Next Article in Journal
Sustainable Materials Containing Biochar Particles: A Review
Previous Article in Journal
Self- and Cross-Fusing of Furan-Based Polyurea Gels Dynamically Cross-Linked with Maleimides
Previous Article in Special Issue
Water-Soluble Visible Light Sensitive Photoinitiating System Based on Charge Transfer Complexes for the 3D Printing of Hydrogels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

High-Performance Photoinitiating Systems for LED-Induced Photopolymerization

1
Centre National de la Recherche Scientifique (CNRS), L’Institut de Science des Matériaux de Mulhouse (IS2M UMR 7361), Université de Haute-Alsace, F-68100 Mulhouse, France
2
Université de Strasbourg, F-67081 Strasbourg, France
3
Research School of Chemistry, Australian National University, Canberra, ACT 2601, Australia
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(2), 342; https://doi.org/10.3390/polym15020342
Submission received: 25 November 2022 / Revised: 29 December 2022 / Accepted: 5 January 2023 / Published: 9 January 2023

Abstract

:
Currently, increasing attention has been focused on light-emitting diodes (LEDs)-induced photopolymerization. The common LEDs (e.g., LED at 365 nm and LED at 405 nm) possess narrow emission bands. Due to their light absorption properties, most commercial photoinitiators are sensitive to UV light and cannot be optimally activated under visible LED irradiation. Although many photoinitiators have been designed for LED-induced free radical polymerization and cationic polymerization, there is still the issue of the mating between photoinitiators and LEDs. Therefore, the development of novel photoinitiators, which could be applied under LED irradiation, is significant. Many photoinitiating systems have been reported in the past decade. In this review, some recently developed photoinitiators used in LED-induced photopolymerization, mainly in the past 5 years, are summarized and categorized as Type Ⅰ photoinitiators, Type Ⅱ photoinitiators, and dye-based photoinitiating systems. In addition, their light absorption properties and photoinitiation efficiencies are discussed.

Graphical Abstract

1. Introduction

Monomers can be transformed into polymers under irradiation (e.g., UV and visible light) in photopolymerization. Photopolymerization demonstrates numerous advantages, such as no VOCs, excellent controllability, and high efficiency [1,2,3,4,5]. Currently, photopolymerization is applied in many fields, including coatings [6,7,8], adhesives [9,10,11], 3D printing [12,13,14,15], biomaterials [16,17,18], the microelectronics industry [19,20,21], etc.
Photoinduced polymerization is a chain reaction. Photoinitiators (PIs) can generate active species (free radicals, cations, etc.) under light irradiation to initiate the polymerization of monomers, so a PI has a crucial influence on polymerization rates and final function conversions (FC) for monomers. A photoinitiating system (PIS) can be composed of a PI or PI/additive. Free radical photopolymerization (FRP) can be initiated by generated free radicals. It should be mentioned that FRP is normally inhibited by oxygen, which can quench the primary initiating and propagating radicals. Cationic photopolymerization (CP) is normally sensitive to the presence of moisture [22,23]. Some essential characteristics, including desirable light absorption properties, high photochemical activity, good solubility, and low toxicity, are the important evaluation standard for PIs [24,25,26,27].
Traditional UV light sources (i.e., a mercury lamp) demonstrate many drawbacks, such as short service life, slow switching times, sometimes long heat-up times, and high energy consumption, compared to modern light-emitting diode (LED) light sources. In addition, significant heat generation by broadband illumination of a mercury lamp can affect the surface properties of products, and the ozone release in operation is harmful to the human body [28,29,30,31]. The inherent disadvantages of UV light sources hinder the development of photopolymerization, albeit UV curing is already well-established in the industry. Recently, LEDs have been applied more and more in photopolymerization because of their low cost and high safety [32,33,34]. Nowadays, the emission bands of LEDs are narrow (FWHM in the range of 10 nm) and focused on 365 nm, 395 nm, 405 nm, and 455 nm normally [35]. A large portion of commercial PIs are UV-sensitive and do not absorb light > 360 nm. The effect of the narrow emission bands and the redshift in wavelengths, for example, is clearly presented in the cited literature for different commercial initiators [35]. Therefore, the development of PIs, which can be activated under LED irradiation, is significant [36,37,38,39,40,41,42,43,44,45,46,47].
To match with LEDs, many PISs was designed and developed in the past decade, such as Type Ⅰ and Type Ⅱ PIs, dye-based PISs, metal-complex-based PISs, polyoxometalate-based PISs, and nanoparticle-based PISs, and so on [43,48,49,50,51,52,53,54]. Herein, some highly interesting works for dye-based PISs and the modification of commercial Type Ⅰ PIs and Type Ⅱ PIs are presented. Their light absorption properties are also characterized by their maximal absorption wavelength (λmax) and molar absorption coefficient (ε) (see the tables below).

2. Dye-Based PISs

Dye molecules have been widely used as PIs in photoinitiation due to their strong absorption in the visible range. Except for dye molecules, the additives, as well as co-initiators, also play important roles in PISs. In the excited states, dye molecules are able to interact with additives through electron transfer reactions to generate active radicals and cations [23,55]. Then, the initial active species can combine with monomers through addition reactions or ring-opening reactions to produce a polymer. Iodonium salts and sulfoniun salts, such as bis(4-tert-butylphenyl)iodonium hexafluorophosphate (Iod) or (sulfanediyldibenzene-4,1-diyl) bis(diphenylsulfonium) bis(hexafluoroantimonate) in propylene carbonate (Sulf) are frequently used as electron acceptors. Amine compounds, including ethyl 4-(dimethylamino)-benzoate (EDB), N-vinylcarbazole (NVK) and N-phenyl glycine (NPG), are excellent electron donors.
The photochemical mechanism of a typical dye/Iod/EDB system is depicted in Scheme 1. In the reductive cycle, the dye molecule can accept an electron from the electron donor EDB to generate free radical, and the dye˙ radical anion is oxidized by Iod to regenerate the dye molecule. In the oxidative cycle, the dye molecule, as an electron donor, transfers an electron to God, producing Ar˙ and dye˙+ to induce the FRP and CP, respectively. Then, the dye˙+ is reduced by EDB to regenerate the dye molecule. The photoredox catalytic cycle slows down the consumption of dye molecules to some degree and accelerates the photopolymerization. Due to their versatile features and good efficiency, numerous dyes have been designed as promising PIs for LED photoinitiation. Here, some representative chemical structures, including carbazole, triphenylamine, naphthalimide, chalcones, and coumarin derivatives, are presented in the following context.

2.1. Carbazole-Based Photoinitiators

Carbazole is a representative scaffold applied in the context of an LED photoinitiator. The benzene rings in the carbazole structure can be functionalized with various groups to obtain the extension of conjugation. The nitrogen atom can be modified with alkyl chains to improve solubility. In addition, the remarkable electron-donating ability and low oxidation potential contribute to the interaction with additives in PISs [56,57,58]. The chemical structures of carbazole-based PIs mentioned in this review are shown in Table 1.
A serious of carbazole derivatives were investigated by Mousawi et al. [59,60]. Four carbazole derivatives, C1–C4, were designed as PIs for both FRP and CP. These compounds exhibit good absorptions in the 350–450 nm range. Under irradiation with a LED at 405 nm, high function conversions of 3,4-Epoxycyclohexylmethyl-3′,4′-epoxycyclohexane carboxylate (EPOX) (C1/Iod: FC = 76%, C2/Iod: FC = 50%, C3/Iod: FC = 58%, C4/Iod: FC = 70%) were found for carbazole derivative/Iod combinations. For C/Iod/EDB PISs, a photoredox catalyst behavior could be found in FRP, and the acrylate function conversions for trimethylolpropane triacrylate (TMPTA) were favorable. 3D objects were obtained by a LED projector using these interesting photoinitiating systems. In addition, other carbazole derivatives A1–A4 (Table 1) with thermally activated delayed fluorescence (TADF) properties were synthesized and proposed. Favorable light absorption properties were found for A1–A4, allowing for the application of LED at 405 nm in photopolymerization. High epoxy function conversions (47–55%) were obtained in CP for A/Iod combinations under LED at 405 nm irradiation. Good performances were also observed in the FRP of TMPTA. Finally, these new photoinitiating systems were successfully applied in photocurable 3D printing experiments. Indeed, the TADF property was helpful to the reaction from the excited singlet state between carbazole derivatives and additives.
Four carbazole-based two-photon initiators, A3–1, A3–2, A3–3 and A3–4 (Table 1), containing conjugation bridges were designed and synthesized by Li et al. [61]. Their absorption peaks were located at 350 nm, 360 nm, 415 nm and 394 nm, respectively. No polymerization of TMPTA was found for resin upon exposure to UV, which demonstrated the good stability of carbazole derivatives under one-photon irradiation. The two-photon polymerization using carbazole derivatives was evaluated by direct laser write setup with an 800 nm pulsed laser. In the photopolymerization experiments, A3–2 and A3–4 demonstrated lower threshold energy compared to benchmark PI 1-benzyl-1-(dimethylamino)propyl 4-morpholinophenyl ketone.
Table 1. The chemical structures and light absorption properties of carbazole-based photoinitiators.
Table 1. The chemical structures and light absorption properties of carbazole-based photoinitiators.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i001
C1
λmax ~ 364
εmax ~ 11,750
ε405nm ~ 2600
[59]
Polymers 15 00342 i002
C2
λmax ~ 374
εmax ~ 11,180
ε405nm ~ 5200
[59]
Polymers 15 00342 i003
C3
λmax ~ 364
εmax ~ 14,000
ε405nm ~ 2450
[59]
Polymers 15 00342 i004
C4
λmax ~ 388
εmax ~ 6000
ε405nm ~ 5200
[59]
Polymers 15 00342 i005
A1
λmax ~ 330
εmax ~ 8800
ε405nm ~ 1350
[60]
Polymers 15 00342 i006
A2
λmax ~ 340
εmax ~ 40,000
ε405nm ~ 7800
[60]
Polymers 15 00342 i007
A3
λmax ~ 333
εmax ~ 33,000
ε405nm ~ 5700
[60]
Polymers 15 00342 i008
A4
λmax ~ 349
εmax ~ 18,000
ε405nm ~ 3300
[60]
Polymers 15 00342 i009
A3–1
λmax ~ 350
εmax ~ 49,000
[61]
Polymers 15 00342 i010
A3–2
λmax ~ 360
εmax ~ 53,000
[61]
Polymers 15 00342 i011
A3–3
λmax ~ 415
εmax ~ 29,000
[61]
Polymers 15 00342 i012
A3–4
λmax ~ 394
εmax ~ 29,000
[61]

2.2. Triphenylamine-Based Photoinitiators

Triphenylamine is widely employed for the design of new PIs due to its excellent electron-donating ability. Chemical modification is usually carried out with various groups to obtain extended conjugation. The chemical structures of triphenylamine-based PIs mentioned in this review are given in Table 2. Han et al. [62] synthesized a blue-green-light-sensitive PI CDM2 based on the triphenylamine-curcumin structure. The combination of triphenylamine and curcumin units made it show strong absorption in the blue and green light regions. The introduction of triphenylamine on both sides of the CMD2 reduced the energy gap, which contributed to the red-shifted absorption band. It was very interesting that the maximum absorption wavelength (λmax) of CDM2 in tetrahydrofuran was 467 nm, which was suitable for long-wavelength irradiation sources. As a result, the CDM2/iodonium salt system demonstrated good initiation efficiency, and the function conversions of diglycidyl ether of bisphenol A could reach up to 56% and 46% under blue LED or green LED, respectively.
Two triphenylamine derivatives, Dye3 and Dye4, were designed as PIs by Abdallah et al. [63]. The molar extinction coefficients at 405 nm (ε405nm) were 4110 M−1 cm−1 and 4740 M−1 cm−1 for Dye3 and Dye4. They were efficient in initiating the CP of EPOX using LED at 405 nm, especially the epoxy function conversion was 66% for Dye3/Iod system. Good performances were shown in FRP of TMPTA for Dye/Iod/NPG PISs, and function conversions were 59% and 57% for Dye3 and Dye4 systems, respectively. The photolysis and fluorescence quenching processes indicated that the reactions for dyes and Iod were effective. Indeed, the electron-donating ability of the triphenylamine core allowed for effective electron transfer reactions.
Three triphenylamine-based hexaarylbiimidazole derivatives (HABI1, HABI2, HABI3) were synthesized by Li [64]. They exhibit favorable absorption bands from 360 nm to 420 nm. The maximum absorption wavelengths (λmax) of HABI1–3 were 383, 385, and 384 nm, respectively. The performance of HABIs was evaluated by differential scanning photocalorimetry (photo-DSC) under a UV lamp (250–450 nm) and LED lamp (380–750 nm), respectively. The favorable values of free energy changes proved to have good electron transfer ability for the HABIs/NPG systems. The final function conversions of TMPTA for the HABI1/NPG system were 80% and 58% upon exposure to UV and LED light, respectively. High conversions demonstrated that the HABI1/NPG system could be applied under both UV and LED irradiation sources.
Jin et al. [65] proposed a conjugated sulfonium-based triphenylamine derivative (PI-PAG) as PI in CP. Triarylsulfonium salts are widely used as cationic PI due to the generation of strong protic acids and thermal stability. However, the absorption peaks of most commercial triarylsulfonium salts are below 300 nm, so they cannot be applied under longer-wavelength LEDs, such as LED at 365 and 405 nm. To extend the light absorption band, the sulfonium salt moiety was associated with a long-wavelength triphenylamine chromophore in PI-PAG to form a π-conjugated structure. Interestingly, PI-PAG had good light absorption properties (λmax = 381 nm, ε365nm = 19,200 M−1 cm−1, ε405nm = 14,900 M−1 cm−1). The photoinitiation ability of PI-PAG (1 wt%) in CP was evaluated upon exposure to 365, 385, and 405 nm LEDs. The final function conversion for EPOX at 5 min was 52.3% for PI-PAG alone using a LED at 365 nm. When Iod1 was added as a co-initiator, the final function conversion of EPOX reached 82% for PI-PAG/Iod1 (1%/3%, w/w). In addition, the CP of other monomers, such as cyclohexene oxide and Triethyleneglycol divinyl ether (DVE-3), was also investigated. For PI-PAG (0.5 wt%) alone, the polymerization of DVE-3 was observed. Furthermore, PI-PAG/Iod1 (0.5%/2%, w/w) had a good performance. Final function conversions of DVE-3 were 90.65% and 90.3% using LED at 365 and 405 nm, respectively.
A photochemical mechanism of PI-PAG was investigated by photolysis and electron spin resonance-spin trapping (ESR-ST) experiments. The S−C bond could break under irradiation; then, the generated sulfur cation radicals further reacted with hydrogen donors to produce the active initiation species H+ for CP. The PI-PAG/Iod1 system could also generate cations through an electron transfer reaction. Due to its good performance, PI-PAG successfully initiated the polymerization under chemiluminescence irradiation; the associated chemical mechanisms are presented in detail in [66]. The light emission at ~430 nm was generated through chemical reaction. The CP of cyclohexene oxide was observed in the presence of PI-PAG under the emitted light. These results indicated that the design strategy of π-conjugated sulfonium salts was effective.
Table 2. The chemical structures and light absorption properties of triphenylamine-based photoinitiators.
Table 2. The chemical structures and light absorption properties of triphenylamine-based photoinitiators.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i013
CDM2
λmax ~ 467
εmax ~ 77,190
ε460nm ~ 75,490
ε520nm ~ 12,000
[62]
Polymers 15 00342 i014
Dye3
λmax ~ 352
εmax ~ 26,610
ε405nm ~ 4110
[63]
Polymers 15 00342 i015
Dye4
λmax ~ 357
εmax ~ 17,700
ε405nm ~ 4740
[63]
Polymers 15 00342 i016
HABI1
λmax ~ 383
εmax ~ 6600
[64]
Polymers 15 00342 i017
HABI2
λmax ~ 385
εmax ~ 12,100
[64]
Polymers 15 00342 i018
HABI3
λmax ~ 384
εmax ~ 14,800
[64]
Polymers 15 00342 i019
PI-PAG
λmax ~ 381
εmax ~ 23,200
ε365nm ~ 19,200
ε405nm ~ 14,900
[65]

2.3. Naphthalimide-Based Photoinitiators

Naphthalimide derivatives are widely used as PIs due to their good stability and ease of synthesis. Additionally, the absorption properties from 400 nm to 600 nm of naphthalimide derivatives can be adjusted by different chemical groups [67,68,69,70]. The chemical structures of naphthalimide-based PIs mentioned in this review are presented in Table 3. A series of naphthalimide derivatives NDP1−NDP7 were proposed as PIs upon LED at 405 nm exposure [71]. The λmax of NDP3 and NDP5 were located below 350 nm. Others demonstrated good light absorption properties because of nitro withdrawing substituent in the naphthalene moiety. Their maximum absorption wavelengths were all located in the range of 417 to 440 nm. The CP of epoxides was evaluated. NDP1/Iod1 and NDP3/Iod1 systems demonstrated low efficiency (FC < 30%) upon exposure to LED at 405 nm. Interestingly, high epoxy function conversions of EPOX 59%, 62% and 63% were obtained in the presence of NDP2/Iod1, NDP4/Iod1, and NDP6/Iod1 systems, respectively, at 800 s. NDP2, NDP4 and NDP6 also exhibit good cationic initiation ability under LED at 455 nm irradiation. The polymerization of TMPTA was evaluated using NDPs/Iod1 and NDPs/Iod1/NVK systems. An excellent performance was observed for NDP2/Iod1 (FC = 52%), NDP2/Iod1/NVK (FC = 63%), NDP4/Iod1 (FC = 52%) and NDP4/Iod1/NVK (FC = 58%). The amino or alkylamino groups contributed to the good light absorption for NDP2, NDP4, and NDP6 structures, which allowed for favorable efficiency in polymerization experiments.
Four naphthalimide derivatives, NDA1–NDA4 with amino or alkylthio substituents, were designed by Xiao et al. [72]. The maximum absorption wavelengths of NDA1, NDA2, NDA3, and NDA4 in acetonitrile are 416 nm, 431 nm, 387 nm, and 439 nm, respectively. Compared to others, the NDA3 exhibited a blue-shifted absorption, which could be attributed to the alkylthio substituent in a naphthalimide skeleton. Steady-state photolysis experiments of NDAs/Iod1 systems exhibit high photochemical reactivity, and favorable electron transfer reaction processes was found. Moreover, the generated phenyl radicals following the NDA2/Iod1 electron transfer reaction were detected in ESR-ST experiments. These naphthalimide derivatives demonstrated good performance in CP, and the epoxy function conversions were all higher than 58% for NDA1/Iod1, NDA2/Iod1, and NDA3/Iod1 systems upon exposure to LED at 405 and 455 nm.
Yu et al. [73] prepared six naphthalimide aryl sulfide derivatives, NAS1–NAS6, and their maximum absorption wavelengths were 389 nm, 385 nm, 340 nm, 387 nm, 391 nm, and 395 nm, respectively. Due to the electron-withdrawing substituents (acetyl, nitro, and fluoro group), blue-shifted absorption was found for NAS2, NAS3, and NAS4, compared to NAS1. The values of free energy changes ΔG for NASs/Iod1 systems were negative, which ensured favorable electron transfer. The CP of the epoxy monomer was carried out for NASs/Iod1 systems using LED at 405 nm. The NAS6/Iod1 system exhibited the best initiation performance, and the epoxy function conversion was 56%. In addition, the C(aryl)-S bond could dissociate to produce active radicals. For different structures, the cleavage of C-S bonds in different positions resulted in the difference in photoinitiation ability. The function conversion of HDDA for NAS6 (0.5 wt%) reached up to 83% upon LED exposure at 405 nm. The electron-rich radicals with isopropyl and methyl groups generated by NAS6 could induce the polymerization of acrylate easily [74]. These naphthalimide aryl sulfide derivatives had high stability under sunlight and demonstrated potential application in the photocuring field.
A PI named Naphth-Iod was designed by Zivic et al. [75]. The λmax of Naphth-Iod was 340 nm, and the absorption band reached up to 390 nm. The CP of EPOX was carried out under LED at 365 nm irradiation. Final epoxy conversions were 40% and 43% in the presence of 1 wt% and 2 wt% Naphth-Iod, respectively, at 800 s. The CP of triethyleneglycol divinyl ether was also carried out on a BaF2 pellet in laminate, and a high vinyl ether function conversion 90% was obtained within 100 s upon exposure to LED at 365 nm at room temperature. In addition, Naphth-Iod could also be used in FRP. The acrylate function conversion of 2,2-bis-[4-(methacryloxy-2-hydroxy-propoxy)-phenyl]-propane (Bis-GMA) and triethylene glycol dimethacrylate (TEGDMA) blend (70/30) was 93% using 2 wt% Naphth-Iod. High-function conversions demonstrated that Naphth-Iod could be used as a versatile PI in CP and FRP. For the photolysis experiment, the absorbance of PI decreased upon exposure to LED at 365 nm. Moreover, the phenyl radical was detected in spin-trapping ESR experiments. Based on these results, a mechanism was proposed. The C-I bond broke to produce a phenyl radical for FRP, and the cation was generated for CP through an in-cage process. In addition, the singlet and triplet excited states energy were higher than the bond dissociation energy (C-I), which was in agreement with the favorable cleavage process.
Table 3. The chemical structures and light absorption properties of naphthalimide-based photoinitiators.
Table 3. The chemical structures and light absorption properties of naphthalimide-based photoinitiators.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i020
NDP1
λmax ~ 421
εmax ~ 620
ε405nm ~560
[71]
Polymers 15 00342 i021
NDP2
λmax ~ 417
εmax ~ 5600
ε405nm ~ 5100
[71]
Polymers 15 00342 i022
NDP3
λmax ~ 334
εmax ~ 13,100
[71]
Polymers 15 00342 i023
NDP4
λmax ~ 426
εmax ~ 9800
ε405nm ~ 8200
[71]
Polymers 15 00342 i024
NDP5
λmax ~ 340
εmax ~ 17,800
[71]
Polymers 15 00342 i025
NDP6
λmax ~ 431
εmax ~ 17,400
ε405nm ~ 12,100
[71]
Polymers 15 00342 i026
NDP7
λmax ~ 440
εmax ~ 11,300
ε405nm ~ 8800
[71]
Polymers 15 00342 i027
NDA1
λmax ~ 416
εmax ~ 4600
ε405nm ~ 4300
ε455nm ~ 1200
[72]
Polymers 15 00342 i028
NDA2
λmax ~ 431
εmax ~ 14,600
ε405nm ~ 10,300
[72]
Polymers 15 00342 i029
NDA3
λmax ~ 387
εmax ~ 18,000
ε405nm ~ 13,000
ε455nm ~ 1000
[72]
Polymers 15 00342 i030
NDA4
λmax ~ 439
εmax ~ 16,300
ε405nm ~ 9300
[72]
Polymers 15 00342 i031
NAS1
λmax ~ 389
ε405nm ~ 10,100
[73]
Polymers 15 00342 i032
NAS2
λmax ~ 385
ε405nm ~ 9300
[73]
Polymers 15 00342 i033
NAS3
λmax ~ 340
ε405nm ~ 6100
[73]
Polymers 15 00342 i034
NAS4
λmax ~ 387
ε405nm ~ 12,800
[73]
Polymers 15 00342 i035
NAS5
λmax ~ 391
ε405nm ~ 12,600
[73]
Polymers 15 00342 i036
NAS6
λmax ~ 395
ε405nm ~ 11,700
[73]
Polymers 15 00342 i037
Naphth-Iod
λmax ~ 340[75]

2.4. Coumarine-Based Photoinitiators

Coumarin derivatives are widely applied in many fields, such as fluorescent bio labels and organic light-emitting diodes. Through chemical modification, a wide variety of electronic, photochemical and photophysical properties can be obtained for coumarin derivatives [76]. Therefore, coumarin is a good scaffold to use in the design of PIs. The chemical structures of coumarin-based PIs mentioned in this review are shown in Table 4. Two coumarin derivatives (CoumA and CoumB) were designed by Abdallah et al. [77]. Good light absorption (CoumA: λmax = 421 nm; CoumB: λmax = 405 nm) ensured the availability of LED at 405 nm. The polymerization of EPOX (25 μm) in the air was studied using LED at 405 nm. A fast polymerization rate and high epoxy function conversion (FC = 80%) were found for the CoumA/Iod system. However, no polymerization of EPOX was observed for the CoumB/Iod system in the same condition. The acrylate function conversions of TMPTA were 81% and 93% for the CoumA/Iod/NPG and CoumB/Iod/NPG systems. Efficient interactions between coumarins and Iod were observed. Indeed, coumarin derivatives as photoredox catalysts were efficient when amine was used as an electron donor and iodonium salt as an electron acceptor. In addition, water-soluble CoumB was also used for hydrogel synthesis. Finally, free radical polymerization in 3D printing was successfully proved due to the high performance of CoumA and CoumB.
Some derivatives bearing an iodonium salt moiety were designed as efficient one-component PIs by Topa et al. [78]. Their maximum absorption wavelengths were 350 nm. The polymerization of EPOX was evaluated using coumarin derivatives alone under air. 7M-P demonstrated the best initiation ability among them, and the epoxy function conversions were 58% and 53% using LED at 365 and 405 nm, respectively. Due to the good performance of coumarin-based PIs in CP and FRP, 3D-printed objects were fabricated successfully using TMPTA/EPOX blends. In steady-state photolysis experiments, the absorbance of coumarin derivatives at maximum absorption wavelength decreased obviously under LED at 365 and 405 nm irradiation. According to laser flash photolysis (LFP) experiments, the iodonium salts dissociated by homolytic or heterolytic cleavage of carbon–iodine bond. The BDE of coumarin moiety-iodine was higher than that of I-Ph-R bond. It indicated the cleavage in I-Ph-R moiety was more favorable. In addition, the aryl radicals were detected by spin-trapping ESR experiments. Based on the above results, a cleavage mechanism was proposed.
Table 4. The chemical structures and light absorption properties of coumarin-based photoinitiators.
Table 4. The chemical structures and light absorption properties of coumarin-based photoinitiators.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i038
CoumA
λmax ~ 421
εmax ~ 35,200
ε405nm ~ 30,600
[77]
Polymers 15 00342 i039
CoumB
λmax ~ 405
εmax ~ 28,100
ε405nm ~ 28,100
[77]
Polymers 15 00342 i040
7M-P
λmax ~ 350
εmax ~ 20,440
ε365nm ~ 16,660
ε405nm ~ 280
[78]
Polymers 15 00342 i041
7M-CN-P
λmax ~ 352
εmax ~ 18,930
ε365nm ~ 16,800
ε405nm ~ 360
[78]
Polymers 15 00342 i042
7M-NO2-P
λmax ~ 351
εmax ~ 19,100
ε365nm ~ 16,770
ε405nm ~ 460
[78]
Polymers 15 00342 i043
7M-Me-P
λmax ~ 349
εmax ~ 22,700
ε365nm ~ 18,200
ε405nm ~ 240
[78]
Polymers 15 00342 i044
7M-iPr-P
λmax ~ 350
εmax ~ 18,440
ε365nm ~ 14,960
ε405nm ~ 220
[78]

2.5. Chalcone-Based Photoinitiators

Chalcone is a natural scaffold that has been found in numerous plants. Chalcones have been applied in many fields, including organogels and organic photovoltaics [79]. Interestingly, chalcones can be synthesized directly by condensation of an aldehyde with acetophenone [80,81]. Markedly, chalcones also undergo a competing [2+2] photodimerization, which can be employed for photocrosslinking. Therefore, as an environment-friendly and available dye, chalcone has been employed to design bioinspired PIs in recent years [82,83,84]. The chemical structures of chalcone-based PIs mentioned in this review are shown in Table 5.
A series of chalcones bearing carbazole or triphenylamine moiety was designed by Chen et al. [85]. Carbazole or triphenylamine were outstanding electron donors. The combination of carbazole or triphenylamine with chalcone formed D-π-A or A-π-D-π-A structures. Except for Chalcones 2, 3, 5, and 6, λmax of all the compounds were located in the visible region (λ > 400 nm). Particularly, high molar extinction coefficients were observed for Chalcone 4, Chalcone 7 and Chalcone 10. PEG-diacrylate (SR 610) was used in FRP under LED at 405 nm irradiation. Chalcones/Iod and Chalcones/Iod/EDB systems exhibit favorable efficiency in thick molds (2 mm) and thin films (0.1 mm). Chalcone 4, 7, and 10 exhibited the best performance in Chalcones/Iod and Chalcones/Iod/EDB systems. High acrylate function conversions of PEG-diacrylate in thick molds were observed for Chalcone 4/Iod/EDB (FC = 95%), Chalcone 7/Iod/EDB (FC = 94%), and Chalcone 10/Iod/EDB (FC = 92%). 3D-printed patterns were successfully obtained for Chalcones/Iod systems through laser write experiments. In addition, the shapes of 3D patterns using the Chalcone 7/Iod system had reversible deformation behavior because of the hydrophilic response.
A series of bis-chalcone compounds were proposed by Chen et al. [86]. These bis-chalcone compounds exhibit favorable absorption. High ε405nm for Bis-chalcone 5 and Bis-chalcone9 were observed. The FRP of PEG-acrylate monomer in laminate (~20 μm) was evaluated using bis-chalcone/Iod/EDB systems upon exposure to LED at 405 nm. The structures Bis-chalcone 5, 7, and 9 exhibited good initiation ability, and the acrylate function conversions were 62%, 95%, and 91%, respectively. Due to the high photoactivity, 3D-printed objects were successfully fabricated in laser write experiments (laser diode at 405 nm) using Bis-chalcone 5 and 9. In addition, other chalcone compounds were synthesized by Chen et al. [87,88]. These studies provide a powerful tool for designing the PISs used in LED photopolymerization.
Table 5. The chemical structures and light absorption properties of chalcone-based photoinitiators.
Table 5. The chemical structures and light absorption properties of chalcone-based photoinitiators.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i045
Chalcone 1
λmax ~ 425
εmax ~ 8930
ε405nm ~ 7450
[85]
Polymers 15 00342 i046
Chalcone 2
λmax ~ 369
εmax ~ 20,520
ε405nm ~ 4830
[85]
Polymers 15 00342 i047
Chalcone 3
λmax ~ 369
εmax ~ 17,740
ε405nm ~ 5960
[85]
Polymers 15 00342 i048
Chalcone 4
λmax ~ 408
εmax ~ 23,900
ε405nm ~ 23,580
[85]
Polymers 15 00342 i049
Chalcone 5
λmax ~ 370
εmax ~ 21,100
ε405nm ~ 7020
[85]
Polymers 15 00342 i050
Chalcone 6
λmax ~ 360
εmax ~ 24,900
ε405nm ~ 3530
[85]
Polymers 15 00342 i051
Chalcone 7
λmax ~ 405
εmax ~ 18,740
ε405nm ~ 18,740
[85]
Polymers 15 00342 i052
Chalcone 8
λmax ~ 430
εmax ~ 7990
ε405nm ~ 6760
[85]
Polymers 15 00342 i053
Chalcone 9
λmax ~ 428
εmax ~ 8540
ε405nm ~ 7200
[85]
Polymers 15 00342 i054
Chalcone 10
λmax ~ 430
εmax ~ 10,500
ε405nm ~ 9020
[85]
Polymers 15 00342 i055
Bis-chalcone 1
λmax ~ 370
εmax ~ 21,900
ε405nm ~ 12,740
[86]
Polymers 15 00342 i056
Bis-chalcone 2
λmax ~ 380
εmax ~ 28,200
ε405nm ~ 19,730
[86]
Polymers 15 00342 i057
Bis-chalcone 3
λmax ~ 332
εmax ~ 18,200
ε405nm ~ 4420
[86]
Polymers 15 00342 i058
Bis-chalcone 4
λmax ~ 369
εmax ~ 4980
ε405nm ~ 2550
[86]
Polymers 15 00342 i059
Bis-chalcone 5
λmax ~ 347
εmax ~ 23,100
ε405nm ~ 6800
[86]
Polymers 15 00342 i060
Bis-chalcone 6
λmax ~ 364
εmax ~ 22,700
ε405nm ~ 10,070
[86]
Polymers 15 00342 i061
Bis-chalcone 7
λmax ~ 430
εmax ~ 38,900
ε405nm ~ 26,420
[86]
Polymers 15 00342 i062
Bis-chalcone8
λmax ~ 330
εmax ~ 19,800
ε405nm ~ 2960
[86]
Polymers 15 00342 i063
Bis-chalcone 9
λmax ~ 370
εmax ~ 24,600
ε405nm ~ 13,670
[86]
Polymers 15 00342 i064
Bis-chalcone 10
λmax ~ 350
εmax ~ 49,300
ε405nm ~ 2220
[86]

3. Type Ⅱ Photoinitiators

Many studies have focused on Type Ⅱ PIs which can interact with hydrogen donors through H-abstraction reaction to generate two free radicals. Some ketone-type compounds, including benzophenone (BP), thioxanthone (TX), and camphorquinone (CQ), are widely used as Type II PIs. Tertiary amines and thiols are normally used as hydrogen donors due to their good reductant behavior [89,90]. After electron transfer and proton transfer, two free radicals are generated. Because of steric hindrance and delocalization of unpaired electrons, ketyl radicals from ketone compounds have no reactivity to double bonds [91]. The absorption for most traditional Type Ⅱ PIs are mainly located in the UV region, so many efforts have been devoted to expanding the light absorption range of Type Ⅱ PIs to match the near UV or visible LEDs. Introducing the Type Ⅱ PI structures onto dye scaffolds to obtain π-conjugated structures is an effective method. The π-conjugated structure is expected to expand the absorption and achieve good H-abstraction ability. The obtained structures can be classified as dye-based PISs. However, compared to the general dye molecules, the obtained structures demonstrate a more obvious feature of Type Ⅱ PI in the presence of hydrogen donors. Therefore, these structures are separately listed as Type Ⅱ PIs in this review. Benzophenone and thioxanthone are the most commonly employed, and some published chemical structures are introduced in the following context.

3.1. Benzophenone Photoinitiators

As a traditional UV-sensitive PI, BP does not absorb light in longer wavelength regions (>380 nm). BP was modified widely to obtain better photoinitiation ability. Many groups, including amino, thioether, and oxime, were induced into the BP skeleton, and these BP derivatives demonstrated good efficiency under UV light or halogen lamp [92,93,94,95,96,97,98,99,100,101,102]. Herein, the BP derivatives proposed in recent years are mainly summarized. The chemical structures of benzophenone PIs mentioned in this review are presented in Table 6.
A Type Ⅱ PI (Py_BP) based on the BP group linked to a pyrene moiety was designed by Tehfe et al. [103]. The maximum absorption wavelength of Py_BP was 348 nm, and high values of ε348nm = 26,000 M−1 cm−1, and ε405nm = 1400 M−1 cm−1 were found. The FRP of TMPTA was carried out upon halogen lamp exposure (λ > 300 nm), and no polymerization was found for BP/amine system. The function conversion of TMPTA reached up to 35% using Py_BP/N-methyldiethanolamine (MDEA) (1%/4%, wt%/wt%) system. Indeed, the good light absorption properties and H-abstraction ability allowed an effective Type Ⅱ behavior for Py_BP. A good polymerization profile (FC = 55% at 800 s) was also observed in the CP of EPOX upon halogen lamp exposure for Py_BP/Iod (0.5%/2%, wt%/wt%) system. All results indicated that Py_BP was a reactive PI for both FRP and CP. Another BP derivative (Py_BP5) with the coupling of BP and pyrene moieties was designed by Telitel et al. [104]. Py_BP5 had a hybrid structure where BP and pyrene were fused via a benzene ring. The maximum absorption wavelength of Py_BP5 was 395 nm. Compared to Py_BP, Py_BP5 had a red-shifted absorption wavelength which could be attributed to the strong molecular orbital coupling. Free radical polymerization was carried out under Xe-Hg lamp with a filter (λ > 340 nm) irradiation. The function conversion of 40% of TMPTA was obtained using Py_BP5/MDEA (1%/2%, wt%/wt%) system at t = 200 s. Indeed, combining BP with a light-absorbing moiety is a good method to improve the light absorption properties of BP derivatives, and it was also reported in other works [105].
Zhang et al. [106] connected the BP group with naphthalimide chromophore to design another Type Ⅱ PI BPND. Good light absorption properties (λmax = 431 nm, ε405nm = 15,700 M−1 cm−1) were obtained. In photolysis experiments, the absorbance of BPND at 431 nm decreased slowly during the irradiation at 405 nm, on account of the H-abstraction for two BPND molecules probably. The generated aminoalkyl radical was detected in spin-trapping ESR experiments. When Iod was added, fast bleaching was observed under irradiation. In FRP experiments, BPND alone could initiate the polymerization of TMPTA as one-component Type Ⅱ PI. Interestingly, BPND/Iod system exhibit excellent and similar photoinitiation ability (55~57%) under LED at 405, 455 and 470 nm irradiation. High function conversions (56~70%) were also obtained in CP for EPOX in the presence of the BPND/Iod system. The high performance of BPND benefits from the good design of the molecular structure. In a word, the ketone-dye-based compounds demonstrate great potential in the design of PIs for LED photoinitiation.
Five visible light benzophenone-based PIs (BP1–BP5) were synthesized by Huang et al. [107]. In these structures, BP was incorporated into different arylamin moieties. The UV absorption spectra of BPs were investigated in a dichloromethane solution. Compared to BP, BP derivatives exhibit red-shifted absorption, and λmax of them were all located in a range of 360 to 375 nm. Due to the good absorption properties and the BP moiety, these PIs were expected to act as efficient visible Type Ⅱ PIs in FRP. The photolysis of BPs/triethylamine (TEA) systems were studied under 365 nm light irradiation. The absorption band at ~360 nm decreased; meanwhile, an increasing peak at ~290 nm was observed. The FRP of TMPTA under a UV lamp was evaluated by photo-DSC in the presence of BPs/TEA systems. The reference BP/TEA system demonstrated the best performance among them. No polymerization was observed for BP3/EDB system. The polymerization of TMPTA was also evaluated under white light LED irradiation (380–750 nm). High function conversions of TMPTA were found for BP-1/TEA and BP/TEA systems, respectively. Indeed, good light absorption properties ensured the efficient photoinitiation ability for BP-1 as a Type Ⅱ PI.
Three compounds named C-DBP, P-DBP and T-DBP were designed as PIs for LED photopolymerization by Jia et al. [108]. These chemical structures incorporated benzophenone as an electron acceptor and carbazole/phenothiazine/triphenylamine as an electron donor. Considering the photo-isomerization of the double bond, the triple bond is a desired π-linker for D-π-A structures to extend the conjugated structures. These PIs had favorable absorption bands in the region of 300 to 450 nm. The FRP of tripropylene glycol diacrylate (TPGDA) were carried out upon exposure to LED at 405 nm using DBP/triethanolamine (TEOA) system. It was interesting that C-DBP/TEOA (0.25%/2%, w/w) exhiexhibitedd photoinitiation ability and the acrylate function conversion reached up to 95%, while P-DBP and T-DBP systems demonstrated poor performance in polymerization experiments. In fluorescence and nanosecond transient absorption experiments, C-DBP demonstrated both BP-like features, contributing to good H-abstraction ability. Thus, C-DBP had good performance in free radical photopolymerization.
A benzophenone derivative named BPN was designed by Xue et al. [109]. BPN had strong adsorption in the range of 320~500 nm (ε405nm = 42,400 M−1 cm−1), indicating BPN could be used as visible light PI under LED illumination. It was interesting that the polymerization of the TPGDA blends were observed at LED at 405 nm without any additional co-initiator. The acrylate function conversion was 80% at 80 s in the presence of 0.5 wt% BPN. It was attributed to dimethyl amine moiety in the structure. Interestingly, the faster polymerization rate and higher function conversion (90%) were obtained using BPN/TEOA system. BPN demonstrated the same performance as benchmark ITX in polymerization experiments. Fast photolysis was observed for BPN alone upon exposure to LED at 405 nm. The reason was that an H-abstraction reaction occurred between benzophenone and amine groups in two BPN molecules. In addition, the absorption of BPN declined fast in the presence of TEOA. Indeed, the H-abstraction reaction mainly occurred in BPN/TEOA bimolecular system. In spin-trapping ESR experiments, the generated BPN(-H) radical and TEOA(-H) radical were detected, which verified the H-abstraction reaction in unimolecular and bimolecular systems, respectively. Because of its high performance, BPN has great potential to be used in photoinitiation.
A series of benzophenone-carbazole compounds (BPC-BPC4) was investigated by Liu et al. [110,111]. Interestingly, the benzoyl substituent is connected with carbazole to generate a benzophenone moiety. The benzophenone-carbazole structure demonstrates the feature of the Type Ⅱ photoinitiator (PI). Good absorption in the UV region was found (εmax = 18,600 M−1 cm−1, λmax = 342 nm) for BPC and the value of ε365nm was favorable. Therefore, their photoinitiation performances of them were studied using a LED at 365 nm. Benzophenone-carbazole PIs could initiate the polymerization of monomers alone. In addition, better performances were found for two-component systems (FC = 60% for BPC/EDB system; FC = 63% for BPC/Iod system). Besides, these PIs demonstrated favorable performance in cationic polymerization of EPOX (FC = 46% for BPC/Iod system).
Some benzophenone-triphenylamine PIs (BT1–BT4) were designed for LED-induced photopolymerization [112]. Interestingly, benzophenone-triphenylamine PIs exhibit a favorable molar extinction coefficient at 405 nm (e.g., ε405nm (BT3) = 6100 M−1 cm−1, ε405nm (BT4) = 6700 M−1 cm−1). Some PIs demonstrated better performances than benchmark isopropylthioxanthone (ITX) in free radical polymerization upon exposure to LED at 405 nm. In cationic polymerization, good polymerization profiles of EPOX were found. Finally, 3D-printed objects were obtained successfully using the developed new PIs, and these chemical structures demonstrated rational design for PIs [112,113].
Table 6. The chemical structures of benzophenone photoinitiators and their absorption properties.
Table 6. The chemical structures of benzophenone photoinitiators and their absorption properties.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i065
BP
λmax ~ 253
εmax ~ 22,000
ε363nm ~250
[104,107]
Polymers 15 00342 i066
BPB
λmax ~ 290
εmax ~ 14,610
[100]
Polymers 15 00342 i067
Py_BP
λmax ~ 348
εmax ~ 26,000
ε405nm ~1400
[103]
Polymers 15 00342 i068
Py_BP5
λmax ~ 395
εmax ~ 4400
[104]
Polymers 15 00342 i069
BPND
λmax ~ 431
εmax ~ 15,700
ε405nm ~ 10,900
ε470nm ~ 5700
[106]
Polymers 15 00342 i070
BP1
λmax ~ 369
εmax ~ 22,300
[107]
Polymers 15 00342 i071
BP2
λmax ~ 341
εmax ~ 13,300
[107]
Polymers 15 00342 i072
BP3
λmax ~ 368
εmax ~ 20,200
[107]
Polymers 15 00342 i073
BP4
λmax ~ 369
εmax ~ 42,900
[107]
Polymers 15 00342 i074
BP5
λmax ~ 374
εmax ~ 27,800
[107]
Polymers 15 00342 i075
A4
λmax ~ 349
εmax ~ 18,000
[108]
Polymers 15 00342 i076
A3–1
λmax ~ 350
εmax ~ 49,000
[108]
Polymers 15 00342 i077
A3–2
λmax ~ 360
εmax ~ 53,000
[108]
Polymers 15 00342 i078
BPN
λmax ~ 400
εmax ~ 43,700
ε405nm ~ 42,400
[109]
Polymers 15 00342 i079
BPC
λmax ~ 342
εmax ~ 18,600
ε365nm ~ 6000
[110]
Polymers 15 00342 i080
BPC1
λmax ~ 334
εmax ~ 13,910
ε365nm ~ 3270
[111]
Polymers 15 00342 i081
BPC2
λmax ~ 325
εmax ~ 13,900
ε365nm ~ 2210
[111]
Polymers 15 00342 i082
BPC3
λmax ~ 334
εmax ~ 13,350
ε365nm ~ 3460
[111]
Polymers 15 00342 i083
BPC4
λmax ~ 325
εmax ~ 12,400
ε365nm ~ 2170
[111]
Polymers 15 00342 i084
BT1
λmax ~ 359
εmax ~ 21,000
ε405nm ~ 1800
[112]
Polymers 15 00342 i085
BT2
λmax ~ 373
εmax ~ 27,200
ε405nm ~ 5000
[112]
Polymers 15 00342 i086
BT3
λmax ~ 370
εmax ~ 41,600
ε405nm ~ 6100
[112]
Polymers 15 00342 i087
BT4
λmax ~ 377
εmax ~ 21,700
ε405nm ~ 6700
[112]

3.2. Thioxanthone Photoinitiators

Thioxanthone (TX) is another efficient Type Ⅱ PI, and the TX derivatives are widely used in FRP and CP. Compared to BP, TX exhibits a longer absorption wavelength in the range of 300–400 nm, so TX derivatives are in a class of potential visible light PIs [114,115,116]. As known, the commercial isopropylthioxanthone (ITX) is widely used as a benchmark Type Ⅱ PI. The chemical structures of thioxanthone PIs mentioned in this review are given in Table 7.
A thioxanthone-carbazole derivative (TX-C) was synthesized by Yilmaz et al. [117,118,119]. TX moiety was in conjunction with carbazole chromophore to obtain light absorption in the visible range. TX-C demonstrated good absorption at a longer wavelength range (400–500 nm), where TX was almost transparent. Indeed, the extended conjugation of TX and carbazole moieties contributed to the longer absorption wavelength. The λmax was 434 nm for TX-C, and a high molar extinction coefficient was found (ε434nm = 2010 M−1 cm−1). In the steady-state photolysis experiments, the absorbance at 434 nm of TX-C decreased. It indicated that an H-abstraction reaction occurred between TX and the amine group in the carbazole moiety. Therefore, TX-C could be used as a one-component Type Ⅱ PI without an additional hydrogen donor. The initiation ability of PI was evaluated upon exposure to visible light (430–490 nm). There is no polymerization of monomers in the presence of the TX/amine system due to the poor absorption property of TX in the visible range. As expected, TX-C could initiate the polymerization of methyl methacrylate (MMA) alone. The excited state TX-C reacted with the ground state TX-C to generate carbazoyl radical, which was active in the polymerization of MMA. When the amine (hydrogen donor) was added, TX-C/amine system demonstrated a faster polymerization rate than one-component TX-C. Indeed, in the presence of amine, H-abstraction occurred easily between TX-C and amine. In summary, TX-C was an excellent visible Type Ⅱ PI and had many targeted applications.
A thioxanthone-anthracene PI (TX-A) was synthesized by Balta et al. [120,121]. The maximum absorption wavelength of TX-A was 368 nm and the absorption band extended to 450 nm. In photolysis experiments, the absorption band of the anthracene moiety almost disappeared with the consumption of TX-A, while the absorption of thioxanthone moiety at 380 nm was still observed. TX-A was used as a PI for the FRP of MMA. It was strange that higher function conversions were found in the presence of TX-A under air, compared to TX-A/amine system. In addition, no polymerization was found under nitrogen. Based on the photolysis and polymerization experiments, the conclusions could be obtained that TX moiety did not likely participate in the H-abstraction reaction, and oxygen played an important role in polymerization. As known, when the triplet oxygen (ground state molecular oxygen) is quenched, singlet oxygen will produce through energy transfer. In an air-saturated TX-A solution, the generation of singlet oxygen was confirmed by the NIR luminescence spectrum. Therefore, singlet oxygen could react with anthracene moiety to generate endoperoxide. Then endoperoxide decomposed into radicals that were active for monomers. The chemical mechanism of TX-A is helpful in overcoming oxygen inhibition in FRP processes [122,123].
A PI named TX-NPG was designed by Tar et al. [124]. TX-NPG exhibit good light absorption in the region of 300–600 nm in N,N-dimethylformamide. Favorable molar extinction coefficients at 392 nm (ε392nm = 1670 M−1 cm−1) and 583 nm (ε583nm = 440 M−1 cm−1) were observed. The red-shifted band (~583 nm) disappeared when triethylamine was added to the solution, which could be attributed to the intramolecular or intermolecular hydrogen bond in TX-NPG solution. This peculiar absorption was also reported in other TX derivatives, such as thioxanthone carboxylic acid and sodium fluorenecarboxylate-thioxanthone [115,125]. Favorable absorption properties made TX-NPG attractive as a visible PI. TX-NPG was investigated as a Type Ⅱ PI in the photopolymerization experiments under irradiation with 392, 473, 532, and 635 nm. In the presence of TX-NPG alone, the function conversions of acrylate monomers under irradiation at 392, 473, 532, and 635 nm were 63%, 21%, 24% and 14%, respectively. This was attributed to the H-abstraction reaction in a one-component system. The addition of MDEA as a hydrogen donor contributed to the H-abstraction reaction, and higher final function conversions were obtained. The performance of TX-NPG demonstrated the possibility of its use under panchromatic irradiation. In addition, several TX derivatives bearing abstractable hydrogen sites (amine group) were also designed as one-component PIs [126,127,128].
Three TX derivatives (TX-2CBZ, TX-2DPA, TX-2PTZ) with D-A-D structures were designed by Mau et al. [129]. In the D-A-D structures, the TX moiety acted as the acceptor and dimethoxyphenylamine, carbazole, phenothiazine moieties acted as electron donors, respectively. These D-A-D structures were expected to obtain enhanced light absorption. Three investigated TX derivatives exhibited a bathochromic shift (λmax = 396 nm for TX-2CBZ; λmax = 478 nm for TX-2DPA; λmax = 415 nm for TX-2PTZ) compared to ITX (λmax = 386 nm). Meanwhile, three compounds demonstrated higher extinction coefficients at 405 nm (ε405nm = 5900 M−1 cm−1 for TX-2CBZ; ε405nm = 2000 M−1 cm−1 for TX-2DPA; ε405nm = 2600 M−1 cm−1 for TX-2PTZ) than ITX (ε405nm = 1000 M−1 cm−1). These PIs were evaluated as Type Ⅱ PIs in the presence of EDB. TX-2CBZ and ITX exhibited good performances in the polymerization of TMPTA, and function conversions were 83% and 84%, respectively. Furthermore, poor efficiencies were found in the presence of TX-2DPA and TX-2PTZ.
To explore the relationships between structure and efficiency, the singlet and triplet states of TX derivatives were detected. The results showed that a thioxanthone triplet state was found in TX-2CBZ and ITX structures, and the lifetime were about 1.0 μs and 5.8 μs, respectively. TX-2CBZ demonstrated strong H-abstraction ability due to the TX moiety in structure, while no significant H-abstraction was observed for TX-2DPA and TX-2PTZ. It could possibly be ascribed to the internal conversion as a deactivation pathway of the excited state, which resulted in the slow reaction with EDB. The TX-2CBZ/Iod system also exhibited good performance in the CP of EPOX (FC = 59%). In the 3D-printing experiments, the pattern was successfully fabricated using a three-component TX-2CBZ/Iod/EDB system with TMPTA/EPOX blend. The PI TX-2CBZ, which had a similar performance to ITX, was applied in 3D printing [130].
Table 7. The chemical structures of thioxanthone photoinitiators and their absorption properties.
Table 7. The chemical structures of thioxanthone photoinitiators and their absorption properties.
Chemical StructuresAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i088
TX
λmax ~ 380
εmax ~ 5300
[131]
Polymers 15 00342 i089
ITX
λmax ~ 386
εmax ~ 6500
ε395nm ~ 3900
[132]
Polymers 15 00342 i090
TX-C
λmax ~ 434
εmax ~ 2010
[118,119]
Polymers 15 00342 i091
TX-A
λmax ~ 368
εmax ~ 14,000
[120,121]
Polymers 15 00342 i092
TX-NPG
λmax ~ 392,583
ε392nm ~ 1670
ε583nm ~ 440
[124]
Polymers 15 00342 i093
TX-MPM
λmax ~ 410
εmax ~ 4390
[126]
Polymers 15 00342 i094
TX-1
λmax ~ 438
εmax ~ 4400
[127]
Polymers 15 00342 i095
TX-2
λmax ~ 438
εmax ~ 4800
[127]
Polymers 15 00342 i096
TX-3
λmax ~ 444
εmax ~ 4100
[127]
Polymers 15 00342 i097
TX-MPA
λmax ~ 407
εmax ~ 3610
[128]
Polymers 15 00342 i098
TX-2DPA
λmax ~ 478
εmax ~ 3400
ε405nm ~ 2000
[129]
Polymers 15 00342 i099
TX-2CBZ
λmax ~ 396
εmax ~ 7900
ε405nm ~ 5900
[129]
Polymers 15 00342 i100
TX-2PTZ
λmax ~ 305,415
ε305nm ~ 2400
ε405nm ~ 2600
[129]

4. Type Ⅰ Photoinitiators

Type Ⅰ PIs are used widely in industry and research laboratories. Oxime esters, acylphosphine oxides, and amino ketones are commonly used in photopolymerization. However, except for several commercial PIs, such as diphenyl(2,4,6-trimethylbenzoyl)phosphine oxide (TPO) and 1-benzyl-1-(dimethylamino)propyl 4-morpholinophenyl ketone (369), most of Type Ⅰ PIs do not absorb light above 400 nm. Therefore, it is important to improve the absorption ability of Type Ⅰ PIs to match the emission spectra of long-wavelength LEDs. Herein, oxime esters and acylphosphine oxides are mainly summarized in this review.

4.1. Oxime Ester Photoinitiators

Oxime esters have been investigated widely as Type Ⅰ PIs due to their high photoactivity [133,134]. Commercial oxime esters, OXE 01 and OXE 02, have been used to successfully produce thin films in color filter resists [135]. Under irradiation, the N−O bond in the oxime ester structure can break to produce iminyl and acyloxy radicals. The decarboxylation reaction occurs in the acyloxy radical to produce CO2 and in another active radical that can induce free radical polymerization. As the OXE 01 and OXE 02 can absorb UV light only, they perform poorly under visible light. Therefore, the design of novel oxime esters which can be used under visible LEDs is important. The chemical structures of oxime ester PIs mentioned in this review are shown in Table 8. Due to the different structures and substituents, the presented oxime esters exhibited very different light absorption properties, as shown in their maximal absorption wavelengths (λmax in Table 8).
A series of oxime ester PIs based on the coumarin chromophore were synthesized by Li et al. [136]. They were designed to study the substituent and electronic effects. These PIs exhibited strong absorption, and the λmax of O-3 and O-4 were 436 nm (ε436nm = 41,690 M−1 cm−1) and 433 nm (ε433nm = 7680 M−1 cm−1), respectively. Compared to O-4, O-3 demonstrated a dramatically increased molar extinction coefficient, which could be ascribed to the substitution position of the oxime-ester moiety. Stronger absorption of O-3 could be ascribed to the planar conformation and increased degree of conjugation [137]. The λmax of O-3F and O-3O were all 436 nm. The photoinitiation ability of the PIs was investigated under a LED at 450 nm. O-3 demonstrated the best performance among them due to its strong absorption. Interestingly, both high-thiol and vinyl double-bond conversions were found using O-3 in thiol-based click polymerization. O-3 exhibit photobleaching property, and the curing depth was 2.6 mm within 10 min irradiation upon exposure to LED at 450 nm (200 mW/cm2). Indeed, the colorless photoreaction products allowed the light to penetrate deeper into the formulation. In addition, some oxime ester PIs based on coumarin chromophores were investigated and applied in 3D printing [138,139].
Hammoud et al. [140] designed some oxime esters based on the coumarin scaffold. All the chemical structures exhibited favorable molar extinction coefficients at 405 nm (i.e., OXE-D: ε405nm = 22,500 M−1 cm−1; OXE-J: ε405nm = 25,000 M−1 cm−1). The Type Ⅰ photoinitiation behaviors of the oxime esters were investigated upon exposure to LED at 405 nm in the FRP of TMPTA. OXE-J and OXE-D exhibited better initiation performances than the others, and the function conversions of TMPTA were 72% and 73%, respectively. According to computational calculations, the enthalpy value of the decarboxylation reaction and the cleavage process for the N–O bond was energetically favorable for OXE-D. In addition, the highest spin density of the methyl radical contributed to the good reactivity of OXE-D in polymerization experiments. Finally, a 3D-printed object was successfully obtained using an OXE-D/Iod system in acrylate monomers.
Two arylaminocarbazole oxime ester PIs (OXE1 and OXE2) were designed by Ma et al. [141]. In addition to Type Ⅰ PIs, OXE1 and OXE2 could also play the role of photosensitizer in multicomponent systems. OXE1 and OXE2 had wide absorption ranges at 200~450 nm. Under irradiation at 405 nm, the absorbance of OXE1 and OXE2 decreased to a certain extent. The results indicated the generation of stable photolysis products. When iodonium salt was added, the photolysis rate of the OXE1/iodonium salt system was faster than that of OXE1 alone. It suggested that the interaction of OXE1/iodonium salt was quick and efficient. The favorable values of free energy changes for OXE1/iodonium salt and OXE2/iodonium salt systems were obtained. In addition, the phenyl radical was detected in OXE1/iodonium salt system by spin-trapping ESR experiments. The FRP of TPGDA was studied upon exposure to the laser diode at 405 nm. TPGDA polymerized because of the oxime ester moiety in these structures. Higher function conversions were found using OXEs/iodonium salt (0.2%/1%, w/w). The photoinitiation performance was in line with the results of photolysis and the ESR experiments. The polymerization of epichlorohydrin was evaluated using OXEs/iodonium salt systems, and favorable function conversions were obtained. Indeed, the OXEs/iodonium salt systems could generate cations through electron transfer reactions to induce epoxide ring-opening polymerization. In conclusion, the arylaminocarbazole oxime esters broaden the application of oxime esters in visible light photopolymerization.
Ding et al. [142] designed two oxime esters named E-FBOXEs. The FRP of TPGDA was investigated upon exposure to LED at 395 nm. High function conversions of 81% and 80% were obtained using E-FBOXE-Me and E-FBOXE-ph, respectively. Interestingly, photochromism of the polymers prepared by E-FBOXEs was found under heating. After polymerization, a brown PTPGDA film was obtained using E-FBOXE-Me. The brown turned colorless when the film was heated at 50 °C. The photochromic mechanism was investigated by ESR and FTIR spectra. The results demonstrated that the N−O bond in E-FBOXE-Me molecules underwent cleavage under LED at 395 nm irradiation. Then, under heating, the generated acetoxy and colorful iminyl radicals could recombine to generate neutral E-FBOXE-Me structures. This special photochemical mechanism of E-FBOXEs makes it possible to design photochromic materials.
Four oxime esters based on carbazole-coumarin fused subunit were designed by Zhou et al. [143]. All the oxime esters exhibit similar absorption spectra with the maximum absorption wavelength λmax = 374 nm. The absorbance of OXE-EM decreased gradually under LED at 365 nm irradiation. Meanwhile, this process was also monitored through 1H NMR. The results demonstrated the dissociation of the N-O bond. In addition, methyl radical was detected in spin-trapping ESR experiments. The FRP of TMPTA was evaluated by Photo-DSC. Under LED at 365, 385 and 405 nm irradiation, the photoinitiation ability of the oxime esters followed the order OXE-EM > OXE-IM > OXE-EP > OXE-IP. It was interesting that these PIs demonstrated better performance under LED at 405 nm than LED at 365 nm, although the molar extinction coefficients at 365 nm were higher. When irradiated at 365 or 405 nm, the photolysis behavior of OXE-EM was similar. A wavelength-dependent photopolymerization mechanism was proposed in this work. In addition, stilbene-based, phenyl thienyl thioether-based, and bicarbazole-based oxime ester PIs were also reported by Jin’s group [144,145,146].
Recently, three oxime esters based on the nitro-carbazole scaffold were designed (namely OXE-M, OXE-V, and OXE-P) [147]. Interestingly, OXE-M (methyl substituent) showed higher efficiency than OXE-P (phenyl substituent). It could be attributed to the decarboxylation reactions. Afterward, a series of oxime ester derivatives (i.e., D1, D2) with different substituents was designed to investigate the relationships between chemical structures and performances [148]. The light absorption properties (i.e., ε405nm (D1) = 5200 M−1 cm−1, ε405nm (D2) = 5400 M−1 cm−1) are good to use a LED at 405 nm. Interestingly, some PIs had better performance than TPO under LED at 405 nm irradiation. The effect of the substituents was studied by theoretical calculations, monitoring of CO2, and the study of free radicals. All results show that substituents have an effect on the performance of PIs in polymerization experiments via a decarboxylation reaction.
Table 8. The chemical structures and light absorption properties of oxime ester photoinitiators.
Table 8. The chemical structures and light absorption properties of oxime ester photoinitiators.
StructureAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i101
O-3
λmax ~ 436
εmax ~ 41,690
ε450nm ~ 37,450
[136]
Polymers 15 00342 i102
O-3F
λmax ~ 436
εmax ~ 29,930
ε450nm ~ 26,630
[136]
Polymers 15 00342 i103
O-3O
λmax ~ 436
εmax ~ 29,950
ε450nm ~ 26,620
[136]
Polymers 15 00342 i104
O-4
λmax ~ 433
εmax ~ 7680
ε450nm ~ 6790
[136]
Polymers 15 00342 i105
DCCA
λmax ~ 436
εmax ~ 51,000
ε450nm ~ 45,000
[138]
Polymers 15 00342 i106
OEC3–1
λmax ~ 406
εmax ~ 51,000
ε450nm ~ 21,000
[139]
Polymers 15 00342 i107
OEC3–2
λmax ~ 405
εmax ~ 42,000
ε450nm ~ 17,000
[139]
Polymers 15 00342 i108
OXE-B
λmax ~ 431
εmax ~ 33,000
ε405nm ~ 22,000
[140]
Polymers 15 00342 i109
OXE-D
λmax ~ 431
εmax ~ 34,000
ε405nm ~ 22,500
[140]
Polymers 15 00342 i110
OXE-E
λmax ~ 437
εmax ~ 28,500
ε405nm ~ 16,500
[140]
Polymers 15 00342 i111
OXE-F
λmax ~ 437
εmax ~ 36,000
ε405nm ~ 21,000
[140]
Polymers 15 00342 i112
OXE-G
λmax ~ 436
εmax ~ 31,500
ε405nm ~ 18,000
[140]
Polymers 15 00342 i113
OXE-H
λmax ~ 435
εmax ~ 26,500
ε405nm ~ 17,000
[140]
Polymers 15 00342 i114
OXE-I
λmax ~ 437[140]
Polymers 15 00342 i115
OXE-J
λmax ~ 441
εmax ~ 50,000
ε405nm ~ 25,000
[140]
Polymers 15 00342 i116
OXE-K
λmax ~ 435
εmax ~ 31,000
ε405nm ~ 18,000
[140]
Polymers 15 00342 i117
OXE1
λmax ~ 350
εmax ~ 46,900
ε405nm ~ 1100
[141]
Polymers 15 00342 i118
OXE2
λmax ~ 357
εmax ~ 29,300
ε405nm ~ 4410
[141]
Polymers 15 00342 i119
E-FBOXE-Me
λmax ~ 260
ε260nm ~ 17,980
ε395nm ~ 20
[142]
Polymers 15 00342 i120
E-FBOXE-Ph
λmax ~ 262
ε260nm ~ 51,410
ε395nm ~ 50
[142]
Polymers 15 00342 i121
OXE-EM
λmax ~ 374
εmax ~ 16,400
ε355nm ~ 12,500
[143]
Polymers 15 00342 i122
OXE-IM
λmax ~ 374
εmax ~ 16,000
ε355nm ~ 12,800
[143]
Polymers 15 00342 i123
OXE-EP
λmax ~ 374
εmax ~ 17,000
ε355nm ~ 13,200
[143]
Polymers 15 00342 i124
OXE-IP
λmax ~ 374
εmax ~ 16,600
ε355nm ~ 13,000
[143]
Polymers 15 00342 i125
OXE-M
λmax ~ 369
εmax ~ 13,000
ε405nm ~ 4100
[147]
Polymers 15 00342 i126
OXE-P
λmax ~ 368
εmax ~ 13,800
ε405nm ~ 4100
[147]
Polymers 15 00342 i127
OXE-V
λmax ~ 369
εmax ~ 12,400
ε405nm ~ 3900
[147]
Polymers 15 00342 i128
D1
λmax ~ 372
εmax ~ 13,000
ε405nm ~ 5200
[148]
Polymers 15 00342 i129
D2
λmax ~ 372
εmax ~ 14,000
ε405nm ~ 5400
[148]

4.2. Acylphosphine Oxide Photoinitiators

Acylphosphine oxides are another class of Type Ⅰ PI. They are widely used in many fields due to their high efficiency and good photobleaching. As the absorption of most acylphosphine oxide PIs is located in the UV range, it is important to broaden their absorption of them in the visible range to match the emission spectra of LEDs. The chemical structures of acylphosphine oxide PIs mentioned in this review are shown in Table 9.
A series of acylphosphine oxide compounds were proposed by Dietlin et al. [149]. Before synthesis, molecular modeling was used to select the potentially reactive compounds. The maximum absorption wavelength was computed to match the selected LEDs. The energy of the triplet state (ET), as well as the bond dissociation energy (BDE), were computed to explore whether the cleavage of the bond was energetically favorable (⊗H < 0). The spin density on the radical center was computed to ensure the efficient addition of the double bond; 7 molecules were selected from 86 molecules based on favorable calculated parameters. The molar extinction coefficients at 395 nm of acylphosphine oxide compounds (except ADPO-6 and ADPO-7) were higher than that of TPO-L. The FRP of TMPTA was performed under LED at 395 nm irradiation using (1 wt%). All the photoinitiators (except ADPO-5) demonstrated higher final function conversions than commercial TPO-L and BAPO. Although ADPO-5 had good absorption properties, the cleavage of the bond took place with difficulty. The rational design of acylphosphine oxide compounds is helpful in selecting an efficient PI to use in LED photopolymerization.
An acylphosphine oxide PI 4-(diethylamino)benzoyldiphenylphosphine oxide (DEAPO) was designed and synthesized by Xie et al. [150]. Compared to TPO (λmax = 380 nm), a red-shifted maximum absorption wavelength was obtained for DEAPO (λmax = 386 nm). Interestingly, the absorption band of DEAPO extended to 440 nm and high molar extinction coefficients of DEAPO was obtained (ε385nm = 43,800 M−1 cm−1; ε420nm = 5950 M−1 cm−1). The good absorption property was ascribed to the diethyl amino moiety in the DEAPO structure (prolonging the conjugation system). DEAPO was evaluated as a Type Ⅰ PI with TMPTA. Upon exposure to LED at 385 nm, the function conversion of the DEAPO system (FC = 68.7%) was higher than the TPO system (FC = 58.2%). In addition, the DEAPO system (FC = 64.3%) demonstrated better performance than the TPO system (FC = 47.1%) under LED at 420 nm irradiation. Indeed, good absorption properties for DEAPO allowed the efficient initiation ability under LEDs. The results demonstrated that DEAPO has the potential to be applied in the field of dental materials or food packaging.
Two carbazolyl-based acylphosphine oxide compounds (ETPO and ALPO) were synthesized by Wu et al. [151]. Favorable molar extinction coefficients were observed for ETPO (ε405nm = 2270 M−1 cm−1) and ALPO (ε405nm = 1300 M−1 cm−1), which could be attributed to the large rigid plane and the strong conjugate system in carbazole moiety. In steady-state photolysis experiments, the absorbance peak of ETPO and ALPO decreased dramatically under LED at 395 nm irradiation, which exhibit high photoactivity for the two compounds. The FRP of TMPTA (~40 μm) was studied upon exposure to LED at 405 nm. Acrylate function conversions for monomers were 44.7%, 75.7%, and 56.7% using ETPO, ALPO and TPO (1 wt%), respectively. Due to the bad solubility in monomers, the efficiency of ETPO in TMPTA was low. ALPO demonstrated better performance than TPO due to its high photoactivity and good absorption properties. ALPO can be used as a visible PI in many fields.
Table 9. The chemical structures and light absorption properties of acylphosphine oxide photoinitiators.
Table 9. The chemical structures and light absorption properties of acylphosphine oxide photoinitiators.
StructureAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i130
TPO
λmax ~ 380
εmax ~ 570
ε385nm ~ 510
ε420nm ~ 20
[150]
Polymers 15 00342 i131
TPO-L
λmax ~ 383
ε395nm ~ 130
[149]
Polymers 15 00342 i132
BAPO
λmax ~ 384
ε395nm ~ 660
[149]
Polymers 15 00342 i133
ADPO-1
λmax ~ 373
ε395nm ~ 300
[149]
Polymers 15 00342 i134
ADPO-2
λmax ~ 379
ε395nm ~ 390
[149]
Polymers 15 00342 i135
ADPO-3
λmax ~ 382
ε395nm ~ 1060
[149]
Polymers 15 00342 i136
ADPO-4
λmax ~ 386
ε395nm ~ 160
[149]
Polymers 15 00342 i137
ADPO-5
λmax ~ 405
ε395nm ~ 26,600
[149]
Polymers 15 00342 i138
ADPO-6
λmax ~ 405[149]
Polymers 15 00342 i139
ADPO-7
λmax ~ 384[149]
Polymers 15 00342 i140
DEAPO
λmax ~ 386
εmax ~ 43,810
ε385nm ~ 43,800
ε420nm ~ 5950
[150]
Polymers 15 00342 i141
ETPO
λmax ~ 366
εmax ~ 13,830
ε395nm ~ 4530
ε405nm ~ 2270
[151]
Polymers 15 00342 i142
ALPO
λmax ~ 362
εmax ~ 12,200
ε395nm ~ 2890
ε405nm ~ 1300
[151]

4.3. Other Type I photoinitiators

An acylstannane-based PI tetrakis(2,4,6-trimethylbenzoyl)stannane (see Table 10) was designed for visible light photopolymerization [152]. Acylstannanes could show Type I PIs features like those in acylgermanes [153,154,155]. The absorption band of this acylstannane can extend to 550 nm. In steady-state photolysis, fast photobleaching was observed for the acylstannane (1 × 10−3 mol/L in acetonitrile) with a LED at 460 nm (1 W/cm2) after 3 min. It was attributed to the cleavage of the Sn-CO bond. The photoinitiation ability was studied under LED at 460 and LED at 522 nm. Upon exposure to a LED at 460 nm, tetrakis(2,4,6-trimethylbenzoyl)stannane demonstrated a similar performance with Ivocerin®, and high acrylate conversion was found. In addition, the acylstannane still showed high reactivity under LED at 522 nm. Interesting, favorable curing depths were obtained in the presence of this acylstannane-based PI. In prospect, it can be applied in many industrial fields.
Some N-hydroxynaphthalimide ester derivatives (namely NPIE1–NPIE9, see Table 10) are proposed as Type I PIs [156]. They demonstrated good light absorption properties, such as ε405nm (NPIE1) = 15,000 M−1 cm−1 and ε405nm (NPIE2) = 14,400 M−1 cm−1. Interestingly, these PIs demonstrated a favorable performance for the FRP of the monomers. NPIE1 (FC = 68%) demonstrated better performance than benchmark structure TPO (FC = 66%) in polymerization experiments of TMPTA. A decarboxylation process was discovered for these structures. The mechanisms N-hydroxynaphthalimide ester derivatives were investigated using a computational procedure, steady-state photolysis, and fluorescence approaches. Finally, the cleavage mechanism (N−O bond) of the Type Ⅰ PI was proposed.
Table 10. The chemical structures and light absorption properties of other Type I photoinitiators.
Table 10. The chemical structures and light absorption properties of other Type I photoinitiators.
StructureAbsorption Properties
(λ/nm, εmax/M−1 cm−1)
Refs.
Polymers 15 00342 i143
Ivocerin®
λmax ~ 408
εmax ~ 711
ε385nm ~ 505
[157]
Polymers 15 00342 i144
tetrakis(2,4,6-trimethylbenzoyl)stannane
[152]
Polymers 15 00342 i145
NPIE1
λmax ~ 397
εmax ~ 16,000
ε405nm ~ 15,000
[156]
Polymers 15 00342 i146
NPIE2
λmax ~ 398
εmax ~ 15,200
ε405nm ~ 14,400
[156]
Polymers 15 00342 i147
NPIE3
λmax ~ 398
εmax ~ 15,600
ε405nm ~ 14,900
[156]
Polymers 15 00342 i148
NPIE4
λmax ~ 397
εmax ~ 14,100
ε405nm ~ 13,100
[156]
Polymers 15 00342 i149
NPIE5
λmax ~ 397
εmax ~ 15,300
ε405nm ~ 14,500
[156]
Polymers 15 00342 i150
NPIE6
λmax ~ 397
εmax ~ 14,700
ε405nm ~ 13,800
[156]
Polymers 15 00342 i151
NPIE7
λmax ~ 397
εmax ~ 13,500
ε405nm ~ 12,800
[156]
Polymers 15 00342 i152
NPIE8
λmax ~ 397
εmax ~ 12,900
ε405nm ~ 12,000
[156]
Polymers 15 00342 i153
NPIE9
λmax ~ 397
εmax ~ 14,900
ε405nm ~ 14,000
[156]

5. Conclusions and Perspectives

In the past few years, photoinitiators for LED photopolymerization have received increasing attention. An increasing number of organic dyes have been used in photoinitiating systems due to their good light absorption properties. Many chemical structures demonstrate good performance in FRP and CP under LED irradiation. The modification for traditional Type I and Type II photoinitiators have also been carried out frequently to expand the range of their applications. For the development of photoinitiators, besides light absorption properties, synthesis procedures, toxicity, and cost should also be taken into account. All the presented PIs remain on the lab scale even if industrial development can be considered. In the future, more efforts should be devoted to addressing biocompatibility, which can provide significant opportunities for photoinitiators in the field of biomaterials and biomedicine. In addition, it is also desirable and crucial to resolving the yellow coloration induced by some photoinitiators during the photopolymerization. Some scaffolds with photo-bleaching characteristics are a potential option for the design of photoinitiators with a favorable curing depth.

Author Contributions

Conceptualization, all authors; validation, all authors; resources, J.L. and P.X.; writing—original draft, S.L.; writing—review and editing, all authors; supervision, J.L. and P.X.; project administration, J.L. and P.X.; funding acquisition, P.X. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research project is supported by the China Scholarship Council (CSC No. 201906880009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Corrigan, N.; Yeow, J.; Judzewitsch, P.; Xu, J.; Boyer, C. Seeing the Light: Advancing Materials Chemistry through Photopolymerization. Angew. Chem. Int. Ed. 2019, 58, 5170–5189. [Google Scholar] [CrossRef] [PubMed]
  2. Andrzejewska, E. Photopolymerization kinetics of multifunctional monomers. Prog. Polym. Sci. 2001, 26, 605–665. [Google Scholar] [CrossRef]
  3. Chen, M.; Zhong, M.; Johnson, J.A. Light-Controlled Radical Polymerization: Mechanisms, Methods, and Applications. Chem. Rev. 2016, 116, 10167–10211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Dumur, F. Recent advances on benzylidene ketones as photoinitiators of polymerization. Eur. Polym. J. 2022, 178, 111500. [Google Scholar] [CrossRef]
  5. Yilmaz, G.; Yagci, Y. Light-induced step-growth polymerization. Prog. Polym. Sci. 2020, 100, 101178. [Google Scholar] [CrossRef]
  6. Khudyakov, I.V. Fast photopolymerization of acrylate coatings: Achievements and problems. Prog. Org. Coat. 2018, 121, 151–159. [Google Scholar] [CrossRef]
  7. Sangermano, M.; Pegel, S.; Pötschke, P.; Voit, B. Antistatic Epoxy Coatings With Carbon Nanotubes Obtained by Cationic Photopolymerization. Macromol. Rapid Commun. 2008, 29, 396–400. [Google Scholar] [CrossRef]
  8. Gam-Derouich, S.; Carbonnier, B.; Turmine, M.; Lang, P.; Jouini, M.; Hassen-Chehimi, D.B.; Chehimi, M.M. Electrografted aryl diazonium initiators for surface-confined photopolymerization: a new approach to designing functional polymer coatings. Langmuir 2010, 26, 11830–11840. [Google Scholar] [CrossRef] [PubMed]
  9. Nunes, T.G.; Ceballos, L.; Osorio, R.; Toledano, M. Spatially resolved photopolymerization kinetics and oxygen inhibition in dental adhesives. Biomaterials 2005, 26, 1809–1817. [Google Scholar] [CrossRef]
  10. Daniloska, V.; Carretero, P.; Tomovska, R.; Asua, J.M. High performance pressure sensitive adhesives by miniemulsion photopolymerization in a continuous tubular reactor. Polymer 2014, 55, 5050–5056. [Google Scholar] [CrossRef]
  11. Besse, V.; Derbanne, M.A.; Pham, T.N.; Cook, W.D.; Le Pluart, L. Photopolymerization study and adhesive properties of self-etch adhesives containing bis(acyl)phosphine oxide initiator. Dent. Mater. 2016, 32, 561–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bagheri, A.; Jin, J. Photopolymerization in 3D Printing. ACS Appl. Polym. Mater. 2019, 1, 593–611. [Google Scholar] [CrossRef] [Green Version]
  13. Layani, M.; Wang, X.; Magdassi, S. Novel Materials for 3D Printing by Photopolymerization. Adv. Mater. 2018, 30, 1706344. [Google Scholar] [CrossRef] [PubMed]
  14. Aduba, D.C.; Margaretta, E.D.; Marnot, A.E.C.; Heifferon, K.V.; Surbey, W.R.; Chartrain, N.A.; Whittington, A.R.; Long, T.E.; Williams, C.B. Vat photopolymerization 3D printing of acid-cleavable PEG-methacrylate networks for biomaterial applications. Mater. Today Commun. 2019, 19, 204–211. [Google Scholar] [CrossRef]
  15. Zhao, X.; Zhao, Y.; Li, M.D.; Li, Z.; Peng, H.; Xie, T.; Xie, X. Efficient 3D printing via photooxidation of ketocoumarin based photopolymerization. Nat. Commun. 2021, 12, 2873. [Google Scholar] [CrossRef]
  16. Jandt, K.D.; Mills, R.W. A brief history of LED photopolymerization. Dent. Mater. 2013, 29, 605–617. [Google Scholar] [CrossRef]
  17. Weems, A.C.; Chiaie, K.R.D.; Yee, R.; Dove, A.P. Selective Reactivity of Myrcene for Vat Photopolymerization 3D Printing and Postfabrication Surface Modification. Biomacromolecules 2020, 21, 163–170. [Google Scholar] [CrossRef]
  18. Yang, H.; Li, G.; Stansbury, J.W.; Zhu, X.; Wang, X.; Nie, J. Smart Antibacterial Surface Made by Photopolymerization. ACS Appl. Mater. Interfaces 2016, 8, 28047–28054. [Google Scholar] [CrossRef] [PubMed]
  19. Gibson, I.; Rosen, D.; Stucker, B. Vat Photopolymerization Processes. Additive Manufacturing Technologies; Springer: New York, NY, USA, 2015; pp. 63–106. [Google Scholar]
  20. Vitale, A.; Priola, A.; Tonelli, C.; Bongiovanni, R. Nanoheterogeneous networks by photopolymerization of perfluoropolyethers and acrylic co-monomers. Polym. Int. 2013, 62, 1395–1401. [Google Scholar] [CrossRef]
  21. Crivello, J.V.; Reichmanis, E. Photopolymer Materials and Processes for Advanced Technologies. Chem. Mater. 2013, 26, 533–548. [Google Scholar] [CrossRef]
  22. Zhou, J.; Allonas, X.; Ibrahim, A.; Liu, X. Progress in the development of polymeric and multifunctional photoinitiators. Prog. Polym. Sci. 2019, 99, 101165. [Google Scholar] [CrossRef]
  23. Dadashi-Silab, S.; Doran, S.; Yagci, Y. Photoinduced Electron Transfer Reactions for Macromolecular Syntheses. Chem. Rev. 2016, 116, 10212–10275. [Google Scholar] [CrossRef] [PubMed]
  24. Christmann, J.; Ley, C.; Allonas, X.; Ibrahim, A.; Croutxé-Barghorn, C. Experimental and theoretical investigations of free radical photopolymerization: Inhibition and termination reactions. Polymer 2019, 160, 254–264. [Google Scholar] [CrossRef]
  25. Lin, J.T.; Cheng, D.C.; Chen, K.T.; Chiu, Y.C.; Liu, H.W. Enhancing UV Photopolymerization by a Red-light Preirradiation: Kinetics and Modeling Strategies for Reduced Oxygen Inhibition. J. Polym. Sci. 2020, 58, 683–691. [Google Scholar] [CrossRef]
  26. Courtecuisse, F.; Karasu, F.; Allonas, X.; Croutxé-Barghorn, C.; van der Ven, L. Confocal Raman microscopy study of several factors known to influence the oxygen inhibition of acrylate photopolymerization under LED. Prog. Org. Coat. 2016, 92, 1–7. [Google Scholar] [CrossRef]
  27. Ma, Q.; Song, J.; Zhang, X.; Jiang, Y.; Ji, L.; Liao, S. Metal-free atom transfer radical polymerization with ppm catalyst loading under sunlight. Nat. Commun. 2021, 12, 429. [Google Scholar] [CrossRef]
  28. Lecamp, L.; Lebaudy, P.; Youssef, B.; Bunel, C. Influence of UV radiation wavelength on conversion and temperature distribution profiles within dimethacrylate thick material during photopolymerization. Polymer 2001, 42, 8541–8547. [Google Scholar] [CrossRef]
  29. Li, Z.; Shen, W.; Liu, X.; Liu, R. Efficient unimolecular photoinitiators for simultaneous hybrid thiol–yne–epoxy photopolymerization under visible LED light irradiation. Polym. Chem. 2017, 8, 1579–1588. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Xu, Y.; Simon-Masseron, A.; Lalevee, J. Radical photoinitiation with LEDs and applications in the 3D printing of composites. Chem. Soc. Rev. 2021, 50, 3824–3841. [Google Scholar] [CrossRef] [PubMed]
  31. 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]
  32. Dietlin, C.; Schweizer, S.; Xiao, P.; Zhang, J.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; Lalevée, J. Photopolymerization upon LEDs: new photoinitiating systems and strategies. Polym. Chem. 2015, 6, 3895–3912. [Google Scholar] [CrossRef]
  33. Kocaarslan, A.; Kütahya, C.; Keil, D.; Yagci, Y.; Strehmel, B. Near-IR and UV-LED Sensitized Photopolymerization with Onium Salts Comprising Anions of Different Nucleophilicities. ChemPhotoChem 2019, 3, 1127–1132. [Google Scholar] [CrossRef] [Green Version]
  34. Zuo, X.; Morlet-Savary, F.; Schmitt, M.; Le Nouën, D.; Blanchard, N.; Goddard, J.-P.; Lalevée, J. Novel applications of fluorescent brighteners in aqueous visible-light photopolymerization: high performance water-based coating and LED-assisted hydrogel synthesis. Polym. Chem. 2018, 9, 3952–3958. [Google Scholar] [CrossRef]
  35. Schmitt, M. Method to analyse energy and intensity dependent photo-curing of acrylic esters in bulk. RSC Adv. 2015, 5, 67284–67298. [Google Scholar] [CrossRef] [Green Version]
  36. Crivello, J.V. A new visible light sensitive photoinitiator system for the cationic polymerization of epoxides. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 866–875. [Google Scholar] [CrossRef]
  37. Oprych, D.; Schmitz, C.; Ley, C.; Allonas, X.; Ermilov, E.; Erdmann, R.; Strehmel, B. Photophysics of Up-Conversion Nanoparticles: Radical Photopolymerization of Multifunctional Methacrylates Comprising Blue- and UV-Sensitive Photoinitiators. ChemPhotoChem 2019, 3, 1119–1126. [Google Scholar] [CrossRef]
  38. Al Mousawi, A.; Poriel, C.; Dumur, F.; Toufaily, J.; Hamieh, T.; Fouassier, J.P.; Lalevée, J. Zinc Tetraphenylporphyrin as High Performance Visible Light Photoinitiator of Cationic Photosensitive Resins for LED Projector 3D Printing Applications. Macromolecules 2017, 50, 746–753. [Google Scholar] [CrossRef]
  39. 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. Dye. Pigment. 2019, 165, 467–473. [Google Scholar] [CrossRef]
  40. Zhang, J.; Frigoli, M.; Dumur, F.; Xiao, P.; Ronchi, L.; Graff, B.; Morlet-Savary, F.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. Design of Novel Photoinitiators for Radical and Cationic Photopolymerizations under Near UV and Visible LEDs (385, 395, and 405 nm). Macromolecules 2014, 47, 2811–2819. [Google Scholar] [CrossRef]
  41. Yang, J.; Liao, W.; Xiong, Y.; Wang, X.; Li, Z.; Tang, H. A multifunctionalized macromolecular silicone-naphthalimide visible photoinitiator for free radical polymerization. Prog. Org. Coat. 2018, 115, 151–158. [Google Scholar] [CrossRef]
  42. 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]
  43. Shi, S.; Croutxé-Barghorn, C.; Allonas, X. Photoinitiating systems for cationic photopolymerization: Ongoing push toward long wavelengths and low light intensities. Prog. Polym. Sci. 2017, 65, 1–41. [Google Scholar] [CrossRef]
  44. Xue, T.; Huang, B.; Li, Y.; Li, X.; Nie, J.; Zhu, X. Enone dyes as visible photoinitiator in radical polymerization: The influence of peripheral N-alkylated (hetero)aromatic amine group. J. Photochem. Photobiol. A Chem. 2021, 419, 113449. [Google Scholar] [CrossRef]
  45. Li, F.; Song, Y.; Yao, M.; Nie, J.; He, Y. Design and properties of novel photothermal initiators for photoinduced thermal frontal polymerization. Polym. Chem. 2020, 11, 3980–3986. [Google Scholar] [CrossRef]
  46. Yao, M.; Liu, S.; Huang, C.; Nie, J.; He, Y. Significantly improve the photoinitiation ability of hydroxyalkyl-derived polymerizable α-hydroxyalkylacetophenone photoinitiators by blocking hyperconjugation. J. Photochem. Photobiol. A Chem. 2021, 419, 113451. [Google Scholar] [CrossRef]
  47. Tehfe, M.-A.; Dumur, F.; Graff, B.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Blue-to-Red Light Sensitive Push–Pull Structured Photoinitiators: Indanedione Derivatives for Radical and Cationic Photopolymerization Reactions. Macromolecules 2013, 46, 3332–3341. [Google Scholar] [CrossRef]
  48. Lalevée, J.; Blanchard, N.; Tehfe, M.-A.; Peter, M.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.P. Efficient dual radical/cationic photoinitiator under visible light: a new concept. Polym. Chem. 2011, 2, 1986–1991. [Google Scholar] [CrossRef]
  49. Breloy, L.; Brezová, V.; Blacha-Grzechnik, A.; Presset, M.; Yildirim, M.S.; Yilmaz, I.; Yagci, Y.; Versace, D.-L. Visible Light Anthraquinone Functional Phthalocyanine Photoinitiator for Free-Radical and Cationic Polymerizations. Macromolecules 2019, 53, 112–124. [Google Scholar] [CrossRef]
  50. 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]
  51. Xiao, P.; Zhang, J.; Dumur, F.; Tehfe, M.A.; Morlet-Savary, F.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Visible light sensitive photoinitiating systems: Recent progress in cationic and radical photopolymerization reactions under soft conditions. Prog. Polym. Sci. 2015, 41, 32–66. [Google Scholar] [CrossRef]
  52. Allushi, A.; Kutahya, C.; Aydogan, C.; Kreutzer, J.; Yilmaz, G.; Yagci, Y. Conventional Type II photoinitiators as activators for photoinduced metal-free atom transfer radical polymerization. Polym. Chem. 2017, 8, 1972–1977. [Google Scholar] [CrossRef]
  53. Schmitt, M.; Dietlin, C.; Lalevée, J. Towards Visible LED Illumination: ZnO-ZnS Nanocomposite Particles. ChemistrySelect 2020, 5, 985–987. [Google Scholar] [CrossRef]
  54. Schmitt, M.; Garra, P.; Lalevée, J. Bulk Polymerization Photo-Initiator ZnO: Increasing of the Benzoyl Formic Acid Concentration and LED Illumination. Macromol. Chem. Phys. 2018, 219, 1800208. [Google Scholar] [CrossRef]
  55. Tasdelen, M.A.; Lalevée, J.; Yagci, Y. Photoinduced free radical promoted cationic polymerization 40 years after its discovery. Polym. Chem. 2020, 11, 1111–1121. [Google Scholar] [CrossRef]
  56. Dumur, F. Recent advances on carbazole-based photoinitiators of polymerization. Eur. Polym. J. 2020, 125, 109503. [Google Scholar] [CrossRef]
  57. 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]
  58. Sun, K.; Liu, S.; Chen, H.; Morlet-Savary, F.; Graff, B.; Pigot, C.; Nechab, M.; Xiao, P.; Dumur, F.; Lalevée, J. N-ethyl carbazole-1-allylidene-based push-pull dyes as efficient light harvesting photoinitiators for sunlight induced polymerization. Eur. Polym. J. 2021, 147, 110331. [Google Scholar] [CrossRef]
  59. Al Mousawi, A.; Dumur, F.; Garra, P.; Toufaily, J.; Hamieh, T.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Carbazole Scaffold Based Photoinitiator/Photoredox Catalysts: Toward New High Performance Photoinitiating Systems and Application in LED Projector 3D Printing Resins. Macromolecules 2017, 50, 2747–2758. [Google Scholar] [CrossRef]
  60. Al Mousawi, A.; Lara, D.M.; Noirbent, G.; Dumur, F.; Toufaily, J.; Hamieh, T.; Bui, T.-T.; Goubard, F.; Graff, B.; Gigmes, D.; et al. Carbazole Derivatives with Thermally Activated Delayed Fluorescence Property as Photoinitiators/Photoredox Catalysts for LED 3D Printing Technology. Macromolecules 2017, 50, 4913–4926. [Google Scholar] [CrossRef]
  61. Li, Z.; Hu, P.; Zhu, J.; Gao, Y.; Xiong, X.; Liu, R. Conjugated Carbazole-Based Schiff Bases as Photoinitiators: From Facile Synthesis to Efficient Two-Photon Polymerization. J. Polym. Sci. Part A Polym. Chem. 2018, 56, 2692–2700. [Google Scholar] [CrossRef]
  62. 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]
  63. Abdallah, M.; Dumur, F.; Graff, B.; Hijazi, A.; Lalevée, J. High performance dyes based on triphenylamine, cinnamaldehyde and indane-1,3-dione derivatives for blue light induced polymerization for 3D printing and photocomposites. Dye. Pigment. 2020, 182, 108580. [Google Scholar] [CrossRef]
  64. Li, Y.-H.; Chen, Y.-C. Triphenylamine-hexaarylbiimidazole derivatives as hydrogen-acceptor photoinitiators for free radical photopolymerization under UV and LED light. Polym. Chem. 2020, 11, 1504–1513. [Google Scholar] [CrossRef]
  65. 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]
  66. 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]
  67. 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]
  68. 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]
  69. Yang, J.; Xu, C.; Xiong, Y.; Wang, X.; Xie, Y.; Li, Z.; Tang, H. A Green and Highly Efficient Naphthalimide Visible Photoinitiator with an Ability Initiating Free Radical Polymerization under Air. Macromol. Chem. Phys. 2018, 219, 1800256. [Google Scholar] [CrossRef]
  70. Kanji, S.; Masamitsu, S. Photobase generators: Recent progress and application trend in polymer systems. Prog. Polym. Sci. 2009, 34, 194–209. [Google Scholar]
  71. Zhang, J.; Dumur, F.; Xiao, P.; Graff, B.; Bardelang, D.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Structure Design of Naphthalimide Derivatives: Toward Versatile Photoinitiators for Near-UV/Visible LEDs, 3D Printing, and Water-Soluble Photoinitiating Systems. Macromolecules 2015, 48, 2054–2063. [Google Scholar] [CrossRef]
  72. 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]
  73. 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]
  74. Suga, T.; Shimazu, S.; Ukaji, Y. Low-Valent Titanium-Mediated Radical Conjugate Addition Using Benzyl Alcohols as Benzyl Radical Sources. Org. Lett. 2018, 20, 5389–5392. [Google Scholar] [CrossRef] [PubMed]
  75. 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]
  76. Dumur, F. Recent advances on coumarin-based photoinitiators of polymerization. Eur. Polym. J. 2022, 163, 110962. [Google Scholar] [CrossRef]
  77. Abdallah, M.; Hijazi, A.; Graff, B.; Fouassier, J.-P.; Rodeghiero, G.; Gualandi, A.; Dumur, F.; Cozzi, P.G.; Lalevée, J. Coumarin derivatives as versatile photoinitiators for 3D printing, polymerization in water and photocomposite synthesis. Polym. Chem. 2019, 10, 872–884. [Google Scholar] [CrossRef]
  78. Topa, M.; Hola, E.; Galek, M.; Petko, F.; Pilch, M.; Popielarz, R.; Morlet-Savary, F.; Graff, B.; Lalevée, J.; Ortyl, J. One-component cationic photoinitiators based on coumarin scaffold iodonium salts as highly sensitive photoacid generators for 3D printing IPN photopolymers under visible LED sources. Polym. Chem. 2020, 11, 5261–5278. [Google Scholar] [CrossRef]
  79. Giacoletto, N.; Dumur, F. Recent Advances in bis-Chalcone-Based Photoinitiators of Polymerization: From Mechanistic Investigations to Applications. Molecules 2021, 26, 3192. [Google Scholar] [CrossRef]
  80. Sharma, V.S.; Sharma, A.S.; Agarwal, N.K.; Shah, P.A.; Shrivastav, P.S. Self-assembled blue-light emitting materials for their liquid crystalline and OLED applications: from a simple molecular design to supramolecular materials. Mol. Syst. Des. Eng. 2020, 5, 1691–1705. [Google Scholar] [CrossRef]
  81. Zhang, Y.-P.; Wang, B.-X.; Yang, Y.-S.; Liang, C.; Yang, C.; Chai, H.-L. Synthesis and self-assembly of chalcone-based organogels. Colloids Surf. A Physicochem. Eng. Asp. 2019, 577, 449–455. [Google Scholar] [CrossRef]
  82. 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]
  83. Ibrahim-Ouali, M.; Dumur, F. Recent advances on chalcone-based photoinitiators of polymerization. Eur. Polym. J. 2021, 158, 110688. [Google Scholar] [CrossRef]
  84. Liu, S.; Zhang, Y.; Sun, K.; Graff, B.; Xiao, P.; Dumur, F.; Lalevée, J. Design of photoinitiating systems based on the chalcone-anthracene scaffold for LED cationic photopolymerization and application in 3D printing. Eur. Polym. J. 2021, 147, 110300. [Google Scholar] [CrossRef]
  85. Chen, H.; Noirbent, G.; Zhang, Y.; Brunel, D.; Gigmes, D.; Morlet-Savary, F.; Graff, B.; Xiao, P.; Dumur, F.; Lalevée, J. Novel D–π-A and A–π-D–π-A three-component photoinitiating systems based on carbazole/triphenylamino based chalcones and application in 3D and 4D printing. Polym. Chem. 2020, 11, 6512–6528. [Google Scholar] [CrossRef]
  86. 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]
  87. 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]
  88. 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. Dye. Pigment. 2021, 188, 109213. [Google Scholar] [CrossRef]
  89. Arsu, N.; Balta, D.K.; Yagci, Y.; Jockusch, S.; Turro, N.J. Mechanistic Study of Photoinitiated Free Radical Polymerization Using Thioxanthone Thioacetic Acid as One-Component Type II Photoinitiator. Macromolecules 2005, 38, 4133–4138. [Google Scholar]
  90. Kork, S.; Yilmaz, G.; Yagci, Y. Poly(vinyl alcohol)-Thioxanthone as One-Component Type II Photoinitiator for Free Radical Polymerization in Organic and Aqueous Media. Macromol. Rapid Commun. 2015, 36, 923–928. [Google Scholar] [CrossRef]
  91. Arsu, N.; Cokbaglan, L.; Yagci, Y.; Jockusch, S.; Turro, N.J. 2-Mercaptothioxanthone as a Novel Photoinitiator for Free Radical Polymerization. Macromolecules 2003, 36, 2649–2653. [Google Scholar]
  92. Griesser, M.; Rosspeintner, A.; Dworak, C.; Hofer, M.; Grabner, G.; Liska, R.; Gescheidt, G. Initiators Based on Benzaldoximes: Bimolecular and Covalently Bound Systems. Macromolecules 2012, 45, 8648–8657. [Google Scholar] [CrossRef] [PubMed]
  93. Allen, N.S.; Lam, E.; Howells, E.M.; Green, P.N.; Green, A.; Catalina, F.; Peinado, C. Photochemistry and photopolymerization activity of novel 4-alkylamino benzophenone initiators-synthesis, characterization, spectroscopic and photopolymerization activity. Eur. Polym. J. 1990, 26, 1345–1353. [Google Scholar] [CrossRef]
  94. Temel, G.; Enginol, B.; Aydin, M.; Balta, D.K.; Arsu, N. Photopolymerization and photophysical properties of amine linked benzophenone photoinitiator for free radical polymerization. J. Photochem. Photobiol. A Chem. 2011, 219, 26–31. [Google Scholar] [CrossRef]
  95. Jauk, S.; Liska, R. Photoinitiators with Functional Groups, 8. Macromol. Rapid Commun. 2005, 26, 1687–1692. [Google Scholar] [CrossRef]
  96. Jauk, S.; Liska, R. Photoinitiators with Functional Groups 9: New Derivatives of Covalently Linked Benzophenone-amine Based Photoinitiators. J. Macromol. Sci. Part A 2008, 45, 804–810. [Google Scholar] [CrossRef]
  97. Pietrzak, M.; Wrzyszczyński, A. Novel sulfur-containing benzophenone derivative as radical photoinitiator for photopolymerization. J. Appl. Polym. Sci. 2011, 122, 2604–2608. [Google Scholar] [CrossRef]
  98. Xiao, P.; Lalevée, J.; Allonas, X.; Fouassier, J.P.; Ley, C.; El-Roz, M.; Shi, S.Q.; Nie, J. Photoinitiation mechanism of free radical photopolymerization in the presence of cyclic acetals and related compounds. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 5758–5766. [Google Scholar] [CrossRef]
  99. Wang, Y.; Jiang, X.; Yin, J. Novel polymeric photoinitiators comprising of side-chain benzophenone and coinitiator amine: Photochemical and photopolymerization behaviors. Eur. Polym. J. 2009, 45, 437–447. [Google Scholar] [CrossRef]
  100. Temel, G.; Esen, D.S.; Arsu, N. One-component benzoxazine type photoinitiator for free radical polymerization. Polym. Eng. Sci. 2012, 52, 133–138. [Google Scholar] [CrossRef]
  101. Yang, J.; Shi, S.; Xu, F.; Nie, J. Synthesis and photopolymerization kinetics of benzophenone sesamol one-component photoinitiator. Photochem. Photobiol. Sci. 2013, 12, 323–329. [Google Scholar] [CrossRef]
  102. Balta, D.K.; Karahan, Ö.; Avci, D.; Arsu, N. Synthesis, photophysical and photochemical studies of benzophenone based novel monomeric and polymeric photoinitiators. Prog. Org. Coat. 2015, 78, 200–207. [Google Scholar] [CrossRef]
  103. Tehfe, M.-A.; Dumur, F.; Graff, B.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Design of new Type I and Type II photoinitiators possessing highly coupled pyrene–ketone moieties. Polym. Chem. 2013, 4, 2313–2324. [Google Scholar] [CrossRef]
  104. Telitel, S.; Dumur, F.; Gigmes, D.; Graff, B.; Fouassier, J.P.; Lalevée, J. New functionalized aromatic ketones as photoinitiating systems for near visible and visible light induced polymerizations. Polymer 2013, 54, 2857–2864. [Google Scholar] [CrossRef]
  105. Lalevee, J.; Tehfe, M.A.; Dumur, F.; Gigmes, D.; Graff, B.; Morlet-Savary, F.; Fouassier, J.P. Light-harvesting organic photoinitiators of polymerization. Macromol. Rapid Commun. 2013, 34, 239–245. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, J.; Zivic, N.; Dumur, F.; Xiao, P.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. A benzophenone-naphthalimide derivative as versatile photoinitiator of polymerization under near UV and visible lights. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 445–451. [Google Scholar] [CrossRef]
  107. Huang, T.L.; Li, Y.-H.; Chen, Y.-C. Benzophenone derivatives as novel organosoluble visible light Type II photoinitiators for UV and LED photoinitiating systems. J. Polym. Sci. 2020, 58, 2914–2925. [Google Scholar] [CrossRef]
  108. Jia, X.; Zhao, D.; You, J.; Hao, T.; Li, X.; Nie, J.; Wang, T. Acetylene bridged D-(π-A)2 type dyes containing benzophenone moieties: Photophysical properties, and the potential application as photoinitiators. Dye. Pigment. 2021, 184, 108583. [Google Scholar] [CrossRef]
  109. Xue, T.; Li, Y.; Zhao, X.; Nie, J.; Zhu, X. A facile synthesized benzophenone Schiff-base ligand as efficient type II visible light photoinitiator. Prog. Org. Coat. 2021, 157, 106329. [Google Scholar] [CrossRef]
  110. 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]
  111. 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]
  112. Liu, S.; Brunel, D.; Noirbent, G.; Mau, A.; Chen, H.; Morlet-Savary, F.; Graff, B.; Gigmes, D.; Xiao, P.; Dumur, F.; et al. New multifunctional benzophenone-based photoinitiators with high migration stability and their applications in 3D printing. Mater. Chem. Front. 2021, 5, 1982–1994. [Google Scholar] [CrossRef]
  113. 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]
  114. 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]
  115. 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]
  116. Esen, D.S.; Temel, G.; Balta, D.K.; Allonas, X.; Arsu, N. One-component thioxanthone acetic acid derivative photoinitiator for free radical polymerization. Photochem. Photobiol. 2014, 90, 463–469. [Google Scholar] [CrossRef] [PubMed]
  117. Yilmaz, G.; Tuzun, A.; Yagci, Y. Thioxanthone-carbazole as a visible light photoinitiator for free radical polymerization. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 5120–5125. [Google Scholar] [CrossRef]
  118. Yilmaz, G.; Beyazit, S.; Yagci, Y. Visible light induced free radical promoted cationic polymerization using thioxanthone derivatives. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 1591–1596. [Google Scholar] [CrossRef]
  119. Karaca, N.; Balta, D.K.; Ocal, N.; Arsu, N. Mechanistic studies of thioxanthone–carbazole as a one-component type II photoinitiator. J. Lumin. 2014, 146, 424–429. [Google Scholar] [CrossRef]
  120. Arsu, N.; Balta, D.K.; 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]
  121. 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]
  122. Balta, D.K.; Temel, G.; Goksu, G.; Ocal, N.; Arsu, N. Thioxanthone–Diphenyl Anthracene: Visible Light Photoinitiator. Macromolecules 2011, 45, 119–125. [Google Scholar] [CrossRef]
  123. Balta, D.K.; Arsu, N. Thioxanthone-ethyl anthracene. J. Photochem. Photobiol. A Chem. 2013, 257, 54–59. [Google Scholar] [CrossRef]
  124. Tar, H.; Esen, D.S.; 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]
  125. Yilmaz, G.; Acik, G.; Yagci, Y. Counteranion Sensitization Approach to Photoinitiated Free Radical Polymerization. Macromolecules 2012, 45, 2219–2224. [Google Scholar] [CrossRef]
  126. Dogruyol, S.K.; Dogruyol, Z.; Arsu, N. A thioxanthone-based visible photoinitiator. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 4037–4043. [Google Scholar] [CrossRef]
  127. Wu, Q.; Xiong, Y.; Liang, Q.; Tang, H. Developing thioxanthone based visible photoinitiators for radical polymerization. RSC Adv. 2014, 4, 52324–52331. [Google Scholar] [CrossRef]
  128. Doğruyol, S.K.; Doğruyol, Z.; Arsu, N. Thioxanthone based 9-[2-(methyl-phenyl-amino)-acetyl]-thia-naphthacene-12-one as a visible photoinitiator. J. Lumin. 2013, 138, 98–104. [Google Scholar] [CrossRef]
  129. Mau, A.; Le, T.H.; Dietlin, C.; Bui, T.-T.; Graff, B.; Dumur, F.; Goubard, F.; Lalevee, J. Donor–acceptor–donor structured thioxanthone derivatives as visible photoinitiators. Polym. Chem. 2020, 11, 7221–7234. [Google Scholar] [CrossRef]
  130. Breloy, L.; Losantos, R.; Sampedro, D.; Marazzi, M.; Malval, J.-P.; Heo, Y.; Akimoto, J.; Ito, Y.; Brezová, V.; Versace, D.-L. Allyl amino-thioxanthone derivatives as highly efficient visible light H-donors and co-polymerizable photoinitiators. Polym. Chem. 2020, 11, 4297–4312. [Google Scholar] [CrossRef]
  131. Rodrigues, M.R.; Neumann, M.G. Mechanistic Study of Tetrahydrofuran Polymerization Photoinitiated by a Sulfonium Salt/Thioxanthone System. Macromol. Chem. Phys. 2001, 202, 2776–2782. [Google Scholar] [CrossRef]
  132. Guo, X.D.; Zhou, H.Y.; Wang, J.X. A novel thioxanthone-hydroxyalkylphenone bifunctional photoinitiator: Synthesis, characterization and mechanism of photopolymerization. Prog. Org. Coat. 2021, 154, 106214. [Google Scholar] [CrossRef]
  133. Dworak, C.; Liska, R.J. Alternative initiators for bimolecular photoinitiating systems. Polym. Sci. Part A Polym. Chem. 2010, 48, 5865–5871. [Google Scholar] [CrossRef]
  134. Dietliker, K.; Hüsler, R.; Birbaum, J.L.; Ilg, S.; Villeneuve, S.; Studer, K.; Jung, T.; Benkhoff, J.; Kura, H.; Matsumoto, A.; et al. Advancements in photoinitiators—Opening up new applications for radiation curing. Prog. Org. Coat. 2007, 58, 146–157. [Google Scholar] [CrossRef]
  135. Fast, D.E.; Lauer, A.; Menzel, J.P.; Kelterer, A.-M.; Gescheidt, G.; Barner-Kowollik, C. Wavelength-Dependent Photochemistry of Oxime Ester Photoinitiators. Macromolecules 2017, 50, 1815–1823. [Google Scholar] [CrossRef]
  136. 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] [PubMed]
  137. 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] [PubMed]
  138. Qiu, W.; Zhu, J.; Dietliker, K.; Li, Z. Polymerizable Oxime Esters: An Efficient Photoinitiator with Low Migration Ability for 3D Printing to Fabricate Luminescent Devices. ChemPhotoChem 2020, 4, 5296–5303. [Google Scholar] [CrossRef]
  139. Qiu, W.; Hu, P.; Zhu, J.; Liu, R.; Li, Z.; Hu, Z.; Chen, Q.; Dietliker, K.; Liska, R. Cleavable Unimolecular Photoinitiators Based on Oxime-Ester Chemistry for Two-Photon Three-Dimensional Printing. ChemPhotoChem 2019, 3, 1090–1094. [Google Scholar] [CrossRef]
  140. Hammoud, F.; Giacoletto, N.; Noirbent, G.; Graff, B.; Hijazi, A.; Nechab, M.; Gigmes, D.; Dumur, F.; Lalevée, J. Substituent effects on the photoinitiation ability of coumarin-based oxime-ester photoinitiators for free radical photopolymerization. Mater. Chem. Front. 2021, 5, 8361–8370. [Google Scholar] [CrossRef]
  141. Ma, X.; Cao, D.; Fu, H.; You, J.; Gu, R.; Fan, B.; Nie, J.; Wang, T. Multicomponent photoinitiating systems containing arylamino oxime ester for visible light photopolymerization. Prog. Org. Coat. 2019, 135, 517–524. [Google Scholar] [CrossRef]
  142. Ding, Y.; Jiang, S.; Gao, Y.; Nie, J.; Du, H.; Sun, F. Photochromic Polymers Based on Fluorophenyl Oxime Ester Photoinitiators as Photoswitchable Molecules. Macromolecules 2020, 53, 5701–5710. [Google Scholar] [CrossRef]
  143. Zhou, R.; Sun, X.; Mhanna, R.; Malval, J.-P.; Jin, M.; Pan, H.; Wan, D.; Morlet-Savary, F.; Chaumeil, H.; Joyeux, C. Wavelength-Dependent, Large-Amplitude Photoinitiating Reactivity within a Carbazole-Coumarin Fused Oxime Esters Series. ACS Appl. Polym. Mater. 2020, 2, 2077–2085. [Google Scholar] [CrossRef]
  144. Chen, S.; Jin, M.; Malval, J.-P.; Fu, J.; Morlet-Savary, F.; Pan, H.; Wan, D. Substituted stilbene-based oxime esters used as highly reactive wavelength-dependent photoinitiators for LED photopolymerization. Polym. Chem. 2019, 10, 6609–6621. [Google Scholar] [CrossRef]
  145. Wang, W.; Jin, M.; Pan, H.; Wan, D. Remote effect of substituents on the properties of phenyl thienyl thioether-based oxime esters as LED-sensitive photoinitiators. Dye. Pigment. 2021, 192, 109435. [Google Scholar] [CrossRef]
  146. Zhou, R.; Pan, H.; Wan, D.; Malval, J.-P.; Jin, M. Bicarbazole-based oxime esters as novel efficient photoinitiators for photopolymerization under UV-Vis LEDs. Prog. Org. Coat. 2021, 157, 106306. [Google Scholar] [CrossRef]
  147. Liu, S.; Graff, B.; Xiao, P.; Dumur, F.; Lalevee, J. Nitro-Carbazole Based Oxime Esters as Dual Photo/Thermal Initiators for 3D Printing and Composite Preparation. Macromol. Rapid Commun. 2021, 42, 2100207. [Google Scholar] [CrossRef]
  148. Liu, S.; Giacoletto, N.; Schmitt, M.; Nechab, M.; Graff, B.; Morlet-Savary, F.; Xiao, P.; Dumur, F.; Lalevée, J. Effect of Decarboxylation on the Photoinitiation Behavior of Nitrocarbazole-Based Oxime Esters. Macromolecules 2022, 55, 2475–2485. [Google Scholar] [CrossRef]
  149. Dietlin, C.; Trinh, T.T.; Schweizer, S.; Graff, B.; Morlet-Savary, F.; Noirot, P.-A.; Lalevée, J. Rational Design of Acyldiphenylphosphine Oxides as Photoinitiators of Radical Polymerization. Macromolecules 2019, 52, 7886–7893. [Google Scholar] [CrossRef]
  150. Xie, C.; Wang, Z.; Liu, Y.; Song, L.; Liu, L.; Wang, Z.; Yu, Q. A novel acyl phosphine compound as difunctional photoinitiator for free radical polymerization. Prog. Org. Coat. 2019, 135, 34–40. [Google Scholar] [CrossRef]
  151. Wu, Y.; Li, R.; Wang, J.; Situ, Y.; Huang, H. A new carbazolyl-basedacylphosphine oxide photoinitiator with high performance and low migration. J. Polym. Sci. 2022, 60, 52–61. [Google Scholar] [CrossRef]
  152. Mitterbauer, M.; Knaack, P.; Naumov, S.; Markovic, M.; Ovsianikov, A.; Moszner, N.; Liska, R. Acylstannanes: Cleavable and Highly Reactive Photoinitiators for Radical Photopolymerization at Wavelengths above 500 nm with Excellent Photobleaching Behavior. Angew. Chem. Int. Ed. Engl. 2018, 57, 12146–12150. [Google Scholar] [CrossRef]
  153. Radebner, J.; Eibel, A.; Leypold, M.; Gorsche, C.; Schuh, L.; Fischer, R.; Torvisco, A.; Neshchadin, D.; Geier, R.; Moszner, N.; et al. Tetraacylgermanes: Highly Efficient Photoinitiators for Visible-Light-Induced Free-Radical Polymerization. Angew. Chem. Int. Ed. Engl. 2017, 56, 3103–3107. [Google Scholar] [CrossRef] [PubMed]
  154. Neshchadin, D.; Rosspeintner, A.; Griesser, M.; Lang, B.; Mosquera-Vazquez, S.; Vauthey, E.; Gorelik, V.; Liska, R.; Hametner, C.; Ganster, B.; et al. Acylgermanes: photoinitiators and sources for Ge-centered radicals. insights into their reactivity. J. Am. Chem. Soc. 2013, 135, 17314–17321. [Google Scholar] [CrossRef] [PubMed]
  155. Ganster, B.; Fischer, U.K.; Moszner, N.; Liska, R. New Photocleavable Structures. Diacylgermane-Based Photoinitiators for Visible Light Curing. Macromolecules 2008, 41, 2394–2400. [Google Scholar] [CrossRef]
  156. 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]
  157. Haslinger, C.; Leutgeb, L.P.; Haas, M.; Baudis, S.; Liska, R. Synthesis and Photochemical Investigation of Tetraacylgermanes. ChemPhotoChem 2022, 6, e202200108. [Google Scholar] [CrossRef]
Scheme 1. Photochemical mechanism of the dye/Iod/EDB system.
Scheme 1. Photochemical mechanism of the dye/Iod/EDB system.
Polymers 15 00342 sch001
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

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. https://doi.org/10.3390/polym15020342

AMA Style

Liu S, Borjigin T, Schmitt M, Morlet-Savary F, Xiao P, Lalevée J. High-Performance Photoinitiating Systems for LED-Induced Photopolymerization. Polymers. 2023; 15(2):342. https://doi.org/10.3390/polym15020342

Chicago/Turabian Style

Liu, Shaohui, Timur Borjigin, Michael Schmitt, Fabrice Morlet-Savary, Pu Xiao, and Jacques Lalevée. 2023. "High-Performance Photoinitiating Systems for LED-Induced Photopolymerization" Polymers 15, no. 2: 342. https://doi.org/10.3390/polym15020342

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