commun 2017, 53, 12560–12563. [DOI] [PubMed] [Google Scholar]

RESOURCES