• Nie Znaleziono Wyników

Development of new high-performance biphenyl and terphenyl derivatives as versatile photoredox photoinitiating systems and their applications in 3D printing photopolymerization processes

N/A
N/A
Protected

Academic year: 2022

Share "Development of new high-performance biphenyl and terphenyl derivatives as versatile photoredox photoinitiating systems and their applications in 3D printing photopolymerization processes"

Copied!
27
0
0

Pełen tekst

(1)

catalysts

Article

Development of New High-Performance Biphenyl and Terphenyl Derivatives as Versatile Photoredox Photoinitiating Systems and Their Applications in 3D Printing Photopolymerization Processes

Wiktoria Tomal1 , Maciej Pilch1, Anna Chachaj-Brekiesz2 and Joanna Ortyl1,3,*

1 Cracow University of Technology, Faculty of Chemical Engineering and Technology, Warszawska 24, 31–155 Cracow, Poland

2 Jagiellonian University, Faculty of Chemistry, Gronostajowa 2, 30–387 Cracow, Poland

3 Photo HiTech Ltd., Bobrzy ´nskiego 14, 30–348 Cracow, Poland

* Correspondence: jortyl@chemia.pk.edu.pl

Received: 2 September 2019; Accepted: 28 September 2019; Published: 1 October 2019 

Abstract:Novel 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were proposed as photosensitizes of iodonium salt for a highly effective bimolecular photoinitiating system upon soft irradiation conditions under long-wave ultraviolet (UV-A) and visible light.

Remarkably, these structures are highly versatile, allowing access to photoinitiating systems for the free-radical polymerization of acrylates, the cationic photopolymerization of epoxides, glycidyl, and vinyl ethers, the synthesis of interpenetrated polymer networks (IPNs) and the thiol-ene photopolymerization processes. Excellent polymerization profiles for all of the monomers, along with the high final conversions, were obtained. The initiation mechanisms of these bimolecular systems based on the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were investigated using the real-time FT-IR technique, steady-state photolysis, fluorescence experiments, theoretical calculations of molecular orbitals, and electrochemical analysis. Moreover, the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were investigated as a type II free-radical photoinitiator with amine. It was confirmed that the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives, in combination with different types of additives, e.g., amine as co-initiator or in the presence of onium salt, can act as a bimolecular photoinitiating system via the photo-reduction or photo-oxidation pathways, respectively.

Keywords: photo-reduction; photo-oxidation; cationic photopolymerization; free-radical photopolymerization; copolymerization; IPNs; 3D printing

1. Introduction

The field of photopolymerization has been one of the most widely and rapidly growing industries in recent years [1,2]. Thus, these materials have found widespread use in the coating industry [3,4], printing inks [5], pressure-sensitive adhesives, including structural adhesives [6,7], photolithography [8,9], medicine [10] for obtaining hydrogel polymeric materials [11,12], and dentistry in photocurable fillings [13,14]. Moreover, the light-induced processes play a key role in the microelectronics [15] and electronics industry by applying photocured conductive adhesives or by creating printed circuits [16].

Photopolymerization is also useful in the automotive industry [17], using light for photocurable coatings applied to selected elements of the car [18]. Another currently developing and interesting use of photoinduced processes of polymerization is the manufacture of (holographic) optical elements [19]

and microfluidic devices [20,21]. Photopolymerization processes have also succeeded in designing and

Catalysts 2019, 9, 827; doi:10.3390/catal9100827 www.mdpi.com/journal/catalysts

(2)

Catalysts 2019, 9, 827 2 of 27

forming three-dimensional (3D) models [22] in 3D printing stereolithography (SLA—stereolithography apparatus) [23], DLP (direct light processing) [24], and the rapidly developing CLIPTM (continuous liquid interface production) method [25]. All of the above-presented examples of applications for photopolymerization are only an indication of the wide spectrum of possibilities for the development of innovative photocurable polymeric materials created by the photochemistry of polymers.

The popularity of photopolymerization in so many areas is principally related to the several advantages of this process. First of all, the main advantage of this process is high speed, as it can be completed even within a few seconds. Additionally, it provides the possibility of conducting photopolymerization processes at ambient temperature, which permits the coating of heat-sensitive substrates such as wood and plastics. Furthermore, the use of solvent-free formulations makes the preparation of polymeric materials by means of the photoinduced polymerization process one of the most efficient photochemical technologies. For these reasons, in recent years, much attention has been paid to the design and manufacture of photoinitiating systems, which will be effective on many levels: They will demonstrate universality, i.e., the effectiveness of the initiation of polymerization occurring according to the free-radical and cationic photopolymerization mechanisms. From a practical point of view, the basic photocurable systems are (meth)acrylic resins, polymerizing according to the free-radical mechanism [26]. It is worth mentioning that, compared to cationic photopolymerization, free-radical initiation is more advanced mainly due to the availability of a wider range of monomers and photoinitiators. Due to the high reactivity of these monomers and the possibility of modification of the ester chain, the materials obtained as a result of the polymerization of these raw materials are characterized by various properties. Photoinitiators of the free-radical polymerization process are compounds that generate radicals to initiate a chemical reaction under the influence of light [27]. From a practical point of view, two types of photoinitiators for free-radical polymerization are used. These two types of initiators differ in their mechanism for the generation of reactive free radicals: Type I radical photoinitiators are unimolecular photoinitiators, while type II are mostly bimolecular radical photoinitiators. Type I photoinitiators undergo homolytic bond cleavage from the triplet state following light excitation, generating two free radicals that start the polymerization. Type II photoinitiators are a combination of two or more component photoinitiating systems with molecules, and form relatively long-lived excited triplet states capable of undergoing hydrogen-absorption or electron/proton transfer reactions with co-initiator molecules. Nevertheless, the main disadvantage of free-radical photopolymerization is oxygen inhibition, which reduces the efficiency of polymerization [28]. The second type of photopolymerization is cationic polymerization, which is a technique that may be used to polymerize important classes of monomers which cannot be polymerized by free-radical means, such as epoxides, vinyl ethers, propenyl ethers, siloxanes, oxetanes, cyclic acetals and formals, cyclic sulfides, lactones, and lactams [29]. These monomers include both unsaturated monomers that undergo chain polymerization through the carbon–carbon double bonds and cyclic monomers that undergo ring-opening polymerization. Nevertheless, cationic photopolymerization is applied mainly to epoxy monomers, vinyl ethers, and their derivatives in the industry [30,31]. In the photochemical industry, onium salts and mainly diaryliodonium and triarylsulfonium salts [32,33], in which protons are the particles initiating the process and active centers are carbocations or onion ions [34], are used as efficient photoinitiators for the cationic photopolymerization process. In contrast to free-radical photopolymerization, which experiences rapid termination of the polymerization process when the light is turned off (due to radical–radical termination reactions), the cationic polymerization processes proceeds long after irradiation has ceased, until nearly all of the monomer has been consumed [35].

Moreover, the indisputable advantage of cationic photopolymerization is its living character and resistance to the inhibitory effect of atmospheric oxygen [36]. These features of cationic polymerization indicate its attractive application prospects. Nevertheless, cationic photopolymerization still has some weaknesses. In contrast to free-radical photoinitiators, fewer cationic photoinitiators are commercially available.

(3)

Catalysts 2019, 9, 827 3 of 27

From an industrial point of view, an interesting variation of the photopolymerization process is a hybrid photochemical-initiated polymerization, which combines different polymerization mechanisms or different types of materials [37]. Essentially, there are two types of hybrid systems, one is a system of monomers (e.g., vinyl and acrylic) polymerizing according to different mechanisms, i.e., radical and cationic, which usually takes place using two different photoinitiators. As a consequence, this leads to so-called interpenetrating polymer networks (IPNs), which are formed as a result of the simultaneous polymerization of two multifunctional monomers [38,39]. The second type of hybrid system is a system of different polymeric monomers with one common mechanism (e.g., vinyl monomer chain and cycloaliphatic epoxy monomer catatonically polymerizing with ring opening) and, thus, use one type of photoinitiator [40]. Indeed, these different photopolymerization modes often use specific or adapted initiating systems. The need to add two different types of photoinitiators often causes problems with the choice of light source for the process of photopolymerization of the IPNs, because different photoinitiators have different characteristics of absorption, which affects the effectiveness of initiation. For example, some radical photoinitiators can initiate in visible light, while cationic photoinitiators are not adapted for this purpose. IPN photopolymerization carried out in this way causes problems at the stage of selecting a photoinitiating system; however, this type of process allows to obtain materials with completely innovative functional, mechanical, and strength properties, which are not available when conducting photopolymerization of one type of monomer. For this reason, the development of efficient versatile photoinitiating systems remains a great challenge.

Taking all of the above into consideration, there is a need to design new high-performance photoinitiating systems with absorption characteristics compatible with the emission characteristics of low power LED light sources to provide photoinitiation of free-radical and cationic polymerization simultaneously, ensuring the versatility of operation regardless of the monomers selected. In this paper, we describe bimolecular photoinitiating systems operating in the long-wave ultraviolet (UV-A) and visible light range, where 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were selected as the effective photosensitizes. In order to verify the applicability of the biphenyl derivatives, their spectroscopic characteristics were determined. More specifically, the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives are proposed as bimolecular photoredox initiators with iodonium salt to induce the formation of radicals and cations as reactive species for both the free-radical polymerization of acrylates and the cationic polymerization of epoxides and glycidyles under UV-A and visible light. The photochemical mechanisms were studied through steady-state photolysis, fluorescence experiments, theoretical calculations of molecular orbitals, and electrochemical analysis. This approach allowed the study of the structure/reactivity/efficiency relationships of the investigated biphenyl derivatives as photosensitizes of iodonium salt in cationic, free-radical, and IPN photopolymerization under low intensity LED light sources. Moreover, the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were investigated as a type II free-radical photoinitiator with amine. It was confirmed that the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives, in combination with different types of additives, e.g., amine as co-initiator or in the presence of onium salt, can act as a bimolecular photoinitiating system via the photo-reduction or photo-oxidation pathways, respectively. Additionally, the use of these new high-performance bimolecular photoinitiating systems in 3D printing was also investigated. Different types of compositions for 3D printing are proposed, including IPN systems and copolymers.

2. Results and Discussion

2.1. Spectroscopic Characteristics of Investigated Compounds

The UV−Vis spectra of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives were recorded in acetonitrile to determine the maxima of the absorption characteristics and the corresponding extinction coefficient at an absorption maximum wavelength (Table1and Figure1a,b). All investigated

(4)

Catalysts 2019, 9, 827 4 of 27

biphenyl derivatives are characterized by good molar extinction coefficients in the UV–Vis range, which reach values around 5000–15,000 [dm3·mol−1·cm−1] at an absorption maximum wavelength. Moreover, the absorption characteristic range of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives is 425 nm, ensuring a great compatibility with the emission spectra of long ultraviolet sources of light (e.g., with λem-maxat 365 nm UV-LED) and visible sources of light (e.g., with λem-maxat 405 nm Vis-LED) (Table1).

Catalysts 2019, 9, x FOR PEER REVIEW 4 of 27

1a,b). All investigated biphenyl derivatives are characterized by good molar extinction coefficients in the UV–Vis range, which reach values around 5000–15,000 [dm3·mol−1·cm−1] at an absorption maximum wavelength. Moreover, the absorption characteristic range of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives is 425 nm, ensuring a great compatibility with the emission spectra of long ultraviolet sources of light (e.g., with λem-max at 365 nm UV-LED) and visible sources of light (e.g., with λem-max at 405 nm Vis-LED) (Table 1).

(a) (b)

(c) (d)

Figure 1. Spectroscopic properties of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives. (a) UV-visible absorption spectra of series 1 in acetonitrile; (b) UV-visible absorption spectra of series 2 in acetonitrile; (c) fluorescence spectra of series 1 in acetonitrile in extinction 320 nm and integration time 1s; (d) fluorescence spectra of series 2 in acetonitrile in extinction 320 nm and integration time 1 s.

Therefore, their extinction coefficient is relatively high, in the 350–425 nm spectral range, ensuring a fairly good overlap with the emission spectra of the LED@365 nm and LED@405 nm used in this work. In order to comprehensively determine the spectroscopic characteristics, the fluorescence spectra of the investigated compounds in acetonitrile were also measured. As given in Figure 1c,d, all proposed compounds exhibit fluorescence in the visible spectrum of a range between 360–700 nm.

0 4000 8000 12000 16000 20000 24000

200 250 300 350 400 450

Molar extinction coefficient [dm3·mol-1·cm-1]

Wavelength [nm]

B1 B2 B3 B4 B5 B6 B7

0 4000 8000 12000 16000 20000 24000

200 250 300 350 400 450

Molar extinction coefficient [dm3·mol-1·cm-1]

Wavelength [nm]

B1A B2A B3A B4A B5A B6A B7A

0 5000 10000 15000 20000 25000 30000

350 425 500 575 650 725 800

Luminescence intensity [j.w.]

Wavelength [nm]

B1 B2 B3 B4 B5 B6 B7

0 5000 10000 15000 20000 25000 30000

360 410 460 510 560

Luminescence intensity [j.w.]

Wavelength [nm]

B1A B2A B3A B4A B5A B6A B7A

Figure 1. Spectroscopic properties of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives. (a) UV-visible absorption spectra of series 1 in acetonitrile; (b) UV-visible absorption spectra of series 2 in acetonitrile; (c) fluorescence spectra of series 1 in acetonitrile in extinction 320 nm and integration time 1s; (d) fluorescence spectra of series 2 in acetonitrile in extinction 320 nm and integration time 1 s.

Therefore, their extinction coefficient is relatively high, in the 350–425 nm spectral range, ensuring a fairly good overlap with the emission spectra of the LED@365 nm and LED@405 nm used in this work.

In order to comprehensively determine the spectroscopic characteristics, the fluorescence spectra of the investigated compounds in acetonitrile were also measured. As given in Figure1c,d, all proposed compounds exhibit fluorescence in the visible spectrum of a range between 360–700 nm.

(5)

Catalysts 2019, 9, 827 5 of 27

Table 1. Spectral characteristics of the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives studied in acetonitrile.

Acronym

λmax-ab [nm]

ε@λmax-ab

[dm3·mol−1·cm−1]

ε@λmax-ab

[dm3·mol−1·cm−1]

ε@λmax-ab

[dm3·mol−1·cm−1]

λmax-fluo in ACN [nm]

Imax-fluo in ACN [a.u.]

Stokes Shift [cm1]

Series1

B1 351 10,790 7000 0 417 9550 4475

B2 351 10,860 7320 210 415 18,040 4372

B3 351 11,350 7970 145 428 23,480 5095

B4 352 9490 6810 160 426 7610 4986

B5 352 5440 3890 110 424 5320 4840

B6 341 14,840 10,410 235 489 28,930 8845

B7 352 5380 3740 200 423 4930 4787

Series2

B1A 344 3200 3130 75 407 10,030 4492

B2A 350 6700 4040 210 405 14,150 3866

B3A 350 7120 4090 183 402 23,380 3645

B4A 353 6720 5730 870 435 6920 5335

B5A 353 4900 3920 210 429 5050 5024

B6A 351 9990 8010 650 413 11,370 4305

B7A 353 6360 4450 85 423 8630 4723

Table1presents a summary of the spectroscopic properties of the investigated 2-amino-4-methyl-6- phenyl-benzene-1,3-dicarbonitrile derivatives. On the basis of the obtained data, Stockes’s shift was calculated. The effects of substituents, as well as of the number of benzene rings on the spectroscopic characteristics of the investigated biphenyl derivatives, are noticeable.

2.2. Performance of the 2-Amino-4-Methyl-6-Phenyl-Benzene-1,3-Dicarbonitrile Derivatives as

Photosensitizers for Iodonium Salt in a Bimolecular Photoinitiating System for Cationic Photopolymerization The cationic photopolymerization process is particularly interesting and relatively widely used in many applications, as it has several essential practical advantages. The living character of cationic photopolymerization guarantees that the reaction continues to be effective even after the turn-off of the radiation source. Additionally, during this process, oxygen inhibition does not occur. For this reason, in this research, the cationic photopolymerization process of 3,4-epoxycyclohexylmethyl-3,4-epoxycyclohexane-carboxylate monomer (CADE) was carried out under air. The evaluation of the epoxy group content was continuously followed at about 790 cm−1, where the area of the band from the epoxy group decreased during the photopolymerization process.

Additionally, a new band ascribed to the formation of the polyether network at ~1080 cm−1appeared during cationic photopolymerization. Upon irradiation with the UV-A LED source under an emission maximum of 365 nm (with I0~3.77 mW/cm2), the ring opening photopolymerization of the epoxy monomer in the presence of two component initiating systems consisting of biphenyl derivatives and iodonium salt (Speedcure 938) (0.1%/1% w/w) was very efficient in terms of the final epoxy group conversion. The conversion time profiles for the cationic photopolymerization process of the CADE monomer under ultraviolet light conditions are given in Figures2a and3a. All investigated biphenyl derivatives can be used as highly effective photosensitizes for iodonium salts during cationic photopolymerization in UV-light conditions. Remarkably, higher values of epoxide monomer conversions were obtained for 2-amino-4-methyl-6-(4-phenylphenyl)-benzene-1,3-dicarbonitrile derivatives (series 1 of biphenyl compounds).

(6)

Catalysts 2019, 9, 827 6 of 27

Catalysts 2019, 9, x FOR PEER REVIEW 6 of 27

(a) (b)

Figure 2. (a) Cationic photopolymerization profiles (epoxy function conversion vs. irradiation time) initiated by a bimolecular photoinitiating system based on Speedcure 938 (1% wt.) and 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (0.1% wt.) for series 1 (a) under irradiation at 365 nm; (b) under irradiation at 405 nm.

Nevertheless, a different situation is observed while carrying out cationic ring opening polymerization, upon irradiation under a visible source of light with the emission maximum at 405 nm (with I0 ~12.07 mW/cm2). Under this irradiation condition, only several derivatives from series 1 exhibit good performance as visible photosensitizes in two component photoinitiating systems based on iodonium salt (Figure 2b). Derivatives of 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile (series 2 of biphenyls) besides the B1A compound, however, can be used as efficient visible photosensitizers (Figure 3b). This effect is dictated by their structure, which is proven by the absorbance characteristics of those compounds.

As shown in Figure 2,3, no conversion of the epoxy group was found when the sample had no biphenyl derivatives acting as photosensitizes of iodonium salt. As Speedcure 938 has a maximum absorbance at 242 nm [41,42], the irradiation wavelength was 365 nm or 405nm. In this case, Speedcure 938 could not decompose in order to form reactive species to induce the polymerization of epoxide groups. The exact values of epoxy group conversions under ultraviolet and visible light conditions are given in Table 2.

(a) (b)

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1+Speedcure 938 B2+Speedcure 938 B3+Speedcure 938 B4+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B2+Speedcure 938 B3+Speedcure 938 B6+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

Figure 2. (a) Cationic photopolymerization profiles (epoxy function conversion vs. irradiation time) initiated by a bimolecular photoinitiating system based on Speedcure 938 (1% wt.) and 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (0.1% wt.) for series 1 (a) under irradiation at 365 nm; (b) under irradiation at 405 nm.

Catalysts 2019, 9, x FOR PEER REVIEW 6 of 27

(a) (b)

Figure 2. (a) Cationic photopolymerization profiles (epoxy function conversion vs. irradiation time) initiated by a bimolecular photoinitiating system based on Speedcure 938 (1% wt.) and 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (0.1% wt.) for series 1 (a) under irradiation at 365 nm; (b) under irradiation at 405 nm.

Nevertheless, a different situation is observed while carrying out cationic ring opening polymerization, upon irradiation under a visible source of light with the emission maximum at 405 nm (with I0 ~12.07 mW/cm2). Under this irradiation condition, only several derivatives from series 1 exhibit good performance as visible photosensitizes in two component photoinitiating systems based on iodonium salt (Figure 2b). Derivatives of 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile (series 2 of biphenyls) besides the B1A compound, however, can be used as efficient visible photosensitizers (Figure 3b). This effect is dictated by their structure, which is proven by the absorbance characteristics of those compounds.

As shown in Figure 2,3, no conversion of the epoxy group was found when the sample had no biphenyl derivatives acting as photosensitizes of iodonium salt. As Speedcure 938 has a maximum absorbance at 242 nm [41,42], the irradiation wavelength was 365 nm or 405nm. In this case, Speedcure 938 could not decompose in order to form reactive species to induce the polymerization of epoxide groups. The exact values of epoxy group conversions under ultraviolet and visible light conditions are given in Table 2.

(a) (b)

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1+Speedcure 938 B2+Speedcure 938 B3+Speedcure 938 B4+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B2+Speedcure 938 B3+Speedcure 938 B6+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60 70 80 90

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

Figure 3.(a) Cationic photopolymerization profiles (epoxy function conversion vs. irradiation time) initiated by bimolecular photoinitiating system based on Speedcure 938 (1% wt.) and 2-amino-4-methyl- 6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (0.1% wt.) for series 2 (a) under irradiation at 365 nm; (b) under irradiation at 405 nm.

Nevertheless, a different situation is observed while carrying out cationic ring opening polymerization, upon irradiation under a visible source of light with the emission maximum at 405 nm (with I0~12.07 mW/cm2). Under this irradiation condition, only several derivatives from series 1 exhibit good performance as visible photosensitizes in two component photoinitiating systems based on iodonium salt (Figure2b). Derivatives of 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile (series 2 of biphenyls) besides the B1A compound, however, can be used as efficient visible photosensitizers (Figure3b). This effect is dictated by their structure, which is proven by the absorbance characteristics of those compounds. As shown in Figures2and3, no conversion of the epoxy group was found when

(7)

Catalysts 2019, 9, 827 7 of 27

the sample had no biphenyl derivatives acting as photosensitizes of iodonium salt. As Speedcure 938 has a maximum absorbance at 242 nm [41,42], the irradiation wavelength was 365 nm or 405nm. In this case, Speedcure 938 could not decompose in order to form reactive species to induce the polymerization of epoxide groups. The exact values of epoxy group conversions under ultraviolet and visible light conditions are given in Table2.

Bisphenol A diglycidyl ether (DGEBA) is an inexpensive and widely used monomer in polymer industry. The price of DGEBA is only a fraction of cycloaliphatic epoxide monomer, 3,4-epoxycyclohexylmethyl-3,4-epoxy-cyclohexane-carboxylate (CADE), widely used in the polymer coating industry. DGEBA is typically employed in condensation network polymerizations with amines [43], amides [44], and anhydrides [45] in order to obtain thermosetting epoxy resins. Bisphenol A diglycidyl ether is also polymerizable by cationic photopolymerization by electron beam sources [46].

This is due to the fact that the cationic photopolymerization of DGEBA is very sluggish [47,48].

Furthermore, the high viscosity of DGEBA epoxides makes their handling difficult during the process of preparation for coating. This is caused by poor mixing with other components, e.g., with photoinitiators or photoinitiating systems. The problem caused by the high viscosity of this resin based on Bisphenol A diglycidyl ether was solved by adding to the composition the low-viscosity 1,2-epoxy-3-phenoxypropane monomer (EPXPROP), which acts as a reactive diluent during the cationic photo-copolymerization process of these two types of glycidyle monomers. The compositions consisting of glycidyl ether monomers based on DGEBA and EPXPROP (70%/30% w/w) was investigated using real-time FT-IR. The cationic photopolymerization of this monomer mixture was carried out under air and in ambient room temperature (25C) upon irradiation with an ultraviolet UV-LED source of light with maximum emission at 365 nm and intensity at the surface of the sample of about ~7.97 mW/cm2. The evaluation of the glycidyl group content was continuously followed at about 915 cm−1, where the area of the band from the epoxy group decreased during the photopolymerization process. The cationic photopolymerization of glycidyl monomers in the presence of bimolecular photoinitiating systems consisting of biphenyl derivatives and iodonium salt (0.2%/2% w/w) was very efficient in terms of the final glycidyl group conversion (Figure4a,b). The values of conversion were in the range of 81–89% (Table2), which is the reason for obtaining a highly cross-linked polymer network.

Catalysts 2019, 9, x FOR PEER REVIEW 7 of 27

Figure 3. (a) Cationic photopolymerization profiles (epoxy function conversion vs. irradiation time) initiated by bimolecular photoinitiating system based on Speedcure 938 (1% wt.) and 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (0.1% wt.) for series 2 (a) under irradiation at 365 nm; (b) under irradiation at 405 nm.

Bisphenol A diglycidyl ether (DGEBA) is an inexpensive and widely used monomer in polymer industry. The price of DGEBA is only a fraction of cycloaliphatic epoxide monomer, 3,4-epoxycyclohexylmethyl-3,4-epoxy-cyclohexane-carboxylate (CADE), widely used in the polymer coating industry. DGEBA is typically employed in condensation network polymerizations with amines [43], amides [44], and anhydrides [45] in order to obtain thermosetting epoxy resins.

Bisphenol A diglycidyl ether is also polymerizable by cationic photopolymerization by electron beam sources [46]. This is due to the fact that the cationic photopolymerization of DGEBA is very sluggish [47,48]. Furthermore, the high viscosity of DGEBA epoxides makes their handling difficult during the process of preparation for coating. This is caused by poor mixing with other components, e.g., with photoinitiators or photoinitiating systems. The problem caused by the high viscosity of this resin based on Bisphenol A diglycidyl ether was solved by adding to the composition the low-viscosity 1,2-epoxy-3-phenoxypropane monomer (EPXPROP), which acts as a reactive diluent during the cationic photo-copolymerization process of these two types of glycidyle monomers. The compositions consisting of glycidyl ether monomers based on DGEBA and EPXPROP (70%/30%

w/w) was investigated using real-time FT-IR. The cationic photopolymerization of this monomer mixture was carried out under air and in ambient room temperature (25 °C) upon irradiation with an ultraviolet UV-LED source of light with maximum emission at 365 nm and intensity at the surface of the sample of about ~7.97 mW/cm2. The evaluation of the glycidyl group content was continuously followed at about 915 cm−1, where the area of the band from the epoxy group decreased during the photopolymerization process. The cationic photopolymerization of glycidyl monomers in the presence of bimolecular photoinitiating systems consisting of biphenyl derivatives and iodonium salt (0.2%/2% w/w) was very efficient in terms of the final glycidyl group conversion (Figure 4a,b). The values of conversion were in the range of 81–89% (Table 2), which is the reason for obtaining a highly cross-linked polymer network.

(a) (b)

Figure 4. Cationic photopolymerization profiles DEGDBA/EPXPROP (70/30% w/w) (glycidyl group function conversion vs. irradiation time) initiated by bimolecular photoinitiating system based on Speedcure 938 (2% wt.) and biphenyl derivatives (0.2% wt.) under light irradiation at 365 nm for (a) biphenyl derivatives from series 1, (b) biphenyl derivatives from series 2.

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1+Speedcure 938 B2+Speedcure 938 B3+Speedcure 938 B4+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX + Speedcure 938

0 10 20 30 40 50 60 70 80 90 100

0 100 200 300 400 500 600 700 800

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

Figure 4.Cationic photopolymerization profiles DEGDBA/EPXPROP (70/30% w/w) (glycidyl group function conversion vs. irradiation time) initiated by bimolecular photoinitiating system based on Speedcure 938 (2% wt.) and biphenyl derivatives (0.2% wt.) under light irradiation at 365 nm for (a) biphenyl derivatives from series 1, (b) biphenyl derivatives from series 2.

(8)

Catalysts 2019, 9, 827 8 of 27

Table 2.Summary of functional group conversions of used monomers for various photopolymerization processes, with the use of the 2-amino-4-methyl-6-phenyl-benzene -1,3-dicarbonitrile derivatives.

Conversion [%]

Cationic Photopolymerization Free-Radical Photopolymerization

Acronim

CADE with Iod at

~790 cm−1

Glycidyl with Iod at

~915 cm−1

TMPTA with Iod at

~1635 cm−1

TMPTA with EDB at

~1635 cm−1

@365 nma @405 nmb @365nmc @365nma @405nmb @365nma @405nmb

Series1

B1 64.9 NPd 83.4 52.0 NPd 58.6 NPd

B2 83.4 38.4 84.2 51.2 35.2 54.2 NPd

B3 81.6 67.3 80.7 53.4 47.7 53.3 55.8

B4 54.4 NPd 83.5 49.8 NPd 52.7 NPd

B5 32.2 NPd 83.5 38.5 33.3 NPd NPd

B6 86.5 71.2 81.0 53.4 47.4 58.6 50.2

B7 48.2 NPd 82.5 53.2 43.7 55.6 53.8

Series2

B1A 39.2 NPd 81.7 31.5 21.3 30.7 NPd

B2A 39.1 49.2 82.1 37.4 23.8 47.3 NPd

B3A 44.2 30.4 87.0 41.9 25.2 48 13.9

B4A 27.1 30.6 84.1 38,7 37.7 48 35.2

B5A 46.5 44.3 89.4 39.7 38.6 45.3 30.5

B6A 61.3 29.3 88.0 51.0 44.0 55.6 40.7

B7A 36.7 29.9 84.8 23.1 24.0 56.1 37.4

a LED @365 nm, ~3.77 mW/cm2; b LED @405 nm, ~12.07 mW/cm2; cLED @365 nm, ~7.97 mW/cm2; d NP, no polymerization.

2.3. Performance of 2-Amino-4-Methyl-6-Phenyl-Benzene-1,3-Dicarbonitrile as Photosensitizer for Iodonium Salt in a Bimolecular Photoinitiating System for Free-Radical Photopolymerization

Radical photopolymerization is one of the most widely used photopolymerizations based on the photochemical initiation of free radicals. The polymerization process under the influence of the applied light is mainly used for acrylate and methacrylate monomers. The reason for their popularity is the availability of different types of (meth)acrylate monomers on the market and their high reactivity, as they polymerize rapidly in the presence of free radicals. Acrylates, however, are not initiated by cationic species, because they have ester groups that act as electron withdrawing groups, decreasing the electron density on the double bond. This leads to a repulsive force between monomers and cationic species. Studies have proved, however, that the developed bimolecular photoinitiating system based on biphenyl derivatives and iodonium salt can act as a dual-type initiating system. In this approach, the dual type means that this type of system can efficiently initiate free-radical and cationic photopolymerization processes, respectively. As confirmation, the free-radical photopolymerization processes of trimethylolpropane triacrylate monomer (TMPTA) were carried out in laminate conditions, to prevent the inhibitory effect of oxygen on the generated radical species. The process was controlled by monitoring the decrease of the maximum area of a band at 1635 cm−1 wavelength, which was responsible for the double bond of the acrylate group content. Conversion time profiles for the two irradiation conditions under long wavelength ultraviolet (UV-A LED @365 nm with the emission intensity on the surface of the sample at about 3.77 mW/cm2) and visible light sources (Vis-LED @405 nm with the emission intensity on the surface of the sample at about 7.97 mW/cm2) are shown in Figures5and6. The final values of the double bond conversion are given in Table2.

(9)

Catalysts 2019, 9, 827 9 of 27

Catalysts 2019, 9, x FOR PEER REVIEW 9 of 27

(a) (b)

Figure 5. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm; (b) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm.

(a) (b)

Figure 6. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm; (b) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm.

From the data obtained during the measurements, it was shown that bimolecular photoinitiating systems, containing biphenyl derivatives and iodonium salt (0.1%/1% w/w), can be used in the free-radical photopolymerization of acrylates, reaching a maximum conversion value

~50%. In conclusion, the photosensitization of iodonium salt using

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1+Speedcure 938 B2+Speedcure 938 B3+Speedcure 938 B4+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B2+Speedcure 938 B3+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

Figure 5. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm; (b) series 1 of 2-amino-4-methyl-6-(4- phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm.

Catalysts 2019, 9, x FOR PEER REVIEW 9 of 27

(a) (b)

Figure 5. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm; (b) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 365 nm.

(a) (b)

Figure 6. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm; (b) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm.

From the data obtained during the measurements, it was shown that bimolecular photoinitiating systems, containing biphenyl derivatives and iodonium salt (0.1%/1% w/w), can be used in the free-radical photopolymerization of acrylates, reaching a maximum conversion value

~50%. In conclusion, the photosensitization of iodonium salt using

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1+Speedcure 938 B2+Speedcure 938 B3+Speedcure 938 B4+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B2+Speedcure 938 B3+Speedcure 938 B5+Speedcure 938 B6+Speedcure 938 B7+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1A+Speedcure 938 B2A+Speedcure 938 B3A+Speedcure 938 B4A+Speedcure 938 B5A+Speedcure 938 B6A+Speedcure 938 B7A+Speedcure 938 Speedcure 938 DETHX+Speedcure 938

Figure 6. Free-radical photopolymerization profiles (acrylate function conversion vs. irradiation time) for TMPTA with the bimolecular photoinitiating system based on (a) series 1 of 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm; (b) series 1 of 2-amino-4-methyl-6-(4- phenylphenyl)benzene-1,3-dicarbonitrile derivatives and Speedcure 938 (0.1%/1% w/w) under visible light LED source 405 nm.

From the data obtained during the measurements, it was shown that bimolecular photoinitiating systems, containing biphenyl derivatives and iodonium salt (0.1%/1% w/w), can be used in the free-radical photopolymerization of acrylates, reaching a maximum conversion value

(10)

Catalysts 2019, 9, 827 10 of 27

~50%. In conclusion, the photosensitization of iodonium salt using 2-amino-4-methyl-6-phenyl- benzene-1,3-dicarbonitrile derivatives in the free-radical polymerization of TMPTA is a highly efficient process at 365 nm and 405 nm. Moreover, the polymerization rate enhancement effect of this process can be easily increased by extending the power of the used LED light sources. It is worth mentioning that diphenyliodonium salt (e.g., Speedcure 938, Figures5and6) alone does not initiate free-radical photopolymerization, indicating the role of biphenyl derivatives as photosensitizers of iodonium salt through the mechanism of photo-oxidizable sensitization of this salt under long-wavelength ultraviolet and visible light regions. This efficient behavior during the free-radical photoinitiation process will be discussed in detail in the mechanistic section of this paper. The usefulness of the developed initiating systems in the radical polymerization process gives us the opportunity to use them as dual initiating systems, which can be utilized for the hybrid photopolymerization processes of different types of monomers. This application will be described in detail in a later part of this manuscript.

2.4. Performance of 2-Amino-4-Methyl-6-Phenyl-Benzene-1,3-Dicarbonitrile as Bimolecular Free-Radical Photoinitiator Type II

It was found that the investigated 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives exhibit the ability to initiate photoinduced free-radical polymerization processes according to the photo-reduction mechanism. This was confirmed by real-time FT-IR studies of the radical polymerization process of acrylate TMPTA using free-radical type II initiators based on biphenyl derivatives and amine (EDB). Type II free-radical photoinitiators generate radicals in the presence of co-initiators in a multi-step reaction mechanism. In this case, an amine ethyl 4-(dimethylamino)benzoate (EDB) was used as a co-initiator. Upon illumination, a photosensitizer in the excited state began to interact with a co-initiator that was an electron donor (EDB). In these bimolecular photoinitiating systems, biphenyl derivatives acted as electron acceptors, whereas the main radicals that initiated polymerization formed on those co-initiator molecules. The combination of biphenyl derivatives as photosensitizers and an amine as an electron donor component in this II type free-radical bimolecular initiator is an effective system. The highest values of conversion were obtained for B6/EDB (0.5/1.5%

w/w) from series 1 and for its analogue from series 2, B6A/EDB (0.5/1.5% w/w) (Figure7). Such dependence was confirmed by the value of the calculated Gibbs free energy for these two biphenyl compounds (Table2).

Catalysts 2019, 9, x FOR PEER REVIEW 10 of 27

2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives in the free-radical polymerization of TMPTA is a highly efficient process at 365 nm and 405 nm. Moreover, the polymerization rate enhancement effect of this process can be easily increased by extending the power of the used LED light sources. It is worth mentioning that diphenyliodonium salt (e.g., Speedcure 938, Figures 5 and 6) alone does not initiate free-radical photopolymerization, indicating the role of biphenyl derivatives as photosensitizers of iodonium salt through the mechanism of photo-oxidizable sensitization of this salt under long-wavelength ultraviolet and visible light regions. This efficient behavior during the free-radical photoinitiation process will be discussed in detail in the mechanistic section of this paper. The usefulness of the developed initiating systems in the radical polymerization process gives us the opportunity to use them as dual initiating systems, which can be utilized for the hybrid photopolymerization processes of different types of monomers.

This application will be described in detail in a later part of this manuscript.

2.4. Performance of 2-Amino-4-Methyl-6-Phenyl-Benzene-1,3-Dicarbonitrile as Bimolecular Free-Radical Photoinitiator Type II

It was found that the investigated 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives exhibit the ability to initiate photoinduced free-radical polymerization processes according to the photo-reduction mechanism. This was confirmed by real-time FT-IR studies of the radical polymerization process of acrylate TMPTA using free-radical type II initiators based on biphenyl derivatives and amine (EDB). Type II free-radical photoinitiators generate radicals in the presence of co-initiators in a multi-step reaction mechanism. In this case, an amine ethyl 4-(dimethylamino)benzoate (EDB) was used as a co-initiator. Upon illumination, a photosensitizer in the excited state began to interact with a co-initiator that was an electron donor (EDB). In these bimolecular photoinitiating systems, biphenyl derivatives acted as electron acceptors, whereas the main radicals that initiated polymerization formed on those co-initiator molecules. The combination of biphenyl derivatives as photosensitizers and an amine as an electron donor component in this II type free-radical bimolecular initiator is an effective system. The highest values of conversion were obtained for B6/EDB (0.5/1.5% w/w) from series 1 and for its analogue from series 2, B6A/EDB (0.5/1.5% w/w) (Figure 7). Such dependence was confirmed by the value of the calculated Gibbs free energy for these two biphenyl compounds (Table 2).

Figure 7. Photopolymerization profiles (double bond function conversion vs. irradiation time) in the presence of photoinitiating system based on biphenyls derivatives and EDB (0.5/1.5% w/w) at UV-A light LED source 365 nm for (a) 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (series 1) and (b) 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives (series 2).

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1+EDB B2+EDB B3+EDB B4+EDB B6+EDB B7+EDB EDB

0 10 20 30 40 50 60

0 100 200 300 400

Conversion [%]

Time [s]

B1A+EDB B2A+EDB B3A+EDB B4A+EDB B5A+EDB B6A+EDB B7A+EDB EDB

Figure 7.Photopolymerization profiles (double bond function conversion vs. irradiation time) in the presence of photoinitiating system based on biphenyls derivatives and EDB (0.5/1.5% w/w) at UV-A light LED source 365 nm for (a) 2-amino-4-methyl-6-(4-phenylphenyl)benzene-1,3-dicarbonitrile derivatives (series 1) and (b) 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives (series 2).

(11)

Catalysts 2019, 9, 827 11 of 27

Table2shows the conversion values obtained during the polymerization of various monomers according to different mechanisms and different light sources.

2.5. Photoinduced Electron Transfer Process with Photo-Oxidation Mechanism Between Biphenyl Derivatives and Iodonium Salt

As shown above, all investigated derivatives of 2-amino-4-methyl-6-phenyl-benzene- 1,3-dicarbonitrile derivatives can be employed as visible light photosensitizers in cationic and free-radical photoinitiated polymerizations. These derivatives initiate a photoinduced electron transfer reaction in the photoexcited state through oxidative processes in conjunction with iodonium salt (e.g., Speedcure 938). In this process, a bimolecular system is involved, in which one molecule, after absorbing the light, becomes an electron donor and is oxidized as a result of electron transfer, while the second molecule is reduced. In this case, the investigated 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives are the photosensitizers which, due to their spectroscopic properties (high molar extinction coefficient for the absorption band in the range of 300–425 nm), can easily absorb light in the UV-A and visible ranges and are easily oxidizable. Diaryliodonium salt (e.g., Speedcure 938) shows strong electron-acceptor properties in its basic state and is reduced as a result of electron transfer. The radicals resulting from the reduction of iodonium cation decay into secondary radicals are capable of further reaction, including initiation of a free-radical photopolymerization process, while the radicals produced from the oxidation of the biphenyl derivatives, in combination with anions from iodonium salts, may decay into strong protonic acids capable of initiating cationic polymerization. Alternatively, they may initiate the process directly. In order to demonstrate further that the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives are suitable for the role of photosensitizers, their oxidation potential was determined by cyclic voltammetry (Figure8a). The results summarized in Table3show that these compounds have sufficiently low oxidation potentials to interact effectively with iodonium salt. Additionally, an important parameter determining the efficiency of molecules in the photoinitiating system operating according to the photo-oxidation mechanism is the free energy change of electron transfer (∆Get) between individual components of the photoinitiating system. This process must be thermodynamically allowed; therefore, the free energy change (∆Get) must have a negative value. For this reason, the values of the parameter (∆Get) from the classical Rehm–Weller equation were calculated. The singlet state energy of sensitizer (ES1) was determined based on excitation and emission spectra (Figure8b).

In turn, triplet state energy (ET1) was calculated from molecular orbital calculations using the density functional theory (DFT) method at a B3LYP/6-31G (d, p) level of theory (contour plots of HOMOs and LUMOs for all biphenyl derivatives structures were optimized by B3LYP/6-31G (d, p) level of theory found in the Supporting Information in the Table S1). Negative values of Gibbs free energy were obtained for both singlet and triplet excited states, which confirmed the ability of electron exchange between the 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives and iodonium salt.

Thereby, iodonium salt, in combination with a biphenyl compound, can act as an effective bimolecular photoinitiating system for cationic and free-radical photopolymerizations during a photoinduced electron transfer process under a photo-oxidation mechanism.

Additionally, in order to fully confirm this mechanism, the steady-state photolysis of derivatives in the presence of iodonium salt (Speedcure 938) was carried out. The interaction of 2-amino-4-methyl-6-phenyl-benzene-1,3-dicarbonitrile derivatives with Speedcure 938 was confirmed by the rapid decrease of intensity at the maximum of absorption (λmax-ab) spectra of the studied biphenyl derivatives over time. A very fast decrease in the absorption spectrum at a wavelength in the range 320–400 nm was observed, together with a noticeable increase in the absorption peak in the long-term spectrum in the range of about 400–500 nm, which indicates the appearance of products of photolysis decay (example of photolysis can be found in the Supporting Information, Figures S99–S102).

Cytaty

Powiązane dokumenty

Skoro logika ma, zdaniem Wittgensteina, kodyfi kować formalne własności języka i świata 47 i jednocześnie poprzedzać wszelkie możliwe doświadczenie „trosz- cząc się sama

180 Warstwy ochronne na bazie metali wysokotopliwych wytwarzane techniką natryskiwania cieplnego.. ciekłe szkło oraz brak reakcji ze szkłem dającej niekorzystne zabarwienie

Dorobek 50 -letniej działalności Sekcji Folklorystycznej i Lu- doznawczej oraz Sekcji Historii Regionu Zarządu Głównego Polskiego Związku Kulturalno- -Oświatowego.

Pod wrażeniem wypadków wileńskich Starzeński złożył na ręce Wałujewa nowy memoriał, w którym powtórnie wskazał na zgubne konsekwencje polityki rosyjskiej na

W stacji sformatowanej dobór zawartości muzycznej ustalany jest odgórnie, czyli 

Високопродуктивні методи моделювання та ідентифікації feedback-впливів компететивної адсорбції поглинутих газових забруднень атмосфери на мікро- і

języka różne sfery, jakich dotyczy problem głodu. W etyce bowiem można ująć zarówno ekonomiczne jak i społeczne tło problemu, a także uwzględnić punkt

Eigensystem o f a Matrix:A powerful tool in both the anal- ysis and implementation of linear equations such as (124) i s the set of eigenvalues and eigenvectors