Next Article in Journal
Enhancing the Piezoelectric Properties of 3D Printed PVDF Using Concurrent Torsional Shear Strain
Next Article in Special Issue
Enhancing Dental Material Performance: Tung Oil-Infused Polyurea Microcapsule Coatings for Self-Healing and Antimicrobial Applications
Previous Article in Journal
Antibacterial Effect and Therapy of Chronic Skin Defects Using the Composite Bioscaffold Polycaprolactone/GelitaSpon/Povidone-Iodine in Domestic Dogs
Previous Article in Special Issue
Curing of Poly(styrene-co-methyl methacrylate-co-2-hydroxyethyl methacrylate) Terpolymers in the Presence of Amino Compounds of Different Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Monocomponent Visible Light Photoinitiating Systems Based on Sulfonium Salts

Aix Marseille Univ, CNRS, ICR, UMR 7273, F-13397 Marseille, France
Polymers 2023, 15(21), 4202; https://doi.org/10.3390/polym15214202
Submission received: 21 September 2023 / Revised: 15 October 2023 / Accepted: 18 October 2023 / Published: 24 October 2023
(This article belongs to the Special Issue Polymer Coatings: Principles, Methods, and Applications)

Abstract

:
During the last decades, multicomponent photoinitiating systems have been the focus of intense research efforts, especially for the design of visible light photoinitiating systems. Although highly reactive three-component and even four-component photoinitiating systems have been designed, the complexity to elaborate such mixtures has incited researchers to design monocomponent Type II photoinitiators. Using this approach, the photosensitizer and the radical/cation generator can be combined within a unique molecule, greatly simplifying the elaboration of the photocurable resins. In this field, sulfonium salts are remarkable photoinitiators but these structures lack absorption in the visible range. Over the years, various structural modifications have been carried out in order to redshift their absorptions in the visible region. In this work, an overview of the different sulfonium salts activable under visible light and reported to date is proposed.

1. Introduction

In recent years, photopolymerization has been the subject of intense research efforts with regards to the numerous applications making use of this polymerization technique [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15]. Among them, applications such as solvent-free paints, adhesives, coatings and varnishes, dental restoration materials, 3D and 4D printing or microelectronics can be cited as relevant examples [9,10,16,17,18,19,20,21,22,23,24,25,26]. With the aim of designing efficient photoinitiating systems, multicomponent systems have been extensively investigated, especially for the design of visible light photoinitiating systems. Indeed, although historically photopolymerization was carried out in the UV range, this approach is nowadays more and more contested due to safety concerns related to UV light (skin cancers and eye damages) [27,28]. Parallel to this, energy sobriety imposed by numerous countries is an important parameter to consider and UV irradiation setups are, most of the time, energy-consuming devices which do not fit with the current energy efficiency action plan proposed by numerous countries. To date, since 2020, Hg lamps cannot be commercialized anymore, resulting from the treaty adopted in 2013 by numerous countries and resulting from the United Nations Environmental Program (UNEP) Minamata Convention on Mercury. Disinterest for UV photoinitiating systems is also related to the higher light penetration in resins that is achievable upon irradiation in the visible range. As shown in Figure 1, it can vary from a few millimeters up to a few centimeters depending on the irradiation wavelength. If this approach is of interest, photons of the visible spectrum exhibit lower energy than UV photons so that visible light photoinitiating systems of higher reactivity should be designed in order to compensate for this reduction of energy. To address this issue, a wide range of structures have been explored as exemplified by camphorquinone [29,30], thiophenes [31], push-pull dyes [2,32,33], anthracenes [34], acridine-1,8-diones [35], benzylidene ketones [36,37,38,39,40,41], chromones and flavones [42,43,44], perylenes [45], diketopyrrolopyrroles [46,47], naphthalimides [48,49,50,51,52,53,54], cyclohexanones [55,56,57,58], chalcones [24,59,60,61,62,63,64], dithienophosphole derivatives [65], thioxanthones [66,67,68,69,70,71,72,73,74,75,76,77,78,79,80], benzophenones [81,82,83,84], cyanines [85], bodipy [66,86,87,88,89,90], NIR dyes [91], anthraquinones [92], curcumin [93,94,95,96], furan derivatives [97], pyrenes [98,99,100,101], iodonium salts [66,102,103,104,105,106,107], carbazoles [108,109,110,111,112,113,114], quinoxalines [115,116,117,118,119,120,121,122,123,124,125,126,127,128], glyoxylates [129], helicenes [130], photochroms [131] and naphthoquinones [132]. Metal complexes were also investigated such as copper complexes [133,134,135], zinc complexes [136], gold complexes [137], iridium complexes [138,139,140] or iron complexes [141,142].
Interestingly, highly efficient photoinitiating systems could be designed with these different chromophores using multicomponent systems. Indeed, as shown in Scheme 1, Type II photoinitiators can interact with hydrogen donors or electron donors/acceptors to generate reactive species. Over the years, reactivity of Type II photoinitiators could be improved by designing multicomponent systems, and four-component photoinitiating systems were even reported in the literature [144]. Considering the complexity of elaborating such photocurable resins, the design of monocomponent systems and especially of Type I photoinitiators is nowadays more and more studied. Among these structures, benzylketals, o-acyl-α-oximino ketones, trichloromethyl-S-triazine glyoxylates, acyloximino esters, hydroxyacetophenones, α-haloacetophenones, hexaaryl biimidazoles (HABIs), benzoin derivatives, oxime esters, phosphine oxides or α-aminoalkylacetophenones were examined for their photoinitiating abilities [145,146,147]. As for specificities, the working principle of Type I photoinitiators is based on the selective cleavage of a specific bond that can homolytically cleave upon photoexcitation. But monocomponent systems can also be developed with Type II photoinitiators. This strategy was notably successfully when applied to the design of monocomponent systems based on diaryliodonium salts [66,148,149,150]. By combining the photosensitizer and the iodonium salt within a unique molecule, efficient photoinitiating systems could be developed. In the field of onium salts, diaryliodonium salts were not the only structures studied for the design of monocomponent systems, and sulfonium salts were also identified as promising candidates [151,152].
Historically, diaryliodonium salts and sulfonium salts were studied starting from the 1970s by Crivello and coworkers for the cationic polymerization of epoxides [103,153,154,155,156,157]. Upon light irradiation, triarylsulfonium salts can release Brönsted acids. However, photosensitivity of triarylsulfonium is centred in the far UV region, rendering their photoactivation with visible light impossible. Sulfonium salts also suffer from a low solubility in numerous monomers, rendering these salts less attractive than iodonium salts for industrial applications [158]. Photoacid generation also strongly depends on the counter-ions used to prepare the salts and the following order of reactivity (SbF6 > AsF6 > PF6 ≫ BF4) could be determined [159]. Recently, new counter-anions were examined such as the tetrakis(perfluoro-tert-butyloxy)aluminate anion or the tris((trifluoromethyl) sulfonyl)methanide anion that makes sulfonium salts more reactive than those prepared with hexafluoroantimonate (SbF6) [160]. By using anions such as 2,2-difluoro-2-sulfoethyl tricyclo[3.3.1.13,7]decane-1-carboxylate or 1,1,2,2,3,3,4,4,4-nonafluoro-1-butanesulfonate anions, hydrophobic sulfonium salts could be prepared [161]. Sulfonium salts bearing perfluorinated tetraphenylborates were also designed as photoinitiators for 3D printing experiments performed at high temperatures which are also named hot lithography [162]. The general initiation mechanism proposed for the cationic polymerization (CP) of epoxides upon irradiation is depicted in Scheme 2 [154,163,164,165]. Upon photoexcitation, an aryl radical and a radical cation can form. In the presence of a hydrogen donor, Bronsted acid can be generated, enabling the promotion of the ring-opening polymerization of epoxides.
Noticeably, sulfonium salts are versatile structures since protons and free radicals can be both formed, enabling the initiation of cationic and free radical polymerization processes [166,167]. Sulfonium salts are also thermally stable structures, facilitating their storage. However, in 1995, an allylic sulfonium salt was reported by Yagci and coworkers as exhibiting a dual thermal and photochemical curing mode [168,169]. But, this dual initiating ability remains limited to a few structures [170]. A few sulfonium salts by the convenient substitution of their scaffolds could also act as latent thermal initiators and the polymerization of glycidyl phenyl ether could proceed efficiently at more than 80 °C and not at less than 60 °C [171]. Starting from the 1980s, numerous sulfonium salts were reported as efficient thermal initiators for the cationic polymerization of epoxides, but these structures were inefficient as photochemical initiators [172,173]. Recently, Allonas and coworkers developed a two-component system comprising a triarylsulfonium salt acting as a photoinitiator and a phenyldialkylsulfonium salt acting as a thermal initiator. By this unique combination, a photochemically induced thermal polymerization of epoxides could be initiated [159]. By the heat released by the photoinduced cationic polymerization of epoxides, thermal decomposition of the phenyldialkylsulfonium salts could be induced so that carbon-fiber-reinforced polymers with a concentration of carbon fibers up to 50 wt% could be prepared upon irradiation at 365 nm. If triarylsulfonium salts were popular photoinitiators, trialkylsulfonium salts were also studied, as exemplified with dimethylphenacylsulfonium salts bearing various borate anions [174]. As a drawback, due to the lack of aromatic rings and therefore lack of chromophores on the sulfonium salts, visible light photoinitiating systems could only be developed by mean of the sensitization by various styryl dyes. Using the sensitization approach, photopolymerization experiments could be performed at 488, 514 and even 647 nm. Still, based on phenacyl sulfonium salts, Yagci and coworkers developed a diphenylphenacyl sulfonium salt exhibiting a dual photoinitiating ability for free radical and cationic polymerizations. Despite the presence of the diphenyl system, the absorption remained centered in the UV range, below 350 nm [175].
Over the years, two factors have been identified as being crucial in order to improve photoacid generation of sulfonium salts. First, the excited-state energy of the chromophore should be sufficiently high to initiate the photocleavage [176]. Indeed, the rate of the C–S bond cleavage was determined as being proportional to the energy gap existing between the LUMO (σ* orbital) of the sulfonium group and the LUMO + 1 (π* orbital) of the chromophore in the excited state [177]. Second, the radical cations and the neutral radicals formed after photolysis should be stable enough to avoid radical recombination [178]. Considering that benchmark sulfonium salts naturally absorb in the UV range [179], visible light photopolymerization can be initiated only if photosensitizers are used [180,181]. However, efficacy of the photoinduced electron transfer relies on a bimolecular process that is strongly influenced by the viscosity of the resin. In this context and to simplify the composition of resins, the design of monocomponent systems in which the photosensitizer can be directly connected to the sulfonium group was examined, offering a unique opportunity to design monocomponent systems. In this review, an overview of the different sulfonium salts activable under visible light is provided.

2. Sulfonium Salts Activable under Visible Light

2.1. Advantages, Disadvantages and Limitations

The design of monocomponent photoinitiating systems based on sulfonium salts and absorbing in the visible range is an interesting approach. Using this approach, an intramolecular sensitization process can occur. In addition, some advantages and limitations can be mentioned concerning the covalent linkage of a chromophore to a sulfonium salt. Notably, to redshift the absorption of salts exhibiting a UV-centered absorption is hard work. Nowadays, Type II photoinitiating systems that are proposed in the literature exhibit chromophores with structures totally disconnected from these historical UV photoinitiators (see the list of chromophores of innovative structures reported during the last 15 years). As a consequence of this, sulfonium salts with visible light absorption are costly and obtained in multistep syntheses. The benefits of intramolecular sensitization vs. the traditional intermolecular sensitization are sometimes low. A series of pros and cons for intramolecular vs. intermolecular sensitization is presented in Table 1.

2.2. 1,3,5-Triphenyl-2-Pyrazoline Derivatives

With the aim of providing photosensitivity in the visible region, a series of dimethylsulfonium salts derived from 1,3,5-triphenyl-2-pyrazoline derivatives were proposed in 2021 by Jin and coworkers as one-photon and two-photon photoinitiators (See Figure 2) [182]. The choice of 1,3,5-triphenyl-2-pyrazoline as the chromophore was notably motivated by the two-photon capability of this group. Indeed, in 2014, a two-photon absorption (TPA) cross-section of 280 GM was notably reported by Mysliwiec and coworkers for a 1,3,5-triphenyl-2-pyrazoline derivative, constituting an excellent basis for the design of TPA photoinitiators [183]. Due to the intramolecular charge transfer existing between the nitrogen atom N1 and the carbon C3 of the pyrazoline ring (see Figure 2 for numbering), 1,3,5-triphenyl-2-pyrazolines also efficiently absorb in the visible range [184], and the presence of a tertiary nitrogen atom can also contribute to stabilizing the radical cations, improving the quantum yield of photoacid generation.
Depending on the substituent, absorption maxima varying between 344 nm for PI-EtO, 347 nm for PI-H and 372 nm for PI-CF3 could be determined in acetonitrile. The most redshifted absorption was found for PI-CF3 bearing a strong electron-donating group (see Figure 3).
Theoretical calculations revealed the S0–S1 and the S0–S2 transitions correspond to a π–σ* and a π–π* transition, respectively. Photolysis of PI-H performed upon irradiation at 365 nm with a LED revealed the photolysis process to be very fast. In particular, the reformation of the PI-H precursor, i.e., pre-H, was evidenced. However, it only corresponded to 20% of the photolysis products and the formation of numerous byproducts formed during photolysis were also detected. Indeed, the methyl radicals thus formed can attack the phenyl rings of pyrazolines, producing different byproducts (see Figure 4). This undesired reaction could be efficiently avoided in a solution by adding an excess of triethylamine [185].
Investigation of the quantum yields of photoacid generation at 365, 385, 395, 405 and 425 nm revealed the quantum yields decrease by irradiating at long wavelengths. In the case of the best photoacid generator, i.e., PI-EtO, the quantum yield decreased from 0.65 at 365 nm to only 0.20 at 425 nm. If PI-CF3 obtained the lowest quantum yield at 365 nm (ϕH+ = 0.60), this compound could still maintain a quantum yield of 0.31 at 425 nm due to its redshifted absorption (See Table 2).
Considering that the efficiency of photoacid generation is strongly dependent on the energy gap between the lowest unoccupied molecular orbital (LUMO) level of the σ* orbital of the sulfonium salts and the LUMO + 1 level of the π* orbital of pyrazolines, these energy values were thus determined. If the LUMO level of sulfonium salts remained almost unchanged by modifying the substituent, conversely, the substitution pattern of the pyrazoline moiety drastically impacted the LUMO + 1 energy levels. Thus, ΔE(LUMO-LUMO + 1) varying between 0.54 eV for PI-EtO, 0.49 eV for PI-H and 0.25 eV for PI-CF3 were calculated, therefore perfectly fitting the quantum yields of photoacid generation. By cyclic voltammetry, the free energy change (ΔGet) could be determined using the Rehm–Weller equation [177,186,187,188,189,190]. Values varying between −21.73 kJ/mol for PI-CF3 and −55.82 kJ/mol for PI-EtO could be obtained. In particular, calculations revealed the electron transfer occurs from the singlet excited state, from the pyrazoline to the C–S σ* orbital, negative values of free energy changes being only obtained by considering the singlet excited state in the calculation and not the triplet state [178]. Photoinitiating abilities of the different sulfonium salts were examined during the cationic polymerization (CP) of (3,4-epoxycyclohexane)methyl-3,4-epoxycyclohexyl-carboxylate (EPOX) upon irradiation at different wavelengths with LEDs. As shown in Table 3, the order of reactivity determined for the quantum yields of photoacid generation was confirmed during the CP of EPOX, with PI-EtO outperforming the other sulfonium salts. Only at 425 nm, PI-CF3 could furnish a higher monomer conversion than PI-EtO and this is directly related to its redshifted absorption compared with that of PI-EtO. Thus, an EPOX conversion of 51.87% could be obtained with PI-CF3, higher than that obtained with PI-EtO (49.57%). Comparisons of the monomer conversion obtained with the benchmark photoinitiator PI-6992M and the different pyrazoline derivatives revealed PI-6992M to provide a higher conversion at 365 nm than the newly developed photoinitiators due to its higher molar extinction coefficient at this wavelength. At 385 nm, an opposite situation was found, resulting from its low molar extinction coefficient. Finally, in the visible range, no monomer conversion could be obtained with PI-6992M whereas the pyrazoline derivatives were still active, even at 425 nm.
Considering that methyl radicals are formed during photolysis, the free radical polymerization (FRP) of trimethylolpropane triacrylate (TMPTA) was also examined and comparisons of the photoinitiating abilities of PI-H and PI-CF3 were compared with that of OXE-02. As shown in Table 4, PI-CF3 and PI-H furnished lower monomer conversions at 365 and 385 nm than OXE-02, which exhibits better light absorption properties at these two wavelengths. However, at 395 and 405 nm, PI-CF3 and PI-H could outperform OXE-02. PI-H proved to be the best candidate for the FRP of TMPTA, the monomer conversion reaching 50.9% at 365 nm.
Finally, two pyrazoline derivatives were investigated as photoinitiators for two-photon polymerization. Two-photon absorption cross-sections of 105 and 75 GM were respectively determined for PI-CF3 and PI-H. A low threshold power was also determined during the two-photon microfabrication of structures under air using pentaerythritol triacrylate (PETA) as the monomer. During the FRP of PETA at 780 nm under air, a threshold value as low as 4.5 mW at 100–150 µm/s was obtained, evidencing the remarkable photoreactivity of these sulfonium salts.

2.3. Phenothiazine Derivatives

Phenothiazine is an interesting scaffold for the design of visible light photoinitiators. Indeed, phenothiazine is a polyaromatic structure absorbing in the visible range, even without substituents, so that various Type I and Type II photoinitiators were designed and synthesized with this structure [191,192,193,194,195,196,197]. This is also an excellent electron donor that was extensively used for the design of push–pull dyes developed for different applications in organic electronics such as energy conversion [198,199,200,201,202,203,204,205,206,207,208,209,210], light-emitting materials for organic light-emitting diodes (OLEDs) [211,212,213,214,215,216,217,218,219,220,221,222,223] or semi-conductors for organic field effect transistors (OFETs) [199,224,225,226,227,228,229]. Phenothiazines were also developed as fluorescent probes for the detection of various molecules and ions in organic and aqueous solutions, food, live cells and animals [230,231,232,233,234,235,236,237,238,239,240]. In 2021, Yagci and coworkers proposed a phenacyl phenothiazinium salt (P-PTh) exhibiting a panchromatic absorption (see Figure 5) [241]. Indeed, as shown in Figure 6, the absorption spectrum of P-PTh in acetonitrile extends between 350 and 900 nm, with an absorption maximum at 514 nm.
Photolysis experiments performed in the solution and in the presence of an acid indicator revealed the release of Bronsted acid. By electron spin resonance (ESR), the formation of radicals resulting from the homolytic cleavage of the C–S bond was demonstrated, leading to the formation of phenacyl radicals and phenothiazine radical cations. In the presence of a hydrogen donor, Bronsted acid can be formed according to the mechanism depicted in Scheme 3. However, based on these experiments, the concomitant occurrence of a homolytic and heterolytic cleavage of the sulfonium salt cannot be ruled out. Laser flash photolysis experiments and additional spectroscopic investigations evidenced the triplet excited state of phenothiazine to play a key role in the mechanism of fragmentation [242].
Examination of the thermal stability of the phenacyl phenothiazinium salt (P-PTh) by TGA measurements revealed the decomposition temperature to be of 170 °C. P-PTh could also act as a thermal initiator, and onset polymerization temperatures of 100 °C in TMPTA and 160 °C in EPOX were respectively determined by DSC. However, upon storage of the resins at a moderate temperature, i.e., 60 °C in the dark for 100 days, a remarkable thermal stability could be determined, which constitutes an interesting feature for practical applications in industries. Polymerization tests performed at 365 and 405 nm revealed the EPOX conversions to be higher than 90% after 2100 s of irradiation with LEDs. A conversion of 60% could also be obtained at 850 nm, evidencing the versatile photoinitiating ability of P-PTh. Investigation of the FRP of TMPTA at 365 and 405 nm (30 mW/cm2) also furnished high monomer conversions, approximately 40% after 600 s of irradiation. This conversion could be greatly improved by using a broad-spectrum light source (260 mW/cm2) and a conversion of approximately 95% could be obtained for TMPTA. Due to the concomitant formation of radicals and cations, the polymerization of a bifunctional monomer, namely 7-oxabicyclo [4.1.0] heptan-3-yl) methyl methacrylate (TTA-15) was examined both at 365 and 405 nm. As shown in Figure 7, hybrid polymerization of TTA-15 furnished a rapid conversion of both functions, evidencing the FRP process to be independent from the CP.

2.4. Triphenylamine-Based Sulfonium Salts

Triphenylamine is an interesting scaffold for the design of visible light photoinitiators due to its excellent light absorption properties in the near-UV/visible range [243,244]. Dyes of high molar extinction coefficients can also be designed with this group [245,246,247,248]. The first triphenylamine-based sulfonium salt used as a two-photon-activatable photoacid generator BSB-S2 was reported as recent as 2002 by Marder and coworkers (see Figure 8) [249].
Interestingly, BSB-S2 exhibited a large two-photon absorption cross-section of 690 GM and a quantum yield of photoacid generation of 0.5. The choice of triphenylamine as the chromophore was also justified by the fact that triaryl radical cations are relatively stable and can favor the C–S bond cleavage. Protonated forms of triarylamines are also strong acid (pKa < −5) so that the amino group cannot inhibit the cationic polymerization of epoxides. To end, the radical cation formed upon homolysis of the carbon-sulfur bond and centered on the triphenylamine moiety can be efficiently stabilized by mesomeric forms, as shown in Scheme 4.
From the absorption viewpoint, the absorption of BSB-S2 was broad, extending between 250 and 450 nm. Two-photon microstructures could be obtained by polymerizing epoxides at 745 nm. Due to the presence of the triphenylamine moieties in BSB-S2, a low polymerization threshold power of 2.4 mW at 710 nm was determined during the CP of the Araldite CY179MA epoxide resin (composed of a mixture of 7-oxabicyclo[4.1.0]heptane-3-carboxylic acid and 7-oxabicyclo[4.1.0]hept-3-ylmethyl ester). This value is greatly lower than that determined for benchmark photoinitiators such as [4-[(2-hydroxytetradecyl) oxy]phenyl]phenyliodonium hexafluoroantimonate (CD1012), isopropylthioxanthone (ITX), diphenyliodonium 9,10-dimethoxyanthracene sulfonate (DPI-DMAS) or triphenylsulfonium hexafluoroantimonate (TPS) (see Figure 9 and Table 5) [250]. BSB-S2 was also an excellent candidate for a chemically amplified resist system designed for two-photon lithography [251].
In the same year, the same authors examined a series of sulfonium salts (SULF5-SULF9) exhibiting near-UV/visible absorption in order to investigate their photophysical properties (see Figure 8) [178]. Interestingly, by varying the counter-anion (CF3SO3, PF6, SbF6) no modification of the quantum yields of photoacid generation was detected and a value of ca. 0.5 was determined for all salts. This is directly related to the fact that the different sulfonium salts exhibit identical absorption spectra due to the presence of the same chromophore. Photolysis experiments of the different sulfonium salts carried out at 294 nm enabled the identification of several photolysis products. Among them, formation of 3-methylthiotriphenylamine (52% of the sample), methyl-substituted 3-methylthiotriphenylamine (10% of the sample) and two products comprising a carbazole moiety and resulting from an intramolecular coupling (18% of the sample) could be identified as the four main products of the photolysis. Compared with triarylsulfonium salts (ϕH+ = 0.53), photoacid generation of SULF-5-SULF9 was not adversely affected by the presence of the triphenylamine moiety, and quantum yields of photoacid generation of 0.48, 0.48, 0.47 and 0.41 were determined for SULF5 and SULF7-SULF9, respectively, using the method reported by Scaiano and coworkers [252]. These results are better than those obtained for the phenyl dimethylsulfonium salt for which no photoacid generation was detected in the same conditions [177]. In this case, a competition between the C–S bond cleavage and the radiative decay from the singlet excited state was suggested as the main reason impeding photoacid generation with this sulfonium salt. However, these photoacid quantum yields are comparable to those obtained for a polyaromatic structure such as the anthracenyl sulfonium salt SULF11 (see Figure 8). In this case, the rate of photocleavage of the C–S bond is higher than that of the fluorescence decay, enabling efficient photoacid generation [176]. Determination of the free energy change (ΔGet) using the Rehm–Weller equation revealed the photoinduced electron transfer from the aromatic amine to the low-lying C–S σ* orbital occurs from the first excited singlet state, for all investigated structures, as previously evidenced by Yagci and coworkers for a phenothiazine-substituted sulfonium salt [241]. As observed for the triphenylamine derivatives, the anthracenyl sulfonium salt SULF11 benefits from the electron-donating ability of the polyaromatic group, activating the dissociation of the S–C bond by stabilizing the resulting anthracenyl methyl sulfide radical cation formed by homolytic cleavage. If the choice of the counter-anion was without influence for photoacid generation, a different trend was found during the photopolymerization experiments. Indeed, polymerization efficiencies of the different sulfonium salts were strongly dependent on the counter-ion. As previously shown in the literature, nucleophilicity of the anion can strongly impact the strength of the ion pair formed by the anion and the cation so that the reactivity of the cation in photopolymerization can be affected. In previous works, the following scale of reactivity depending on the counter-anions used was established for iodonium salts: BF4 < PF6 < AsF6 < SbF6 < B(C6F5)4 < Ga(C6F5)4 [253]. In the present case, the role of the anion on the polymerization efficiency of the sulfonium salts was clearly evidenced and determined as being comparable to that observed for iodonium salts. Thus, if no polymerization of cyclohexene oxide (CHO) could be detected with SULF5 even after one hour of irradiation at 300 nm, 90% conversions could be obtained within 130 s with SULF8 and SULF9. In addition, a higher polymerization rate was determined for SULF9. To support these experimental results, the benzyl-substituted sulfonium salt SULF9 was determined as undergoing a heterolytic bond cleavage of the benzylic bond in the excited state in addition to the electron-transfer-mediated photoacid generation observed for SULF8, yielding additional benzylic cations that constitute another source of cations and thus can initiate the CP of CHO more efficiently [254].
In 2002, BSB-S2 was reported as the first highly efficient photoacid generator activable in the near-infrared range and based on a symmetrical D–π–D structure. A two-photon sensitivity of 345 GM was measured at 710 nm, constituting a remarkable result. However, numerous works have also demonstrated that the extension of the π-conjugation of the chromophore could help to increase the two-photon sensitivity but also adversely affect the quantum yield of acid generation. This point was notably extensively studied by Belfield and coworkers [177,255,256,257]. In 2013, Malval and coworkers developed a series of asymmetrically substituted sulfonium salts PAG3-PAG6 on the basis of a D–π–A structure (see Figure 10) [258]. The different structures could act as efficient photoacid generators by one-photon excitation at 405 nm or by two-photon excitation at 800 nm. In order to tune the photoreactivity of the sulfonium salts, the influence of the para-to-meta substitution was examined. A comparison between dimethylsulfonium salts and methyl benzyl sulfonium salts was also carried out, previous works evidencing that benzyl cations could participate in the polymerization process.
From the absorption viewpoint, the presence of the triphenylamine group acting as an electron donor contributed to the redshift of the absorptions of PAG3-PAG6 compared with that of PAG1 and PAG2 bearing a weaker electron donor, i.e., an ethoxy group. Thus, if absorption maxima located at 346 and 324 nm were determined for PAG1 and PAG2, these values increased to 395 and 400 nm for PAG3 and PAG5, respectively. A redshift of the absorption by ca 20 nm was also determined for all the para-substituted sulfonium salts compared with the meta-substituted ones, resulting from a better conjugation between the donor and the acceptor (See Table 6). Absorption spectra of all asymmetric sulfonium salts were characterized by two absorption bands corresponding to aryl π–π* transitions at high energy and a π–σ* transition at lower energy. Examination of their photoacid generation ability at 405 nm revealed all triphenylamine-substituted sulfonium salts to be better photoacid generators than the ethoxy-substituted salts, except PAG3. Examination of the photoluminescence properties of PAG1-PAG6 revealed PAG3 to exhibit a photoluminescence quantum yield of 0.16 whereas values lower than 0.01 were found for all the other sulfonium salts. It was thus concluded that the excellent photoacid generation abilities of PAG4-PAG6 could be assigned to an efficient electron transfer from the triphenylamino group towards the sulfonium salt, resulting in an efficient quenching of luminescence. On the other hand, for PAG3, photoluminescence adversely competes with the photoinduced electron transfer, resulting in a low photoacid generation ability. Comparison of the photoacid generation ability of PAG3 and PAG4 with that of PAG5 and PAG6 revealed that PAG5 and PAG6 exhibit higher quantum yields of photoacid generation, attributable to the presence of better leaving groups, namely benzyl groups compared with the methyl groups. Remarkably, all meta-substituted photoacids proved to be better photoacids than their para-substituted analogues, due to their blue-shifted absorptions compared with that of their para-substituted analogues. This trend is consistent with previous works demonstrating that a redshift of the absorption of photoacids was accompanied by a reduction of photoacid generation ability. Benefits of the para-to-meta substitution in the photoacid generation of sulfonium salts was notably demonstrated with PAG1 and PAG2 [259]. In this case, an increase in the photoacid generation quantum yield by a factor 2.4 was determined. Among the most interesting findings, PAG3-PAG6 showed larger two-photon absorption cross-sections than PAG1 and PAG2. If values of 73 and 68 GM were determined for PAG1 and PAG2, respectively, these values increased up to 650 and 680 GM for PAG4 and PAG5.
The interest of these sulfonium salts was demonstrated during the cationic polymerization (CP) of (3,4-epoxycyclohexane)methyl-3,4-epoxycyclohexylcarboxylate (EPOX), and a diepoxide resin, i.e., SU-8 at 405 and 800 nm [185]. As shown in Figure 11, photopolymerization experiments performed at 405 nm (I = 5 mW/cm2) during the CP of EPOX confirmed the higher photoacid generation ability of the meta-substituted sulfonium salts compared with the para-substituted ones. Conversions of 58 and 52% could be respectively obtained with the meta-substituted salts PAG6 and PAG4 (1 wt%). Conversely, the para-substituted PAG3 furnished the lowest EPOX conversion, approximately 35% after 300 s of irradiation. These conversions could be greatly improved by using CHO as the monomer. In this case, conversions higher than 90% could be obtained with all photoacid generators except for PAG4, for which a conversion of 70% was determined.
Photopolymerization of the diepoxide resin SU-8 at 800 nm confirmed the trend established at 405 nm. Thus, high two-photon polymerization thresholds of 25 and 22 µJ were determined for PAG3 and PAG5, respectively, whereas these values decreased to only 10 µJ for the more reactive meta-substituted PAG4 and PAG6.
Finally, the high reactivity of PAG5 and PAG6 was demonstrated during the photopolymerization experiments performed under sunlight. Using CHO as the monomer, epoxide conversions higher than 80% could be determined after 20 min of sunlight exposure (I = 1.2 mW/cm2). However, it has to be noticed that this remarkable polymerization efficiency could only be obtained while increasing the photoinitiator content to 12 wt% and not 1 wt% with the artificial light sources.
In the previous examples, only mono-sulfonium salts have been designed and synthesized. An extension of the former strategy was proposed in 2016 with the design of bifunctional sulfonium salts [260]. A comparison between mono and bifunctional salts could thus be established (see Figure 12).
This strategy was pertinent since the di-substituted sulfonium salts showed photoacid generation abilities twice as high as that of their mono-substituted counterparts as well as improved two-photon absorption cross-sections. From the absorption viewpoint, the introduction of sulfonium groups redshifted the absorption of Mono-Para/Mono-Meta, Bi-Para/Bi-Meta compared with their parent structures Pre-Para/Pre-Meta (see Figure 13 and Table 7). Noticeably, this redshift was more important for the para-substituted sulfonium salts, attributable to a better conjugation of the sulfonium group with the rest of the structure. Benefits of the multi-branch approach were evidenced during the two-photon absorption measurements. Indeed, higher two-photon absorption cross-section (δ) values could be determined for the bifunctional PAGs compared with the monofunctional ones. Notably, the two-photon absorption cross-section δ value increased from 234 GM for Mono-Meta up to 745 GM for Bi-Meta. As anticipated during the design of these bifunctional sulfonium salts, the best photoacid generation efficiency was determined for Bi-Para and Bi-Meta, these sulfonium salts produce twice more Bronsted acids than the mono-substituted ones. Influence of the solvent used for photolysis was also demonstrated. Thus, higher photoacid generation efficiencies were determined in dichloromethane than in the more polar solvent, i.e., acetonitrile. Using triethyleneglycol divinyl ether (DVE-3) as the monomer, conversions higher than 90% could be determined with the four sulfonium salts upon irradiation at 405 nm. In this series of sulfonium salts, all the meta-substituted PAGs exhibited faster polymerization rates than the para-substituted ones. However, similar monomer conversions could be determined after 10 min of irradiation (see Figure 14).
This difference of reactivity was confirmed during the two-photon polymerization experiments performed at 780 nm using SU-8 as the resin.
In 2015, Malval and coworkers examined another strategy to extend the π-conjugation, and a series of oligomeric sulfonium salts differing by the number of phenyl rings used as the spacer between the triphenylamine and the sulfonium moieties were designed and synthesized (see Figure 15) [261]. 4-Cyanobenzyl groups were also introduced on the sulfonium moiety as photocleavable groups.
Due to the presence of consecutive aromatic rings between the donor and the acceptor, the electronic delocalization was strongly limited so that the different sulfonium salts exhibited a strongly UV-centered absorption. If a significant redshift of the absorption maximum was found between Mono-Ben and Bi-Ben (292 and 344 nm, respectively), almost no additional redshift was detected for Tri-Ben and Tetra-Ben (349 and 351 nm, respectively), evidencing that the biarylic group was sufficient to interrupt the π-conjugation between the triphenylamine and the sulfonium moieties. An absorption extending up to 420 nm could be determined for Bi-Ben, Tri-Ben and Tetra-Ben, enabling the photopolymerization experiments to be performed under visible light (see Figure 16).
Photoacid generation was examined at 365 nm in acetonitrile. The mechanism of photoacid generation is depicted in Scheme 5.
In this series, photoacid generation ability increased from Mono-Ben to Bi-Ben and Tri-Ben (ϕH+ = 0.60, 0.69 and 0.73, respectively) and then decreased for Tetra-Ben (ϕH+ = 0.42). In the case of Mono-Ben, a higher photoacid generation ability than that of SULF9 (previously reported by Marder and coworkers) was demonstrated (ϕH+ = 0.60 vs. 0.41 for SULF9), resulting from the substitution of the benzylic group by a cyano group in Mono-Ben [178]. This was assigned to the better leaving propensity of the 4-cyanobenzyl group, facilitating Bronsted acid generation. In the case of Tetra-Ben, a reduction of the photoacid generation ability was determined (ϕH+ = 0.42). Furthermore, it remains greatly higher than that reported for anthracenyl and naphthacenyl sulfonium salts [176]. In terms of PAG design, it was thus concluded that the introduction of cyanobenzyl groups and benzene oligomers was beneficial for improving the photoacid generation ability. An optimal structure was thus identified with Tri-Ben. The superiority of Tri-Ben over the other PAGs was confirmed during the CP of EPOX and the FRP of TMPTA using 1 wt% photoinitiators (see Figure 17). Upon irradiation at 365 nm (I = 15 mW/cm2), conversions of 90, 40 and 30% were, respectively, determined for tri(ethylene glycol)divinyl ether (DVE-3), TMPTA and EPOX used as the monomers. Excellent results were also obtained during the CP of SU-8.
Examination of the thermal properties revealed the four derivatives to exhibit decomposition temperatures higher than 150 °C, which is sufficient for numerous practical applications. The different PAGs could also be used as two-photon initiators. Indeed, two-photon absorption cross-sections of 184, 230, 474 and 393 GM were determined at 700 nm. The highest two-photon absorption cross-section was determined for Tri-Ben which is therefore a versatile photoinitiator as it can efficiently initiate one-photon and two-photon polymerizations. By laser-scanning lithography (LSL) at 780 nm, the SU-8 resin containing 1 wt% Tri-Ben could efficiently polymerize and only near the focus of the laser beam. The 3D patterns exhibiting excellent spatial resolution could be obtained.
Following this work, Malval and coworkers examined a series of para-substituted derivatives in which the position of the double bond as well as the number of consecutive aromatic rings were modified (see Figure 18) [262].
Compared with the previous meta-substituted series, a redshift of the absorption spectra was determined for all derivatives, consistent with a better conjugation between the donor and the acceptor. For the most conjugated structure, namely PAG-S, an absorption maximum peaking at 400 nm was determined in acetonitrile. Upon the addition of a biarylic system, a blueshift of the absorption maximum was determined for PAG-PS and PAG-SP, the highest blueshift being determined for PAG-PS (380 nm vs. 290 nm for PAG-SP). In the case of PAG-P, an absorption located at 360 nm was detected, resulting from the presence of the biarylic systems (see Figure 19). Logically, the redshift of the absorptions was accompanied by a reduction of the photoacid generation quantum yields. Thus, quantum yields of 0.58, 0.50, 0.48 and 0.32 were, respectively, determined for PAG-P, PAG-PS, PAG-SP and PAG-S, consistent with the trend previously determined for other sulfonium salts.
Photolysis of the sulfonium salts resulted in the generation of byproducts that often absorb at similar wavelengths than the parent sulfonium salt. Malval and coworkers advantageously used the absorption of these byproducts to induce the decomposition of an iodonium salt, namely diphenyliodonium hexafluorophosphate (Iod) (see Figure 20) [167]. In this aim, PI-PAG was designed and synthesized.
Benefiting from an extended π-conjugated system but also from a biarylic system, an interesting absorption extending up to 450 nm could be obtained. A slight redshift of the absorption was obtained by converting the methylsulfanyl group in the precursor pre-H as a sulfonium group in PI-PAG (381 nm for PI-PAG vs. 374 nm for pre-H) (see Figure 21).
Photopolymerization experiments could thus be carried out at 365, 385, 405 and 425 nm. Upon irradiation at 385 nm, a quantum yield of photoacid generation of 0.44 could be determined for PI-PAG. Polymerization tests performed at 365 nm (I = 40 mW/cm2) in the presence of Iod revealed an improvement of the EPOX conversion from 52% with PI-PAG (1 wt%) up to 82% upon addition of Iod (3 wt%) (see Figure 22). It was thus concluded that the chromophore in PI-PAG could also sensitize the photodecomposition of Iod.
This point was confirmed during the polymerization tests carried out with the precursor pre-H which was used for the synthesis of PI-PAG. Using the two-component precursor pre-H/Iod (1%/3%, w/w) system, a conversion of 30% could be determined after 300 s of irradiation, evidencing that precursor pre-H could act as a photosensitizer for Iod. Ability of precursor pre-H but also of the byproducts formed during the photolysis of PI-PAG to sensitize the decomposition of Iod was confirmed at 385, 405 and 425 nm. In these different cases, an improvement of the monomer conversion was still obtained with the two-component systems compared with PI-PAG used alone (see Figure 23). A reduction of the monomer conversion was detected at long wavelengths, consistent with a decrease in the molar extinction coefficients.
Photolysis experiments also revealed that the photolysis of PI-PAG ended within 5 s so that the contribution of the photolysis byproducts to sensitize Iod was clearly demonstrated. The following mechanism was suggested to support the improvement of monomer conversion in the presence of Iod (see Equations (1)–(6)). In this mechanism, the key step is the photoinduced electron transfer between the excited dye and the electron deficient Iod (2). Dye•+ is thus formed, constituting an initiating species. In the presence of the monomer, a hydrogen abstraction reaction can occur, generating Bronsted acid (see Equations (3) and (4)). Parallel to this, the phenyl radical Ph can also react with the monomer, generating M+.
Dye ⟶ Dye* (hν)
Dye* + Ph2I+ ⟶ Dye•+ + Ph + Ph-I
M-H + Dye•+ ⟶ Dye+-H + M
Dye+-H ⟶ Dye + H+
M-H + Ph ⟶ Ph-H + M
M + Ph2I+ ⟶ M+ + Ph-I + Ph
Estimation of the Gibbs free energy change (ΔGet) using the Rehm-Weller equation revealed that the photoinduced electron transfer proceeds from the singlet excited state for both PI-PAG and precursor pre-H. Notably, in the case of the two-component precursor pre-H/Iod system, ΔGet = −180 kJ/mol was determined, constituting a high driving force for the photoinduced electron transfer. Occurrence of a photoinduced electron transfer from precursor pre-H to Iod was confirmed by fluorescence quenching measurements, with a complete quenching of the fluorescence of the precursor in the presence of Iod. In 2020, Yagci and coworkers revisited PI-PAG in the context of photopolymerization experiments carried out under low light intensities since the weak chemiluminescence of 9,10-diphenylanthracene (DPA) was used as an inner light source [263]. An emission centered at 430 nm could be obtained with DPA. Due to the weak light intensity, 20 min were required to obtain an efficient conversion of monomers such as N-vinyl carbazole (NVK) and n-butyl vinyl ether (BVE) bearing electron-donating moieties. In the case of cyclohexene oxide (CHO), 100 min were necessary to obtain acceptable monomer conversions. To support the high monomer conversions obtained with NVK and BVE, the concomitant occurrence of a cationic polymerization process issued from Bronsted acids and a free-radical-promoted cation polymerization process resulting from the oxidation of monomer-derived radicals by the sulfonium salt was suggested. Among the most interesting results, an efficient post-curing process could be evidenced. Thus, an increase in the CHO conversion from 20% could be obtained in the dark after 40 min after the irradiation has ceased. This is directly related to the formation of Bronsted acids that permit this efficient dark curing process.
In the same spirit, up-conversion nanoparticles (UCNPs) were recently used to activate sulfonium salts in the near-UV visible range while irradiating in the near-infrared range (see Figure 24) [264]. To date, only a few works have been reported in the literature concerning the use of UCNPs for developing safe polymerization processes [264,265,266,267,268,269,270,271]. Taking advantage of the simultaneous formation of Bronsted acids and radicals, hybrid polymerizations could be carried out, providing access to interpenetrated polymer networks (IPNs) [135,272,273,274,275,276,277,278]. Lanthanide-doped UCNPs β-NaYF4: 18% Yb, 0.5% Tm nanoparticles were used for up-conversion so that, upon irradiation at 980 nm, a blue light with emission peaks at 345 and 261 nm could be produced. Due to the better adequation of the absorptions of GR55 and CSS with that of the blue light, faster photolyses were obtained compared with GR61, for which a mismatch of the absorption was detected. Indeed, absorption of GR61 is strongly UV-centered and the use of a photosensitizer is commonly used to develop visible light photoinitiating systems with this mixture of sulfonium salts [279]. The high reactivity of GR55 was confirmed during the polymerization experiments, the best CHO conversion being obtained for sulfonium salt.
Here again, an important dark curing was evidenced since the CHO conversion increased from ca 20% after one hour in the dark, comparable to the results previously reported during the chemiluminescence-assisted cationic polymerization. Examination of the depth of the cure obtained during the radical/cationic hybrid polymerization of a TMPTA/EPOX (1/1) blend revealed that the curing depth reachs 3.9 cm after 2 min of irradiation at 980 nm (I = 11.6 W) and by using 0.4 wt% of UCNPs and 0.3 wt% of GR55. By elongating the irradiation time to 5 min, the remarkable curing depth of 11.7 cm could be evidenced (see Figure 25).
In 2022, the benchmark mixture of sulfonium salts, i.e., GR61, enabled the designing of antimicrobial coatings using a biobased epoxy monomer [280]. By increasing the vanillin alcohol diglycidyl ether (VDGE) content within the interpenetrated polymer networks prepared with glycerol dimethacrylate (GDMK), an improvement of the antimicrobial activities of the polymer films was determined (see Figure 26). Antibacterial and antifungal activities of the polymer blends comprising 66, 50 and 33% of VDGE were investigated by examining the viability of E. coli and St. aureus in contact with the polymer films. After 5 h, the viability of E. coli was reduced to 0% with all polymer films. In the case of St. aureus, 24 h of incubation were required to reduce the cell viability to 0%. In addition, as shown in Figure 27, the fastest reduction of E. coli and St. aureus cells were obtained for polymer films comprising 66 and 30% of VDGE. Comparison of the antibacterial activities of chitosan and hydroxyethyl starch with that of the newly developed sulfonium salts revealed the different sulfonium salts require longer exposure times than chitosan (approximately 24 h for a complete inhibition of the bacterial viability of St. aureus whereas two hours are only required with chitosan films). However, compared with hydroxyethyl starch, a better activity was found with the different sulfonium salts.

2.5. Bopidy-Based Sulfonium Salts

Bodipys, also named boron dipyrromethenes or 4,4-difluoro-4-bora-3a,4a-diaza-s-indecenes, have been identified, starting from 2000, as potential candidates for photoinitiation [281]. However, the introduction of the bodipy moiety onto sulfonium salts was only examined in 2020 by Zhang and coworkers that designed PAG-1 and PAG-2 (see Figure 28) [282]. The two sulfonium salts differing by the substitution pattern showed absorption maxima at 509 and 566 nm, respectively, in methanol. Due to the more extended structure of PAG-2 compared with that of PAG-1, a redshift of the absorption spectrum was logically detected. Noticeably, a low photoluminescence quantum yield was determined for the two salts, namely 0.02 for PAG-2 and 0.11 for PAG-1. Examination of their photoacid generation quantum yields revealed PAG-1 to be a better candidate since a yield of 7% was determined, higher than that obtained for PAG-2, namely 4%.
Polymerization experiments were carried out at 595 nm (I = 80 mW/cm2) using cyclohexane oxide (CHO) as the monomer and 0.8 wt% photoinitiators. The different experiments only revealed slow polymerization processes. After 30 min, 21% of conversion could only be determined consistent with their low photoacid generation abilities. Despite these low monomer conversions, these two structures can be cited as the sulfonium salts displaying the most redshifted absorption ever reported to date.

2.6. Charge Transfer Complexes

Sulfonium salts suffer from a UV-centered absorption so that these structures are inactive under visible light [159,283,284]. In order to extend their spectral activities, an interesting approach was proposed by Yagci and coworkers, consisting in generating charge transfer complexes (CTCs) [285,286]. The interest of the charge transfer approach relies on the fact that the combination of an electron donor and an electron acceptor not absorbing in the visible range can produce, once the CTC forms, a new species absorbing in the visible range. In 2019, an isopropylthioxanthone-based sulfonium salt (ITXPhenS) (see Figure 29) was notably examined as an electron-accepting group for an electron-rich amine, i.e., N,N-dimethylaniline (DMA), so that an interesting charge transfer complex (CTC) absorbing in the visible range could be produced [287].
As shown in Figure 30, absorbance of the ITXPhenS/DMACTC could be controlled by modifying the ITXPhenS to DMA ratio. By replacing a ITXPhenS/DMACTC 1:1 ratio for a 1:2 ratio, a redshift of the absorbance could be obtained. Thus, if an absorption band ranging between 425–500 nm was determined for the ITXPhenS/DMACTC 1:1 ratio, this absorption band was shifted to 500–625 nm for the ITXPhenS/DMACTC 1:2 ratio. Noticeably, the possibility to form CTCs with other electron-rich amines was examined such as triethylamine, triphenylamine and pyridine. However, the formation of CTC was not detected with these different amines. The different ITXPhenS/DMACTC proved to be versatile photoinitiators since photopolymerization experiments could be carried out at 350, 450 nm and even under sunlight. Using the sun as the light source, excellent monomer conversions could be obtained using isobutyl vinyl ether (IBVE) or methyl methacrylate (MMA) as the monomers, and a concentration of (5 × 10−2 M) for the CTC. Comparison of the polymerization profiles obtained during the photopolymerization of the difunctional triethylene glycol divinylether (TEGDVE) and triethyleneglycol dimethacrylate (TEGDMA) revealed that much faster polymerizations could take place with TEGDVE than with TEGDMA. Considering that TEGDVE and TEGDMA, respectively, polymerize by means of a cationic and a free radical mechanism, the initiation mechanism of ITXPhenS/DMACTC was presumed to occur by means of a homolytic and a heterolytic cleavage of the phenacyl moiety in the excited state, followed by an electron transfer. as classically observed for sulfonium salts (see Figure 31).
More recently, the possibility of designing new CTCs with ITXPhenS was examined [170]. Notably, CTCs could be prepared using 4-N,N-Trimethylaniline (TMA) or 2-(diphenylphosphino)benzoic acid (2DPPBA) as the electron donors (see Figure 32). Interestingly, ITXPhenS/TMACTC (0.5%/0.25% w/w) and ITXPhenS/2DPPBACTC (1%/1% w/w) exhibited a dual curing mode since the FRP of a mixture of BisGMA/TEGDMA (70/30) could be photochemically and thermally initiated. As thermal initiator, the ITXPhenS/TMACTC could initiate the FRP of the BisGMA/TEGDMA blend with a maximum exothermic peak at approximately 120 °C whereas this maximal temperature increased up to 176 °C for the ITXPhenS/2DPPBACTC. For comparison, thermal polymerization using ITXPhenS as the initiator could be initiated in almost similar conditions than that observed for the ITXPhenS/2DPPBACTC since the maximum exothermic peak was approximately 186 °C.
An excellent storage stability was determined for the two CTC-based resins since no modification of the polymerization profiles with that of freshly prepared resins was detected, even after 45 days of storage. Due to the photo/thermal dual curing abilities of the two CTCs, composites could be prepared. Using light, the surface of the composites could be efficiently polymerized, whereas the polymerization in bulk could be obtained by thermal curing (see Figure 33).

2.7. Fluorene-Based Sulfonium Salts

Different strategies were developed to redshift the absorption of the sulfonium salts, and an efficient strategy was proposed by using fluorene as a π-conjugated spacer [288]. Fluorene is an excellent electron-donating group that was notably used for the design of chromophores for solar cells [289,290,291,292,293,294,295,296,297,298,299], light emitters for OLEDs [300,301,302,303,304,305,306] and charge transport materials for OFETs [307,308,309,310,311,312,313,314,315] and as fluorescent probes [316,317,318,319,320,321,322,323]. The fluorene scaffold was notably used by Yagci and coworkers to design thioxanthone exhibiting a high molar extinction coefficient and a redshifted absorption compared with PI-PAG [324]. Influence of the substitution pattern was also examined, since Flu-MS and Flu-PS vary from each other by the position of the sulfonium moiety in meta and para positions, respectively (see Figure 34). As anticipated, Flu-PS exhibited a redshifted absorption compared with Flu-MS (378 nm for Flu-PS vs. 368 nm for Flu-MS), resulting from an extended conjugation in Flu-PS (see Figure 35). Jointly, a slight increase in the molar extinction coefficient could be evidenced for Flu-PS compared with Flu-MS (36 900 for Flu-MS vs. 39 400 M−1·cm−1). Irrespective of the substitution pattern, the two salts absorb up to 450 nm so that polymerization experiments could be carried out at 365, 385 and 405 nm using LEDs of low light intensity (2 mW/cm2). Photoacid generation quantum yields of 0.43 and 0.63 were, respectively, determined for Flu-PS and Flu-MS upon irradiation at 365 nm. In the case of Flu-MS, this value is higher than that previously reported for Mono-Ben (0.60 in Ref. [261] and 0.48 in Ref. [178]), not comprising a conjugated spacer. It therefore suggests that the elongation of the π-conjugation facilitates the dissociation of the S–C bond by stabilizing the intermediate radical cation.
In order to investigate the origin of the photoinduced electron transfer in Flu-MS and Flu-PS, the Gibbs free energy change (ΔGet) for an electron transfer occurring from the singlet excited state was determined for the two structures. Thus, values of −111.61 and −71.92 kJ/mol were, respectively, determined for Flu-MS and Flu-PS. These results are consistent with the photoacid generation quantum yields determined for the two structures, the highest photoacid generation quantum yield being obtained for the structure exhibiting the most negative ΔGet.
The higher reactivity of Flu-MS was confirmed during the CP of DVE-3, CHO, EPOX, the FRP of TMPTA, hybrid polymerization of an EPOX/TMPTA blend and the thiol-ene polymerization of trithiol/TMPTA performed upon irradiation at 365 nm and by using 1 wt% photoinitiators (I = 2 or 4 mW/cm2, see Table 8). During cationic polymerizations, the highest monomer conversions were obtained with DVE-3, with conversions higher than 98%. During the concomitant polymerization of an EPOX/TMPTA blend (50%/50% w/w), a higher TMPTA conversion was obtained when blended (70%) than during the FRP of TMPTA alone (56%), resulting from a synergetic effect between the CP of EPOX and the FRP of TMPTA. Among the most interesting results, extremely fast thiol-ene polymerization processes could be evidenced with Flu-MS and Flu-PS since the maximum conversions for TMPTA and trithiol could be obtained within 10 s.
Examination of the monomer conversions obtained for different monomers at longer wavelengths revealed the two sulfonium salts to be capable of maintaining a higher monomer conversion, even upon irradiation at 450 nm (See Table 9).
The design of more extended structures was also examined by Belfield and coworkers who developed a series of two-photon photoinitiators based on fluorenes (see Figure 36) [256]. These structures were notably proposed as photoinitiators for two-photon 3D write-once read-many (WORM) optical data storage systems.

2.8. Anthracene-Based Sulfonium Salts

Sulfonium salts dramatically lack absorption above 320 nm, but their spectral response can be significantly improved by the covalent linkage of the sulfonium group to anthracene (see Figure 37) [325]. Anthracene is an excellent photosensitizer that was used numerous times in photopolymerization [34,69,326,327,328,329,330,331,332,333,334]. Using this strategy, absorption of the different sulfonium salts could range between 330 and 410 nm, benefiting from the strong absorption of anthracene in this spectral range. Consequently, anthracene could act as a photosensitizer for the sulfonium salts, enabling the generation of mono-component photoinitiating systems. For comparison, a series of sulfonium salts were designed with a phenyl group (see compounds 6a6c).
Photoinitiating ability of the different structures was investigated during the cationic polymerization of a mixture of three monomers comprising tris(4-hydroxyphenyl) methane triglycidyl ether (TPGE), SU-8 and EPOX upon irradiation with a medium pressure mercury lamp (500 mJ/cm2) and using a 2.5 wt% photoinitiator. Interestingly, during the CP of the blend of three monomers (SU-8, TPGE and EPOX), 2 and 4a showed a photosensitivity six times higher than that of 4b, bearing only one aromatic ring on the sulfonium side. In the case of the alkyl-substituted sulfonium salts, 4c and 4d, no photoinitiating activity was detected. Comparison with a benchmark photoinitiator composed of a mixture of diphenyl-4-thiophenoxyphenylsulfonium hexafluoroantimonate and bis[4-(diphenylsulfonio)phenyl]sulfide bis-hexafluoroantimonate (GR61) revealed 2 and 4a (and even 4b) to exhibit a higher photosensitivity than the benchmark photoinitiator GR61 when sensitized by 9-methylanthracene, evidencing the interest of the intramolecular approach and the covalent linkage of the aryl-substituted sulfonium groups by the anthracene chromophore. Photoacid generation quantum yields ϕH+ of 2 and 4b in acetonitrile using a merocyanine indicator dye furnished values of 0.32 and 0.21, respectively, consistent with the order of photoactivity determined during the photopolymerization experiments. Examination of the photoproducts formed during photolysis of the anthracene/GR61 combination revealed phenylated anthracene to be formed in 55% yield, with 9-phenylanthracene as the main side-product (73%). The formation of 9-phenylanthracene was assigned to a photoinduced electron transfer between the excited anthracene (in its singlet excited state) and the electrodeficient sulfonium salt [163,335]. Then, by an in-cage combination of an aryl radical and the radical cation of anthracene, 9-phenylanthracene can be formed subsequent to a proton elimination (see Scheme 6).
Based on the previous work of Sharma and coworkers, the authors suggested the good polymerization efficiency of 2 and 4a is unlikely to arise from an inefficient back electron transfer [336]. Indeed, back electron transfer in photopolymerization is one of the major drawbacks of the photosensitization approach [337,338,339]. Although well-known, this undesired back electron transfer cannot be anticipated. More likely, the occurrence of an efficient electron-transfer induced concerted bond cleavage and a cation radical/radical coupling mechanism was suggested as a plausible explanation to support the polymerization efficiency of 2 and 4a, in light of the previous work carried out by Luss and their coworkers on intramolecular photoinduced rearrangements of para-cyanobenzylated anthracenes [166]. Photolysis experiments performed in the solution confirmed the good reactivity of 4a, with a decrease in the absorption intensity of 55% after 30 min of irradiation in acetonitrile. Fluorescence experiments revealed that 2, 4a and 4b exhibit a low fluorescence intensity, indicative of an intramolecular electron transfer from the excited singlet state of anthracene to the sulfonium group. Conversely, fluorescence quenching of anthracene was not detected for 4c and 4d, evidencing the lack of photoinduced electron transfer. Using the Rehm–Weller equation, the free energy change ΔG for 4a, 4b and 4d could be determined. ΔG values of −107, −49 and −20 kJ/mol were determined, confirming that the thermodynamic was unfavorable for an intramolecular electron transfer for 4d.

2.9. Coumarin-Based Sulfonium Salts

Coumarins are biosourced or bioinspired structures [340,341,342] that have been extensively used for the design of photoinitiating systems [343,344,345,346,347,348]. Coumarins, by their absorptions located in the near-UV-visible range, are appropriate photoinitiators for photopolymerization experiments performed at 405 nm, which is the wavelength used in 3D printers [349,350,351,352,353,354]. In 2023, a series of coumarin-substituted sulfonium salts bearing long alkyl chains for solubility was proposed by Liu and coworkers and examined in cationic, free radical and hybrid photopolymerizations (see Figure 38) [355]. From the absorption viewpoint, the introduction of a methoxy group in p-OMe-Me CSS, p-OMe-Bu CSS and p-OMe-Hep CSS contributed to the redshift of the absorption of the different dyes and increase in the molar extinction coefficient by favoring the electronic delocalization. Thus, if an absorption maximum at approximately 380–390 nm was found for p-OMe-Me CSS, p-OMe-Bu CSS and p-OMe-Hep CSS in acetonitrile, an absorption maxima at 300–320 nm was determined for p-H-Me CSS, p-H-Bu CSS and p-H-Hep CSS (see Figure 39). The crucial role of the electron-donating methoxy group on the optical properties of the sulfonium salts was thus demonstrated. The beneficial effect of the methoxy group was also determined during the steady-state photolysis experiments performed upon irradiation at 365 and 405 nm. Thus, all seven substituted sulfonium salts could decompose faster than the unsubstituted ones, resulting in their redshifted absorptions better fitting with the emission of LEDs.
By ESR experiments, formation of the acetyl coumarin radical and dimethyl sulfide radical cations could be demonstrated, resulting from the homolytic cleavage of the C–S bond. By GC-MS analysis, formation of acetylcoumarin could also be evidenced, confirming the ESR results. In light of these results, the mechanism of photoinitiation could be determined and the mechanism that was proposed by the authors is depicted in Figure 40.
The investigation of their photoinitiating abilities during the CP of EPOX and the FRP of TMPTA revealed the good adequation between molar extinction at the irradiation wavelength and monomer conversions (see Table 10). Control experiments were carried out with bis[4-(diphenylsulfonio)phenyl] sulfide bis(hexafluoroantimonate) (201s) for the CP of EPOX and ITX for the FRP of TMPTA. Photopolymerization tests were carried out at 365 nm (I = 70 mW/cm2) and 405 nm (I = 70 mW/cm2), using a concentration of 5.27 × 10−5 mol/g photoinitiators.
Interestingly, by varying the length of the alkyl chains on the sulfonium group from one to seven carbons, an optimal situation was found for p-H-Bu CSS and p-OMe-Bu CSS, both bearing butyl chains. As can be seen in Table 10, the best monomer conversions obtained at 365 and 405 nm during the CP of EPOX and the FRP of TMPTA were obtained for these two photoinitiators. These results are remarkable considering that monomer conversions outperforming that obtained with 201s and ITX could be determined.
Considering the high monomer conversions obtained during the cationic polymerization and the free radical polymerization, hybrid polymerization of tripropylene glycol diacrylate (TPGDA) and EPOX was investigated. Noticeably, the FRP of TPGDA proceeded faster than the CP of EPOX, the radicals being directly formed upon photoexcitation of the photoinitiators, contrarily to the cation species that are formed in successive reactions (see Figure 41). This point was especially evidenced with p-OMe-Me CSS.

3. Conclusions

Sulfonium salts are efficient photoinitiators suffering from a weak absorption in the visible range. Over the years, this issue was addressed by connecting the sulfonium groups to different chromophores such as Bodipy, 1,3,5-triphenyl-2-pyrazolines, triphenylamines, fluorenes, anthracene or coumarins. The choice of these chromophores for generating monocomponent photoinitiating systems is directly related to the recent advances in the design of Type II photoinitiating systems where these structures have emerged as promising scaffolds. As specificities, these different chromophore-bonded sulfonium salts have been obtained by means of multistep syntheses. If the resulting structures are costly, the efficiency of the intramolecular electron transfer vs. the intermolecular one has been clearly demonstrated, evidencing the pertinence of this approach. The formation of charge transfer complexes of sulfonium salts with different electron-rich amines is an interesting approach enabling the minimization of the synthetic efforts. Indeed, such structures can be directly prepared from commercially available sulfonium salts and aromatic amines. Among the most recent advances, up-conversion nanoparticles have been used to promote the photocleavage of sulfonium salts while exciting in the near-infrared range, i.e., in safe irradiation conditions. As other advantages of these mono-component systems, chromophores bonded to the sulfonium group greatly contributed to the generation of sulfonium salts soluble in resins, which constitutes a major drawback of most of the benchmark sulfonium salts. At present, only a few sulfonium salts showed sufficient reactivity to initiate polymerization processes under sunlight. Furthermore, such structures would be of interest for numerous outdoor applications. At present, water-soluble sulfonium salts activable in the visible range have not been reported yet, to the best of our knowledge. In addition, photopolymerization in water could become possible, enabling polymerization in green conditions. Sensitivity to sunlight and the development of water-soluble sulfonium salts will certainly be the focus of intense works in the coming years.

Funding

This research received no external funding.

Data Availability Statement

No data available.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Allegrezza, M.L.; DeMartini, Z.M.; Kloster, A.J.; Digby, Z.A.; Konkolewicz, D. Visible and Sunlight Driven RAFT Photopolymerization Accelerated by Amines: Kinetics and Mechanism. Polym. Chem. 2016, 7, 6626–6636. [Google Scholar] [CrossRef]
  2. Sun, K.; Chen, H.; Zhang, Y.; Morlet-Savary, F.; Graff, B.; Xiao, P.; Dumur, F.; Lalevée, J. High-Performance Sunlight Induced Polymerization Using Novel Push-Pull Dyes with High Light Absorption Properties. Eur. Polym. J. 2021, 151, 110410. [Google Scholar] [CrossRef]
  3. Ciftci, M.; Tasdelen, M.A.; Yagci, Y. Sunlight Induced Atom Transfer Radical Polymerization by Using Dimanganese Decacarbonyl. Polym. Chem. 2014, 5, 600–606. [Google Scholar] [CrossRef]
  4. Decker, C.; Bendaikha, T. Interpenetrating Polymer Networks. II. Sunlight-Induced Polymerization of Multifunctional Acrylates. J. Appl. Polym. Sci. 1998, 70, 2269–2282. [Google Scholar] [CrossRef]
  5. Tehfe, M.-A.; Lalevée, J.; Gigmes, D.; Fouassier, J.P. Green Chemistry: Sunlight-Induced Cationic Polymerization of Renewable Epoxy Monomers Under Air. Macromolecules 2010, 43, 1364–1370. [Google Scholar] [CrossRef]
  6. Wang, J.; Rivero, M.; Muñoz Bonilla, A.; Sanchez-Marcos, J.; Xue, W.; Chen, G.; Zhang, W.; Zhu, X. Natural RAFT Polymerization: Recyclable-Catalyst-Aided, Opened-to-Air, and Sunlight-Photolyzed RAFT Polymerizations. ACS Macro Lett. 2016, 5, 1278–1282. [Google Scholar] [CrossRef] [PubMed]
  7. Yu, J.; Gao, Y.; Jiang, S.; Sun, F. Naphthalimide Aryl Sulfide Derivative Norrish Type I Photoinitiators with Excellent Stability to Sunlight under Near-UV LED. Macromolecules 2019, 52, 1707–1717. [Google Scholar] [CrossRef]
  8. Bao, Y. Recent Trends in Advanced Photoinitiators for Vat Photopolymerization 3D Printing. Macromol. Rapid Commun. 2022, 43, 2200202. [Google Scholar] [CrossRef]
  9. Lalevée, J.; Fouassier, J.-P. Dyes and Chromophores in Polymer Science; ISTE Ltd.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2015; ISBN 978-1-84821-742-3. [Google Scholar]
  10. Fouassier, J.P.; Lalevée, J. Three-Component Photoinitiating Systems: Towards Innovative Tailor Made High Performance Combinations. RSC Adv. 2012, 2, 2621–2629. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Xu, Y.; Simon-Masseron, A.; Lalevée, J. Radical Photoinitiation with LEDs and Applications in the 3D Printing of Composites. Chem. Soc. Rev. 2021, 50, 3824–3841. [Google Scholar] [CrossRef]
  12. Lalevée, J.; Allonas, X.; Fouassier, J.P. Tris(Trimethylsilyl)Silane (TTMSS)-Derived Radical Reactivity toward Alkenes:  A Combined Quantum Mechanical and Laser Flash Photolysis Study. J. Org. Chem. 2007, 72, 6434–6439. [Google Scholar] [CrossRef] [PubMed]
  13. Chatgilialoglu, C.; Lalevée, J. Recent Applications of the (TMS)3SiH Radical-Based Reagent. Molecules 2012, 17, 527–555. [Google Scholar] [CrossRef] [PubMed]
  14. Garra, P.; Dietlin, C.; Morlet-Savary, F.; Dumur, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Redox Two-Component Initiated Free Radical and Cationic Polymerizations: Concepts, Reactions and Applications. Prog. Polym. Sci. 2019, 94, 33–56. [Google Scholar] [CrossRef]
  15. Lalevée, J.; Mokbel, H.; Fouassier, J.-P. Recent Developments of Versatile Photoinitiating Systems for Cationic Ring Opening Polymerization Operating at Any Wavelengths and under Low Light Intensity Sources. Molecules 2015, 20, 7201–7221. [Google Scholar] [CrossRef]
  16. Jasinski, F.; Zetterlund, P.B.; Braun, A.M.; Chemtob, A. Photopolymerization in Dispersed Systems. Prog. Polym. Sci. 2018, 84, 47–88. [Google Scholar] [CrossRef]
  17. Noè, C.; Hakkarainen, M.; Sangermano, M. Cationic UV-Curing of Epoxidized Biobased Resins. Polymers 2021, 13, 89. [Google Scholar] [CrossRef] [PubMed]
  18. Yuan, Y.; Li, C.; Zhang, R.; Liu, R.; Liu, J. Low Volume Shrinkage Photopolymerization System Using Hydrogen-Bond-Based Monomers. Prog. Org. Coat. 2019, 137, 105308. [Google Scholar] [CrossRef]
  19. Khudyakov, I.V.; Legg, J.C.; Purvis, M.B.; Overton, B.J. Kinetics of Photopolymerization of Acrylates with Functionality of 1–6. Ind. Eng. Chem. Res. 1999, 38, 3353–3359. [Google Scholar] [CrossRef]
  20. Dickens, S.H.; Stansbury, J.W.; Choi, K.M.; Floyd, C.J.E. Photopolymerization Kinetics of Methacrylate Dental Resins. Macromolecules 2003, 36, 6043–6053. [Google Scholar] [CrossRef]
  21. Maffezzoli, A.; Pietra, A.D.; Rengo, S.; Nicolais, L.; Valletta, G. Photopolymerization of Dental Composite Matrices. Biomaterials 1994, 15, 1221–1228. [Google Scholar] [CrossRef]
  22. Dikova, T.; Maximov, J.; Todorov, V.; Georgiev, G.; Panov, V. Optimization of Photopolymerization Process of Dental Composites. Processes 2021, 9, 779. [Google Scholar] [CrossRef]
  23. Andreu, A.; Su, P.-C.; Kim, J.-H.; Ng, C.S.; Kim, S.; Kim, I.; Lee, J.; Noh, J.; Subramanian, A.S.; Yoon, Y.-J. 4D Printing Materials for Vat Photopolymerization. Addit. Manuf. 2021, 44, 102024. [Google Scholar] [CrossRef]
  24. Chen, H.; Noirbent, G.; Zhang, Y.; Sun, K.; Liu, S.; Brunel, D.; Gigmes, D.; Graff, B.; Morlet-Savary, F.; Xiao, P.; et al. Photopolymerization and 3D/4D Applications Using Newly Developed Dyes: Search around the Natural Chalcone Scaffold in Photoinitiating Systems. Dyes Pigment. 2021, 188, 109213. [Google Scholar] [CrossRef]
  25. Bagheri, A.; Jin, J. Photopolymerization in 3D Printing. ACS Appl. Polym. Mater. 2019, 1, 593–611. [Google Scholar] [CrossRef]
  26. Tomal, W.; Kiliclar, H.C.; Fiedor, P.; Ortyl, J.; Yagci, Y. Visible Light Induced High Resolution and Swift 3D Printing System by Halogen Atom Transfer. Macromol. Rapid Commun. 2023, 44, 2200661. [Google Scholar] [CrossRef]
  27. Armstrong, B.K.; Kricker, A. The Epidemiology of UV Induced Skin Cancer. J. Photochem. Photobiol. B Biol. 2001, 63, 8–18. [Google Scholar] [CrossRef] [PubMed]
  28. de Gruijl, F.R. Skin Cancer and Solar UV Radiation. Eur. J. Cancer 1999, 35, 2003–2009. [Google Scholar] [CrossRef] [PubMed]
  29. Kamoun, E.A.; Winkel, A.; Eisenburger, M.; Menzel, H. Carboxylated Camphorquinone as Visible-Light Photoinitiator for Biomedical Application: Synthesis, Characterization, and Application. Arab. J. Chem. 2016, 9, 745–754. [Google Scholar] [CrossRef]
  30. Santini, A.; Gallegos, I.T.; Felix, C.M. Photoinitiators in Dentistry: A Review. Prim. Dent. J. 2013, 2, 30–33. [Google Scholar] [CrossRef]
  31. Dumur, F. Recent Advances on Visible Light Thiophene-Based Photoinitiators of Polymerization. Eur. Polym. J. 2022, 169, 111120. [Google Scholar] [CrossRef]
  32. Helmy, S.; Oh, S.; Leibfarth, F.A.; Hawker, C.J.; Read de Alaniz, J. Design and Synthesis of Donor–Acceptor Stenhouse Adducts: A Visible Light Photoswitch Derived from Furfural. J. Org. Chem. 2014, 79, 11316–11329. [Google Scholar] [CrossRef] [PubMed]
  33. Pigot, C.; Noirbent, G.; Brunel, D.; Dumur, F. Recent Advances on Push–Pull Organic Dyes as Visible Light Photoinitiators of Polymerization. Eur. Polym. J. 2020, 133, 109797. [Google Scholar] [CrossRef]
  34. Dumur, F. Recent Advances on Anthracene-Based Photoinitiators of Polymerization. Eur. Polym. J. 2022, 169, 111139. [Google Scholar] [CrossRef]
  35. Abdallah, M.; Le, H.; Hijazi, A.; Schmitt, M.; Graff, B.; Dumur, F.; Bui, T.-T.; Goubard, F.; Fouassier, J.-P.; Lalevée, J. Acridone Derivatives as High Performance Visible Light Photoinitiators for Cationic and Radical Photosensitive Resins for 3D Printing Technology and for Low Migration Photopolymer Property. Polymer 2018, 159, 47–58. [Google Scholar] [CrossRef]
  36. Bao, B.; You, J.; Li, D.; Zhan, H.; Zhang, L.; Li, M.; Wang, T. Double Benzylidene Ketones as Photoinitiators for Visible Light Photopolymerization. J. Photochem. Photobiol. A Chem. 2022, 429, 113938. [Google Scholar] [CrossRef]
  37. Fu, H.; Qiu, Y.; You, J.; Hao, T.; Fan, B.; Nie, J.; Wang, T. Photopolymerization of Acrylate Resin and Ceramic Suspensions with Benzylidene Ketones under Blue/Green LED. Polymer 2019, 184, 121841. [Google Scholar] [CrossRef]
  38. Zhao, Y.; Wang, W.; Wu, F.; Zhou, Y.; Huang, N.; Gu, Y.; Zou, Q.; Yang, W. Polyethylene Glycol-Functionalized Benzylidene Cyclopentanone Dyes for Two-Photon Excited Photodynamic Therapy. Org. Biomol. Chem. 2011, 9, 4168–4175. [Google Scholar] [CrossRef] [PubMed]
  39. Fang, Y.; Liu, T.; Zou, Q.; Zhao, Y.; Wu, F. Cationic Benzylidene Cyclopentanone Photosensitizers for Selective Photodynamic Inactivation of Bacteria over Mammalian Cells. RSC Adv. 2015, 5, 56067–56074. [Google Scholar] [CrossRef]
  40. Xue, J.; Zhao, Y.; Wu, F.; Fang, D.-C. Effect of Bridging Position on the Two-Photon Polymerization Initiating Efficiencies of Novel Coumarin/Benzylidene Cyclopentanone Dyes. J. Phys. Chem. A 2010, 114, 5171–5179. [Google Scholar] [CrossRef]
  41. Egorov, A.E.; Kostyukov, A.A.; Shcherbakov, D.A.; Kolymagin, D.A.; Chubich, D.A.; Matital, R.P.; Arsenyev, M.V.; Burtsev, I.D.; Mestergazi, M.G.; Zhiganshina, E.R.; et al. Benzylidene Cyclopentanone Derivative Photoinitiator for Two-Photon Photopolymerization-Photochemistry and 3D Structures Fabrication for X-Ray Application. Polymers 2023, 15, 71. [Google Scholar] [CrossRef]
  42. Tehfe, M.-A.; Dumur, F.; Xiao, P.; Graff, B.; Morlet-Savary, F.; Fouassier, J.-P.; Gigmes, D.; Lalevée, J. New Chromone Based Photoinitiators for Polymerization Reactions under Visible Light. Polym. Chem. 2013, 4, 4234–4244. [Google Scholar] [CrossRef]
  43. You, J.; Fu, H.; Zhao, D.; Hu, T.; Nie, J.; Wang, T. Flavonol Dyes with Different Substituents in Photopolymerization. J. Photochem. Photobiol. A Chem. 2020, 386, 112097. [Google Scholar] [CrossRef]
  44. Al Mousawi, A.; Garra, P.; Schmitt, M.; Toufaily, J.; Hamieh, T.; Graff, B.; Fouassier, J.P.; Dumur, F.; Lalevée, J. 3-Hydroxyflavone and N-Phenylglycine in High Performance Photoinitiating Systems for 3D Printing and Photocomposites Synthesis. Macromolecules 2018, 51, 4633–4641. [Google Scholar] [CrossRef]
  45. Xiao, P.; Dumur, F.; Frigoli, M.; Graff, B.; Morlet-Savary, F.; Wantz, G.; Bock, H.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. Perylene Derivatives as Photoinitiators in Blue Light Sensitive Cationic or Radical Curable Films and Panchromatic Thiol-Ene Polymerizable Films. Eur. Polym. J. 2014, 53, 215–222. [Google Scholar] [CrossRef]
  46. Xiao, P.; Hong, W.; Li, Y.; Dumur, F.; Graff, B.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. Green Light Sensitive Diketopyrrolopyrrole Derivatives Used in Versatile Photoinitiating Systems for Photopolymerizations. Polym. Chem. 2014, 5, 2293–2300. [Google Scholar] [CrossRef]
  47. Xiao, P.; Hong, W.; Li, Y.; Dumur, F.; Graff, B.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. Diketopyrrolopyrrole Dyes: Structure/Reactivity/Efficiency Relationship in Photoinitiating Systems upon Visible Lights. Polymer 2014, 55, 746–751. [Google Scholar] [CrossRef]
  48. Zivic, N.; Bouzrati-Zerrelli, M.; Villotte, S.; Morlet-Savary, F.; Dietlin, C.; Dumur, F.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. A Novel Naphthalimide Scaffold Based Iodonium Salt as a One-Component Photoacid/Photoinitiator for Cationic and Radical Polymerization under LED Exposure. Polym. Chem. 2016, 7, 5873–5879. [Google Scholar] [CrossRef]
  49. Xiao, P.; Dumur, F.; Zhang, J.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Naphthalimide Derivatives: Substituent Effects on the Photoinitiating Ability in Polymerizations under Near UV, Purple, White and Blue LEDs (385, 395, 405, 455, or 470 nm). Macromol. Chem. Phys. 2015, 216, 1782–1790. [Google Scholar] [CrossRef]
  50. Xiao, P.; Dumur, F.; Zhang, J.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Naphthalimide-Phthalimide Derivative Based Photoinitiating Systems for Polymerization Reactions under Blue Lights. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 665–674. [Google Scholar] [CrossRef]
  51. Zhang, J.; Zivic, N.; Dumur, F.; Xiao, P.; Graff, B.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. N-[2-(Dimethylamino)Ethyl]-1,8-Naphthalimide Derivatives as Photoinitiators under LEDs. Polym. Chem. 2018, 9, 994–1003. [Google Scholar] [CrossRef]
  52. Noirbent, G.; Dumur, F. Recent Advances on Naphthalic Anhydrides and 1,8-Naphthalimide-Based Photoinitiators of Polymerization. Eur. Polym. J. 2020, 132, 109702. [Google Scholar] [CrossRef]
  53. Zivic, N.; Zhang, J.; Bardelang, D.; Dumur, F.; Xiao, P.; Jet, T.; Versace, D.-L.; Dietlin, C.; Morlet-Savary, F.; Graff, B.; et al. Novel Naphthalimide–Amine Based Photoinitiators Operating under Violet and Blue LEDs and Usable for Various Polymerization Reactions and Synthesis of Hydrogels. Polym. Chem. 2015, 7, 418–429. [Google Scholar] [CrossRef]
  54. Liu, S.; Giacoletto, N.; Graff, B.; Morlet-Savary, F.; Nechab, M.; Xiao, P.; Dumur, F.; Lalevée, J. N-Naphthalimide Ester Derivatives as Type Ⅰ Photoinitiators for LED Photopolymerization. Mater. Today Chem. 2022, 26, 101137. [Google Scholar] [CrossRef]
  55. Li, J.; Zhang, X.; Ali, S.; Akram, M.Y.; Nie, J.; Zhu, X. The Effect of Polyethylene Glycoldiacrylate Complexation on Type II Photoinitiator and Promotion for Visible Light Initiation System. J. Photochem. Photobiol. A Chem. 2019, 384, 112037. [Google Scholar] [CrossRef]
  56. Li, J.; Li, S.; Li, Y.; Li, R.; Nie, J.; Zhu, X. In Situ Monitoring of Photopolymerization by Photoinitiator with Luminescence Characteristics. J. Photochem. Photobiol. A Chem. 2020, 389, 112225. [Google Scholar] [CrossRef]
  57. Li, J.; Hao, Y.; Zhong, M.; Tang, L.; Nie, J.; Zhu, X. Synthesis of Furan Derivative as LED Light Photoinitiator: One-Pot, Low Usage, Photobleaching for Light Color 3D Printing. Dyes Pigment. 2019, 165, 467–473. [Google Scholar] [CrossRef]
  58. Xu, Y.; Noirbent, G.; Brunel, D.; Ding, Z.; Gigmes, D.; Graff, B.; Xiao, P.; Dumur, F.; Lalevée, J. Novel Ketone Derivative-Based Photoinitiating Systems for Free Radical Polymerization under Mild Conditions and 3D Printing. Polym. Chem. 2020, 11, 5767–5777. [Google Scholar] [CrossRef]
  59. Chen, H.; Noirbent, G.; Sun, K.; Brunel, D.; Gigmes, D.; Morlet-Savary, F.; Zhang, Y.; Liu, S.; Xiao, P.; Dumur, F.; et al. Photoinitiators Derived from Natural Product Scaffolds: Monochalcones in Three-Component Photoinitiating Systems and Their Applications in 3D Printing. Polym. Chem. 2020, 11, 4647–4659. [Google Scholar] [CrossRef]
  60. Chen, H.; Noirbent, G.; Liu, S.; Brunel, D.; Graff, B.; Gigmes, D.; Zhang, Y.; Sun, K.; Morlet-Savary, F.; Xiao, P.; et al. Bis-Chalcone Derivatives Derived from Natural Products as near-UV/Visible Light Sensitive Photoinitiators for 3D/4D Printing. Mater. Chem. Front. 2021, 5, 901–916. [Google Scholar] [CrossRef]
  61. Sun, K.; Xu, Y.; Dumur, F.; Morlet-Savary, F.; Chen, H.; Dietlin, C.; Graff, B.; Lalevée, J.; Xiao, P. In Silico Rational Design by Molecular Modeling of New Ketones as Photoinitiators in Three-Component Photoinitiating Systems: Application in 3D Printing. Polym. Chem. 2020, 11, 2230–2242. [Google Scholar] [CrossRef]
  62. Yen, S.-C.; Ni, J.-S.; Chen, Y.-C. Triphenylamine-Functionalized Chalcones as One-Component Type II Visible-Light-Absorbing Photoinitiators for Free Radical Photopolymerization. Eur. Polym. J. 2023, 187, 111885. [Google Scholar] [CrossRef]
  63. Gao, Y.; Qu, J. New Long-Wavelength D–π-A–π-D Chalcone Photoinitiator for Visible Light Polymerization with Photobleaching and Biocompatibility Properties. Polym. Chem. 2023, 14, 952–962. [Google Scholar] [CrossRef]
  64. Deng, L.; Qu, J. Synthesis and Properties of Novel Bis-Chalcone-Based Photoinitiators for LED Polymerization with Photobleaching and Low Migration. Prog. Org. Coat. 2023, 174, 107240. [Google Scholar] [CrossRef]
  65. Al Mousawi, A.; Garra, P.; Sallenave, X.; Dumur, F.; Toufaily, J.; Hamieh, T.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. π-Conjugated Dithienophosphole Derivatives as High Performance Photoinitiators for 3D Printing Resins. Macromolecules 2018, 51, 1811–1821. [Google Scholar] [CrossRef]
  66. Topa-Skwarczyńska, M.; Galek, M.; Jankowska, M.; Morlet-Savary, F.; Graff, B.; Lalevée, J.; Popielarz, R.; Ortyl, J. Development of the First Panchromatic BODIPY-Based One-Component Iodonium Salts for Initiating the Photopolymerization Processes. Polym. Chem. 2021, 12, 6873–6893. [Google Scholar] [CrossRef]
  67. Karaca, N.; Ocal, N.; Arsu, N.; Jockusch, S. Thioxanthone-Benzothiophenes as Photoinitiator for Free Radical Polymerization. J. Photochem. Photobiol. A Chem. 2016, 331, 22–28. [Google Scholar] [CrossRef]
  68. Balta, D.K.; Cetiner, N.; Temel, G.; Turgut, Z.; Arsu, N. An Annelated Thioxanthone as a New Type II Initiator. J. Photochem. Photobiol. A Chem. 2008, 199, 316–321. [Google Scholar] [CrossRef]
  69. Balta, D.K.; Temel, G.; Goksu, G.; Ocal, N.; Arsu, N. Thioxanthone–Diphenyl Anthracene: Visible Light Photoinitiator. Macromolecules 2012, 45, 119–125. [Google Scholar] [CrossRef]
  70. Dadashi-Silab, S.; Aydogan, C.; Yagci, Y. Shining a Light on an Adaptable Photoinitiator: Advances in Photopolymerizations Initiated by Thioxanthones. Polym. Chem. 2015, 6, 6595–6615. [Google Scholar] [CrossRef]
  71. Eren, T.N.; Yasar, N.; Aviyente, V.; Morlet-Savary, F.; Graff, B.; Fouassier, J.P.; Lalevee, J.; Avci, D. Photophysical and Photochemical Studies of Novel Thioxanthone-Functionalized Methacrylates through LED Excitation. Macromol. Chem. Phys. 2016, 217, 1501–1512. [Google Scholar] [CrossRef]
  72. Qiu, J.; Wei, J. Thioxanthone Photoinitiator Containing Polymerizable N-Aromatic Maleimide for Photopolymerization. J. Polym. Res. 2014, 21, 559. [Google Scholar] [CrossRef]
  73. Tar, H.; Sevinc Esen, D.; Aydin, M.; Ley, C.; Arsu, N.; Allonas, X. Panchromatic Type II Photoinitiator for Free Radical Polymerization Based on Thioxanthone Derivative. Macromolecules 2013, 46, 3266–3272. [Google Scholar] [CrossRef]
  74. Wu, Q.; Wang, X.; Xiong, Y.; Yang, J.; Tang, H. Thioxanthone Based One-Component Polymerizable Visible Light Photoinitiator for Free Radical Polymerization. RSC Adv. 2016, 6, 66098–66107. [Google Scholar] [CrossRef]
  75. Wu, Q.; Tang, K.; Xiong, Y.; Wang, X.; Yang, J.; Tang, H. High-Performance and Low Migration One-Component Thioxanthone Visible Light Photoinitiators. Macromol. Chem. Phys. 2017, 218, 1600484. [Google Scholar] [CrossRef]
  76. Wu, X.; Jin, M.; Malval, J.-P.; Wan, D.; Pu, H. Visible Light-Emitting Diode-Sensitive Thioxanthone Derivatives Used in Versatile Photoinitiating Systems for Photopolymerizations. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 4037–4045. [Google Scholar] [CrossRef]
  77. Esen, D.S.; Karasu, F.; Arsu, N. The Investigation of Photoinitiated Polymerization of Multifunctional Acrylates with TX-BT by Photo-DSC and RT-FTIR. Prog. Org. Coat. 2011, 70, 102–107. [Google Scholar] [CrossRef]
  78. Lalevée, J.; Blanchard, N.; Tehfe, M.A.; Fries, C.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.P. New Thioxanthone and Xanthone Photoinitiators Based on Silyl Radical Chemistry. Polym. Chem. 2011, 2, 1077–1084. [Google Scholar] [CrossRef]
  79. Gencoglu, T.; Eren, T.N.; Lalevée, J.; Avci, D. A Water Soluble, Low Migration, and Visible Light Photoinitiator by Thioxanthone-Functionalization of Poly(Ethylene Glycol)-Containing Poly(β-Amino Ester). Macromol. Chem. Phys. 2022, 223, 2100450. [Google Scholar] [CrossRef]
  80. Wang, Y.; Chen, R.; Liu, D.; Peng, C.; Wang, J.; Dong, X. New Functionalized Thioxanthone Derivatives as Type I Photoinitiators for Polymerization under UV-Vis LEDs. New J. Chem. 2023, 47, 5330–5337. [Google Scholar] [CrossRef]
  81. Liu, S.; Chen, H.; Zhang, Y.; Sun, K.; Xu, Y.; Morlet-Savary, F.; Graff, B.; Noirbent, G.; Pigot, C.; Brunel, D.; et al. Monocomponent Photoinitiators Based on Benzophenone-Carbazole Structure for LED Photoinitiating Systems and Application on 3D Printing. Polymers 2020, 12, 1394. [Google Scholar] [CrossRef]
  82. Liu, S.; Brunel, D.; Sun, K.; Zhang, Y.; Chen, H.; Xiao, P.; Dumur, F.; Lalevée, J. Novel Photoinitiators Based on Benzophenone-Triphenylamine Hybrid Structure for LED Photopolymerization. Macromol. Rapid Commun. 2020, 41, 2000460. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, S.; Brunel, D.; Sun, K.; Xu, Y.; Morlet-Savary, F.; Graff, B.; Xiao, P.; Dumur, F.; Lalevée, J. A Monocomponent Bifunctional Benzophenone–Carbazole Type II Photoinitiator for LED Photoinitiating Systems. Polym. Chem. 2020, 11, 3551–3556. [Google Scholar] [CrossRef]
  84. Lin, J.-T.; Lalevee, J. Efficacy Modeling of New Multi-Functional Benzophenone-Based System for Free-Radical/Cationic Hybrid Photopolymerization Using 405 nm LED. J. Polym. Res. 2021, 29, 100. [Google Scholar] [CrossRef]
  85. Launay, V.; Wolf, R.; Dumur, F.; Lalevée, J. Photothermal Activation in the near Infrared Range for 4-Dimensional Printing Using Relevant Organic Dyes. Addit. Manuf. 2022, 58, 103031. [Google Scholar] [CrossRef]
  86. Xu, Y.; Feng, T.; Yang, T.; Wei, H.; Yang, H.; Li, G.; Zhao, M.; Liu, S.; Huang, W.; Zhao, Q. Utilizing Intramolecular Photoinduced Electron Transfer to Enhance Photothermal Tumor Treatment of Aza-BODIPY-Based Near-Infrared Nanoparticles. ACS Appl. Mater. Interfaces 2018, 10, 16299–16307. [Google Scholar] [CrossRef] [PubMed]
  87. Skotnicka, A.; Kabatc, J. New BODIPY Dyes Based on Benzoxazole as Photosensitizers in Radical Polymerization of Acrylate Monomers. Materials 2022, 15, 662. [Google Scholar] [CrossRef] [PubMed]
  88. Lu, P.; Chung, K.-Y.; Stafford, A.; Kiker, M.; Kafle, K.; Page, Z.A. Boron Dipyrromethene (BODIPY) in Polymer Chemistry. Polym. Chem. 2021, 12, 327–348. [Google Scholar] [CrossRef]
  89. Telitel, S.; Blanchard, N.; Schweizer, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; Lalevée, J. BODIPY Derivatives and Boranil as New Photoinitiating Systems of Cationic Polymerization Exhibiting a Tunable Absorption in the 400–600 nm Spectral Range. Polymer 2013, 54, 2071–2076. [Google Scholar] [CrossRef]
  90. Telitel, S.; Lalevée, J.; Blanchard, N.; Kavalli, T.; Tehfe, M.-A.; Schweizer, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P. Photopolymerization of Cationic Monomers and Acrylate/Divinylether Blends under Visible Light Using Pyrromethene Dyes. Macromolecules 2012, 45, 6864–6868. [Google Scholar] [CrossRef]
  91. Bonardi, A.; Bonardi, F.; Noirbent, G.; Dumur, F.; Dietlin, C.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Different NIR Dye Scaffolds for Polymerization Reactions under NIR Light. Polym. Chem. 2019, 10, 6505–6514. [Google Scholar] [CrossRef]
  92. Zhang, J.; Lalevée, J.; Zhao, J.; Graff, B.; Stenzel, M.H.; Xiao, P. Dihydroxyanthraquinone Derivatives: Natural Dyes as Blue-Light-Sensitive Versatile Photoinitiators of Photopolymerization. Polym. Chem. 2016, 7, 7316–7324. [Google Scholar] [CrossRef]
  93. Zhao, J.; Lalevée, J.; Lu, H.; MacQueen, R.; Kable, S.H.; Schmidt, T.W.; Stenzel, M.H.; Xiao, P. A New Role of Curcumin: As a Multicolor Photoinitiator for Polymer Fabrication under Household UV to Red LED Bulbs. Polym. Chem. 2015, 6, 5053–5061. [Google Scholar] [CrossRef]
  94. Crivello, J.V.; Bulut, U. Curcumin: A Naturally Occurring Long-Wavelength Photosensitizer for Diaryliodonium Salts. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 5217–5231. [Google Scholar] [CrossRef]
  95. Han, W.; Fu, H.; Xue, T.; Liu, T.; Wang, Y.; Wang, T. Facilely Prepared Blue-Green Light Sensitive Curcuminoids with Excellent Bleaching Properties as High Performance Photosensitizers in Cationic and Free Radical Photopolymerization. Polym. Chem. 2018, 9, 1787–1798. [Google Scholar] [CrossRef]
  96. Mishra, A.; Daswal, S. Curcumin, A Novel Natural Photoinitiator for the Copolymerization of Styrene and Methylmethacrylate. Null 2005, 42, 1667–1678. [Google Scholar] [CrossRef]
  97. Dumur, F. Recent Advances on Furan-Based Visible Light Photoinitiators of Polymerization. Catalysts 2023, 13, 493. [Google Scholar] [CrossRef]
  98. Telitel, S.; Dumur, F.; Faury, T.; Graff, B.; Tehfe, M.-A.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. New Core-Pyrene π Structure Organophotocatalysts Usable as Highly Efficient Photoinitiators. Beilstein J. Org. Chem. 2013, 9, 877–890. [Google Scholar] [CrossRef] [PubMed]
  99. Uchida, N.; Nakano, H.; Igarashi, T.; Sakurai, T. Nonsalt 1-(Arylmethyloxy)Pyrene Photoinitiators Capable of Initiating Cationic Polymerization. J. Appl. Polym. Sci. 2014, 131, 40510. [Google Scholar] [CrossRef]
  100. Mishra, A.; Daswal, S. 1-(Bromoacetyl)Pyrene, a Novel Photoinitiator for the Copolymerization of Styrene and Methylmethacrylate. Radiat. Phys. Chem. 2006, 75, 1093–1100. [Google Scholar] [CrossRef]
  101. Dumur, F. Recent Advances on Pyrene-Based Photoinitiators of Polymerization. Eur. Polym. J. 2020, 126, 109564. [Google Scholar] [CrossRef]
  102. Tasdelen, M.A.; Kumbaraci, V.; Jockusch, S.; Turro, N.J.; Talinli, N.; Yagci, Y. Photoacid Generation by Stepwise Two-Photon Absorption:  Photoinitiated Cationic Polymerization of Cyclohexene Oxide by Using Benzodioxinone in the Presence of Iodonium Salt. Macromolecules 2008, 41, 295–297. [Google Scholar] [CrossRef]
  103. Crivello, J.V.; Lam, J.H.W. Diaryliodonium Salts. A New Class of Photoinitiators for Cationic Polymerization. Macromolecules 1977, 10, 1307–1315. [Google Scholar] [CrossRef]
  104. He, Y.; Zhou, W.; Wu, F.; Li, M.; Wang, E. Photoreaction and Photopolymerization Studies on Squaraine Dyes/Iodonium Salts Combination. J. Photochem. Photobiol. A Chem. 2004, 162, 463–471. [Google Scholar] [CrossRef]
  105. Jun, L.I.; Miaozhen, L.I.; Huaihai, S.; Yongyuan, Y.; Erjian, W. Photopolymerization Initiated by Dimethylaminochalcone/Diphenyliodonium Salt Combination System Sensitive to Visible Light. Chin. J. Polym. Sci. 1993, 11, 163–170. [Google Scholar]
  106. Petko, F.; Galek, M.; Hola, E.; Topa-Skwarczyńska, M.; Tomal, W.; Jankowska, M.; Pilch, M.; Popielarz, R.; Graff, B.; Morlet-Savary, F.; et al. Symmetric Iodonium Salts Based on Benzylidene as One-Component Photoinitiators for Applications in 3D Printing. Chem. Mater. 2022, 34, 10077–10092. [Google Scholar] [CrossRef]
  107. Petko, F.; Galek, M.; Hola, E.; Popielarz, R.; Ortyl, J. One-Component Cationic Photoinitiators from Tunable Benzylidene Scaffolds for 3D Printing Applications. Macromolecules 2021, 54, 7070–7087. [Google Scholar] [CrossRef]
  108. Zhang, J.; Campolo, D.; Dumur, F.; Xiao, P.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. The Carbazole-Bound Ferrocenium Salt as a Specific Cationic Photoinitiator upon near-UV and Visible LEDs (365–405 nm). Polym. Bull. 2016, 73, 493–507. [Google Scholar] [CrossRef]
  109. Al Mousawi, A.; Garra, P.; Dumur, F.; Bui, T.-T.; Goubard, F.; Toufaily, J.; Hamieh, T.; Graff, B.; Gigmes, D.; Fouassier, J.P.; et al. Novel Carbazole Skeleton-Based Photoinitiators for LED Polymerization and LED Projector 3D Printing. Molecules 2017, 22, 2143. [Google Scholar] [CrossRef]
  110. Dumur, F. Recent Advances on Carbazole-Based Photoinitiators of Polymerization. Eur. Polym. J. 2020, 125, 109503. [Google Scholar] [CrossRef]
  111. Xu, C.; Gong, S.; Wu, X.; Wu, Y.; Liao, Q.; Xiong, Y.; Li, Z.; Tang, H. High-Efficient Carbazole-Based Photo-Bleachable Dyes as Free Radical Initiators for Visible Light Polymerization. Dyes Pigment. 2022, 198, 110039. [Google Scholar] [CrossRef]
  112. Dumur, F. Recent Advances on Carbazole-Based Oxime Esters as Photoinitiators of Polymerization. Eur. Polym. J. 2022, 175, 111330. [Google Scholar] [CrossRef]
  113. Liao, W.; Liao, Q.; Xiong, Y.; Li, Z.; Tang, H. Design, Synthesis and Properties of Carbazole-Indenedione Based Photobleachable Photoinitiators for Photopolymerization. J. Photochem. Photobiol. A Chem. 2023, 435, 114297. [Google Scholar] [CrossRef]
  114. Bin, F.-C.; Guo, M.; Li, T.; Zheng, Y.-C.; Dong, X.-Z.; Liu, J.; Jin, F.; Zheng, M.-L. Carbazole-Based Anion Ionic Water-Soluble Two-Photon Initiator for Achieving 3D Hydrogel Structures. Adv. Funct. Mater. 2023, 33, 2300293. [Google Scholar] [CrossRef]
  115. Corakci, B.; Hacioglu, S.O.; Toppare, L.; Bulut, U. Long Wavelength Photosensitizers in Photoinitiated Cationic Polymerization: The Effect of Quinoxaline Derivatives on Photopolymerization. Polymer 2013, 54, 3182–3187. [Google Scholar] [CrossRef]
  116. Pyszka, I.; Skowroński, Ł.; Jędrzejewska, B. Study on New Dental Materials Containing Quinoxaline-Based Photoinitiators in Terms of Exothermicity of the Photopolymerization Process. Int. J. Mol. Sci. 2023, 24, 2752. [Google Scholar] [CrossRef] [PubMed]
  117. Pyszka, I.; Jędrzejewska, B. Photoinitiation Abilities of Indeno- and Indoloquinoxaline Derivatives and Mechanical Properties of Dental Fillings Based on Multifunctional Acrylic Monomers and Glass Ionomer. Polymer 2023, 266, 125625. [Google Scholar] [CrossRef]
  118. Arsu, N.; Aydın, M. Photoinduced Free Radical Polymerization Initiated with Quinoxalines. Angew. Makromol. Chem. 1999, 270, 1–4. [Google Scholar] [CrossRef]
  119. Aydın, M.; Arsu, N. Photoinitiated Free Radical Polymerization of Methylmethacrylate by Using of Quinoxalines in the Presence of Aldehydes. Prog. Org. Coat. 2006, 56, 338–342. [Google Scholar] [CrossRef]
  120. Bulut, U.; Gunbas, G.E.; Toppare, L. A Quinoxaline Derivative as a Long Wavelength Photosensitizer for Diaryliodonium Salts. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 209–213. [Google Scholar] [CrossRef]
  121. Bulut, U.; Kolay, M.; Tarkuc, S.; Udum, Y.A.; Toppare, L. Quinoxaline Derivatives as Long Wavelength Photosensitizers in Photoinitiated Cationic Polymerization of Diaryliodonium Salts. Prog. Org. Coat. 2012, 73, 215–218. [Google Scholar] [CrossRef]
  122. Cao, X.; Jin, F.; Li, Y.-F.; Chen, W.-Q.; Duan, X.-M.; Yang, L.-M. Triphenylamine-Modified Quinoxaline Derivatives as Two-Photon Photoinitiators. New J. Chem. 2009, 33, 1578–1582. [Google Scholar] [CrossRef]
  123. Karaca Balta, D.; Keskin, S.; Karasu, F.; Arsu, N. Quinoxaline Derivatives as Photoinitiators in UV-Cured Coatings. Prog. Org. Coat. 2007, 60, 207–210. [Google Scholar] [CrossRef]
  124. Kucybała, Z.; Pyszka, I.; Pączkowski, J. Development of New Dyeing Photoinitiators for Free Radical Polymerization Based on the 1H-Pyrazolo[3,4-b]Quinoxaline Skeleton. Part 2. J. Chem. Soc. Perkin Trans. 2 2000, 7, 1559–1567. [Google Scholar] [CrossRef]
  125. Podsiadły, R.; Szymczak, A.M.; Podemska, K. The Synthesis of Novel, Visible-Wavelength, Oxidizable Polymerization Sensitizers Based on the 8-Halogeno-5,12-Dihydroquinoxalino[2,3-b]Quinoxaline Skeleton. Dyes Pigment. 2009, 82, 365–371. [Google Scholar] [CrossRef]
  126. Yao, J.-Y.; Hou, H.-H.; Ma, X.-D.; Xu, H.-J.; Shi, Z.-X.; Yin, J.; Jiang, X.-S. Combining Photo-Cleavable and Hydrogen-Abstracting Groups in Quinoxaline with Thioether Bond as Hybrid Photoinitiator. Chin. Chem. Lett. 2017, 28, 6–12. [Google Scholar] [CrossRef]
  127. Sun, L.; Jiang, X.; Yin, J. Study of Methoxyphenylquinoxalines (MOPQs) as Photoinitiators in the Negative Photo-Resist. Prog. Org. Coat. 2010, 67, 225–232. [Google Scholar] [CrossRef]
  128. Ercan, B.T.; Gultekin, S.S.; Yesil, T.; Dincalp, H.; Koyuncu, S.; Yagci, Y.; Zafer, C. Highly Conjugated Isoindigo and Quinoxaline Dyes as Sunlight Photosensitizers for Onium Salt-Photoinitiated Cationic Polymerization of Epoxy Resins. Polym. Int. 2022, 71, 867–873. [Google Scholar] [CrossRef]
  129. Dumur, F. Recent Advances on Glyoxylates and Related Structures as Photoinitiators of Polymerization. Macromol 2023, 3, 149–174. [Google Scholar] [CrossRef]
  130. Mousawi, A.A.; Dumur, F.; Garra, P.; Toufaily, J.; Hamieh, T.; Goubard, F.; Bui, T.-T.; Graff, B.; Gigmes, D.; Fouassier, J.P.; et al. Azahelicenes as Visible Light Photoinitiators for Cationic and Radical Polymerization: Preparation of Photoluminescent Polymers and Use in High Performance LED Projector 3D Printing Resins. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 1189–1199. [Google Scholar] [CrossRef]
  131. Noirbent, G.; Xu, Y.; Bonardi, A.-H.; Duval, S.; Gigmes, D.; Lalevée, J.; Dumur, F. New Donor-Acceptor Stenhouse Adducts as Visible and Near Infrared Light Polymerization Photoinitiators. Molecules 2020, 25, 2317. [Google Scholar] [CrossRef]
  132. Dumur, F. Recent Advances on Naphthoquinone-Based Photoinitiators of Polymerization. Eur. Polym. J. 2023, 193, 112120. [Google Scholar] [CrossRef]
  133. Garra, P.; Dumur, F.; Gigmes, D.; Al Mousawi, A.; Morlet-Savary, F.; Dietlin, C.; Fouassier, J.P.; Lalevée, J. Copper (Photo)Redox Catalyst for Radical Photopolymerization in Shadowed Areas and Access to Thick and Filled Samples. Macromolecules 2017, 50, 3761–3771. [Google Scholar] [CrossRef]
  134. Garra, P.; Dumur, F.; Morlet-Savary, F.; Dietlin, C.; Fouassier, J.P.; Lalevée, J. A New Highly Efficient Amine-Free and Peroxide-Free Redox System for Free Radical Polymerization under Air with Possible Light Activation. Macromolecules 2016, 49, 6296–6309. [Google Scholar] [CrossRef]
  135. Lin, J.-T.; Lalevee, J.; Cheng, D.-C. Efficacy Analysis of New Copper Complex for Visible Light (455, 530 nm) Radical/Cationic Photopolymerization: The Synergic Effects and Catalytic Cycle. PLoS ONE 2022, 17, e0270679. [Google Scholar] [CrossRef] [PubMed]
  136. Tehfe, M.-A.; Dumur, F.; Telitel, S.; Gigmes, D.; Contal, E.; Bertin, D.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; Lalevée, J. Zinc-Based Metal Complexes as New Photocatalysts in Polymerization Initiating Systems. Eur. Polym. J. 2013, 49, 1040–1049. [Google Scholar] [CrossRef]
  137. Nzulu, F.; Telitel, S.; Stoffelbach, F.; Graff, B.; Morlet-Savary, F.; Lalevée, J.; Fensterbank, L.; Goddard, J.-P.; 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] [CrossRef]
  138. Tehfe, M.-A.; Lepeltier, M.; Dumur, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Structural Effects in the Iridium Complex Series: Photoredox Catalysis and Photoinitiation of Polymerization Reactions under Visible Lights. Macromol. Chem. Phys. 2017, 218, 1700192. [Google Scholar] [CrossRef]
  139. Dumur, F.; Bertin, D.; Gigmes, D. Iridium (III) Complexes as Promising Emitters for Solid–State Light–Emitting Electrochemical Cells (LECs). Int. J. Nanotechnol. 2012, 9, 377–395. [Google Scholar] [CrossRef]
  140. Dumur, F.; Nasr, G.; Wantz, G.; Mayer, C.R.; Dumas, E.; Guerlin, A.; Miomandre, F.; Clavier, G.; Bertin, D.; Gigmes, D. Cationic Iridium Complex for the Design of Soft Salt-Based Phosphorescent OLEDs and Color-Tunable Light-Emitting Electrochemical Cells. Org. Electron. 2011, 12, 1683–1694. [Google Scholar] [CrossRef]
  141. Fouassier, J.P.; Morlet-Savary, F.; Yamashita, K.; Imahashi, S. The Role of the Dye/Iron Arene Complex/Amine System as a Photoinitiator for Photopolymerization Reactions. Polymer 1997, 38, 1415–1421. [Google Scholar] [CrossRef]
  142. Telitel, S.; Dumur, F.; Campolo, D.; Poly, J.; Gigmes, D.; Fouassier, J.P.; Lalevée, 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] [CrossRef]
  143. Bonardi, A.H.; Dumur, F.; Grant, T.M.; Noirbent, G.; Gigmes, D.; Lessard, B.H.; Fouassier, J.-P.; Lalevée, J. High Performance Near-Infrared (NIR) Photoinitiating Systems Operating under Low Light Intensity and in the Presence of Oxygen. Macromolecules 2018, 51, 1314–1324. [Google Scholar] [CrossRef]
  144. Bonardi, A.-H.; Bonardi, F.; Noirbent, G.; Dumur, F.; Gigmes, D.; Dietlin, C.; Lalevée, J. Free-Radical Polymerization upon near-Infrared Light Irradiation, Merging Photochemical and Photothermal Initiating Methods. J. Polym. Sci. 2020, 58, 300–308. [Google Scholar] [CrossRef]
  145. Lago, M.A.; de Quirós, A.R.-B.; Sendón, R.; Bustos, J.; Nieto, M.T.; Paseiro, P. Photoinitiators: A Food Safety Review. Food Addit. Contam. Part A 2015, 32, 779–798. [Google Scholar] [CrossRef] [PubMed]
  146. Hammoud, F.; Hijazi, A.; Schmitt, M.; Dumur, F.; Lalevée, J. A Review on Recently Proposed Oxime Ester Photoinitiators. Eur. Polym. J. 2023, 188, 111901. [Google Scholar] [CrossRef]
  147. Sun, K.; Xiao, P.; Dumur, F.; Lalevée, J. Organic Dye-Based Photoinitiating Systems for Visible-Light-Induced Photopolymerization. J. Polym. Sci. 2021, 59, 1338–1389. [Google Scholar] [CrossRef]
  148. Mokbel, H.; Toufaily, J.; Hamieh, T.; Dumur, F.; Campolo, D.; Gigmes, D.; Fouassier, J.P.; Ortyl, J.; Lalevée, J. Specific Cationic Photoinitiators for near UV and Visible LEDs: Iodonium versus Ferrocenium Structures. J. Appl. Polym. Sci. 2015, 132, 42759. [Google Scholar] [CrossRef]
  149. Dumur, F. Recent Advances on Diaryliodonium-Based Monocomponent Photoinitiating Systems. Eur. Polym. J. 2023, 195, 112193. [Google Scholar] [CrossRef]
  150. Boeira, P.O.; Meereis, C.T.W.; Suárez, C.E.C.; de Almeida, S.M.; Piva, E.; da Silveira Lima, G. Coumarin-Based Iodonium Hexafluoroantimonate as an Alternative Photoinitiator for Experimental Dental Adhesives Resin. Appl. Adhes. Sci. 2017, 5, 2. [Google Scholar] [CrossRef]
  151. Gómez, M.L.; Previtali, C.M.; Montejano, H.A. Two- and Three-Component Visible Light Photoinitiating Systems for Radical Polymerization Based on Onium Salts: An Overview of Mechanistic and Laser Flash Photolysis Studies. Int. J. Photoenergy 2012, 2012, 260728. [Google Scholar] [CrossRef]
  152. Prabhakaran, P.; Lee, K.-S. Photo-Polymerization. In Functional Polymers; Jafar Mazumder, M.A., Sheardown, H., Al-Ahmed, A., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 1–52. ISBN 978-3-319-95987-0. [Google Scholar]
  153. Crivello, J.V. The Discovery and Development of Onium Salt Cationic Photoinitiators. J. Polym. Sci. Part A Polym. Chem. 1999, 37, 4241–4254. [Google Scholar] [CrossRef]
  154. Crivello, J.V.; Lam, J.H.W. Photoinitiated Cationic Polymerization with Triarylsulfonium Salts. J. Polym. Sci. Polym. Chem. Ed. 1979, 17, 977–999. [Google Scholar] [CrossRef]
  155. Crivello, J.V.; Lam, J.H.W. A New Preparation of Triarylsulfonium and -Selenonium Salts via the Copper(II)-Catalyzed Arylation of Sulfides and Selenides with Diaryliodonium Salts. J. Org. Chem. 1978, 43, 3055–3058. [Google Scholar] [CrossRef]
  156. Sangermano, M.; Roppolo, I.; Chiappone, A. New Horizons in Cationic Photopolymerization. Polymers 2018, 10, 136. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, S.; Borjigin, T.; Schmitt, M.; Morlet-Savary, F.; Xiao, P.; Lalevée, J. High-Performance Photoinitiating Systems for LED-Induced Photopolymerization. Polymers 2023, 15, 342. [Google Scholar] [CrossRef] [PubMed]
  158. Martin, C.J.; Rapenne, G.; Nakashima, T.; Kawai, T. Recent Progress in Development of Photoacid Generators. J. Photochem. Photobiol. C Photochem. Rev. 2018, 34, 41–51. [Google Scholar] [CrossRef]
  159. Gachet, B.; Lecompère, M.; Croutxé-Barghorn, C.; Burr, D.; L’Hostis, G.; Allonas, X. Highly Reactive Photothermal Initiating System Based on Sulfonium Salts for the Photoinduced Thermal Frontal Cationic Polymerization of Epoxides: A Way to Create Carbon-Fiber Reinforced Polymers. RSC Adv. 2020, 10, 41915–41920. [Google Scholar] [CrossRef] [PubMed]
  160. Klikovits, N.; Knaack, P.; Bomze, D.; Krossing, I.; Liska, R. Novel Photoacid Generators for Cationic Photopolymerization. Polym. Chem. 2017, 8, 4414–4421. [Google Scholar] [CrossRef]
  161. Niu, X.-Z.; Pepel, R.D.; Paniego, R.; Field, J.A.; Chorover, J.; Abrell, L.; Sáez, A.E.; Sierra-Alvarez, R. Photochemical Fate of Sulfonium Photoacid Generator Cations under Photolithography Relevant UV Irradiation. J. Photochem. Photobiol. A Chem. 2021, 416, 113324. [Google Scholar] [CrossRef]
  162. Taschner, R.; Koch, T.; Wolff, R.; Stampfl, J.; Liska, R.; Knaack, P. Evaluation of Sulfonium Borate Initiators for Cationic Photopolymerization and Their Application in Hot Lithography. ACS Appl. Polym. Mater. 2023, 5, 3023–3033. [Google Scholar] [CrossRef]
  163. Peter Pappas, S.; Pappas, B.C.; Gatechair, L.R.; Jilek, J.H.; Schnabel, W. Photoinitiation of Cationic Polymerization. IV. Direct and Sensitized Photolysis of Aryl Iodonium and Sulfonium Salts. Polym. Photochem. 1984, 5, 1–22. [Google Scholar] [CrossRef]
  164. Crivello, J.V. Cationic Polymerization—Iodonium and Sulfonium Salt Photoinitiators. In Proceedings of the Initiators—Poly-Reactions—Optical Activity; Springer: Berlin/Heidelberg, Germany, 1984; pp. 1–48. [Google Scholar]
  165. Crivello, J.V.; Lee, J.L. Photosensitized Cationic Polymerizations Using Dialkylphenacylsulfonium and Dialkyl(4-Hydroxyphenyl)Sulfonium Salt Photoinitiators. Macromolecules 1981, 14, 1141–1147. [Google Scholar] [CrossRef]
  166. Saeva, F.D.; Breslin, D.T.; Luss, H.R. Intramolecular Photoinduced Rearrangements via Electron-Transfer-Induced, Concerted Bond Cleavage and Cation Radical/Radical Coupling. J. Am. Chem. Soc. 1991, 113, 5333–5337. [Google Scholar] [CrossRef]
  167. Jin, M.; Wu, X.; Malval, J.P.; Wan, D.; Pu, H. Dual Roles for Promoting Monomers to Polymers: A Conjugated Sulfonium Salt Photoacid Generator as Photoinitiator and Photosensitizer in Cationic Photopolymerization. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 2722–2730. [Google Scholar] [CrossRef]
  168. Denizligil, S.; Yagci, Y.; McArdle, C. Photochemically and Thermally Induced Radical Promoted Cationic Polymerization Using an Allylic Sulfonium Salt. Polymer 1995, 36, 3093–3098. [Google Scholar] [CrossRef]
  169. Denizligil, S.; Resul, R.; Yagci, Y.; Mc Ardle, C.; Fouassier, J.-P. Photosensitized Cationic Polymerization Using Allyl Sulfonium Salt. Macromol. Chem. Phys. 1996, 197, 1233–1240. [Google Scholar] [CrossRef]
  170. Wang, D.; Kaya, K.; Garra, P.; Fouassier, J.-P.; Graff, B.; Yagci, Y.; Lalevée, J. Sulfonium Salt Based Charge Transfer Complexes as Dual Thermal and Photochemical Polymerization Initiators for Composites and 3D Printing. Polym. Chem. 2019, 10, 4690–4698. [Google Scholar] [CrossRef]
  171. Hamazu, F.; Akashi, S.; Koizumi, T.; Takata, T.; Endo, T. Novel Benzyl Sulfonium Salt Having an Aromatic Group on Sulfur Atom as a Latent Thermal Initiator. J. Polym. Sci. Part A Polym. Chem. 1991, 29, 1675–1680. [Google Scholar] [CrossRef]
  172. Pappas, S.P. Kinetic Parameter Considerations for Maximizing Stability and Minimizing Cure Temperature of Thermo-Setting Coatings: Sulfonium Salts as Latent Thermal Initiators for Cationic Polymerization. J. Coat. Technol. 1981, 53, 43–51. [Google Scholar]
  173. Endo, T.; Arita, H. Polymerization of Spiroorthocarbonates by Sulfonium Salts, as Latent Thermal Catalysts. Makromol. Chem. Rapid Commun. 1985, 6, 137–139. [Google Scholar] [CrossRef]
  174. Toba, Y.; Yasuike, M.; Usui, Y. Visible laser polymerizations with the sulfonium borate styryl dyes as new photoinitiator systems. J. Photosci. Int. J. Off. Organ Korean Soc. Photosci. 1998, 5, 63–67. [Google Scholar]
  175. Kaya, K.; Kreutzer, J.; Yagci, Y. Diphenylphenacyl Sulfonium Salt as Dual Photoinitiator for Free Radical and Cationic Polymerizations. J. Polym. Sci. Part A Polym. Chem. 2018, 56, 451–457. [Google Scholar] [CrossRef]
  176. Saeva, F.D.; Garcia, E.; Martic, P.A. Comparative Photochemical Behavior of Some Anthracenyl and Naphthacenyl Sulfonium Salt Derivatives. J. Photochem. Photobiol. A Chem. 1995, 86, 149–154. [Google Scholar] [CrossRef]
  177. Saeva, F.D.; Breslin, D.T.; Martic, P.A. The Effect of the Nature of the Lowest Unoccupied Molecular Orbital (LUMO) of Some Arylmethylsulfonium Salt Derivatives on Photochemical and Redox Behavior. J. Am. Chem. Soc. 1989, 111, 1328–1330. [Google Scholar] [CrossRef]
  178. Zhou, W.; Kuebler, S.M.; Carrig, D.; Perry, J.W.; Marder, S.R. Efficient Photoacids Based upon Triarylamine Dialkylsulfonium Salts. J. Am. Chem. Soc. 2002, 124, 1897–1901. [Google Scholar] [CrossRef] [PubMed]
  179. Zivic, N.; Kuroishi, P.K.; Dumur, F.; Gigmes, D.; Dove, A.P.; Sardon, H. Recent Advances and Challenges in the Design of Organic Photoacid and Photobase Generators for Polymerizations. Angew. Chem. Int. Ed. 2019, 58, 10410–10422. [Google Scholar] [CrossRef] [PubMed]
  180. Dadashi-Silab, S.; Doran, S.; Yagci, Y. Photoinduced Electron Transfer Reactions for Macromolecular Syntheses. Chem. Rev. 2016, 116, 10212–10275. [Google Scholar] [CrossRef]
  181. Ichimura, K.; Horie, S.; Nagano, S. Solid-State Photoelectron Transfer in Powdery Nanocomposites Comprised of a Sensitiser, Photoacid Generators and Silica Nanoparticles. Phys. Chem. Chem. Phys. 2011, 13, 5974–5980. [Google Scholar] [CrossRef]
  182. Chen, S.; Cao, C.; Shen, X.; Qiu, Y.; Kuang, C.; Wan, D.; Jin, M. One/Two-Photon Sensitive Sulfonium Salt Photoinitiators Based on 1,3,5-Triphenyl-2-Pyrazoline. Eur. Polym. J. 2021, 153, 110525. [Google Scholar] [CrossRef]
  183. Mysliwiec, J.; Szukalski, A.; Sznitko, L.; Miniewicz, A.; Haupa, K.; Zygadlo, K.; Matczyszyn, K.; Olesiak-Banska, J.; Samoc, M. Synthesis, Optical and Nonlinear Optical Properties of New Pyrazoline Derivatives. Dyes Pigment. 2014, 102, 63–70. [Google Scholar] [CrossRef]
  184. Mori, T.; Saomoto, H.; Machitani, K.; Inoue, K.; Aoki, Y.; Koshitani, T.; Koumura, N.; Murakami, T.N. Influence of the Non-Conjugated 5-Position Substituent of 1,3,5-Triaryl-2-Pyrazoline-Based Photosensitizers on the Photophysical Properties and Performance of a Dye-Sensitized Solar Cell. RSC Adv. 2016, 6, 13964–13970. [Google Scholar] [CrossRef]
  185. Jin, M.; Hong, H.; Xie, J.; Malval, J.-P.; Spangenberg, A.; Soppera, O.; Wan, D.; Pu, H.; Versace, D.-L.; Leclerc, T.; et al. π-Conjugated Sulfonium-Based Photoacid Generators: An Integrated Molecular Approach for Efficient One and Two-Photon Polymerization. Polym. Chem. 2014, 5, 4747–4755. [Google Scholar] [CrossRef]
  186. Rehm, D.; Weller, A. Kinetics of Fluorescence Quenching by Electron and H-Atom Transfer. Isr. J. Chem. 1970, 8, 259–271. [Google Scholar] [CrossRef]
  187. Suppan, P.; Vauthey, E. The Energy Balance of Photoinduced Electron Transfer Reactions. J. Photochem. Photobiol. A Chem. 1989, 49, 239–248. [Google Scholar] [CrossRef]
  188. Verma, M.; Chaudhry, A.F.; Fahrni, C.J. Predicting the Photoinduced Electron Transfer Thermodynamics in Polyfluorinated 1,3,5-Triarylpyrazolines Based on Multiple Linear Free Energy Relationships. Org. Biomol. Chem. 2009, 7, 1536–1546. [Google Scholar] [CrossRef] [PubMed]
  189. Farid, S.; Dinnocenzo, J.P.; Merkel, P.B.; Young, R.H.; Shukla, D.; Guirado, G. Reexamination of the Rehm–Weller Data Set Reveals Electron Transfer Quenching That Follows a Sandros–Boltzmann Dependence on Free Energy. J. Am. Chem. Soc. 2011, 133, 11580–11587. [Google Scholar] [CrossRef] [PubMed]
  190. Taherpour, A.A.; Talebi-Haftadori, Z. Free Energies, Kinetics, and Photoelectron-Transfer Properties, and Theoretical and Quantitative Structural Relationship Studies of [SWCNT(5,5)-Armchair-CnH20][R] (R = H2-CmPd(Dppf), H2-CmPd(Dppr), and H2-CmPd(Dppcym)2, n = 20 to 300 and m = 60 and 70) Nanostructure Complexes. Int. Nano Lett. 2013, 3, 22. [Google Scholar] [CrossRef]
  191. Abdallah, M.; Bui, T.-T.; Goubard, F.; Theodosopoulou, D.; Dumur, F.; Hijazi, A.; Fouassier, J.-P.; Lalevée, J. Phenothiazine Derivatives as Photoredox Catalysts for Cationic and Radical Photosensitive Resins for 3D Printing Technology and Photocomposite Synthesis. Polym. Chem. 2019, 10, 6145–6156. [Google Scholar] [CrossRef]
  192. Dumur, F. Recent Advances on Visible Light Phenothiazine-Based Photoinitiators of Polymerization. Eur. Polym. J. 2022, 165, 110999. [Google Scholar] [CrossRef]
  193. Zhang, Y.; Morlet-Savary, F.; Schmitt, M.; Graff, B.; Rico, A.; Ibrahim-Ouali, M.; Dumur, F.; Lalevée, J. Photoinitiation Behavior of Phenothiazines Containing Two Oxime Ester Functionalities (OXEs) in Free Radical Photopolymerization and 3D Printing Application. Dyes Pigment. 2023, 215, 111202. [Google Scholar] [CrossRef]
  194. Deng, L.; Tang, L.; Qu, J. Novel Chalcone-Based Phenothiazine Derivative Photoinitiators for Visible Light Induced Photopolymerization with Photobleaching and Good Biocompatibility. Prog. Org. Coat. 2022, 167, 106859. [Google Scholar] [CrossRef]
  195. Gomurashvili, Z.; Crivello, J.V. Phenothiazine Photosensitizers for Onium Salt Photoinitiated Cationic Polymerization. In Photoinitiated Polymerization; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2003; Volume 847, pp. 231–241. ISBN 978-0-8412-3813-8. [Google Scholar]
  196. Gomurashvili, Z.; Crivello, J.V. Phenothiazine Photosensitizers for Onium Salt Photoinitiated Cationic Polymerization. J. Polym. Sci. Part A Polym. Chem. 2001, 39, 1187–1197. [Google Scholar] [CrossRef]
  197. Li, M.; Bao, B.; You, J.; Du, Y.; Li, D.; Zhan, H.; Zhang, L.; Wang, T. Phenothiazine Derivatives Based on Double Benzylidene Ketone Structure: Application to Red Light Photopolymerization. Prog. Org. Coat. 2022, 173, 107217. [Google Scholar] [CrossRef]
  198. Al-Ghamdi, S.N.; Al-Ghamdi, H.A.; El-Shishtawy, R.M.; Asiri, A.M. Advances in Phenothiazine and Phenoxazine-Based Electron Donors for Organic Dye-Sensitized Solar Cells. Dyes Pigment. 2021, 194, 109638. [Google Scholar] [CrossRef]
  199. Gangadhar, P.S.; Reddy, G.; Prasanthkumar, S.; Giribabu, L. Phenothiazine Functional Materials for Organic Optoelectronic Applications. Phys. Chem. Chem. Phys. 2021, 23, 14969–14996. [Google Scholar] [CrossRef] [PubMed]
  200. Huang, Z.-S.; Meier, H.; Cao, D. Phenothiazine-Based Dyes for Efficient Dye-Sensitized Solar Cells. J. Mater. Chem. C 2016, 4, 2404–2426. [Google Scholar] [CrossRef]
  201. Buene, A.F.; Almenningen, D.M. Phenothiazine and Phenoxazine Sensitizers for Dye-Sensitized Solar Cells—An Investigative Review of Two Complete Dye Classes. J. Mater. Chem. C 2021, 9, 11974–11994. [Google Scholar] [CrossRef]
  202. Mao, L.; Wu, Y.; Jiang, J.; Guo, X.; Heng, P.; Wang, L.; Zhang, J. Rational Design of Phenothiazine-Based Organic Dyes for Dye-Sensitized Solar Cells: The Influence of π-Spacers and Intermolecular Aggregation on Their Photovoltaic Performances. J. Phys. Chem. C 2020, 124, 9233–9242. [Google Scholar] [CrossRef]
  203. Kumar, A.; Kumar, P. Prediction of Power Conversion Efficiency of Phenothiazine-Based Dye-Sensitized Solar Cells Using Monte Carlo Method with Index of Ideality of Correlation. SAR QSAR Environ. Res. 2021, 32, 817–834. [Google Scholar] [CrossRef]
  204. Buene, A.F.; Christensen, M.; Hoff, B.H. Effect of Auxiliary Donors on 3,8-Phenothiazine Dyes for Dye-Sensitized Solar Cells. Molecules 2019, 24, 4485. [Google Scholar] [CrossRef]
  205. Slodek, A.; Zych, D.; Szafraniec-Gorol, G.; Gnida, P.; Vasylieva, M.; Schab-Balcerzak, E. Investigations of New Phenothiazine-Based Compounds for Dye-Sensitized Solar Cells with Theoretical Insight. Materials 2020, 13, 2292. [Google Scholar] [CrossRef] [PubMed]
  206. TamilSelvan, S.; Prakasam, A.; Venkatesh, G.; Kamal, C.; Sheena Mary, Y.; Parveen Banu, S.; Vennila, P.; Shyma Mary, Y. Synthesis, Spectral Characterizations, Molecular Geometries and Electronic Properties of Phenothiazine Based Organic Dyes for Dye-Sensitized Solar Cells. Z. Phys. Chem. 2021, 235, 1355–1380. [Google Scholar] [CrossRef]
  207. Pan, B.; Zhu, Y.-Z.; Qiu, C.; Wang, B.; Zheng, J.-Y. Synthesis of Phenothiazine Dyes Featuring Benzothiadiazole Unit for Efficient Dye-Sensitized Solar Cells. Acta Chim. Sin. 2018, 76, 215. [Google Scholar] [CrossRef]
  208. Nagane, S.; Ichake, A.; Agrawal, I.; Sadhanala, A.; Friend, R.H.; Ogale, S.; Wadgaonkar, P.P. Phenothiazine-Based D–A–π–A Dyes for Highly Efficient Dye-Sensitized Solar Cells: Effect of Internal Acceptor and Non-Conjugated π-Spacer on Device Performance. ChemPlusChem 2017, 82, 280–286. [Google Scholar] [CrossRef] [PubMed]
  209. Kafafy, H.; Wu, H.; Peng, M.; Hu, H.; Yan, K.; El-Shishtawy, R.M.; Zou, D. Steric and Solvent Effect in Dye-Sensitized Solar Cells Utilizing Phenothiazine-Based Dyes. Int. J. Photoenergy 2014, 2014, 548914. [Google Scholar] [CrossRef]
  210. Chiu, K.Y.; Tseng, Z.-L.; Chen, L.-C.; Shaw, H.; Chen, C.-J.; Chen, Y.-C.; Yang, S.-H.; Wu, W.-T.; Su, Y.O. D-π-A Type Phenothiazine Organic Dyes for Dye-Sensitized TiO2 Solar Cells. Sci. Adv. Mater. 2018, 10, 801–807. [Google Scholar] [CrossRef]
  211. Salunke, J.K.; Wong, F.L.; Feron, K.; Manzhos, S.; Lo, M.F.; Shinde, D.; Patil, A.; Lee, C.S.; Roy, V.A.L.; Sonar, P.; et al. Phenothiazine and Carbazole Substituted Pyrene Based Electroluminescent Organic Semiconductors for OLED Devices. J. Mater. Chem. C 2016, 4, 1009–1018. [Google Scholar] [CrossRef]
  212. He, J.; Xu, Y.; Luo, S.; Miao, J.; Cao, X.; Zou, Y. Phenoxazine and Phenothiazine Embedded Multi-Resonance Emitters for Highly Efficient Pure-Red OLEDs with Improved Color Purity. Chem. Eng. J. 2023, 471, 144565. [Google Scholar] [CrossRef]
  213. Khan, F.; Mahmoudi, M.; Volyniuk, D.; Grazulevicius, J.V.; Misra, R. Stimuli-Responsive Phenothiazine-S,S-Dioxide-Based Nondoped OLEDs with Color-Changeable Electroluminescence. J. Phys. Chem. C 2022, 126, 15573–15586. [Google Scholar] [CrossRef]
  214. Lade, J.J.; Chiou, S.-S.; Tseng, Y.-H.; Chiu, M.-H.; Lin, Y.-F.; Chang, C.-H.; Chetti, P.; Chaskar, A.C. Design and Synthesis of Novel Phenothiazine-Benzothiadiazine-1,1-Dioxide Hybrid Organic Material for OLED Applications. ChemistrySelect 2021, 6, 11029–11038. [Google Scholar] [CrossRef]
  215. Kim, B.; Lee, J.; Park, Y.; Lee, C.; Park, J.W. Highly Efficient New Hole Injection Materials for Organic Light Emitting Diodes Base on Phenothiazine Derivatives. J. Nanosci. Nanotechnol. 2014, 14, 6404–6408. [Google Scholar] [CrossRef] [PubMed]
  216. Qu, B.; Chen, Z.; Liu, Y.; Cao, H.; Xu, S.; Cao, S.; Lan, Z.; Wang, Z.; Gong, Q. Orange and Red Emitting OLEDs Based on Phenothiazine Polymers. J. Phys. D Appl. Phys. 2006, 39, 2680. [Google Scholar] [CrossRef]
  217. Aizawa, N.; Tsou, C.-J.; Park, I.S.; Yasuda, T. Aggregation-Induced Delayed Fluorescence from Phenothiazine-Containing Donor–Acceptor Molecules for High-Efficiency Non-Doped Organic Light-Emitting Diodes. Polym. J. 2017, 49, 197–202. [Google Scholar] [CrossRef]
  218. Baraket, F.; Pedras, B.; Torres, É.; Brites, M.J.; Dammak, M.; Berberan-Santos, M.N. Novel Phenoxazine-Benzonitrile and Phenothiazine-Benzonitrile Donor-Acceptor Molecules with Thermally Activated Delayed Fluorescence (TADF). Dyes Pigment. 2020, 175, 108114. [Google Scholar] [CrossRef]
  219. May, L.; Müller, T.J.J. Dithieno[1,4]Thiazines and Bis[1]Benzothieno[1,4]Thiazines—Organometallic Synthesis and Functionalization of Electron Density Enriched Congeners of Phenothiazine. Molecules 2020, 25, 2180. [Google Scholar] [CrossRef] [PubMed]
  220. Xie, F.-M.; Ou, Q.; Zhang, Q.; Zhang, J.-K.; Dai, G.-L.; Zhao, X.; Wei, H.-X. Two Novel Blue Phosphorescent Host Materials Containing Phenothiazine-5,5-Dioxide Structure Derivatives. Beilstein J. Org. Chem. 2018, 14, 869–874. [Google Scholar] [CrossRef] [PubMed]
  221. Park, Y.; Kim, B.; Lee, C.; Lee, J.; Lee, J.-H.; Park, J. High Efficiency New Hole Injection Materials for Organic Light Emitting Diodes Based on Dimeric Phenothiazine and Phenoxazine Moiety Derivatives. J. Nanosci. Nanotechnol. 2012, 12, 4356–4360. [Google Scholar] [CrossRef] [PubMed]
  222. Li, J.; Yang, S.; Deng, Z.; Islam, A.; Wu, S.; He, J.; Ni, S.; Dang, L.; Li, M.-D. Uncovering the Substituted-Position Effect on Excited-State Evolution of Benzophenone-Phenothiazine Dyads. J. Chem. Phys. 2023, 159, 144502. [Google Scholar] [CrossRef]
  223. Can, M.; Ali Havare, K. OLED Application of Pi-Conjugated Phenylimino Carboxylic Acid Organic Semiconductor Material. Eur. Phys. J. Appl. Phys. 2022, 97, 33. [Google Scholar] [CrossRef]
  224. Truong, N.T.T.; Nguyen, L.T.; Mai, H.L.T.; Doan, B.K.; Tran, D.H.; Truong, K.T.; Nguyen, V.Q.; Nguyen, L.-T.T.; Hoang, M.H.; Van Pham, T.; et al. Phenothiazine Derivatives, Diketopyrrolopyrrole-Based Conjugated Polymers: Synthesis, Optical and Organic Field Effect Transistor Properties. J. Polym. Res. 2020, 27, 223. [Google Scholar] [CrossRef]
  225. Al-Busaidi, I.J.; Haque, A.; Al Rasbi, N.K.; Khan, M.S. Phenothiazine-Based Derivatives for Optoelectronic Applications: A Review. Synth. Metals 2019, 257, 116189. [Google Scholar] [CrossRef]
  226. Lee, M.; Park, J.E.; Park, C.; Choi, H.C. Synthesis of a P-Type Semiconducting Phenothiazine Exfoliatable Layered Crystal. Langmuir 2013, 29, 9967–9971. [Google Scholar] [CrossRef] [PubMed]
  227. May, L.; Müller, T.J.J. Electron-Rich Phenothiazine Congeners and Beyond: Synthesis and Electronic Properties of Isomeric Dithieno[1,4]Thiazines. Chem. A Eur. J. 2020, 26, 12111–12118. [Google Scholar] [CrossRef] [PubMed]
  228. Liu, B.; Fan, S.; Huang, R.; Kawamoto, T.; Mori, T. Transistor Properties of Charge-Transfer Complexes─Combined Requirements from Energy Levels and Orbital Symmetry. J. Phys. Chem. C 2023, 127, 5125–5133. [Google Scholar] [CrossRef]
  229. Hwang, D.-H.; Kim, S.-K.; Park, M.-J.; Lee, J.-H.; Koo, B.-W.; Kang, I.-N.; Kim, S.-H.; Zyung, T. Conjugated Polymers Based on Phenothiazine and Fluorene in Light-Emitting Diodes and Field Effect Transistors. Chem. Mater. 2004, 16, 1298–1303. [Google Scholar] [CrossRef]
  230. Hu, Y.; Shang, Z.; Wang, J.; Hong, M.; Zhang, R.; Meng, Q.; Zhang, Z. A Phenothiazine-Based Turn-on Fluorescent Probe for the Selective Detection of Hydrogen Sulfide in Food, Live Cells and Animals. Analyst 2021, 146, 7528–7536. [Google Scholar] [CrossRef] [PubMed]
  231. Yan, L.; Li, D.; Le, Y.; Dong, P.; Liu, L. Phenothiazine-Based Fluorescent Probe for Fluoride Ions and Its Applications in Rapid Detection of Endemic Disease. Dyes Pigment. 2022, 201, 110200. [Google Scholar] [CrossRef]
  232. Zhang, C.; Wang, Y.; Li, X.; Nie, S.; Liu, C.; Zhang, Y.; Guo, J. A Fluorescent Probe Based on Phenothiazine for Detection of ClO− with Naked-Eye Color Change Properties. Anal. Biochem. 2023, 670, 115131. [Google Scholar] [CrossRef]
  233. Liu, Q.; Li, X.; Xiao, M.; Ai, Y.; Liu, G.; Ding, H.; Pu, S. A “Turn-on” Fluorescent Probe Based on Phenothiazine for Selectively Recognizing ClO− and Its Practical Applications. J. Fluoresc. 2023. [Google Scholar] [CrossRef]
  234. Zheng, Y.; Hou, P.; Li, Y.; Sun, J.; Cui, H.; Zhang, H.; Chen, S. A Phenothiazine-HPQ Based Fluorescent Probe with a Large Stokes Shift for Sensing Biothiols in Living Systems. Molecules 2021, 26, 2337. [Google Scholar] [CrossRef]
  235. Christian, S.T. The Interaction of Phenothiazine Derivatives with Synaptosomal Membranes: The Use of Fluorescent Probes as an Indicator of Drug-Induced Conformational Perturbations. Int. J. Neurosci. 1973, 6, 57–65. [Google Scholar] [CrossRef] [PubMed]
  236. Zhang, C.; Wang, Y.; Zhang, L.; Li, X.; Nie, S.; Liu, C. A Near-Infrared Fluorescent Probe Based on Phenothiazine for Rapid Detecting of CN− and ClO−. Opt. Mater. 2022, 133, 112959. [Google Scholar] [CrossRef]
  237. Al-Zahrani, F.A.M.; El-Shishtawy, R.M.; Asiri, A.M.; Al-Soliemy, A.M.; Mellah, K.A.; Ahmed, N.S.E.; Jedidi, A. A New Phenothiazine-Based Selective Visual and Fluorescent Sensor for Cyanide. BMC Chem. 2020, 14, 2. [Google Scholar] [CrossRef] [PubMed]
  238. Chen, W.; Yue, X.; Li, W.; Hao, Y.; Zhang, L.; Zhu, L.; Sheng, J.; Song, X. A Phenothiazine Coumarin-Based Red Emitting Fluorescent Probe for Nanomolar Detection of Thiophenol with a Large Stokes Shift. Sens. Actuators B Chem. 2017, 245, 702–710. [Google Scholar] [CrossRef]
  239. Rout, Y.; Montanari, C.; Pasciucco, E.; Misra, R.; Carlotti, B. Tuning the Fluorescence and the Intramolecular Charge Transfer of Phenothiazine Dipolar and Quadrupolar Derivatives by Oxygen Functionalization. J. Am. Chem. Soc. 2021, 143, 9933–9943. [Google Scholar] [CrossRef] [PubMed]
  240. Son, M.S.; Chang, S.H. Synthesis and Fluorescent Properties of New Rhodamine B Containing Phenothiazine Moiety. J. Environ. Sci. Int. 2015, 24, 955–960. [Google Scholar] [CrossRef]
  241. Zhu, Y.; Xu, D.; Zhang, Y.; Zhou, Y.; Yagci, Y.; Liu, R. Phenacyl Phenothiazinium Salt as a New Broad-Wavelength-Absorbing Photoinitiator for Cationic and Free Radical Polymerizations. Angew. Chem. Int. Ed. 2021, 60, 16917–16921. [Google Scholar] [CrossRef] [PubMed]
  242. Jockusch, S.; Yagci, Y. The Active Role of Excited States of Phenothiazines in Photoinduced Metal Free Atom Transfer Radical Polymerization: Singlet or Triplet Excited States? Polym. Chem. 2016, 7, 6039–6043. [Google Scholar] [CrossRef]
  243. Ladika, D.; Noirbent, G.; Dumur, F.; Gigmes, D.; Mourka, A.; Barmparis, G.D.; Farsari, M.; Gray, D. Synthesis and Application of Triphenylamine-Based Aldehydes as Photo-Initiators for Multi-Photon Lithography. Appl. Phys. A 2022, 128, 745. [Google Scholar] [CrossRef]
  244. Dumur, F. Recent Advances on Visible Light Triphenylamine-Based Photoinitiators of Polymerization. Eur. Polym. J. 2022, 166, 111036. [Google Scholar] [CrossRef]
  245. Mahmood, A. Triphenylamine Based Dyes for Dye Sensitized Solar Cells: A Review. Solar Energy 2016, 123, 127–144. [Google Scholar] [CrossRef]
  246. Holliman, P.J.; Mohsen, M.; Connell, A.; Kershaw, C.P.; Meza-Rojas, D.; Jones, E.W.; Geatches, D.; Sen, K.; Hsiao, Y.-W. Double Linker Triphenylamine Dyes for Dye-Sensitized Solar Cells. Energies 2020, 13, 4637. [Google Scholar] [CrossRef]
  247. Mahmoud, S.E.; Fadda, A.A.; Abdel-Latif, E.; Elmorsy, M.R. Synthesis of Novel Triphenylamine-Based Organic Dyes with Dual Anchors for Efficient Dye-Sensitized Solar Cells. Nanoscale Res. Lett. 2022, 17, 71. [Google Scholar] [CrossRef] [PubMed]
  248. Pakravesh, F.; Izadyar, M.; Arkan, F. Molecular Engineering of Triphenylamine-Based Metal-Free Organic Dyes for Dye-Sensitized Solar Cells. Int. J. Quantum Chem. 2021, 121, e26620. [Google Scholar] [CrossRef]
  249. Zhou, W.; Kuebler, S.M.; Braun, K.L.; Yu, T.; Cammack, J.K.; Ober, C.K.; Perry, J.W.; Marder, S.R. An Efficient Two-Photon-Generated Photoacid Applied to Positive-Tone 3D Microfabrication. Science 2002, 296, 1106–1109. [Google Scholar] [CrossRef] [PubMed]
  250. Kuebler, S.M.; Braun, K.L.; Zhou, W.; Cammack, J.K.; Yu, T.; Ober, C.K.; Marder, S.R.; Perry, J.W. Design and Application of High-Sensitivity Two-Photon Initiators for Three-Dimensional Microfabrication. J. Photochem. Photobiol. A Chem. 2003, 158, 163–170. [Google Scholar] [CrossRef]
  251. Yu, T.; Ober, C.K.; Kuebler, S.M.; Zhou, W.; Marder, S.R.; Perry, J.W. Chemically Amplified Positive Resists for Two-Photon Three-Dimensional Microfabrication. Adv. Mater. 2003, 15, 517–521. [Google Scholar] [CrossRef]
  252. Pohlers, G.; Scaiano, J.C.; Sinta, R. A Novel Photometric Method for the Determination of Photoacid Generation Efficiencies Using Benzothiazole and Xanthene Dyes as Acid Sensors. Chem. Mater. 1997, 9, 3222–3230. [Google Scholar] [CrossRef]
  253. Gu, H.; Ren, K.; Grinevich, O.; Malpert, J.H.; Neckers, D.C. Characterization of Iodonium Salts Differing in the Anion. J. Org. Chem. 2001, 66, 4161–4164. [Google Scholar] [CrossRef]
  254. Park, J.; Kihara, N.; Ikeda, T.; Endo, T. Photo-Induced Cationic Ring-Opening Polymerization of 2-Alkenyl-4-Methylene-1,3-Dioxolanes by Benzylsulfonium Salt. Macromolecules 1997, 30, 3414–3416. [Google Scholar] [CrossRef]
  255. Yanez, C.O.; Andrade, C.D.; Belfield, K.D. Characterization of Novel Sulfonium Photoacid Generators and Their Microwave-Assisted Synthesis. Chem. Commun. 2009, 827–829. [Google Scholar] [CrossRef] [PubMed]
  256. Yanez, C.O.; Andrade, C.D.; Yao, S.; Luchita, G.; Bondar, M.V.; Belfield, K.D. Photosensitive Polymeric Materials for Two-Photon 3D WORM Optical Data Storage Systems. ACS Appl. Mater. Interfaces 2009, 1, 2219–2229. [Google Scholar] [CrossRef] [PubMed]
  257. Steidl, L.; Jhaveri, S.J.; Ayothi, R.; Sha, J.; McMullen, J.D.; Ng, S.Y.C.; Zipfel, W.R.; Zentel, R.; Ober, C.K. Non-Ionic Photo-Acid Generators for Applications in Two-Photon Lithography. J. Mater. Chem. 2009, 19, 505–513. [Google Scholar] [CrossRef]
  258. Jin, M.; Xu, H.; Hong, H.; Malval, J.-P.; Zhang, Y.; Ren, A.; Wan, D.; Pu, H. Design of D–π–A Type Photoacid Generators for High Efficiency Excitation at 405 nm and 800 nm. Chem. Commun. 2013, 49, 8480–8482. [Google Scholar] [CrossRef] [PubMed]
  259. Xia, R.; Malval, J.-P.; Jin, M.; Spangenberg, A.; Wan, D.; Pu, H.; Vergote, T.; Morlet-Savary, F.; Chaumeil, H.; Baldeck, P.; et al. Enhancement of Acid Photogeneration Through a Para-to-Meta Substitution Strategy in a Sulfonium-Based Alkoxystilbene Designed for Two-Photon Polymerization. Chem. Mater. 2012, 24, 237–244. [Google Scholar] [CrossRef]
  260. Wu, X.; Jin, M.; Xie, J.; Malval, J.P.; Wan, D. One/Two-Photon Cationic Polymerization in Visible and near Infrared Ranges Using Two-Branched Sulfonium Salts as Efficient Photoacid Generators. Dyes Pigment. 2016, 133, 363–371. [Google Scholar] [CrossRef]
  261. Jin, M.; Wu, X.; Xie, J.; Malval, J.P.; Wan, D. One/Two-Photon-Sensitive Photoacid Generators Based on Benzene Oligomer-Containing D–π–A-Type Aryl Dialkylsulfonium Salts. RSC Adv. 2015, 5, 55340–55347. [Google Scholar] [CrossRef]
  262. Wua, X.; Jin, M.; Xie, J.; Wan, D.; Malval, J.P. Effects of Conjugated Systems on UV-Visible Light-Sensitive D-π-A Type Sulfonium Salt Photoacid Generators. Chin. J. Polym. Sci. 2016, 34, 1456–1468. [Google Scholar] [CrossRef]
  263. Wang, C.; Meng, X.; Li, Z.; Li, M.; Jin, M.; Liu, R.; Yagci, Y. Chemiluminescence Induced Cationic Photopolymerization Using Sulfonium Salt. ACS Macro Lett. 2020, 9, 471–475. [Google Scholar] [CrossRef]
  264. Meng, X.; Li, L.; Huang, Y.; Deng, X.; Liu, X.; Li, Z. Upconversion Nanoparticle-Assisted Cationic and Radical/Cationic Hybrid Photopolymerization Using Sulfonium Salts. Polym. Chem. 2021, 12, 7005–7009. [Google Scholar] [CrossRef]
  265. Méndez-Ramos, J.; Ruiz-Morales, J.C.; Acosta-Mora, P.; Khaidukov, N.M. Infrared-Light Induced Curing of Photosensitive Resins through Photon up-Conversion for Novel Cost-Effective Luminescent 3D-Printing Technology. J. Mater. Chem. C 2016, 4, 801–806. [Google Scholar] [CrossRef]
  266. Hu, P.; Xu, H.; Pan, Y.; Sang, X.; Liu, R. Upconversion Particle-Assisted NIR Polymerization Enables Microdomain Gradient Photopolymerization at Inter-Particulate Length Scale. Nat. Commun. 2023, 14, 3653. [Google Scholar] [CrossRef] [PubMed]
  267. Rocheva, V.V.; Koroleva, A.V.; Savelyev, A.G.; Khaydukov, K.V.; Generalova, A.N.; Nechaev, A.V.; Guller, A.E.; Semchishen, V.A.; Chichkov, B.N.; Khaydukov, E.V. High-Resolution 3D Photopolymerization Assisted by Upconversion Nanoparticles for Rapid Prototyping Applications. Sci. Rep. 2018, 8, 3663. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, Q.; Boniface, A.; Parashar, V.K.; Gijs, M.A.M.; Moser, C. Multi-Photon Polymerization Using Upconversion Nanoparticles for Tunable Feature-Size Printing. Nanophotonics 2023, 12, 1527–1536. [Google Scholar] [CrossRef]
  269. Zhakeyev, A.; Abdulrhman, M.; Zhang, Y.; Li, F.; Chen, G.; Marques-Hueso, J. Upconversion 3D Printing Enables Single-Immersion Multi-Material Stereolithography. Appl. Mater. Today 2023, 32, 101854. [Google Scholar] [CrossRef]
  270. Hermes, P.; Hermsen, A.; Jäger, M.; Gutmann, J.S.; Strehmel, V.; Strehmel, B. Challenges and Limits of Upconversion Nanoparticles for Cationic Photopolymerization with UV Initiators Excited at 980 nm. Polym. Chem. 2022, 13, 4879–4886. [Google Scholar] [CrossRef]
  271. Limberg, D.K.; Kang, J.-H.; Hayward, R.C. Triplet–Triplet Annihilation Photopolymerization for High-Resolution 3D Printing. J. Am. Chem. Soc. 2022, 144, 5226–5232. [Google Scholar] [CrossRef] [PubMed]
  272. Tehfe, M.A.; Louradour, F.; Lalevée, J.; Fouassier, J.-P. Photopolymerization Reactions: On the Way to a Green and Sustainable Chemistry. Appl. Sci. 2013, 3, 490–514. [Google Scholar] [CrossRef]
  273. Hola, E.; Morlet-Savary, F.; Lalevée, J.; Ortyl, J. Photoinitiator or Photosensitizer? Dual Behaviour of m-Terphenyls in Photopolymerization Processes. Eur. Polym. J. 2023, 189, 111971. [Google Scholar] [CrossRef]
  274. Samoilenko, T.F.; Iarova, N.V.; Ostapiuk, S.M.; Tkalich, M.H.; Vorontsova, L.O.; Klymchuk, D.O.; Brovko, O.O. Influence of N-Vynilcarbazole on the Photopolymerization Process and Properties of Epoxy-Acrylate Interpenetrating Polymer Networks. e-Polymers 2016, 16, 429–435. [Google Scholar] [CrossRef]
  275. Gallastegui, A.; Dominguez-Alfaro, A.; Lezama, L.; Alegret, N.; Prato, M.; Gómez, M.L.; Mecerreyes, D. Fast Visible-Light Photopolymerization in the Presence of Multiwalled Carbon Nanotubes: Toward 3D Printing Conducting Nanocomposites. ACS Macro Lett. 2022, 11, 303–309. [Google Scholar] [CrossRef] [PubMed]
  276. Xiao, P.; Lalevée, J.; Zhao, J.; Stenzel, M.H. N-Vinylcarbazole as Versatile Photoinaddimer of Photopolymerization under Household UV LED Bulb (392 nm). Macromol. Rapid Commun. 2015, 36, 1675–1680. [Google Scholar] [CrossRef] [PubMed]
  277. Lalevée, J.; Dirani, A.; El-Roz, M.; Allonas, X.; Fouassier, J.P. Germanes as Efficient Coinitiators in Radical and Cationic Photopolymerizations. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 3042–3047. [Google Scholar] [CrossRef]
  278. Fouassier, J.P.; Lalevée, J. Photochemical Production of Interpenetrating Polymer Networks; Simultaneous Initiation of Radical and Cationic Polymerization Reactions. Polymers 2014, 6, 2588–2610. [Google Scholar] [CrossRef]
  279. Li, Y.; Shaukat, U.; Schlögl, S.; Xue, T.; Li, J.; Nie, J.; Zhu, X. A Pyrrole–Carbazole Photoinitiator for Radical and Cationic Visible Light LED Photopolymerization. Eur. Polym. J. 2023, 182, 111700. [Google Scholar] [CrossRef]
  280. Motiekaityte, G.; Navaruckiene, A.; Raudoniene, V.; Bridziuviene, D.; Jaras, J.; Kantminiene, K.; Ostrauskaite, J. Antimicrobial Dual-Cured Photopolymers of Vanillin Alcohol Diglycidyl Ether and Glycerol Dimethacrylate. J. Appl. Polym. Sci. 2023, 140, e53289. [Google Scholar] [CrossRef]
  281. Urano, T.; Ohno-Okumura, E.; Sakamoto, K.; Suzuki, S.; Yamaoka, T. Effect of Ethyl Group in Sensitization Mechanisms of Pyrromethene Dyes in Photopolymer Coating Layer. J. Photopolym. Sci. Technol. 2000, 13, 673–677. [Google Scholar] [CrossRef]
  282. Sambath, K.; Wan, Z.; Wang, Q.; Chen, H.; Zhang, Y. BODIPY-Based Photoacid Generators for Light-Induced Cationic Polymerization. Org. Lett. 2020, 22, 1208–1212. [Google Scholar] [CrossRef]
  283. Zhang, L.; Li, L.; Chen, Y.; Pi, J.; Liu, R.; Zhu, Y. Recent Advances and Challenges in Long Wavelength Sensitive Cationic Photoinitiating Systems. Polymers 2023, 15, 2524. [Google Scholar] [CrossRef]
  284. Wang, X.; Saeva, F.D.; Kampmeier, J.A. Photosensitized Reduction of Sulfonium Salts:  Evidence for Nondissociative Electron Transfer. J. Am. Chem. Soc. 1999, 121, 4364–4368. [Google Scholar] [CrossRef]
  285. Garra, P.; Fouassier, J.P.; Lakhdar, S.; Yagci, Y.; Lalevée, J. Visible Light Photoinitiating Systems by Charge Transfer Complexes: Photochemistry without Dyes. Prog. Polym. Sci. 2020, 107, 101277. [Google Scholar] [CrossRef]
  286. Garra, P.; Graff, B.; Morlet-Savary, F.; Dietlin, C.; Becht, J.-M.; Fouassier, J.-P.; Lalevée, J. Charge Transfer Complexes as Pan-Scaled Photoinitiating Systems: From 50 μm 3D Printed Polymers at 405 nm to Extremely Deep Photopolymerization (31 cm). Macromolecules 2018, 51, 57–70. [Google Scholar] [CrossRef]
  287. Kaya, K.; Kreutzer, J.; Yagci, Y. A Charge-Transfer Complex of Thioxanthonephenacyl Sulfonium Salt as a Visible-Light Photoinitiator for Free Radical and Cationic Polymerizations. ChemPhotoChem 2019, 3, 1187–1192. [Google Scholar] [CrossRef]
  288. Wu, X.; Malval, J.; Wan, D.; Jin, M. D-π-A-Type Aryl Dialkylsulfonium Salts as One-Component Versatile Photoinitiators under UV/Visible LEDs Irradiation. Dyes Pigment. 2016, 132, 128–135. [Google Scholar] [CrossRef]
  289. Jeon, N.J.; Na, H.; Jung, E.H.; Yang, T.-Y.; Lee, Y.G.; Kim, G.; Shin, H.-W.; Il Seok, S.; Lee, J.; Seo, J. A Fluorene-Terminated Hole-Transporting Material for Highly Efficient and Stable Perovskite Solar Cells. Nat. Energy 2018, 3, 682–689. [Google Scholar] [CrossRef]
  290. Suman; Bagui, A.; Datt, R.; Gupta, V.; Singh, S.P. A Simple Fluorene Core-Based Non-Fullerene Acceptor for High Performance Organic Solar Cells. Chem. Commun. 2017, 53, 12790–12793. [Google Scholar] [CrossRef] [PubMed]
  291. Ma, Z.; Dong, Y.; Su, Y.-J.; Yu, R.; Gao, H.; Gong, Y.; Lee, Z.-Y.; Yang, C.; Hsu, C.-S.; Tan, Z. Morphological Stabilization in Organic Solar Cells via a Fluorene-Based Crosslinker for Enhanced Efficiency and Thermal Stability. ACS Appl. Mater. Interfaces 2022, 14, 1187–1194. [Google Scholar] [CrossRef]
  292. Hua, Y.; Zhang, J.; Xu, B.; Liu, P.; Cheng, M.; Kloo, L.; Johansson, E.M.J.; Sveinbjörnsson, K.; Aitola, K.; Boschloo, G.; et al. Facile Synthesis of Fluorene-Based Hole Transport Materials for Highly Efficient Perovskite Solar Cells and Solid-State Dye-Sensitized Solar Cells. Nano Energy 2016, 26, 108–113. [Google Scholar] [CrossRef]
  293. Suman; Kovvuri, J.; Islavath, N. Molecular Modifications in Fluorene Core for Efficient Organic Photovoltaic Cells. J. Photochem. Photobiol. A Chem. 2024, 446, 115162. [Google Scholar] [CrossRef]
  294. Nagarjuna, P.; Chaturvedi, N.; Bagui, A.; Runjhun, R.; Garg, A.; Singh, S.P. Solution-Processed Organic Solar Cells Using New Electron Acceptor Derived from Naphthalene and Fluorene Unit. ChemistrySelect 2017, 2, 7913–7917. [Google Scholar] [CrossRef]
  295. Kumar, D.; Justin Thomas, K.R.; Lee, C.-P.; Ho, K.-C. Organic Dyes Containing Fluorene Decorated with Imidazole Units for Dye-Sensitized Solar Cells. J. Org. Chem. 2014, 79, 3159–3172. [Google Scholar] [CrossRef]
  296. Kacimi, R.; Abram, T.; Bourass, M.; Bejjit, L.; Alimi, K.; Bouachrine, M. Molecular Design of D–A–D Conjugated Molecules Based on Fluorene for Organic Solar Cells. Opt. Quantum Electron. 2019, 51, 76. [Google Scholar] [CrossRef]
  297. Gayathri, R.D.; Gokulnath, T.; Park, H.-Y.; Xie, Z.; Jin, S.-H.; Han, S.C.; Lee, J.W. Facile and Stable Fluorene Based Organic Hole Transporting Materials for Efficient Perovskite Solar Cells. Macromole. Res. 2022, 30, 745–750. [Google Scholar] [CrossRef]
  298. Justin Thomas, K.R.; Baheti, A. Fluorene Based Organic Dyes for Dye Sensitised Solar Cells: Structure–Property Relationships. Mater. Technol. 2013, 28, 71–87. [Google Scholar] [CrossRef]
  299. Zhang, X.; Chen, X.; Huang, S.; Peng, X. Fluorene Dimers as the Cathode Interlayers in Organic Solar Cells. Synth. Metals 2019, 253, 110–115. [Google Scholar] [CrossRef]
  300. Kim, K.-S.; Jeon, Y.-M.; Kim, J.-W.; Lee, C.-W.; Gong, M.-S. Blue Light-Emitting OLED Using New Spiro[Fluorene-7,9′-Benzofluorene] Host and Dopant Materials. Org. Electron. 2008, 9, 797–804. [Google Scholar] [CrossRef]
  301. Li, G.; Zheng, J.; Klimes, K.; Zhu, Z.-Q.; Wu, J.; Zhu, H.; Li, J. Novel Carbazole/Fluorene-Based Host Material for Stable and Efficient Phosphorescent OLEDs. ACS Appl. Mater. Interfaces 2019, 11, 40320–40331. [Google Scholar] [CrossRef]
  302. Sun, W.; Zhou, N.; Xiao, Y.; Wang, S.; Li, X. Novel Carbazolyl-Substituted Spiro[Acridine-9,9′-Fluorene] Derivatives as Deep-Blue Emitting Materials for OLED Applications. Dyes Pigment. 2018, 154, 30–37. [Google Scholar] [CrossRef]
  303. Lin, Y.H.; Lin, W.-H.; Huang, Y.-S.; Wu, C.-H.; Gnanasekaran, P.; Chang, Y.-M.; Teng, S.-W.; Lu, C.-W.; Chang, C.-H.; Chang, Y.J. Highly Efficient (EQE > 27%) Yellow OLEDs Using Spiro[Fluorene-9,9′-Phenanthren-10′-One]-Carbazole-Based Donor–Acceptor–Donor Host Materials. J. Mater. Chem. C 2023, 11, 3101–3111. [Google Scholar] [CrossRef]
  304. Lozano-Hernández, L.-A.; Cameron, J.; Riggs, C.J.; Reineke, S.; Skabara, P.J. A Study on Varying the Number of Fluorene Units in 2, 1, 3-Benzothiadiazole-Containing Oligomers and the Effect on OLED Performance and Stability. ChemPhotoChem 2023, 7, e202200256. [Google Scholar] [CrossRef]
  305. Lee, J.; Kim, B.; Park, Y.; Kim, S.; Park, J. Fluorine Effects in New Indenofluorenedione Derivatives for Electron Transporting Layer in OLED Devices. J. Nanosci. Nanotechnol. 2014, 14, 6431–6434. [Google Scholar] [CrossRef] [PubMed]
  306. Braveenth, R.; Jung, H.; Kim, K.; Kim, B.M.; Bae, I.-J.; Kim, M.; Chai, K.Y. Fluorene–Triphenylamine-Based Bipolar Materials: Fluorescent Emitter and Host for Yellow Phosphorescent OLEDs. Appl. Sci. 2020, 10, 519. [Google Scholar] [CrossRef]
  307. Kim, D.; Reddy, M.R.; Cho, K.; Ho, D.; Kim, C.; Seo, S. Solution-Processable Fluorene Derivative for Organic Thin-Film Transistors. Org. Electron. 2020, 76, 105464. [Google Scholar] [CrossRef]
  308. Lee, C.-H.; Lai, Y.-Y.; Cao, F.-Y.; Hsu, J.-Y.; Lin, Z.-L.; Jeng, U.-S.; Su, C.-J.; Cheng, Y.-J. Synthesis, Molecular and Photovoltaic/Transistor Properties of Heptacyclic Ladder-Type Di(Thienobenzo)Fluorene-Based Copolymers. J. Mater. Chem. C 2016, 4, 11427–11435. [Google Scholar] [CrossRef]
  309. Dong, L.; Sun, H.-S.; Wang, J.-T.; Lee, W.-Y.; Chen, W.-C. Fluorene Based Donor-Acceptor Polymer Electrets for Nonvolatile Organic Transistor Memory Device Applications. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 602–614. [Google Scholar] [CrossRef]
  310. Cho, K.; Reddy, M.R.; Kim, D.; Ho, D.; Yun, C.; Seo, S.; Kim, C. Synthesis and Characterization of Fluorene Derivatives as Organic Semiconductors for Organic Field-Effect Transistor. Mol. Cryst. Liquid Cryst. 2019, 690, 56–66. [Google Scholar] [CrossRef]
  311. Koshiba, Y.; Kato, T.; Misaki, M.; Ishida, K.; Torii, M.; Kato, T.; Tsutsui, K.; Tanigaki, N.; Yase, K.; Ueda, Y. Fabrication of One-Dimensionally Oriented Fluorene–Thiophene Copolymer Thin Films and Anisotropic Transistor Characteristics. Jpn. J. Appl. Phys. 2010, 49, 01AE13. [Google Scholar] [CrossRef]
  312. Kim, Z.-S.; Lim, S.C.; Kim, S.H.; Yang, Y.S.; Hwang, D.-H. Biotin-Functionalized Semiconducting Polymer in an Organic Field Effect Transistor and Application as a Biosensor. Sensors 2012, 12, 11238–11248. [Google Scholar] [CrossRef]
  313. Noh, Y.-Y.; Kim, D.-Y.; Yoshida, Y.; Yase, K.; Jung, B.-J.; Lim, E.; Shim, H.-K.; Azumi, R. Keto Defect Sites in Fluorene-Based Organic Field-Effect Transistors: The Origin of Rapid Degradation on the Performance of the Device. J. Appl. Phys. 2005, 97, 104504. [Google Scholar] [CrossRef]
  314. Lee, W.-Y.; Chen, C.-W.; Chueh, C.-C.; Yang, C.-C.; Chen, W.-C. Synthesis of New Fluorene-Indolocarbazole Alternating Copolymers for Light-Emitting Diodes and Field Effect Transistors. Polym. J. 2008, 40, 249–255. [Google Scholar] [CrossRef]
  315. Heeney, M.; Bailey, C.; Giles, M.; Shkunov, M.; Sparrowe, D.; Tierney, S.; Zhang, W.; McCulloch, I. Alkylidene Fluorene Liquid Crystalline Semiconducting Polymers for Organic Field Effect Transistor Devices. Macromolecules 2004, 37, 5250–5256. [Google Scholar] [CrossRef]
  316. Schafer Hales, K.J.; Belfield, K.D.; Yao, S.; Frederiksen, P.K.; Hales, J.M.; Kolattukudy, P.E. Fluorene-Based Fluorescent Probes with High Two-Photon Action Cross-Sections for Biological Multiphoton Imaging Applications. J. Biomed. Opt. 2005, 10, 051402. [Google Scholar] [CrossRef] [PubMed]
  317. Shaya, J.; Corridon, P.R.; Al-Omari, B.; Aoudi, A.; Shunnar, A.; Mohideen, M.I.H.; Qurashi, A.; Michel, B.Y.; Burger, A. Design, Photophysical Properties, and Applications of Fluorene-Based Fluorophores in Two-Photon Fluorescence Bioimaging: A Review. J. Photochem. Photobiol. C Photochem. Rev. 2022, 52, 100529. [Google Scholar] [CrossRef]
  318. Yang, T.; Zuo, Y.; Zhang, Y.; Gou, Z.; Wang, X.; Lin, W. Novel Fluorene-Based Fluorescent Probe with Excellent Stability for Selective Detection of SCN− and Its Applications in Paper-Based Sensing and Bioimaging. J. Mater. Chem. B 2019, 7, 4649–4654. [Google Scholar] [CrossRef] [PubMed]
  319. Chen, H.; Wang, X.; Ren, H.; Lu, M.; Cui, X.; Wang, L.; Wang, H.; Tang, Q. Fluorene-Based Polymers of Intrinsic Microporosity as Fluorescent Probes for Metal Ions. React. Funct. Polym. 2022, 181, 105431. [Google Scholar] [CrossRef]
  320. Xochitiotzi-Flores, E.; Jiménez-Sánchez, A.; García-Ortega, H.; Sánchez-Puig, N.; Romero-Ávila, M.; Santillan, R.; Farfán, N. Optical Properties of Two Fluorene Derived BODIPY Molecular Rotors as Fluorescent Ratiometric Viscosity Probes. New J. Chem. 2016, 40, 4500–4512. [Google Scholar] [CrossRef]
  321. Morales, A.R.; Schafer-Hales, K.J.; Marcus, A.I.; Belfield, K.D. Amine-Reactive Fluorene Probes: Synthesis, Optical Characterization, Bioconjugation, and Two-Photon Fluorescence Imaging. Bioconjug. Chem. 2008, 19, 2559–2567. [Google Scholar] [CrossRef]
  322. Lala, A.K.; Koppaka, V. Transverse Location of New Fluorene Based Depth Dependent Fluorescent Probes in Membranes—Quenching Studies with 9,10-Dibromostearic Acid. Photochem. Photobiol. 1989, 49, 763–767. [Google Scholar] [CrossRef]
  323. Hosseinzadeh, R.; Mohadjerani, M.; Pooryousef, M. A New Selective Fluorene-Based Fluorescent Internal Charge Transfer (ICT) Sensor for Sugar Alcohols in Aqueous Solution. Anal. Bioanal. Chem. 2016, 408, 1901–1908. [Google Scholar] [CrossRef]
  324. Yilmaz, G.; Aydogan, B.; Temel, G.; Arsu, N.; Moszner, N.; Yagci, Y. Thioxanthone−Fluorenes as Visible Light Photoinitiators for Free Radical Polymerization. Macromolecules 2010, 43, 4520–4526. [Google Scholar] [CrossRef]
  325. Peter Pappas, S.; Tilley, M.G.; Pappas, B.C. Anthracene-Bound Sulfonium Salts: Highly Efficient Photoinitiators for Cationic Polymerization: A New Synthesis of Sulfonium Salts Which Avoids the Use of Silver Salts. J. Photochem. Photobiol. A Chem. 2003, 159, 161–171. [Google Scholar] [CrossRef]
  326. Balta, D.K.; Arsu, N.; Yagci, Y.; Jockusch, S.; Turro, N.J. Thioxanthone−Anthracene:  A New Photoinitiator for Free Radical Polymerization in the Presence of Oxygen. Macromolecules 2007, 40, 4138–4141. [Google Scholar] [CrossRef]
  327. Balta, D.K.; Arsu, N.; Yagci, Y.; Sundaresan, A.K.; Jockusch, S.; Turro, N.J. Mechanism of Photoinitiated Free Radical Polymerization by Thioxanthone−Anthracene in the Presence of Air. Macromolecules 2011, 44, 2531–2535. [Google Scholar] [CrossRef]
  328. Balta, D.K.; Arsu, N. Photodimerization of Thioxanthone–Anthracene: Formation of Monochromophoric Type II Initiator. Des. Monomers Polym. 2016, 19, 222–226. [Google Scholar] [CrossRef]
  329. Balta, D.K.; Arsu, N. Thioxanthone-Ethyl Anthracene. J. Photochem. Photobiol. A Chem. 2013, 257, 54–59. [Google Scholar] [CrossRef]
  330. Crivello, J.V.; Jang, M. Anthracene Electron-Transfer Photosensitizers for Onium Salt Induced Cationic Photopolymerizations. J. Photochem. Photobiol. A Chem. 2003, 159, 173–188. [Google Scholar] [CrossRef]
  331. Toba, Y. Anthracene-Sensitized Polymerization of Vinyl Ethers by Onium Tetrakis(Pentafluorophenyl)Borate Initiators. J. Polym. Sci. Part A Polym. Chem. 2000, 38, 982–987. [Google Scholar] [CrossRef]
  332. Cho, J.-D.; Kim, H.-K.; Kim, Y.-S.; Hong, J.-W. Dual Curing of Cationic UV-Curable Clear and Pigmented Coating Systems Photosensitized by Thioxanthone and Anthracene. Polym. Test. 2003, 22, 633–645. [Google Scholar] [CrossRef]
  333. Cho, J.-D.; Kim, E.-O.; Kim, H.-K.; Hong, J.-W. An Investigation of the Surface Properties and Curing Behavior of Photocurable Cationic Films Photosensitized by Anthracene. Polym. Test. 2002, 21, 781–791. [Google Scholar] [CrossRef]
  334. Iwaki, J.; Suzuki, S.; Park, C.; Miyagawa, N.; Takahara, S.; Yamaoka, T. Sensitization Mechanism of the Anthracene Derivatives/Photoacid Generator Photoinitiating System. J. Photopolym. Sci. Technol. 2004, 17, 123–124. [Google Scholar] [CrossRef]
  335. Pappas, S.P.; Pappas, B.C.; Gatechair, L.R.; Schnabel, W. Photoinitation of Cationic Polymerization. II. Laser Flash Photolysis of Diphenyliodonium Salts. J. Polym. Sci. Polym. Chem. Ed. 1984, 22, 69–76. [Google Scholar] [CrossRef]
  336. DeVoe, R.J.; Sahyun, M.R.V.; Schmidt, E.; Serpone, N.; Sharma, D.K. Electron Transfer Sensitized Photolysis of ’onium Salts. Can. J. Chem. 1988, 66, 319–324. [Google Scholar] [CrossRef]
  337. Runemark, A.; Sundén, H. Overcoming Back Electron Transfer in the Electron Donor–Acceptor Complex-Mediated Visible Light-Driven Generation of α-Aminoalkyl Radicals from Secondary Anilines. J. Org. Chem. 2023, 88, 462–474. [Google Scholar] [CrossRef] [PubMed]
  338. Pączkowski, J.; Pietrzak, M.; Paczkowska, B. Does the Back Electron Transfer Affect the Rate of Photoinduced Free Radical Polymerization? Polimery 2022, 48, 765–773. [Google Scholar] [CrossRef]
  339. Christmann, J.; Allonas, X.; Ley, C.; Ibrahim, A.; Croutxé-Barghorn, C. Triazine-Based Type-II Photoinitiating System for Free Radical Photopolymerization: Mechanism, Efficiency, and Modeling. Macromol. Chem. Phys. 2017, 218, 1600597. [Google Scholar] [CrossRef]
  340. Borges, F.; Roleira, F.; Milhazes, N.; Santana, L.; Uriarte, E. Simple Coumarins and Analogues in Medicinal Chemistry: Occurrence, Synthesis and Biological Activity. Curr. Med. Chem. 2005, 12, 887–916. [Google Scholar] [CrossRef]
  341. Gnonlonfin, G.J.B.; Sanni, A.; Brimer, L. Review Scopoletin—A Coumarin Phytoalexin with Medicinal Properties. Null 2012, 31, 47–56. [Google Scholar] [CrossRef]
  342. Lončarić, M.; Gašo-Sokač, D.; Jokić, S.; Molnar, M. Recent Advances in the Synthesis of Coumarin Derivatives from Different Starting Materials. Biomolecules 2020, 10, 151. [Google Scholar] [CrossRef]
  343. Chen, Q.; Yang, Q.; Gao, P.; Chi, B.; Nie, J.; He, Y. Photopolymerization of Coumarin-Containing Reversible Photoresponsive Materials Based on Wavelength Selectivity. Ind. Eng. Chem. Res. 2019, 58, 2970–2975. [Google Scholar] [CrossRef]
  344. Li, Z.; Zou, X.; Zhu, G.; Liu, X.; Liu, R. Coumarin-Based Oxime Esters: Photobleachable and Versatile Unimolecular Initiators for Acrylate and Thiol-Based Click Photopolymerization under Visible Light-Emitting Diode Light Irradiation. ACS Appl. Mater. Interfaces 2018, 10, 16113–16123. [Google Scholar] [CrossRef]
  345. Rahal, M.; Mokbel, H.; Graff, B.; Toufaily, J.; Hamieh, T.; Dumur, F.; Lalevée, J. Mono vs. Difunctional Coumarin as Photoinitiators in Photocomposite Synthesis and 3D Printing. Catalysts 2020, 10, 1202. [Google Scholar] [CrossRef]
  346. Rajeshirke, M.; Sreenath, M.C.; Chitrambalam, S.; Joe, I.H.; Sekar, N. Enhancement of NLO Properties in OBO Fluorophores Derived from Carbazole–Coumarin Chalcones Containing Carboxylic Acid at the N-Alykl Terminal End. J. Phys. Chem. C 2018, 122, 14313–14325. [Google Scholar] [CrossRef]
  347. Dumur, F. Recent Advances on Coumarin-Based Photoinitiators of Polymerization. Eur. Polym. J. 2022, 163, 110962. [Google Scholar] [CrossRef]
  348. Liu, Z.; Dumur, F. Recent Advances on Visible Light Coumarin-Based Oxime Esters as Initiators of Polymerization. Eur. Polym. J. 2022, 177, 111449. [Google Scholar] [CrossRef]
  349. Camposeo, A.; Arkadii, A.; Romano, L.; D’Elia, F.; Fabbri, F.; Zussman, E.; Pisignano, D. Impact of Size Effects on Photopolymerization and Its Optical Monitoring In-Situ. Addit. Manuf. 2022, 58, 103020. [Google Scholar] [CrossRef]
  350. Breloy, L.; Brezová, V.; Barbieriková, Z.; Ito, Y.; Akimoto, J.; Chiappone, A.; Abbad-Andaloussi, S.; Malval, J.-P.; Versace, D.-L. Methacrylated Quinizarin Derivatives for Visible-Light Mediated Photopolymerization: Promising Applications in 3D-Printing Biosourced Materials under LED@405 nm. ACS Appl. Polym. Mater. 2022, 4, 210–228. [Google Scholar] [CrossRef]
  351. Chao, P.; Gu, R.; Ma, X.; Wang, T.; Zhao, Y. Thiophene-Substituted Phenothiazine-Based Photosensitisers for Radical and Cationic Photopolymerization Reactions under Visible Laser Beams (405 and 455 nm). Polym. Chem. 2016, 7, 5147–5156. [Google Scholar] [CrossRef]
  352. Sun, K.; Peng, X.; Gan, Z.; Chen, W.; Li, X.; Gong, T.; Xiao, P. 3D Printing/Vat Photopolymerization of Photopolymers Activated by Novel Organic Dyes as Photoinitiators. Catalysts 2022, 12, 1272. [Google Scholar] [CrossRef]
  353. Shah, M.; Ullah, A.; Azher, K.; Rehman, A.U.; Juan, W.; Aktürk, N.; Tüfekci, C.S.; Salamci, M.U. Vat Photopolymerization-Based 3D Printing of Polymer Nanocomposites: Current Trends and Applications. RSC Adv. 2023, 13, 1456–1496. [Google Scholar] [CrossRef]
  354. Pagac, M.; Hajnys, J.; Ma, Q.-P.; Jancar, L.; Jansa, J.; Stefek, P.; Mesicek, J. A Review of Vat Photopolymerization Technology: Materials, Applications, Challenges, and Future Trends of 3D Printing. Polymers 2021, 13, 598. [Google Scholar] [CrossRef]
  355. Zhu, Y.; Li, L.; Zhang, Y.; Ou, Y.; Zhang, J.; Yagci, Y.; Liu, R. Broad Wavelength Sensitive Coumarin Sulfonium Salts as Photoinitiators for Cationic, Free Radical and Hybrid Photopolymerizations. Prog. Org. Coat. 2023, 174, 107272. [Google Scholar] [CrossRef]
Figure 1. Penetration of light determined for a polystyrene latex (average diameter around 112 nm). Reprinted with permission from Ref. [143].
Figure 1. Penetration of light determined for a polystyrene latex (average diameter around 112 nm). Reprinted with permission from Ref. [143].
Polymers 15 04202 g001
Scheme 1. Comparison between Type I and Type II photoinitiators for radical generation.
Scheme 1. Comparison between Type I and Type II photoinitiators for radical generation.
Polymers 15 04202 sch001
Scheme 2. Mechanism of cationic polymerization of epoxides using triarylsulfonium salts. (* corresponds to the excited state).
Scheme 2. Mechanism of cationic polymerization of epoxides using triarylsulfonium salts. (* corresponds to the excited state).
Polymers 15 04202 sch002
Figure 2. Chemical structures of 1,3,5-triphenyl-2-pyrazoline derivatives, different monomers and additives.
Figure 2. Chemical structures of 1,3,5-triphenyl-2-pyrazoline derivatives, different monomers and additives.
Polymers 15 04202 g002
Figure 3. UV-visible absorption spectra of 1,3,5-triphenyl-2-pyrazoline derivatives in acetonitrile. Reproduced with permission of Ref. [182].
Figure 3. UV-visible absorption spectra of 1,3,5-triphenyl-2-pyrazoline derivatives in acetonitrile. Reproduced with permission of Ref. [182].
Polymers 15 04202 g003
Figure 4. The different byproducts identified during the photolysis of PI-H in acetonitrile upon irradiation at 365 nm with an LED. Reproduced with permission of Ref. [182]. (* corresponds to the excited state).
Figure 4. The different byproducts identified during the photolysis of PI-H in acetonitrile upon irradiation at 365 nm with an LED. Reproduced with permission of Ref. [182]. (* corresponds to the excited state).
Polymers 15 04202 g004
Figure 5. Chemical structures of P-PTh and TTA-15 used as bifunctional monomers.
Figure 5. Chemical structures of P-PTh and TTA-15 used as bifunctional monomers.
Polymers 15 04202 g005
Figure 6. UV-visible absorption spectrum of P-PTh in acetonitrile. Reproduced with permission of Ref. [241].
Figure 6. UV-visible absorption spectrum of P-PTh in acetonitrile. Reproduced with permission of Ref. [241].
Polymers 15 04202 g006
Scheme 3. Photochemical mechanism involved in the decomposition of the phenacyl phenothiazinium salt (P-PTh).
Scheme 3. Photochemical mechanism involved in the decomposition of the phenacyl phenothiazinium salt (P-PTh).
Polymers 15 04202 sch003
Figure 7. Polymerization profiles of TTA-15 upon irradiation at 365 nm (A) and 405 nm (B) with LEDs (I = 30 mW/cm2) [photoinitiator] = 1.72 × 10−6 mol/g resin. Reproduced with permission of Ref. [241].
Figure 7. Polymerization profiles of TTA-15 upon irradiation at 365 nm (A) and 405 nm (B) with LEDs (I = 30 mW/cm2) [photoinitiator] = 1.72 × 10−6 mol/g resin. Reproduced with permission of Ref. [241].
Polymers 15 04202 g007
Figure 8. Different sulfonium salts based on triphenylamines as the chromophores.
Figure 8. Different sulfonium salts based on triphenylamines as the chromophores.
Polymers 15 04202 g008
Scheme 4. Interest of triphenylamine for the design of sulfonium salts.
Scheme 4. Interest of triphenylamine for the design of sulfonium salts.
Polymers 15 04202 sch004
Figure 9. Chemical structures of benchmark two-photon initiators.
Figure 9. Chemical structures of benchmark two-photon initiators.
Polymers 15 04202 g009
Figure 10. Examples of asymmetrically substituted sulfonium salts comprising a triphenylamine unit.
Figure 10. Examples of asymmetrically substituted sulfonium salts comprising a triphenylamine unit.
Polymers 15 04202 g010
Figure 11. Photopolymerization profiles of EPOX (A) and CHO (B) upon irradiation at 405 nm (I = 5 mW/cm2) using 1 wt% photoacid generator. Reproduced with permission of Ref. [185].
Figure 11. Photopolymerization profiles of EPOX (A) and CHO (B) upon irradiation at 405 nm (I = 5 mW/cm2) using 1 wt% photoacid generator. Reproduced with permission of Ref. [185].
Polymers 15 04202 g011
Figure 12. Chemical structures of mono and bifunctional sulfonium salts prepared by Malval and coworkers.
Figure 12. Chemical structures of mono and bifunctional sulfonium salts prepared by Malval and coworkers.
Polymers 15 04202 g012
Figure 13. UV-visible absorption spectra of photoinitiators differing by the substitution pattern and recorded in acetonitrile. Reproduced with permission of Ref. [260] (a) UV-visible absorption of Pre-Para, Mono-Para, Bi-Para and (b) UV-visible absorption spectra of Pre-Meta, Mono-Meta and Bi-Meta.
Figure 13. UV-visible absorption spectra of photoinitiators differing by the substitution pattern and recorded in acetonitrile. Reproduced with permission of Ref. [260] (a) UV-visible absorption of Pre-Para, Mono-Para, Bi-Para and (b) UV-visible absorption spectra of Pre-Meta, Mono-Meta and Bi-Meta.
Polymers 15 04202 g013
Figure 14. Polymerization profiles determined under air during the CP of DVE-3 by irradiating at 405 nm with a LED (5 mW/cm2), 1 wt% photoinitiator. Reproduced with permission of Ref. [260].
Figure 14. Polymerization profiles determined under air during the CP of DVE-3 by irradiating at 405 nm with a LED (5 mW/cm2), 1 wt% photoinitiator. Reproduced with permission of Ref. [260].
Polymers 15 04202 g014
Figure 15. Chemical structures of oligomeric sulfonium salts.
Figure 15. Chemical structures of oligomeric sulfonium salts.
Polymers 15 04202 g015
Figure 16. UV-visible absorption spectra of Mono-Ben, Bi-Ben, Tri-Ben and Tetra-Ben in acetonitrile. Reproduced with permission of Ref. [261].
Figure 16. UV-visible absorption spectra of Mono-Ben, Bi-Ben, Tri-Ben and Tetra-Ben in acetonitrile. Reproduced with permission of Ref. [261].
Polymers 15 04202 g016
Scheme 5. Mechanism of photoacid generation determined for oligomeric PAGs. (* correspond to the excited state).
Scheme 5. Mechanism of photoacid generation determined for oligomeric PAGs. (* correspond to the excited state).
Polymers 15 04202 sch005
Figure 17. Photopolymerization profiles obtained with Tri-Ben as the photoinitiator: DVE-3 (1 wt% Tri-Ben, curve 1), TMPTA (Tri-Ben/MEDA, 1%/2%, w/w, curve 2) and EPOX (Tri-Ben/NVK, 1%/3%, w/w, curve 3) upon irradiation at 365 nm (15 mW/cm2). Reproduced with permission of Ref. [261].
Figure 17. Photopolymerization profiles obtained with Tri-Ben as the photoinitiator: DVE-3 (1 wt% Tri-Ben, curve 1), TMPTA (Tri-Ben/MEDA, 1%/2%, w/w, curve 2) and EPOX (Tri-Ben/NVK, 1%/3%, w/w, curve 3) upon irradiation at 365 nm (15 mW/cm2). Reproduced with permission of Ref. [261].
Polymers 15 04202 g017
Figure 18. Chemical structures of different sulfonium salts combining double bonds and consecutive aromatic rings in their scaffolds.
Figure 18. Chemical structures of different sulfonium salts combining double bonds and consecutive aromatic rings in their scaffolds.
Polymers 15 04202 g018
Figure 19. UV-visible absorption spectra of different sulfonium salts combining double bonds and consecutive aromatic rings in their scaffolds. Reproduced with permission of Ref. [262].
Figure 19. UV-visible absorption spectra of different sulfonium salts combining double bonds and consecutive aromatic rings in their scaffolds. Reproduced with permission of Ref. [262].
Polymers 15 04202 g019
Figure 20. Sulfonium salt PI-PAG with the dual role of cationic initiator and photosensitizer for an iodonium salt (Iod). Chemical structure of precursor pre-H used for the synthesis of PI-PAG is also presented.
Figure 20. Sulfonium salt PI-PAG with the dual role of cationic initiator and photosensitizer for an iodonium salt (Iod). Chemical structure of precursor pre-H used for the synthesis of PI-PAG is also presented.
Polymers 15 04202 g020
Figure 21. UV-visible absorption spectrum of PI-PAG and its corresponding precursor pre-H in acetonitrile. Reproduced with permission of Ref. [167].
Figure 21. UV-visible absorption spectrum of PI-PAG and its corresponding precursor pre-H in acetonitrile. Reproduced with permission of Ref. [167].
Polymers 15 04202 g021
Figure 22. Photopolymerization profiles obtained during the ring-opening polymerization of EPOX under air upon irradiation at 365 nm (I = 40 mW/cm2). Iod (3 wt%), precursor pre-H (1 wt%), PI-PAG (1 wt%), PI-PAG/Iod (1%/3%, w/w) and precursor pre-H/Iod (1%/3%, w/w). Reproduced with permission of Ref. [167].
Figure 22. Photopolymerization profiles obtained during the ring-opening polymerization of EPOX under air upon irradiation at 365 nm (I = 40 mW/cm2). Iod (3 wt%), precursor pre-H (1 wt%), PI-PAG (1 wt%), PI-PAG/Iod (1%/3%, w/w) and precursor pre-H/Iod (1%/3%, w/w). Reproduced with permission of Ref. [167].
Polymers 15 04202 g022
Figure 23. Monomer conversions determined during the CP of EPOX at different wavelengths using Iod (3 wt%), precursor pre-H (1 wt%), PI-PAG (1 wt%), PI-PAG/Iod (1%/3%, w/w) and precursor pre-H/Iod (1%/3%, w/w), I = 40 mW/cm2 for all LEDs. Reproduced with permission of Ref. [167].
Figure 23. Monomer conversions determined during the CP of EPOX at different wavelengths using Iod (3 wt%), precursor pre-H (1 wt%), PI-PAG (1 wt%), PI-PAG/Iod (1%/3%, w/w) and precursor pre-H/Iod (1%/3%, w/w), I = 40 mW/cm2 for all LEDs. Reproduced with permission of Ref. [167].
Polymers 15 04202 g023
Figure 24. Chemical structures of sulfonium salts used during UCNPS-mediated photopolymerization.
Figure 24. Chemical structures of sulfonium salts used during UCNPS-mediated photopolymerization.
Polymers 15 04202 g024
Figure 25. Curing depth determined during the hybrid polymerization of a TMPTA/EPOX blend upon irradiation of the samples at 980 nm with a laser (I = 11.6 W) for different irradiation times. Reproduced with permission of Ref. [264].
Figure 25. Curing depth determined during the hybrid polymerization of a TMPTA/EPOX blend upon irradiation of the samples at 980 nm with a laser (I = 11.6 W) for different irradiation times. Reproduced with permission of Ref. [264].
Polymers 15 04202 g025
Figure 26. Chemical structures of VDGE and GDMK.
Figure 26. Chemical structures of VDGE and GDMK.
Polymers 15 04202 g026
Figure 27. Reduction of E. coli (a) and S. aureus (b) cells during 24 h of contact time with films of vanillin-based polymers, chitosan and hydroxyethyl starch. Reproduced with permission of Ref. [280].
Figure 27. Reduction of E. coli (a) and S. aureus (b) cells during 24 h of contact time with films of vanillin-based polymers, chitosan and hydroxyethyl starch. Reproduced with permission of Ref. [280].
Polymers 15 04202 g027
Figure 28. Chemical structures of bodipy-based sulfonium salts.
Figure 28. Chemical structures of bodipy-based sulfonium salts.
Polymers 15 04202 g028
Figure 29. Chemical structures of ITXPhenS, different monomers and additives.
Figure 29. Chemical structures of ITXPhenS, different monomers and additives.
Polymers 15 04202 g029
Figure 30. UV-visible absorption spectra of ITX, ITXPhenS and CTCs comprising different ITXPhenS/DMA ratios. Reproduced with permission of Ref. [287].
Figure 30. UV-visible absorption spectra of ITX, ITXPhenS and CTCs comprising different ITXPhenS/DMA ratios. Reproduced with permission of Ref. [287].
Polymers 15 04202 g030
Figure 31. Mechanism of photoinitiation with the charge transfer complex based on ITXPhenS. Reproduced with permission of Ref. [287]. (* corresponds to the excited state).
Figure 31. Mechanism of photoinitiation with the charge transfer complex based on ITXPhenS. Reproduced with permission of Ref. [287]. (* corresponds to the excited state).
Polymers 15 04202 g031
Figure 32. Chemical structures of BisGMA, TEGDMA, TMA and 2DPPBA.
Figure 32. Chemical structures of BisGMA, TEGDMA, TMA and 2DPPBA.
Polymers 15 04202 g032
Figure 33. The strategy developed by Yagci and coworkers to prepare composites using first a photochemical curing followed in a second step by a thermal curing. Reproduced with permission of Ref. [287].
Figure 33. The strategy developed by Yagci and coworkers to prepare composites using first a photochemical curing followed in a second step by a thermal curing. Reproduced with permission of Ref. [287].
Polymers 15 04202 g033
Figure 34. Chemical structures of Flu-MS and Flu-PS.
Figure 34. Chemical structures of Flu-MS and Flu-PS.
Polymers 15 04202 g034
Figure 35. UV-visible absorption spectra of Flu-MS and Flu-PS in acetonitrile. Reproduced with permission of Ref. [288].
Figure 35. UV-visible absorption spectra of Flu-MS and Flu-PS in acetonitrile. Reproduced with permission of Ref. [288].
Polymers 15 04202 g035
Figure 36. Chemical structures of fluorene-based photoinitiators developed by Belfield and coworkers.
Figure 36. Chemical structures of fluorene-based photoinitiators developed by Belfield and coworkers.
Polymers 15 04202 g036
Figure 37. Chemical structures of anthracene-substituted sulfonium salts, different reference compounds and the epoxy monomer SU-8 and TPGE.
Figure 37. Chemical structures of anthracene-substituted sulfonium salts, different reference compounds and the epoxy monomer SU-8 and TPGE.
Polymers 15 04202 g037
Scheme 6. Mechanism of formation of 9-phenylanthracene starting from anthracene and triphenylsulfonium. (* corresponds to the excited state).
Scheme 6. Mechanism of formation of 9-phenylanthracene starting from anthracene and triphenylsulfonium. (* corresponds to the excited state).
Polymers 15 04202 sch006
Figure 38. Chemical structures of coumarin-based sulfonium salts, TPGDA and the sulfonium salt 201s.
Figure 38. Chemical structures of coumarin-based sulfonium salts, TPGDA and the sulfonium salt 201s.
Polymers 15 04202 g038
Figure 39. UV-visible absorption spectra of coumarin-based sulfonium salts in acetonitrile. Reproduced with permission from Ref. [355].
Figure 39. UV-visible absorption spectra of coumarin-based sulfonium salts in acetonitrile. Reproduced with permission from Ref. [355].
Polymers 15 04202 g039
Figure 40. Mechanism of photoinitiation with coumarin-based sulfonium salts. Reproduced with permission from Ref. [355]. (* corresponds to the excited state).
Figure 40. Mechanism of photoinitiation with coumarin-based sulfonium salts. Reproduced with permission from Ref. [355]. (* corresponds to the excited state).
Polymers 15 04202 g040
Figure 41. Hybrid polymerization of TPGDA and EPOX using p-OMe-Me CSS and p-OMe-Bu CSS upon irradiation at 365 nm (I = 70 mW/cm2, sulfonium salt concentration: 5.27 × 10−5 mol/g). Reproduced with permission from Ref. [355].
Figure 41. Hybrid polymerization of TPGDA and EPOX using p-OMe-Me CSS and p-OMe-Bu CSS upon irradiation at 365 nm (I = 70 mW/cm2, sulfonium salt concentration: 5.27 × 10−5 mol/g). Reproduced with permission from Ref. [355].
Polymers 15 04202 g041
Table 1. Comparisons between sulfonium-bounded sulfonium salts and the other sulfonium salts that are commercially available and used in intermolecular sensitization systems.
Table 1. Comparisons between sulfonium-bounded sulfonium salts and the other sulfonium salts that are commercially available and used in intermolecular sensitization systems.
ParametersMonocomponent Sulfonium SaltsOther Sulfonium Salts
Cost/synthesisCompounds are often prepared in multistep syntheses rendering these compounds expensive.Numerous sulfonium salts are commercially available and can be purchased at low cost.
Environmental impactMost synthetic dyes are often prepared in an organic solvent. Purification by column chromatography is also required.The same comments can be performed concerning benchmark sulfonium salts.
Photochemical stabilityChromophore-substituted sulfonium salts are stable compounds. Therefore, their photochemical stability is good.Sulfonium salts are highly stable, notably when comprising aromatic rings.
Absorption rangeChromophores bounded to sulfonium salts exhibit a broad absorption and their absorption spectra can be finely tuned from the near-UV visible range up to the near-infrared rangeBenchmark sulfonium salts exhibit a UV-centered absorption, with no absorption in the visible range.
Photoinitiating abilityChromophores are ideal candidates for photoinitiation in the visible range and even in the near-infrared region. Ability of chromophores to release or accept electrons is well-documented in the literature.Good photoinitiating ability of sulfonium salts in the UV range
AvailabilityAvailability can be restricted, especially when multistep syntheses are used. Expensive compounds, which can limit their scope of applications.Easy availability. Benchmark photoinitiators can be purchased in the Kg scale at low cost. Numerous applications of benchmark sulfonium salts in industry.
Table 2. Quantum yields of photoacid generation (ϕH+) upon irradiation with different LEDs.
Table 2. Quantum yields of photoacid generation (ϕH+) upon irradiation with different LEDs.
365 nm385 nm395 nm405 nm425 nm
PI-CF30.600.560.610.590.31
PI-H0.620.590.620.630.26
PI-EtO0.650.610.640.460.20
Table 3. EPOX conversions (%) determined by photo-DSC upon irradiation at different wavelengths and by using 1 wt% photoinitiator.
Table 3. EPOX conversions (%) determined by photo-DSC upon irradiation at different wavelengths and by using 1 wt% photoinitiator.
365 nm385 nm395 nm405 nm425 nm
PI-CF349.1550.6352.6655.1851.87
PI-H52.8454.0656.3759.7440.25
PI-EtO55.0656.4758.8261.9749.57
PI-6992M59.3143.63---
Table 4. TMPTA conversions (%) determined by photo-DSC upon irradiation at different wavelengths and by using 1 wt% photoinitiator.
Table 4. TMPTA conversions (%) determined by photo-DSC upon irradiation at different wavelengths and by using 1 wt% photoinitiator.
365 nm385 nm395 nm405 nm
PI-CF343.144.941.335.3
PI-H50.948.945.542.4
OXE-0265.661.544.427.0
Table 5. Polymerization thresholds (mW) determined during the polymerization of Araldite CY179MA epoxide resins upon irradiation at 710 nm.
Table 5. Polymerization thresholds (mW) determined during the polymerization of Araldite CY179MA epoxide resins upon irradiation at 710 nm.
Initiatorλ = 710 nm
BSB-S22.4
CD1012/ITX44
CD1012212
TPS>317 a
DPI-DMAS>317 a
a: no polymerization detected at the maximum power, i.e., 317 mW.
Table 6. Photophysical properties of PAG1-PAG6 determined in acetonitrile.
Table 6. Photophysical properties of PAG1-PAG6 determined in acetonitrile.
Initiatorλmax (nm)εmax (M−1·cm−1)ϕH+ aδmax (GM)
PAG134636,1000.1073 (760 nm)
PAG232426,0000.2468 (710 nm)
PAG339534,3000.05643 (870 nm)
PAG438123,7000.44650 (800 nm)
PAG540034,4000.31680 (880 nm)
PAG638025,2000.50648 (800 nm)
a determined at 405 nm.
Table 7. Photophysical properties of a series of multibranch sulfonium salts.
Table 7. Photophysical properties of a series of multibranch sulfonium salts.
Compoundλabs (nm)εmax (M−1·cm−1)ϕH+ aϕH+ bδ780 nm (GM)ϕH+
Pre-Para39453,200----
Mono-Para40440,6000.0020.025281.0
Bi-Para41337,0000.0040.048163.2
Pre-Meta38537,000----
Mono-Meta39232,0000.200.4123446.8
Bi-Meta39930,8000.400.49745298
a Quantum yield of photoacid generation at 405 nm in acetonitrile. b Quantum yield of photoacid generation at 405 nm in dichloromethane.
Table 8. Monomer conversions determined with Flu-MS and Flu-PS using different monomers or monomer blends upon irradiation at 365 nm using 1 wt% photoinitiator.
Table 8. Monomer conversions determined with Flu-MS and Flu-PS using different monomers or monomer blends upon irradiation at 365 nm using 1 wt% photoinitiator.
MonomerPhotoinitiatorIntensity (mW/cm2)Conversion (%)
DVE-3Flu-MS298
DVE-3Flu-PS299
CHOFlu-MS284.5
CHOFlu-PS279
EPOXFlu-MS470
EPOXFlu-PS459
TMPTAFlu-MS458
TMPTAFlu-PS461
EPOX/TMPTAFlu-MS239/71
EPOX/TMPTAFlu-PS243/71
trithiol/TMPTAFlu-MS250/82
trithiol/TMPTAFlu-PS247/85
Table 9. Conversions of EPOX, CHO, TMPTA and EPOX/TMPTA at different irradiation wavelengths and intensities using Flu-MS or Flu-PS (1 wt%).
Table 9. Conversions of EPOX, CHO, TMPTA and EPOX/TMPTA at different irradiation wavelengths and intensities using Flu-MS or Flu-PS (1 wt%).
MonomersPIsLight SourceIntensity (mW/cm2)Conversion (%)
EPOXFlu-MS385 nm465.6
EPOXFlu-MS405 nm473.0
EPOXFlu-MS425 nm463.2
EPOXFlu-PS385 nm463.3
EPOXFlu-PS405 nm468.6
EPOXFlu-PS425 nm464.3
CHOFlu-MS385 nm281.1
CHOFlu-MS405 nm264.3
CHOFlu-MS450 nm1665.0
CHOFlu-PS385 nm276.3
CHOFlu-PS405 nm276.9
CHOFlu-PS450 nm1685.4
TMPTAFlu-MS385 nm448.5
TMPTAFlu-MS405 nm442.8
TMPTAFlu-PS385 nm460.8
TMPTAFlu-PS405 nm456.2
EPOX/TMPTAFlu-MS365 nm222.8/5.3
EPOX/TMPTAFlu-PS365 nm213.1/13.3
Table 10. Monomer conversions (%) obtained during the CP of EPOX and the FRP of TMPTA upon irradiation at 365 and 405 nm (I = 70 mW/cm2) with LEDs.
Table 10. Monomer conversions (%) obtained during the CP of EPOX and the FRP of TMPTA upon irradiation at 365 and 405 nm (I = 70 mW/cm2) with LEDs.
365 nmp-H-Me CSSp-H-Bu CSSp-H-Hep CSSp-OMe-Me CSSp-OMe-Bu CSSp-OMe-Hep CSS
EPOX18643718553255 (201s)
TMPTA44695448625468 (ITX)
405 nmp-H-Me CSSp-H-Bu CSSp-H-Hep CSSp-OMe-Me CSSp-OMe-Bu CSSp-OMe-Hep CSS201s
EPOX12534415694512 (201s)
TMPTA40584339605265 (ITX)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dumur, F. Recent Advances in Monocomponent Visible Light Photoinitiating Systems Based on Sulfonium Salts. Polymers 2023, 15, 4202. https://doi.org/10.3390/polym15214202

AMA Style

Dumur F. Recent Advances in Monocomponent Visible Light Photoinitiating Systems Based on Sulfonium Salts. Polymers. 2023; 15(21):4202. https://doi.org/10.3390/polym15214202

Chicago/Turabian Style

Dumur, Frédéric. 2023. "Recent Advances in Monocomponent Visible Light Photoinitiating Systems Based on Sulfonium Salts" Polymers 15, no. 21: 4202. https://doi.org/10.3390/polym15214202

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

Article Metrics

Back to TopTop