Next Article in Journal
Modified Flory–Rehner Theory Describes Thermotropic Swelling Transition of Smart Copolymer Microgels
Next Article in Special Issue
Bio-Inspired 4D Printing of Dynamic Spider Silks
Previous Article in Journal
Candidate Acetic Acid Bacteria Strains for Levan Production
Previous Article in Special Issue
Integration of Biofunctional Molecules into 3D-Printed Polymeric Micro-/Nanostructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Copper Complexes as Visible Light Photoinitiators for the Synthesis of Interpenetrating Polymer Networks (IPNs)

1
Université de Haute-Alsace, CNRS, IS2M UMR 7361, F-68100 Mulhouse, France
2
Université de Strasbourg, F-67087 Strasbourg, France
3
Laboratory of Materials, Catalysis, Environment and Analytical Methods (MCEMA) and LEADDER Laboratory, Faculty of Sciences, Doctoral School of Sciences and Technology (EDST), Lebanese University, Beirut 6573-14, Lebanon
4
Aix Marseille Univ, CNRS, ICR UMR 7273, F-13397 Marseille, France
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(10), 1998; https://doi.org/10.3390/polym14101998
Submission received: 22 April 2022 / Revised: 7 May 2022 / Accepted: 9 May 2022 / Published: 13 May 2022
(This article belongs to the Special Issue Advanced Materials in 3D/4D Printing Technology)

Abstract

:
This work is devoted to the study of two copper complexes (Cu) bearing pyridine ligands, which were synthesized, evaluated and tested as new visible light photoinitiators for the free radical photopolymerization (FRP) of acrylates functional groups in thick and thin samples upon light-emitting diodes (LED) at 405 and 455 nm irradiation. These latter wavelengths are considered to be safe to produce polymer materials. The photoinitiation abilities of these organometallic compounds were evaluated in combination with an iodonium (Iod) salt and/or amine (e.g., N-phenylglycine—NPG). Interestingly, high final conversions and high polymerization rates were obtained for both compounds using two and three-component photoinitiating systems (Cu1 (or Cu2)/Iodonium salt (Iod) (0.1%/1% w/w) and Cu1 (or Cu2)/Iod/amine (0.1%/1%/1% w/w/w)). The new proposed copper complexes were also used for direct laser write experiments involving a laser diode at 405 nm, and for the photocomposite synthesis with glass fibers using a UV-conveyor at 395 nm. To explain the obtained polymerization results, different methods and characterization techniques were used: steady-state photolysis, real-time Fourier transform infrared spectroscopy (RT-FTIR), emission spectroscopy and cyclic voltammetry.

Graphical Abstract

1. Introduction

The elaboration of polymers by photochemical means, such as free radical photopolymerization (FRP) and cationic photopolymerization (CP), have been mainly based on the use of metal-free organic dyes and photoinitiators at the industrial and academic levels [1,2,3,4,5,6,7,8,9,10,11,12], and these synthetic processes (FRP and CP) are widely used in different fields, e.g., dentistry [13,14,15,16,17,18,19,20,21,22,23], adhesives [24,25,26,27,28], coatings [29,30,31,32,33], composites [34], medicine [35,36,37,38,39,40], direct laser write, 3D and 4D printing [41,42,43,44,45,46,47,48,49,50], etc. On the other hand, organometallic compounds are not really used in industry; in other words, manufacturers avoid incorporating metallic compounds in their synthetic formulations due to their potential toxicity and price [51,52,53,54,55,56,57,58,59,60,61,62,63,64]. With their photochemical properties, such as high-absorption properties in the near-UV and visible range [65,66,67,68], long-lived excited states [69,70,71,72,73,74], suitable redox potentials [75,76,77,78,79,80,81,82,83,84,85,86,87,88,89], copper complexes can be used as photoinitiators (PIs)/photoredox catalysts able to produce active species, according to a catalytic cycle [90,91]. Therefore, it is very important to develop new metal-free photoinitiators or low-cost organometallic-based complexes [92,93,94,95].
In fact, copper complexes have attracted much attention and intense efforts have been devoted in recent years to the development of new copper complexes of improved photosensitivity, due to their competitive costs compared to other metal complexes. Copper complexes bearing a pyridine-based chelate ligand showed excellent photochemical properties for photocatalysis process, such as high-oxidation potential in the excited state [96,97,98,99], long-excited-state lifetime, high-emission quantum yields and high-absorption properties in the UV-visible region. Furthermore, copper complex derivatives have already been tested as PIs for FRP, CP, as well as IPN synthesis [100,101,102,103,104].
In this paper, two new copper complexes (Cu1–Cu2) (Figure 1) were synthesized and investigated as visible light photoinitiators upon exposure to LEDs at 405 and 455 nm for FRP, CP and the synthesis of interpenetrating polymer networks (IPNs) of acrylate/epoxy monomer blends. These compounds will be incorporated in two (Cu1 (or Cu2)/Iod (0.1%/1% w/w)) and three-component (Cu1 (or Cu2)/Iod/NPG (0.1%/1%/1% w/w/w)) photoinitiating systems (PISs) to produce polymer materials by free radical photopolymerization and the polymerization of acrylate/epoxy blend (IPNs). The photoinitiating ability of copper complexes will also be explained based on the interaction of Cu1 (or Cu2)/Iod and Cu1 (or Cu2)/Iod/NPG, which can be studied using different techniques and characterization processes, e.g., steady-state photolysis, cyclic voltammetry, fluorescence quenching and electron spin resonance spin trapping. Finally, to demonstrate the effectiveness of these new copper complex-based photoinitiators, experiments using direct laser writing (DLW), 3D printing and photocomposites synthesis were carried out in this work using different irradiation sources.

2. Materials and Methods

2.1. Synthesis of Chalcones, Ligands and Copper Complexes

Experimental conditions and acquisition conditions have been detailed elsewhere [13,92,100]. The two chalcones used for the design of ligands L1 and L2 were then engaged in a cyclization reaction with β-aminocrotonitrile according to a reaction reported in 1992 by Masaki Matsui [105,106]. Bis(2-isocyanophenyl) phenylphosphonate (binc) was synthesized by adapting a literature procedure [107,108].
Synthesis of bis(2-isocyanophenyl) phenylphosphonate (binc)
Polymers 14 01998 i001
Benzoxazole (9.1 g, 76.3 mmol, 1.0 equiv.) was dissolved in dry THF (200 mL). The solution was cooled at −78 °C and n-BuLi (2.5 M in hexane, 32.0 mL, 80.0 mmol, 1.05 equiv.) was added. Stirring was maintained for 1.5 h at this temperature. Phenyl phosphonic dichloride (5.7 mL, 4.04 mmol, 0.53 equiv.) was added and the solution could warm to room temperature. The solution was poured in Et2O:NaHCO3 (2:1, 150 mL). The organic phase was washed with water several times, dried over magnesium sulfate and the solvent removed under reduced pressure. The residue was crystallized in pentane/ethyl acetate (4/1) to provide the ligand (55% yield) as a light brown solid. 1H NMR (400 MHz, CDCl3) δ(ppm): 8.23–8.11 (m, 2H), 7.69 (td, J = 7.4, 1.3 Hz, 1H), 7.58 (dt, J = 12.5, 6.3 Hz, 2H), 7.48 (d, J = 8.4 Hz, 2H), 7.39 (d, J = 7.9 Hz, 2H), 7.34 (td, J = 8.1, 1.6 Hz, 2H), 7.18 (t, J = 7.7 Hz, 2H); 13C NMR (101 MHz, CDCl3) δ(ppm): 169.7, 145.7, 145.7, 134.5, 132.9, 132.8, 130.7, 130.7, 129.3, 129.1, 128.3, 125.9, 124.1, 124.1, 121.7, 121.7. (isonitrile carbons not detected); HRMS (ESI MS) m/z: theor: 360.0664 found: 360.0666 (M+. detected).
Synthesis of (E)-3-(4-(dimethylamino)phenyl)-1-(pyridin-2-yl)prop-2-en-1-one
Polymers 14 01998 i002
Next, 4-(Dimethylamino)benzaldehyde (1.49 g, 10.0 mmol, M = 149.19 g/mol) and 1-(pyridin-2-yl)ethan-1-one (1.21 g, 10.0 mmol, M = 121.14 g/mol) were suspended in ethanol (50 mL) and aq. KOH (40%) (10 mL) was added. After stirring overnight, the solid was filtered off, washed with ethanol and water, and dried under vacuum. The product was purified by filtration on a plug of SiO2 using dichloromethane (DCM) as the eluent (2.32 g, 92% yield). 1H NMR (400 MHz, CDCl3) δ(ppm): 8.61–8.49 (m, 1H), 8.01 (d, J = 7.4 Hz, 1H), 7.92 (d, J = 15.8 Hz, 1H), 7.77 (d, J = 15.8 Hz, 1H), 7.66 (t, J = 7.0 Hz, 1H), 7.45 (d, J = 7.7 Hz, 2H), 7.31–7.22 (m, 1H), 6.49 (d, J = 7.8 Hz, 2H), 2.83 (s, 6H); 13C NMR (101 MHz, CDCl3) δ(ppm): 189.16, 155.00, 152.11, 148.69, 145.94, 136.89, 130.90, 126.40, 122.99, 122.72, 115.47, 111.74, 40.06; HRMS (ESI MS) m/z: theor: 253.1296 found: 253.1299 ([M + H]+ detected).
Synthesis of (E)-3-(4-(dimethylamino)phenyl)-1-(6-methylpyridin-2-yl)prop-2-en-1-one
Polymers 14 01998 i003
Following this, 4-(Dimethylamino)benzaldehyde (2.21 g, 14.8 mmol, M = 149.19 g/mol) and 1-(6-methylpyridin-2-yl)ethan-1-one (2.00 g, 14.8 mmol, M = 135.17 g/mol) were dissolved in ethanol (50 mL) and aq. KOH (40%) (15 mL) was added. After stirring the solution overnight, the resulting solid was filtered off. It was purified by filtration on a plug of SiO2 using DCM as the eluent. For a higher purity, the solid was first dissolved in DCM and precipitated by addition of pentane (3.51 g, 89% yield). 1H NMR (400 MHz, CDCl3) δ(ppm): 8.02 (d, J = 15.8 Hz, 1H), 7.90 (d, J = 7.5 Hz, 1H), 7.84 (d, J = 15.9 Hz, 1H), 7.65 (t, J = 7.7 Hz, 1H), 7.58–7.53 (m, 2H), 7.23 (d, J = 7.4 Hz, 1H), 6.66–6.60 (m, 2H), 2.97 (s, 6H), 2.60 (s, 3H); 13C NMR (101 MHz, CDCl3) δ(ppm): 189.62, 157.67, 154.57, 152.04, 145.63, 136.95, 130.84, 126.02, 123.26, 119.84, 115.94, 111.77, 40.13, 24.54; HRMS (ESI MS) m/z: theor: 267.1453 found: 267.1451 ([M + H]+ detected).
Synthesis of 4-(4-(dimethylamino)phenyl)-6-methyl-[2,2′-bipyridine]-5-carbonitrile
Polymers 14 01998 i004
Potassium tert-butoxide (1.2 g) and β-aminocrotonitrile (4.92 g, 60 mmol, M = 82.10 g/mol) were dissolved in acetonitrile (300 mL) and the solution was heated at 35 °C for 15 min. Chalcone (E)-3-(4-(dimethylamino)phenyl)-1-(pyridin-2-yl)prop-2-en-1-one (2.52 g, 10.0 mmol, M = 252.32 g/mol) was added and stirring was maintained for three days. The solid was filtered off and washed with ethanol and water. It was purified by filtration on a plug of SiO2 using DCM as the eluent (2.83 g, 90% yield). 1H NMR (400 MHz, CDCl3) δ(ppm): 8.70 (ddd, J = 4.8, 1.7, 0.9 Hz, 1H), 8.50 (dt, J = 8.0, 1.0 Hz, 1H), 8.40 (s, 1H), 7.85 (td, J = 7.8, 1.8 Hz, 1H), 7.72–7.64 (m, 2H), 7.36 (ddd, J = 7.5, 4.8, 1.2 Hz, 1H), 6.85–6.77 (m, 2H), 3.05 (s, 6H), 2.89 (s, 3H); 13C NMR (101 MHz, CDCl3) δ(ppm): 162.47, 157.15, 155.03, 154.08, 151.47, 149.38, 137.02, 129.74, 129.74, 124.47, 123.44, 122.10, 118.22, 117.68, 111.97, 111.97, 106.10, 40.19, 40.19, 24.38; HRMS (ESI MS) m/z: theor: 315.1565 found: 315.1564 ([M + H]+ detected).
Synthesis of 4-(4-(dimethylamino)phenyl)-6,6′-methyl-[2,2′-bipyridine]-5-carbonitrile
Polymers 14 01998 i005
Potassium tert-butoxide (1.2 g) and β-aminocrotonitrile (4.92 g, 60 mmol, M = 82.10 g/mol) were dissolved in acetonitrile (300 mL) and the solution was heated at 35 °C for 15 min. Chalcone (E)-3-(4-(dimethylamino)phenyl)-1-(6-methylpyridin-2-yl)prop-2-en-1-one (L3) (2.66 g, 10.0 mmol, M = 266.34 g/mol) was added and stirring was maintained for three days. The solid was filtered off and washed with ethanol and water. It was purified by filtration on a plug of SiO2 using DCM as the eluent (2.76 g, 84% yield). 1H NMR (400 MHz, CDCl3) δ(ppm): 8.40 (s, 1H), 8.27 (d, J = 7.8 Hz, 1H), 7.72 (t, J = 7.8 Hz, 1H), 7.69–7.63 (m, 2H), 7.21 (d, J = 7.6 Hz, 1H), 6.85–6.79 (m, 2H), 3.05 (s, 6H), 2.88 (s, 3H), 2.63 (s, 3H); 13C NMR (101 MHz, CDCl3) δ(ppm): 162.36, 158.24, 157.56, 154.37, 154.00, 151.43, 137.13, 129.70, 129.70, 124.09, 123.68, 119.11, 118.27, 117.78, 112.01, 112.01, 105.95, 40.21, 40.21, 24.61, 24.38; HRMS (ESI MS) m/z: theor: 329.1722 found: 329.1719 ([M + H]+ detected).
Synthesis of Cu1
Polymers 14 01998 i006
Tetrakis(acetonitrile)copper(I) hexafluorophosphate (372 mg, 1.0 mmol, M = 372.72 g/mol), bis(2-isocyanophenyl)phenylphosphonate (binc) (360 mg, 1.0 mmol, M = 360.31 g/mol) and 4-(4-(dimethylamino)phenyl)-6-methyl-[2,2′-bipyridine]-5-carbonitrile (314 mg, 1.0 mmol, M = 314.39 g/mol) were dissolved in DCM (100 mL) and the solution was stirred at 25 °C for 2 h. The solution was concentrated to ca. 5 mL. Diethyl ether was added, as the product was a solid (866 mg, 98% yield). 1H NMR (400 MHz, CDCl3) δ(ppm): 8.96 (s, 1H), 8.51 (s, 1H), 8.26 (s, 1H), 8.14 (dd, J = 14.2, 7.6 Hz, 2H), 7.85–7.70 (m, 4H), 7.70–7.59 (m, 4H), 7.52–7.35 (m, 5H), 7.24 (t, J = 8.1 Hz, 2H), 6.94 (s (br), 2H), 3.21 (s, 3H), 3.11 (s, 6H); HRMS (ESI MS) m/z: theor: 737.1486 found: 737.1481 (M+. detected); Anal. Calc. for C40H31CuF6N6O3P2: C, 54.4; H, 3.5; O, 5.4; Found: C, 54.6; H, 3.4; O, 5.5%.
Synthesis of Cu2
Polymers 14 01998 i007
Tetrakis(acetonitrile)copper(I) hexafluorophosphate (372 mg, 1.0 mmol, M = 372.72 g/mol), bis(2-isocyanophenyl)phenyl phosphonate (binc) (360 mg, 1.0 mmol, M = 360.31 g/mol) and 4-(4-(dimethylamino)phenyl)-6,6′-methyl-[2,2′-bipyridine]-5-carbonitrile (328 mg, 1.0 mmol, M = 328.42 g/mol) were dissolved in DCM (100 mL) and the solution was stirred at 25 °C for 2 h. The solution was concentrated to ca. 5 mL. Addition of diethyl ether precipitated the product as a yellow solid (883 mg, 98% yield). 1H NMR (400 MHz, DMSO) δ(ppm): 8.76 (s, 1H), 8.62 (s, 1H), 8.21 (s, 1H), 8.02 (dd, J = 13.2, 7.2 Hz, 2H), 7.79 (s, 6H), 7.67 (s, 2H), 7.56 (t, J = 7.9 Hz, 2H), 7.35 (dd, J = 21.3, 8.0 Hz, 5H), 6.92 (d, J = 8.8 Hz, 2H), 3.14 (s, 3H), 3.06 (s, 6H), 2.93 (s, 3H); HRMS (ESI MS) m/z: theor: 751.1642 found: 751.1639 (M+. detected); Anal. Calc. for C41H33CuF6N6O3P2: C, 54.9; H, 3.7; O, 5.3; Found: C, 55.0; H, 3.4; O, 5.5%.

2.2. Other Chemicals

Chemical structure of the different monomers and additives are presented in Figure 2. Di-tert-butyl-diphenyl iodonium hexafluorophosphate (Iod) and ethyl 4-(dimethylamino)benzoate (EDB) were obtained from Lambson Ltd. (UK). Di(trimethylolpropane) tetraacrylate (TA), trimethylolpropane triacrylate (TMPTA), (3,4-epoxycyclohexane)methyl 3,4-epoxycyclohexylcarboxylate (EPOX; Uvacure 1500), N-phenylglycine (NPG), N-vinylcarbazole (NVK) and N,N-dimethyl-p-toluidine (TMA) were obtained from Allnex or Sigma Aldrich. TA, TMPTA and EPOX were chosen as acrylic and cationic monomers for the radical and cationic polymerizations.

2.3. Irradiation Sources: Light-Emitting Diodes

All the irradiation sources used during these experiments are based on light-emitting diodes (LEDs) and used as safe sources: (1) LED at 375 nm (I0 = 75 mW·cm−2) for the photolysis experiments, (2) LED at 405 nm (I0 = 110 mW·cm−2) and 455 nm (I0 = 75 mW·cm−2) for the photopolymerization experiments, (3) LED conveyor at 395 nm (I0 = 4 W·cm−2) for the photocomposite synthesis.

2.4. Photopolymerization Kinetics Determination by Real-Time Fourier Transform Infrared Spectroscopy (RT-FTIR)

In the present work, copper derivatives were used in two and three-component PISs for both FRP and CP under irradiation with LEDs at 405 and 455 nm. PISs were mainly based on two-component Cu1 (or Cu2)/Iod (0.1%-0.2%-0.5%/1% w/w) and three-component Cu1 (or Cu2)/Iod/amine (NPG, NVK) (0.1%/1%/1% w/w/w) systems. The weight percOKent of the photoinitiating (PI, co-initiator and amine) system was calculated from the global monomer content. Firstly, two different samples were studied for each photosensitive formulation in (i) thick (thickness = 1.4 mm) and (ii) thin samples (thickness = 25 µm). The epoxy and acrylate conversions were continuously followed by RT-FTIR using a JASCO FTIR 6600 (JASCO France, Lisses, France), so it was possible to determine the final conversion of reactive functions and to calculate the polymerization rate of each kinetic. Acrylate functions in thick and thin samples show peaks towards 6160 cm−1 and 1630 cm−1, respectively, and the epoxide functions show peaks around 3600 cm−1 and 790 cm−1 for the thick and thin samples, respectively.

2.5. Redox Potentials: Electrochemical Properties

Redox potentials of copper derivatives were determined in DCM by cyclic voltammetry using tetrabutylammonium hexafluorophosphate as the supporting electrolyte (potentials vs. saturated calomel electrode (SCE)). Free energy change (ΔGet) for an electron transfer reaction was calculated from Equation (1) [109], where Eox, Ered, E* and C represent the oxidation potential of the electron donor, the reduction potential of the electron acceptor, the excited-state energy level (determined from fluorescence experiments) and the coulombic term for the initially formed ion pair, respectively. Here, C is neglected as is usually the case for polar solvents.
ΔGet = Eox − Ered − E* + C

2.6. UV-Visible Absorption, Steady-State Photolysis and Luminescence Experiments

Acquisition conditions have been detailed elsewhere [13,92,100].

2.7. Computational Procedure

Computational conditions have been detailed elsewhere [13,92,100,110,111]. Simulated absorption spectra of copper complexes were obtained by time-dependent density functional theory at the MPW1PW91/6-31G* level of theory on the relaxed geometries calculated at the UB3LYP/6-31G* level of theory.

2.8. Photocomposite Access Using a Near-UV Conveyor

Photocomposite materials were obtained using a Dymax-UV conveyor at 395 nm. Firstly, photosensitive resins were deposited on the glass fibers (reinforcement), then, this mixture was cured using an LED conveyor @395 nm (I = 4 W·cm−2). Distance between the belt and the LED was fixed to 15 mm, and the belt speed was fixed at 2 m/min.

2.9. Direct Laser Write (DLW) Experiment

The photosensitive formulation was deposited on a glass slide and 3D patterns were obtained under air using a computer-controlled diode laser at 405 nm (spot size = 50 μm). Analysis of the 3D patterns was carried out using a numerical optical microscope (DSX-HRSU from OLYMPUS Corporation, Rungis, France) [112].

3. Results

Light-absorption properties, initiation ability and applications (photocomposite synthesis and direct laser write) of the investigated copper complexes will be studied in this section.

3.1. Synthetic Routes to Copper Complexes Cu1 and Cu2

Copper complexes bearing a pyridine-based chelate ligand showed excellent photochemical properties for photocatalysis processes, such as a high-oxidation potential in the excited state [96,97,98,99], a long-excited-state lifetime, high-emission quantum yields and high-absorption properties in the UV-visible region. In this work, two new copper complexes have been developed, allowing, by the convenient choice of the ligands, a shift in the absorption properties in the visible range, while maintaining high efficiency.
To allow such a shift of the absorption properties towards the visible, two bipyridine ligands were synthesized starting from a chalcone. Structures of these chalcones and, therefore, of the corresponding ligands have been selected in order to induce a significant shift in the absorption spectrum of the copper complexes towards the visible range.
For the synthesis of the two chalcones, a Claisen-Schmidt condensation reaction under basic conditions between 2-acetylpyridine A1 or 2-acetyl-6-methylpyridine A2 and aldehyde A3 was carried out (See Scheme 1) [113,114,115,116,117,118,119,120,121,122,123,124,125].
These two chalcones engage in a cyclization reaction with β-aminocrotonitrile, allowing the ligands to be formed. This reaction and its corresponding mechanism were described in the literature in 1992 by Masaki Matsui [105,106].
As shown in Scheme 2, the mechanism proposed by Masaki Matsui involves the formation of two (L1 or L2) ligands. β-Aminocrotononitrile can exist as amino (1) and imino (2) isomers in solution. A Michael addition of thae carbanion of imino isomer 3 to C1 or C2 can give intermediate 4, which, after an intramolecular cyclization and dehydration reaction, can provide intermediate 6. At room temperature, dehydrogenation of 6 can give 7.
Starting from the conditions described by Masaki Matsui, the ligands were indeed obtained. The synthesis conditions were then optimized by using reduced quantities of solvent, base and β-aminocrotonitrile, and with a simplified reaction treatment due to the precipitation of the ligand at the end of the reaction, while maintaining a good yield. The optimized synthesis of the ligands is detailed in the synthetic procedure detailed above.
Heteroleptic copper complexes bearing a pyridine-based chelate ligand and a diphosphine ligand, such as 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene (or Xantphos) and bis[2-(diphenylphosphino)phenyl]ether (or DPEphos), have previously been reported in the literature. In this work, this second ligand was replaced by a bisisonitrile ligand, namely bis(2-isocyanophenyl)phenylphosphonate. Indeed, as previously mentioned in a study reported by Matthias Knorn [107], copper complexes bearing this ligand showed a higher photocatalytic activity than their counterpart comprising a bisphosphine ligand. The lower activity of copper complexes comprising bisphosphine ligands can be assigned to the tendency of heteroleptic complexes to form an equilibrium with their homoleptic forms in solution [108], especially for heteroleptic copper complexes combining biphosphine and phenanthroline ligands. In contrast, in the case of the bisisonitrile ligand, studies have revealed the low ability of heteroleptic complexes to undergo ligand exchanges. The ligand was synthesized following the procedure described in the literature. Using these two ligands, the pyridine ligands and the bisisonitrile ligand in a one-step complexation reaction, the two desired copper complexes were obtained.

3.2. UV-Visible Absorption Spectra of Cu1 and Cu2

Ground-state absorption spectra of the new studied copper derivatives were determined in DCM and the results are presented in Figure 3. Extinction coefficients at different emission wavelengths used in photopolymerization experiments are reported in Table 1. New complexes are characterized by a broad absorption band, which extends between 380 nm and 650 nm, and high-extinction coefficients in the blue region, e.g., ε = 7570 M−1·cm−1, 7040 M−1·cm−1 @400 nm for Cu1 and Cu2, respectively. These compounds also have high-extinction coefficients at the emission wavelengths of LEDs (at 405 nm and 455 nm) used in different experiments achieved in this work, for example, ε@405nm = 8460 and 7950 M−1·cm−1 for Cu1 and Cu2, respectively. Remarkably, a bathochromic shift in the absorption spectra of Cu2 was observed compared to that of Cu1. This effect could be related to the presence of a methyl group, which is considered as an electron-donating group (inductive effect), on the pyridine ligands of Cu2 (λmax = 445 nm for Cu1 and 441 nm for Cu2). This difference could also be explained by the optimized geometries, as well as the frontier orbitals (highest occupied molecular orbital—HOMO—and lowest unoccupied molecular orbital—LUMO), which are shown in Figure 4.

3.3. Photopolymerization Experiments

3.3.1. Free Radical Photopolymerization Using TA as a Benchmark Monomer

Due to their good light-absorption properties in the visible range, copper complexes were tested as photoinitiators for the FRP of acrylate-based monomers upon exposure to LEDs at 405 nm (I = 110 mW∙cm−2) and 455 nm (I = 75 mW∙cm−2).
In fact, photoinitiators (0.1% or 0.2%) were dissolved and mixed into the TA acrylate monomer in combination with Iod salt (1%) in order to make two-component photoinitiating systems, on the one hand, and in combination with Iod/amine (1%/1% w/w) to form three-component photoinitiating systems, on the other hand. Interestingly, these dyes exhibit excellent free radical photopolymerization conversions in thick and thin samples. The related results are gathered in Figure 5 and the data are summarized in Table 2. Remarkably, copper complexes alone, Iod and amine alone cannot polymerize the sample. Iod salt and amine are used as co-initiators in this work because they do not absorb visible light. It is important to introduce the dyes (i.e., the copper complexes) into the photosensitive formulations in order to obtain a good light absorption at 405 nm and 455 nm. The obtained results using copper derivatives in two-component PISs showed that Cu2 was more efficient than Cu1 for the FRP of TA using different PI percentages, e.g., FC ~ 64% for Cu1/Iod (0.1%/1% w/w) vs. 70% for Cu2/Iod (0.1%/1% w/w) (Figure 5A curve 1 vs. 2), and FC ~ 62% for Cu1/Iod (0.2%/1% w/w) vs. 85% for Cu2/Iod (0.2%/1% w/w) (Figure 5A curve 3 vs. 4).
Furthermore, Iod/NPG couple showed a weak polymerization initiation ability upon exposure to LEDs at 405 nm and 455 nm after 60 s (e.g., FC ~ 10% @405 nm). Interestingly, a greater efficiency was observed when NPG was incorporated into the formulation. Compared to their two-component system analogues, the different three-component PISs showed a better final conversion of reactive functions and a higher polymerization rate upon irradiation with LEDs at 405 nm or 455 nm (for example, an FC up to 86% is obtained with Cu1/Iod/NPG (0.1%/1%/1% w/w/w), and 88% using Cu2/Iod/NPG (0.1%/1%/1% w/w/w) with a LED @455 nm).

3.3.2. Cationic Polymerization and IPN Synthesis

Typical epoxide function conversion-time profiles for Cu1 and Cu2-based photoinitiating systems are given in Figure 6 and the data are gathered in Table 3. In fact, the cationic polymerization of the epoxide functions was carried out under air and upon irradiation at 405 nm. Indeed, the cationic polymerization is insensitive to oxygen. As expected, copper complexes alone and the additives alone were not able to initiate the CP in these irradiation conditions. The addition of Iod salt or Iod/NVK into the formulation containing the PI induced good photopolymerization profiles, i.e., the combination Cu/Iod/NVK (0.1%/2%/3% w/w/w) is very efficient to produce polymer materials in terms of Rp and final epoxy function conversion compared to Cu/Iod (0.1%/1% w/w), e.g., (FC ~ 50% for Cu1/Iod/NVK (0.1%/2%/3% w/w/w) vs. 27% for Cu1/Iod (0.1%/1% w/w)). The consumption of epoxide functions was accompanied by the formation of a polyether network (appearance of peak at ~1080 cm−1), characterizing the obtained polymer.
On the other hand, IPNs syntheses were also carried out in this work and polymerization tests were performed in thick and thin samples using LEDs at 405 nm and 455 nm (See Table 4 and Table 5). Photopolymerization profiles for the IPN formation are presented in Figure 7. For example, the acrylic network formation was very fast with a high final conversion (98%) for Cu2/Iod/NPG (0.1%/1%/1% w/w/w) in TA/EPOX (50%/50%) upon irradiation at 455 nm, and the formation of the epoxy network was also efficient (high final conversion and Rp) using this system (FC ~ 55%).

3.4. Photocomposites Synthesis

Nowadays, many of our modern technologies require materials with enhanced properties. This is particularly true for materials used in aerospace, underwater and transportation applications. For example, for aeronautical applications, engineers research materials with properties of low density, rigid, solid, impact resistance, temperature and pressure resistance and obviously materials that do not easily corrode. For this purpose, composite materials have been used for different applications. By definition, a composite material is composed of a least two components that results in better properties than those of the individual components used alone: matrix (monomer blends) and reinforcement. The main advantages of composite materials are their high stiffness, strength and low density. The introduction of light for the synthesis of photocomposites will make the manufacture of these materials more ecological.
In this study, the matrix is based on acrylic monomers, such as TMPTA or TA, and the second component (reinforcement) is based on glass fibers. Firstly, the acrylic resins were deposed on the reinforcement (50%/50% w/w) and the mixtures were irradiated using an LED conveyor at 395 nm. Interestingly, a very fast polymerization on the surface and the bottom was observed with tack-free surfaces, after one pass only using one layer of glass fibers (1 mm). Increasing the reinforcement thickness by adding several layers, the polymerization on the surface is always fast and takes place after one pass, but the curing on the bottom is more complicated and will be done after several passes using Cu1 or Cu2/Iod/NPG (0.1%/1%/1% w/w/w) as PISs. The curing photocomposite results are depicted in Figure 8 and Table 6.

3.5. Direct Laser Write (DLW)

The new copper complexes were tested in some direct laser write experiments for the FRP of TMPTA or TA using a laser diode at 405 nm (spot size: 50 μm). The obtained 3D patterns were carried out under air and using different PISs based on Cu1/Iod/TMA, Cu2/Iod/TMA in TA or TMPTA (Figure 9). Due to their high ability to initiate the FRP of acrylates, these systems were able to generate high-spatial-resolution 3D patterns with a great thickness of curing (~2500 μm) in the irradiated area. As such, 3D patterns were generated with very short irradiation times (2–3 min) and they were characterized by numerical microscopy.

3.6. Mechanical Properties: Tensile Test Measurements

The tensile strength of IPNs synthesized using different compositions of the TA/EPOX mixture are presented in Table 7. The results show that with the increase in the percentage of acrylic monomer, the tensile strength increases, which may be due to the rigid character of the acrylates (e.g., 7.2 MPa for Cu2/Iod/NPG in TA/EPOX (30%/70%) vs. 37.2 MPa for the same system in TA/EPOX (70%/30%)).

4. Discussion

In order to explain the initiating ability of the organometallic complexes, their photochemical and photophysical properties were studied using different characterization techniques, allowing for the characterization of the associated chemical mechanisms.

4.1. Steady-State Photolysis of the Investigated Compounds

Photolyses of Cu1 and Cu2 dyes in DCM were investigated upon irradiation at 375 nm and 405 nm, and the related results are shown in Figure 10. First of all, no photolysis occurred for Cu1 and Cu2 alone (0% consumption) upon irradiation at 375 nm and 405 nm, but the incorporation of the iodonium salt into the photosensitive solution could promote the degradation of the dyes, so that a strong decrease in the absorbance band intensity was observed by increasing the irradiation time, e.g., consumption ~ 80% @375 nm and 82% for Cu2/Iod at 375 nm and 405 nm, respectively (Figure 11B). It is important to note that the photolysis of Cu1 in the presence of Iod salt involved the formation of a photoproduct after 60 s of irradiation in the solution, which had an absorption band more shifted in the visible (bathochromic effect) spectrum, then this photoproduct degrades under the effect of the irradiation (Figure 10A).
This difference between these two consumption percentages may be due to the high light-absorption ability of Cu2 at 405 nm, as well as the highest intensity of the LED at 405 nm (110 mW·cm−2), compared to that at 375 nm (75 mW·cm−2). Furthermore, in the case of three-component PISs, the consumption of Cu2 was lower compared to that of the two-component Cu2/Iod system. It can be confidently assigned to the regeneration of Cu2 in the three-component system due to the presence of the sacrificial amine or the formation of new photoproducts (%consumption ~ 31%).

4.2. Photoluminescence and Electrochemical Properties

Fluorescence emission spectra, fluorescence quenching (measured using a JASCO FP-6200 spectrofluorimeter, Lisses, France) and the oxidation potential (measured in DCM by cyclic voltammetry, OrigaLys, Rillieux-la-Pape, France) results of the different Cu derivatives are gathered in Figure 12 and Table 8. The excited-state energy was calculated from the crossing point of the emission and absorption spectra. Using these different values, the free-energy change (ΔG) could be calculated; this parameter reflects the reactivity between Cu and Iod. In fact, a slight decrease in the fluorescence intensity was observed for Cu1/Iod, but this emission spectra showed a strong decrease for Cu2 upon addition of Iod. These behaviors explain the high reactivity of Cu2/Iod compared to Cu1/Iod e.g., ϕ = 0.55 for Cu1/Iod vs. 0.74 for Cu2/Iod. In addition, the ΔG value is negative for both complexes, so that the photo-oxidation interaction Cu1 (or Cu2)/Iod is favorable in both cases, with a superiority observed for Cu2 (ΔG = −0.74 and 0.62 eV for Cu2 and Cu1, respectively). It, therefore, explains the high-photoinitiation ability of Cu2 compared to Cu1.
Finally, the initiation ability of the new copper complexes could be explained by different characterization techniques, which allowed us to propose a chemical photoinitiation mechanism. Firstly, Cu is excited upon irradiation at 405 or 455 nm and interacts with Iod to generate aryl radical (Ar) and radical Cu●+ [r1–r2]. A charge transfer complex CTC can be formed after adding NPG into the photosensitive formulation. This complex is able to produce aryl radicals as active species for the radical photopolymerization [r3–r4]. Then, 1,3Cu could react with NPG and generate two radicals (NPG-H, Cu-H) [r5], and the first radical can undergo a decarboxylation and produce active radicals (NPG(-H,-CO2)) [r6]. This radical can also lead to the formation of two active species after interaction with Iod salt (NPG(-H,-CO2)+, Ar) [r7]. Lastly, copper complex derivatives are regenerated [r8–r9] (See Scheme 3).

5. Conclusions

In the present paper, new copper complexes were synthesized and tested as photoinitiators. These compounds have strong visible-light absorption and are able to initiate both the free radical photopolymerization and cationic polymerization. IPN synthesis through the simultaneous polymerization of acrylate/epoxy monomer blends was performed under air upon irradiation at 405 nm and 455 nm, using a very low quantity of copper complex, in two or three-component PIS. Cu2 showed a very interesting photoinitiation capacity compared to Cu1 in terms of final conversions of reactive functions and polymerization rates. The high reactivity of these compounds was demonstrated through some direct laser write experiments, where high-spatial-resolution 3D patterns were obtained. In addition, the synthesis of thick glass fiber photocomposites was possible. This work paves the way for the development of new organometallic photoinitiators.

Author Contributions

Conceptualization, F.D. and J.L.; methodology, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; software, B.G.; validation, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; formal analysis, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; investigation, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; resources, F.D. and J.L.; data curation, F.D. and J.L.; writing—original draft preparation, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; writing—review and editing, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; visualization, F.D.; supervision, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; project administration, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L.; funding acquisition, M.R., G.N., B.G., J.T., T.H., D.G., F.D. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Aix Marseille University and the Centre National de la Recherche Scientifique.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fouassier, J.P.; Lalevée, J. Photoinitiators for Polymer Synthesis, Scope, Reactivity, and Efficiency; Wiley-VCH Verlag: Weinheim, Germany, 2012. [Google Scholar]
  2. Fouassier, J.P. Photoinitiator, Photopolymerization and Photocuring: Fundamentals and Applications; Hanser Publishers: Munich, Germany, 1995. [Google Scholar]
  3. Dietliker, K.A. Compilation of Photoinitiators Commercially Available for UV Today; Sita Technology Ltd.: London, UK, 2002. [Google Scholar]
  4. Davidson, S. Exploring the Science, Technology and Application of UV and EB Curing; Sita Technology Ltd.: London, UK, 1999. [Google Scholar]
  5. Crivello, J.V.; Dietliker, K.; Bradley, G. Photoinitiators for Free Radical Cationic & Anionic Photopolymerisation; John Wiley & Sons: Chichester, UK, 1999. [Google Scholar]
  6. Cunningham, A.F.; Desobry, V. Radiation Curing in Polymer Science and Technology; Fouassier, J.P., Rabek, J.F., Eds.; Elsevier: Barking, UK, 1993; Volume 2, p. 323. [Google Scholar]
  7. Andrzejewska, E. Chapter 2—Free Radical Photopolymerization of Multifunctional Monomers. In Three-Dimensional Microfabrication Using Two-Photon Polymerization; Baldacchini, T., Ed.; Micro and Nano Technologies; William Andrew Publishing: Oxford, UK, 2016; pp. 62–81. ISBN 978-0-323-35321-2. [Google Scholar]
  8. Luu, T.T.H.; Jia, Z.; Kanaev, A.; Museur, L. Effect of Light Intensity on the Free-Radical Photopolymerization Kinetics of 2-Hydroxyethyl Methacrylate: Experiments and Simulations. J. Phys. Chem. B 2020, 124, 6857–6866. [Google Scholar] [CrossRef] [PubMed]
  9. Andrezajewska, E.; Grajek, K. Recent Advances in Photo-Induced Free-Radical Polymerization. MOJ Polym. Sci. 2017, 1, 58–60. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Z.; Zou, X.; Shi, F.; Liu, R.; Yagci, Y. Highly Efficient Dandelion-like near-Infrared Light Photoinitiator for Free Radical and Thiol-Ene Photopolymerizations. Nat. Commun. 2019, 10, 3560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Sangermano, M.; Roppolo, I.; Chiappone, A. New Horizons in Cationic Photopolymerization. Polymers 2018, 10, 136. [Google Scholar] [CrossRef] [Green Version]
  12. Zhou, Y.; Liao, W.; Ni, X. Cationic Photopolymerization Initiated by a Photocatalytic Complex Sensitive to Visible Light at 520 Nm. Catal. Lett. 2021, 151, 1766–1775. [Google Scholar] [CrossRef]
  13. Bouzrati-Zerelli, M.; Maier, M.; Dietlin, C.; Morlet-Savary, F.; Fouassier, J.P.; Klee, J.E.; Lalevée, J. A novel photoinitiating system producing germyl radicals for the polymerization of representative methacrylate resins: Camphorquinone/R3GeH/iodonium salt. Dent. Mater. 2016, 32, 1226. [Google Scholar] [CrossRef]
  14. Abdul-Monem, M.M. Naturally Derived Photoinitiators for Dental and Biomaterials Applications. Eur. Dent. Res. Biomater. J. 2021, 1, 72–78. [Google Scholar] [CrossRef]
  15. Alizadehgharib, S.; Östberg, A.-K.; Dahlstrand Rudin, A.; Dahlgren, U.; Christenson, K. The Effects of the Dental Methacrylates TEGDMA, Bis-GMA, and UDMA on Neutrophils in Vitro. Clin. Exp. Dent. Res. 2020, 6, 439–447. [Google Scholar] [CrossRef]
  16. Arikawa, H.; Takahashi, H.; Kanie, T.; Ban, S. Effect of Various Visible Light Photoinitiators on the Polymerization and Color of Light-Activated Resins. Dent. Mater. J. 2009, 28, 454–460. [Google Scholar] [CrossRef] [Green Version]
  17. Catarzi, D.; Cecchi, L.; Colotta, V.; Melani, F.; Filacchioni, G.; Martini, C.; Giusti, L.; Lucacchini, A. Structure-Activity Relationships of 1,2,4-Triazolo [1,5-a]Quinoxalines and Their 1-Deaza Analogs Imidazo [1,2-a]Quinoxalines at the Benzodiazepine Receptor. J. Med. Chem. 1994, 37, 2846–2850. [Google Scholar] [CrossRef]
  18. Boeira, P.O.; Meereis, C.T.W.; Suárez, C.E.C.; de Almeida, S.M.; Piva, E.; da Silveira Lima, G. Coumarin-Based Iodonium Hexafluoroantimonate as an Alternative Photoinitiator for Experimental Dental Adhesives Resin. Appl. Adhes. Sci. 2017, 5, 2. [Google Scholar] [CrossRef] [Green Version]
  19. Cadenaro, M.; Antoniolli, F.; Codan, B.; Agee, K.; Tay, F.R.; Dorigo, E.D.S.; Pashley, D.H.; Breschi, L. Influence of Different Initiators on the Degree of Conversion of Experimental Adhesive Blends in Relation to Their Hydrophilicity and Solvent Content. Dent. Mater. 2010, 26, 288–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Dickens, S.H.; Stansbury, J.W.; Choi, K.M.; Floyd, C.J.E. Photopolymerization Kinetics of Methacrylate Dental Resins. Macromolecules 2003, 36, 6043–6053. [Google Scholar] [CrossRef]
  21. Ikemura, K.; Endo, T. A Review of the Development of Radical Photopolymerization Initiators Used for Designing Light-Curing Dental Adhesives and Resin Composites. Dent. Mater. J. 2010, 29, 481–501. [Google Scholar] [CrossRef] [Green Version]
  22. Maffezzoli, A.; Pietra, A.D.; Rengo, S.; Nicolais, L.; Valletta, G. Photopolymerization of Dental Composite Matrices. Biomaterials 1994, 15, 1221–1228. [Google Scholar] [CrossRef]
  23. Neumann, M.G.; Schmitt, C.C.; Ferreira, G.C.; Corrêa, I.C. The Initiating Radical Yields and the Efficiency of Polymerization for Various Dental Photoinitiators Excited by Different Light Curing Units. Dent. Mater. 2006, 22, 576–584. [Google Scholar] [CrossRef]
  24. Maruno, T.; Murata, N. Properties of a UV-Curable, Durable Precision Adhesive. J. Adhes. Sci. Technol. 1995, 9, 1343. [Google Scholar] [CrossRef]
  25. 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] [Green Version]
  26. Gziut, K.; Kowalczyk, A.; Schmidt, B.; Kowalczyk, K.; Weisbrodt, M. Epoxy-Based Structural Self-Adhesive Tapes Modified with Acrylic Syrups Prepared via a Free Radical Photopolymerization Process. Polymers 2021, 13, 189. [Google Scholar] [CrossRef]
  27. Zhu, M.; Cao, Z.; Zhou, H.; Xie, Y.; Li, G.; Wang, N.; Liu, Y.; He, L.; Qu, X. Preparation of Environmentally Friendly Acrylic Pressure-Sensitive Adhesives by Bulk Photopolymerization and Their Performance. RSC Adv. 2020, 10, 10277–10284. [Google Scholar] [CrossRef]
  28. Kilponen, L.; Uusitalo, E.; Tolvanen, M.; Varrela, J.; Vallittu, P.K. Photopolymerization of Light Curing Adhesives Used with Metal Orthodontic Brackets and Matrices. J. Biomater. Tissue Eng. 2016, 6, 659–664. [Google Scholar] [CrossRef]
  29. Wu, L.; Baghdachi, J. Functional Polymer Coatings: Principles, Methods, and Applications; Wiley Series on Polymer Engineering and Technology: New York, NY, USA, 2015. [Google Scholar]
  30. Al Rashid, A.; Ahmed, W.; Khalid, M.Y.; Koç, M. Vat Photopolymerization of Polymers and Polymer Composites: Processes and Applications. Addit. Manuf. 2021, 47, 102279. [Google Scholar] [CrossRef]
  31. Pagac, M.; Hajnys, J.; Ma, Q.-P.; Jancar, L.; Jansa, J.; Stefek, P.; Mesicek, J. A Review of Vat Photopolymerization Technology: Materials, Applications, Challenges, and Future Trends of 3D Printing. Polymers 2021, 13, 598. [Google Scholar] [CrossRef] [PubMed]
  32. Noè, C.; Hakkarainen, M.; Malburet, S.; Graillot, A.; Adekunle, K.; Skrifvars, M.; Sangermano, M. Frontal-Photopolymerization of Fully Biobased Epoxy Composites. Macromol. Mater. Eng. 2021, 2100864. [Google Scholar] [CrossRef]
  33. Dikova, T.; Maximov, J.; Todorov, V.; Georgiev, G.; Panov, V. Optimization of Photopolymerization Process of Dental Composites. Processes 2021, 9, 779. [Google Scholar] [CrossRef]
  34. Mokbel, H.; Anderson, D.; Plenderleith, R.; Dietlin, C.; Morlet-Savary, F.; Dumur, F.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Copper Photoredox Catalyst “G1”: A New High Performance Photoinitiator for near-UV and Visible LEDs. Polym. Chem. 2017, 8, 5580. [Google Scholar] [CrossRef]
  35. Pereira, R.F.; Bártolo, P.J. Photopolymerizable hydrogels in regenerative medicine and drug delivery. In Hot Topics in Biomaterials; Future Science Book Series Volume 6; Future Science Ltd.: London, UK, 2014. [Google Scholar]
  36. Xu, X.; Awad, A.; Robles-Martinez, P.; Gaisford, S.; Goyanes, A.; Basit, A.W. Vat Photopolymerization 3D Printing for Advanced Drug Delivery and Medical Device Applications. J. Control. Release 2021, 329, 743–757. [Google Scholar] [CrossRef]
  37. Chiulan, I.; Heggset, E.B.; Voicu, Ş.I.; Chinga-Carrasco, G. Photopolymerization of Bio-Based Polymers in a Biomedical Engineering Perspective. Biomacromolecules 2021, 22, 1795–1814. [Google Scholar] [CrossRef]
  38. Elisseeff, J.; Anseth, K.; Sims, D.; McIntosh, W.; Randolph, M.; Langer, R. Transdermal Photopolymerization for Minimally Invasive Implantation. Proc. Natl. Acad. Sci. USA 1999, 96, 3104–3107. [Google Scholar] [CrossRef] [Green Version]
  39. Bao, Y.; Paunović, N.; Leroux, J.-C. Challenges and Opportunities in 3D Printing of Biodegradable Medical Devices by Emerging Photopolymerization Techniques. Adv. Funct. Mater. 2022, 32, 2109864. [Google Scholar] [CrossRef]
  40. Baroli, B. Photopolymerization of Biomaterials: Issues and Potentialities in Drug Delivery, Tissue Engineering, and Cell Encapsulation Applications. J. Chem. Technol. Biotechnol. 2006, 81, 491–499. [Google Scholar] [CrossRef]
  41. Lee, J.Y.; An, J.; Chua, C.K. Fundamentals and Applications of 3D Printing for Novel Materials. Appl. Mater. Today. 2017, 7, 120. [Google Scholar] [CrossRef]
  42. Andreu, A.; Su, P.-C.; Kim, J.-H.; Ng, C.S.; Kim, S.; Kim, I.; Lee, J.; Noh, J.; Subramanian, A.S.; Yoon, Y.-J. 4D Printing Materials for Vat Photopolymerization. Addit. Manuf. 2021, 44, 102024. [Google Scholar] [CrossRef]
  43. Bagheri, A.; Jin, J. Photopolymerization in 3D Printing. ACS Appl. Polym. Mater. 2019, 1, 593–611. [Google Scholar] [CrossRef] [Green Version]
  44. Tang, Y.; Dai, B.; Su, B.; Shi, Y. Recent Advances of 4D Printing Technologies Toward Soft Tactile Sensors. Front. Mater. 2021, 8, 658046. [Google Scholar] [CrossRef]
  45. Imrie, P.; Jin, J. Polymer 4D Printing: Advanced Shape-Change and Beyond. J. Polym. Sci. 2022, 60, 149–174. [Google Scholar] [CrossRef]
  46. Zhang, Z.; Corrigan, N.; Bagheri, A.; Jin, J.; Boyer, C. A Versatile 3D and 4D Printing System through Photocontrolled RAFT Polymerization. Angew. Chem. 2019, 131, 18122–18131. [Google Scholar] [CrossRef]
  47. Schwartz, J.J.; Boydston, A.J. Multimaterial Actinic Spatial Control 3D and 4D Printing. Nat. Commun. 2019, 10, 791. [Google Scholar] [CrossRef]
  48. Jeong, H.Y.; Woo, B.H.; Kim, N.; Jun, Y.C. Multicolor 4D Printing of Shape-Memory Polymers for Light-Induced Selective Heating and Remote Actuation. Sci. Rep. 2020, 10, 6258. [Google Scholar] [CrossRef]
  49. Shan, W.; Chen, Y.; Hu, M.; Qin, S.; Liu, P. 4D Printing of Shape Memory Polymer via Liquid Crystal Display (LCD) Stereolithographic 3D Printing. Mater. Res. Express 2020, 7, 105305. [Google Scholar] [CrossRef]
  50. Choong, Y.Y.C.; Maleksaeedi, S.; Eng, H.; Yu, S.; Wei, J.; Su, P.-C. High Speed 4D Printing of Shape Memory Polymers with Nanosilica. Appl. Mater. Today 2020, 18, 100515. [Google Scholar] [CrossRef]
  51. Tedla, G.; Jarabek, A.M.; Byrley, P.; Boyes, W.; Rogers, K. Human Exposure to Metals in Consumer-Focused Fused Filament Fabrication (FFF)/ 3D Printing Processes. Sci. Total Environ. 2022, 814, 152622. [Google Scholar] [CrossRef]
  52. Pierau, L.; Elian, C.; Akimoto, J.; Ito, Y.; Caillol, S.; Versace, D.-L. Bio-Sourced Monomers and Cationic Photopolymerization–The Green Combination towards Eco-Friendly and Non-Toxic Materials. Prog. Polym. Sci. 2022, 127, 101517. [Google Scholar] [CrossRef]
  53. Leonhardt, S.; Klare, M.; Scheer, M.; Fischer, T.; Cordes, B.; Eblenkamp, M. Biocompatibility of Photopolymers for Additive Manufacturing. Curr. Dir. Biomed. Eng. 2016, 2, 113–116. [Google Scholar] [CrossRef]
  54. Klikovits, N.; Knaack, P.; Bomze, D.; Krossing, I.; Liska, R. Novel Photoacid Generators for Cationic Photopolymerization. Polym. Chem. 2017, 8, 4414–4421. [Google Scholar] [CrossRef] [Green Version]
  55. Mueller, M.; Bandl, C.; Kern, W. Surface-Immobilized Photoinitiators for Light Induced Polymerization and Coupling Reactions. Polymers 2022, 14, 608. [Google Scholar] [CrossRef]
  56. 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]
  57. Tomal, W.; Ortyl, J. Water-Soluble Photoinitiators in Biomedical Applications. Polymers 2020, 12, 1073. [Google Scholar] [CrossRef]
  58. Photopolymerization Using Metal Oxide Semiconducting Nanoparticles for Epoxy-Based Coatings and Patterned Films. ACS Appl. Nano Mater. 2020, 3, 2875–2880. Available online: https://pubs-acs-org.lama.univ-amu.fr/doi/10.1021/acsanm.0c00147 (accessed on 24 April 2022). [CrossRef]
  59. Breimer, M.A.; Yevgeny, G.; Sy, S.; Sadik, O.A. Incorporation of Metal Nanoparticles in Photopolymerized Organic Conducting Polymers:  A Mechanistic Insight. Nano Lett. 2001, 1, 305–308. [Google Scholar] [CrossRef]
  60. Guo, J.; Jian, J.; Wang, M.; Tomita, Y.; Cao, L.; Wang, D.; Zhang, X. Ag Nanoparticle-Enhanced Alkyl Radical Generation in Photopolymerization for Holographic Recording. Nanophotonics 2019, 8, 1795–1802. [Google Scholar] [CrossRef]
  61. He, C.; Feng, Z.; Shan, S.; Wang, M.; Chen, X.; Zou, G. Highly Enantioselective Photo-Polymerization Enhanced by Chiral Nanoparticles and in Situ Photopatterning of Chirality. Nat. Commun. 2020, 11, 1188. [Google Scholar] [CrossRef] [PubMed]
  62. Ibn-El-Ahrach, H.; Bachelot, R.; Lérondel, G.; Vial, A.; Grimault, A.-S.; Plain, J.; Royer, P.; Soppera, O. Controlling the Plasmon Resonance of Single Metal Nanoparticles by Near-Field Anisotropic Nanoscale Photopolymerization. J. Microsc. 2008, 229, 421–427. [Google Scholar] [CrossRef]
  63. Ibn El Ahrach, H.; Bachelot, R.; Vial, A.; Lérondel, G.; Plain, J.; Royer, P.; Soppera, O. Spectral Degeneracy Breaking of the Plasmon Resonance of Single Metal Nanoparticles by Nanoscale Near-Field Photopolymerization. Phys. Rev. Lett. 2007, 98, 107402. [Google Scholar] [CrossRef] [PubMed]
  64. Cai, X.; Anyaogu, K.C.; Neckers, D.C. Photopolymerization of Gold Nanoparticles:  Size-Related Charge Separation and Emission. J. Am. Chem. Soc. 2007, 129, 11324–11325. [Google Scholar] [CrossRef] [PubMed]
  65. Lv, J.; Lu, Y.; Wang, J.; Zhao, F.; Wang, Y.; He, H.; Wu, Y. Schiff Base-Type Copper(I) Complexes Exhibiting High Molar Extinction Coefficients: Synthesis, Characterization and DFT Studies. J. Mol. Struct. 2022, 1249, 131638. [Google Scholar] [CrossRef]
  66. Kuang, D.; Ito, S.; Wenger, B.; Klein, C.; Moser, J.-E.; Humphry-Baker, R.; Zakeeruddin, S.M.; Grätzel, M. High Molar Extinction Coefficient Heteroleptic Ruthenium Complexes for Thin Film Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2006, 128, 4146–4154. [Google Scholar] [CrossRef]
  67. Sciortino, G.; Maréchal, J.-D.; Fábián, I.; Lihi, N.; Garribba, E. Quantitative Prediction of Electronic Absorption Spectra of Copper(II)–Bioligand Systems: Validation and Applications. J. Inorg. Biochem. 2020, 204, 110953. [Google Scholar] [CrossRef]
  68. Faggi, E.; Gavara, R.; Bolte, M.; Fajarí, L.; Juliá, L.; Rodríguez, L.; Alfonso, I. Copper(II) Complexes of Macrocyclic and Open-Chain Pseudopeptidic Ligands: Synthesis, Characterization and Interaction with Dicarboxylates. Dalton Trans. 2015, 44, 12700–12710. [Google Scholar] [CrossRef] [Green Version]
  69. Iwamura, M.; Takeuchi, S.; Tahara, T. Ultrafast Excited-State Dynamics of Copper(I) Complexes. Acc. Chem. Res. 2015, 48, 782–791. [Google Scholar] [CrossRef]
  70. Cuttell, D.G.; Kuang, S.-M.; Fanwick, P.E.; McMillin, D.R.; Walton, R.A. Simple Cu(I) Complexes with Unprecedented Excited-State Lifetimes. J. Am. Chem. Soc. 2002, 124, 6–7. [Google Scholar] [CrossRef] [PubMed]
  71. Giereth, R.; Reim, I.; Frey, W.; Junge, H.; Tschierlei, S.; Karnahl, M. Remarkably Long-Lived Excited States of Copper Photosensitizers Containing an Extended π-System Based on an Anthracene Moiety. Sustain. Energy Fuels 2019, 3, 692–700. [Google Scholar] [CrossRef]
  72. Egly, J.; Bissessar, D.; Achard, T.; Heinrich, B.; Steffanut, P.; Mauro, M.; Bellemin-Laponnaz, S. Copper(I) Complexes with Remotely Functionalized Phosphine Ligands: Synthesis, Structural Variety, Photophysics and Effect onto the Optical Properties. Inorg. Chim. Acta 2021, 514, 119971. [Google Scholar] [CrossRef]
  73. Bergmann, L.; Hedley, G.J.; Baumann, T.; Bräse, S.; Samuel, I.D.W. Direct Observation of Intersystem Crossing in a Thermally Activated Delayed Fluorescence Copper Complex in the Solid State. Sci. Adv. 2016, 2, e1500889. [Google Scholar] [CrossRef] [Green Version]
  74. Ochiai, H.; Miura, T.; Ikoma, T.; Minoura, M.; Nakano, H.; Matano, Y. Copper(II) Complexes of 10,20-Diaryl-5,15-Diazaporphyrin: Alternative Synthesis, Excited State Dynamics, and Substituent Effect on the 1O2-Generation Efficiency. BCSJ 2022, 95, 427–432. [Google Scholar] [CrossRef]
  75. Fors, B.P.; Hawker, C.J. Control of a Living Radical Polymerization of Methacrylates by light. Angew. Chem. Int. Ed. 2012, 51, 8850. [Google Scholar] [CrossRef]
  76. Ohtsuki, A.; Goto, A.; Kaji, H. Visible-Light-Induced Reversible Complexation Mediated Living Radical Polymerization of Methacrylates with Organic Catalysts. Macromolecules 2013, 46, 96. [Google Scholar] [CrossRef]
  77. Cope, J.D.; Valle, H.U.; Hall, R.S.; Riley, K.M.; Goel, E.; Biswas, S.; Hendrich, M.P.; Wipf, D.O.; Stokes, S.L.; Emerson, J.P. Tuning the Copper(II)/Copper(I) Redox Potential for More Robust Copper-Catalyzed C–N Bond Forming Reactions. Eur. J. Inorg. Chem. 2020, 2020, 1278–1285. [Google Scholar] [CrossRef]
  78. Tano, T.; Okubo, Y.; Kunishita, A.; Kubo, M.; Sugimoto, H.; Fujieda, N.; Ogura, T.; Itoh, S. Redox Properties of a Mononuclear Copper(II)-Superoxide Complex. Inorg. Chem. 2013, 52, 10431–10437. [Google Scholar] [CrossRef]
  79. Asahi, M.; Yamazaki, S.; Itoh, S.; Ioroi, T. Electrochemical Reduction of Dioxygen by Copper Complexes with Pyridylalkylamine Ligands Dissolved in Aqueous Buffer Solution: The Relationship between Activity and Redox Potential. Dalton Trans. 2014, 43, 10705–10709. [Google Scholar] [CrossRef]
  80. Nishikawa, M.; Kakizoe, D.; Saito, Y.; Ohishi, T.; Tsubomura, T. Redox Properties of Copper(I) Complex Bearing 4,7-Diphenyl-2,9-Dimethyl-1,10-Phenanthroline and 1,4-Bis(Diphenylphosphino)Butane Ligands and Effects of Light in the Presence of Chloroform. BCSJ 2017, 90, 286–288. [Google Scholar] [CrossRef]
  81. Araya, L.M.; Vargas, J.A.; Costamagna, J.A. Ligand Influence on the Redox Properties of Some Copper(II) Complexes with Schiff Bases Derived from Bromosalicylaldehydes and Methyl or Chloro-Substituted Anilines. Transit. Met. Chem. 1986, 11, 312–316. [Google Scholar] [CrossRef]
  82. Das, A.; Ren, Y.; Hessin, C.; Murr, M.D.-E. Copper Catalysis with Redox-Active Ligands. Beilstein J. Org. Chem. 2020, 16, 858–870. [Google Scholar] [CrossRef] [PubMed]
  83. Yang, K.; Yang, X.; Deng, Z.; Zhang, L.; An, J. Copper Piperazine Complex with a High Diffusion Coefficient for Dye-Sensitized Solar Cells. ACS Appl. Energy Mater. 2021, 4, 14004–14013. [Google Scholar] [CrossRef]
  84. Dragonetti, C.; Magni, M.; Colombo, A.; Fagnani, F.; Roberto, D.; Melchiorre, F.; Biagini, P.; Fantacci, S. Towards Efficient Sustainable Full-Copper Dye-Sensitized Solar Cells. Dalton Trans. 2019, 48, 9703–9711. [Google Scholar] [CrossRef]
  85. Giordano, M.; Volpi, G.; Bonomo, M.; Mariani, P.; Garino, C.; Viscardi, G. Methoxy-Substituted Copper Complexes as Possible Redox Mediators in Dye-Sensitized Solar Cells. New J. Chem. 2021, 45, 15303–15311. [Google Scholar] [CrossRef]
  86. Conradie, J. Polypyridyl Copper Complexes as Dye Sensitizer and Redox Mediator for Dye-Sensitized Solar Cells. Electrochem. Commun. 2022, 134, 107182. [Google Scholar] [CrossRef]
  87. Magni, M.; Biagini, P.; Colombo, A.; Dragonetti, C.; Roberto, D.; Valore, A. Versatile Copper Complexes as a Convenient Springboard for Both Dyes and Redox Mediators in Dye Sensitized Solar Cells. Coord. Chem. Rev. 2016, 322, 69–93. [Google Scholar] [CrossRef]
  88. Michaels, H.; Benesperi, I.; Edvinsson, T.; Muñoz-Garcia, A.B.; Pavone, M.; Boschloo, G.; Freitag, M. Copper Complexes with Tetradentate Ligands for Enhanced Charge Transport in Dye-Sensitized Solar Cells. Inorganics 2018, 6, 53. [Google Scholar] [CrossRef] [Green Version]
  89. Rui, H.; Shen, J.; Yu, Z.; Li, L.; Han, H.; Sun, L. Stable Dye-Sensitized Solar Cells Based on Copper(II/I) Redox Mediators Bearing a Pentadentate Ligand. Angew. Chem. Int. Ed. 2021, 60, 16156–16163. [Google Scholar] [CrossRef]
  90. Dumur, F. Recent advances on visible light metal-based photocatalysts for polymerization under low light intensity. Catalysts 2019, 9, 736. [Google Scholar] [CrossRef] [Green Version]
  91. Noirbent, G.; Dumur, F. Recent advances on copper complexes as visible light photoinitiators and (photo)redox initiators of polymerization. Catalysts 2020, 10, 953. [Google Scholar] [CrossRef]
  92. Zivic, N.; Kuroishi, P.K.; Dumur, F.; Gigmes, D.; Dove, A.P.; Sardon, H. Recent advances and challenges in the design of organic photoacid and photobase generators for polymerizations. Angew. Chem. Int. Ed. 2019, 58, 10410–10422. [Google Scholar] [CrossRef] [PubMed]
  93. Tehfe, M.-A.; Lalevée, J.; Dumur, F.; Telitel, S.; Gigmes, D.; Contal, E.; Bertin, D.; Fouassier, J.-P. Zinc-based metal complexes as new photocatalysts in polymerization initiating systems. Eur. Polym. J. 2013, 49, 1040–1049. [Google Scholar] [CrossRef]
  94. Lalevée, J.; Telitel, S.; Xiao, P.; Lepeltier, M.; Dumur, F.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P. Metal and metal free photocatalysts: Mechanistic approach and application as photoinitiators of photopolymerization. Beilstein J. Org. Chem. 2014, 10, 863–876. [Google Scholar] [CrossRef] [Green Version]
  95. Baralle, A.; Fensterbank, L.; Goddard, J.P.; Olivier, C. Aryl Radical Formation by Copper (I) Photocatalyzed Reduction of Diaryliodonium Salts. NMR Evidences for a Cu (II)/Cu (I) Mechanism. Chem. Eur. J. 2013, 19, 10809. [Google Scholar] [CrossRef]
  96. Paria, S.; Reiser, O. Copper in Photocatalysis. ChemCatChem 2014, 6, 2477. [Google Scholar] [CrossRef]
  97. Reiser, O. Shining Light on Copper: Unique Opportunities for Visible-Light-Catalyzed Atom Transfer Radical Addition Reactions and Related Processes. Acc. Chem. Res. 2016, 49, 1990. [Google Scholar] [CrossRef]
  98. Hernandez-Perez, A.C.; Collins, S.K. Heteroleptic Cu-Based Sensitizers in Photoredox Catalysis. Acc. Chem. Res. 2016, 49, 1557. [Google Scholar] [CrossRef]
  99. Hernandez-Perez, A.C.; Collins, S.K. A visible-light-mediated synthesis of carbazoles. Angew. Chem. Int. Ed. 2013, 52, 12696. [Google Scholar] [CrossRef]
  100. Mau, A.; Dietlin, C.; Dumur, F.; Lalevée, J. Concomitant initiation of radical and cationic polymerisations using new copper complexes as photoinitiators: Synthesis and characterisation of acrylate/epoxy interpenetrated polymer networks. Eur. Polym. J. 2021, 152, 110457. [Google Scholar] [CrossRef]
  101. Mau, A.; Noirbent, G.; Dietlin, C.; Graff, B.; Gigmes, D.; Dumur, F.; Lalevée, J. Panchromatic Copper Complexes for Visible Light Photopolymerization. Photochem 2021, 1, 167–189. [Google Scholar] [CrossRef]
  102. Mokbel, H.; Anderson, D.; Plenderleith, R.; Dietlin, C.; Morlet-Savary, F.; Dumur, F.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Simultaneous initiation of radical and cationic polymerization reactions using the “G1” copper complex as photoredox catalyst: Applications of free radical/cationic hybrid photopolymerization in the composites and 3D printing fields. Prog. Org. Coat. 2019, 132, 50. [Google Scholar] [CrossRef]
  103. AL Mousawi, A.; Kermagoret, A.; Versace, D.L.; Toufaily, J.; Hamieh, T.; Graff, B.; Dumur, F.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Copper photoredox catalysts for polymerization upon near UV or visible light: Structure/reactivity/efficiency relationships and use in LED projector 3D printing resins. Polym. Chem. 2017, 8, 568. [Google Scholar] [CrossRef]
  104. Xiao, P.; Dumur, F.; Zhang, J.; Fouassier, J.-P.; Gigmes, D.; Lalevée, J. Copper complexes in radical photoinitiating systems: Applications to free radical and cationic polymerization under visible lights. Macromolecules 2014, 47, 3837–3844. [Google Scholar] [CrossRef]
  105. Matsui, M.; Matsumoto, K.; Shibata, K.; Muramatsu, H. Synthesis and Characterization of 5-Cyano-6-Methyl-2,2′-Bipyridine Metal-Complex Dyes. Dyes Pigments 1992, 18, 47–55. [Google Scholar] [CrossRef]
  106. Matsui, M.; Oji, A.; Hiramatsu, K.; Shibata, K.; Muramatsu, H. Synthesis and Characterization of Fluorescent 4,6-Disubstituted-3-Cyano-2-Methylpyridines. J. Chem. Soc. Perkin Trans. 2 1992, 2, 201–206. [Google Scholar] [CrossRef]
  107. Knorn, M.; Rawner, T.; Czerwieniec, R.; Reiser, O. [Copper(phenanthroline)(bisisonitrile)]+-Complexes for the Visible-Light-Mediated Atom Transfer Radical Addition and Allylation Reactions. ACS Catal. 2015, 5, 5186–5193. [Google Scholar] [CrossRef]
  108. Kaeser, A.; Mohankumar, M.; Mohanraj, J.; Monti, F.; Holler, M.; Cid, J.J.; Moudam, O.; Nierengarten, I.; Karmazin-Brelot, L.; Duhayon, C.; et al. Heteroleptic Copper(I) Complexes Prepared from Phenanthroline and Bis-Phosphine Ligands. Inorg. Chem. 2013, 52, 12140–12151. [Google Scholar] [CrossRef]
  109. Wang, X.; Bai, X.; Su, D.; Zhang, Y.; Li, P.; Lu, S.; Gong, Y.; Zhang, W.; Tang, B. Simultaneous Fluorescence Imaging Reveals N-Methyl-d-aspartic Acid Receptor Dependent Zn2+/H+ Flux in the Brains of Mice with Depression. Anal. Chem. 2020, 92, 4101. [Google Scholar] [CrossRef]
  110. Foresman, J.B.; Frisch, A. Exploring Chemistry with Electronic Structure Methods, 2nd ed.; Gaussian Inc.: Pittsburgh, PA, USA, 1996. [Google Scholar]
  111. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Zakrzewski, V.G.; Montgomery, J.A.; Stratmann, J.R.E.; Burant, J.C.; et al. Gaussian 03, Revision B-2; Gaussian Inc.: Pittsburgh, PA, USA, 2003. [Google Scholar]
  112. Al Mousawi, A.; Dumur, F.; Garra, P.; Toufaily, J.; Hamieh, T.; Goubard, F.; Bui, T.T.; Graff, B.; Gigmes, D.; Fouassier, J.P.; et al. Azahelicenes as visible light photoinitiators for cationic and radical polymerization: Preparation of photoluminescent polymers and use in high performance LED projector 3D printing resins. J. Polym. Sci. A Polym. Chem. 2017, 55, 1189. [Google Scholar] [CrossRef]
  113. Dong, F.; Jian, C.; Zhenghao, F.; Kai, G.; Zuliang, L. Synthesis of Chalcones via Claisen–Schmidt Condensation Reaction Catalyzed by Acyclic Acidic Ionic Liquids. Catal. Commun. 2008, 9, 1924–1927. [Google Scholar] [CrossRef]
  114. Rafiee, E.; Rahimi, F. A Green Approach to the Synthesis of Chalcones via Claisen-Schmidt Condensation Reaction Using Cesium Salts of 12-Tungstophosphoric Acid as a Reusable Nanocatalyst. Monatsh. Chem. 2013, 144, 361–367. [Google Scholar] [CrossRef]
  115. Mousavi, S.R. Claisen–Schmidt Condensation: Synthesis of (1S,6R)/(1R,6S)-2-Oxo-N,4,6-Triarylcyclohex-3-Enecarboxamide Derivatives with Different Substituents in H2O/EtOH. Chirality 2016, 28, 728–736. [Google Scholar] [CrossRef]
  116. Kumar, D.; Suresh; Sandhu, J.S. An Efficient Green Protocol for the Synthesis of Chalcones by a Claisen–Schmidt Reaction Using Bismuth(III)Chloride as a Catalyst under Solvent-Free Condition. Green Chem. Lett. Rev. 2010, 3, 283–286. [Google Scholar] [CrossRef]
  117. Ekanayake, U.G.M.; Weerathunga, H.; Weerasinghe, J.; Waclawik, E.R.; Sun, Z.; MacLeod, J.M.; O’Mullane, A.P.; Ostrikov, K. (Ken) Sustainable Claisen-Schmidt Chalcone Synthesis Catalysed by Plasma-Recovered MgO Nanosheets from Seawater. Sustain. Mater. Technol. 2022, 32, e00394. [Google Scholar] [CrossRef]
  118. Parekh, A.K.; Desai, K.K. Synthesis and Antibacterial Activity of Chalcones and Pyrimidine-2-Ones. E-J. Chem. 1900, 2, 134830. [Google Scholar] [CrossRef] [Green Version]
  119. Jesus, A.R.; Marques, A.P.; Rauter, A.P. An Easy Approach to Dihydrochalcones via Chalcone in Situ Hydrogenation. Pure Appl. Chem. 2016, 88, 349–361. [Google Scholar] [CrossRef]
  120. Julakanti, S.R.; Patel, M.; Ponneri, V. Highly Efficient Synthesis of Chalcones from Poly Carbonyl Aromatic Compounds Using BF3–Et2O via a Regioselective Condensation Reaction. Chem. Pharm. Bull. 2016, 64, 570–576. [Google Scholar] [CrossRef] [Green Version]
  121. Winter, C.; Caetano, J.N.; Araújo, A.B.C.; Chaves, A.R.; Ostroski, I.C.; Vaz, B.G.; Pérez, C.N.; Alonso, C.G. Activated Carbons for Chalcone Production: Claisen-Schmidt Condensation Reaction. Chem. Eng. J. 2016, 303, 604–610. [Google Scholar] [CrossRef]
  122. Qian, H.; Liu, D.; Lv, C. Synthesis of Chalcones via Claisen-Schmidt Reaction Catalyzed by Sulfonic Acid-Functional Ionic Liquids. Ind. Eng. Chem. Res. 2011, 50, 1146–1149. [Google Scholar] [CrossRef]
  123. Calvino, V.; Picallo, M.; López-Peinado, A.J.; Martín-Aranda, R.M.; Durán-Valle, C.J. Ultrasound Accelerated Claisen–Schmidt Condensation: A Green Route to Chalcones. Appl. Surf. Sci. 2006, 252, 6071–6074. [Google Scholar] [CrossRef]
  124. Li, J.-T.; Yang, W.-Z.; Wang, S.-X.; Li, S.-H.; Li, T.-S. Improved Synthesis of Chalcones under Ultrasound Irradiation. Ultrason. Sonochem. 2002, 9, 237–239. [Google Scholar] [CrossRef]
  125. Bui, T.H.; Nguyen, N.T.; Dang, P.H.; Nguyen, H.X.; Nguyen, M.T.T. Design and Synthesis of Chalcone Derivatives as Potential Non-Purine Xanthine Oxidase Inhibitors. SpringerPlus 2016, 5, 1789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Copper complexes used in this work as PIs.
Figure 1. Copper complexes used in this work as PIs.
Polymers 14 01998 g001
Figure 2. Other organic compounds used.
Figure 2. Other organic compounds used.
Polymers 14 01998 g002
Scheme 1. Synthesis and chemical structures of 2-carbonylpyridine-based chalcones.
Scheme 1. Synthesis and chemical structures of 2-carbonylpyridine-based chalcones.
Polymers 14 01998 sch001
Scheme 2. Probable mechanism described by Masaki Matsui.
Scheme 2. Probable mechanism described by Masaki Matsui.
Polymers 14 01998 sch002
Figure 3. UV-visible absorption spectra of (1) Cu1 and (2) Cu2 in DCM.
Figure 3. UV-visible absorption spectra of (1) Cu1 and (2) Cu2 in DCM.
Polymers 14 01998 g003
Figure 4. HOMO and LUMO for Cu1 and Cu2 at the UB3LYP/6–31G* level.
Figure 4. HOMO and LUMO for Cu1 and Cu2 at the UB3LYP/6–31G* level.
Polymers 14 01998 g004
Figure 5. Free radical photopolymerization profiles of acrylate functions vs. irradiation time in: (A) thick samples @455 nm, (B) thin sample @455 nm, (C) thick sample @405 nm and (D) thin sample @405 nm: (1) Cu1/Iod (0.1%/1% w/w), (2) Cu2/Iod (0.1%/1% w/w), (3) Cu1/Iod (0.2%/1% w/w), (4) Cu2/Iod (0.2%/1% w/w), (5) Cu1/Iod/EDB (0.1%/1%/1% w/w/w), (6) Cu2/Iod/EDB (0.1%/1%/1% w/w/w), (7) Cu1/Iod/NPG (0.1%/1%/1% w/w/w), (8) Cu2/Iod/NPG (0.1%/1%/1% w/w/w), (9) Iod/EDB (1%/1% w/w) and (10) Iod/NPG (1%/1% w/w). Irradiation starts at t = 10 s.
Figure 5. Free radical photopolymerization profiles of acrylate functions vs. irradiation time in: (A) thick samples @455 nm, (B) thin sample @455 nm, (C) thick sample @405 nm and (D) thin sample @405 nm: (1) Cu1/Iod (0.1%/1% w/w), (2) Cu2/Iod (0.1%/1% w/w), (3) Cu1/Iod (0.2%/1% w/w), (4) Cu2/Iod (0.2%/1% w/w), (5) Cu1/Iod/EDB (0.1%/1%/1% w/w/w), (6) Cu2/Iod/EDB (0.1%/1%/1% w/w/w), (7) Cu1/Iod/NPG (0.1%/1%/1% w/w/w), (8) Cu2/Iod/NPG (0.1%/1%/1% w/w/w), (9) Iod/EDB (1%/1% w/w) and (10) Iod/NPG (1%/1% w/w). Irradiation starts at t = 10 s.
Polymers 14 01998 g005
Figure 6. (A) Cationic polymerization (CP) profiles of epoxy functions (thin sample) vs. irradiation time upon exposure to LED at 405 nm: (1) Cu1/Iod (0.1%/1% w/w), (2) Cu2/Iod (0.1%/1% w/w), (3) Cu1/Iod/NVK (0.1%/2%/3% w/w/w) and (4) Cu2/Iod/NVK (0.1%/2%/3% w/w/w). (B) IR spectra recorded before and after polymerization. Irradiation starts at t = 10 s.
Figure 6. (A) Cationic polymerization (CP) profiles of epoxy functions (thin sample) vs. irradiation time upon exposure to LED at 405 nm: (1) Cu1/Iod (0.1%/1% w/w), (2) Cu2/Iod (0.1%/1% w/w), (3) Cu1/Iod/NVK (0.1%/2%/3% w/w/w) and (4) Cu2/Iod/NVK (0.1%/2%/3% w/w/w). (B) IR spectra recorded before and after polymerization. Irradiation starts at t = 10 s.
Polymers 14 01998 g006
Figure 7. Photopolymerization profiles of acrylate functions (A) (thick sample), (B) (thin sample)) upon irradiation at 455 nm, and epoxide function ((C) (thick sample), (D) (thin sample)) upon exposure to LED at 405 nm of TA/EPOX using Cu1/Iod/NPG (0.1%/1%/1% w/w/w): (1) 30%/70%, (2) 50%/50%, (3) 70%/30%, and Cu2/Iod/NPG (0.1%/1%/1% w/w/w): (4) 30%/70%, (5) 50%/50%, (6) 70%/30%. The irradiation starts at t = 10 s.
Figure 7. Photopolymerization profiles of acrylate functions (A) (thick sample), (B) (thin sample)) upon irradiation at 455 nm, and epoxide function ((C) (thick sample), (D) (thin sample)) upon exposure to LED at 405 nm of TA/EPOX using Cu1/Iod/NPG (0.1%/1%/1% w/w/w): (1) 30%/70%, (2) 50%/50%, (3) 70%/30%, and Cu2/Iod/NPG (0.1%/1%/1% w/w/w): (4) 30%/70%, (5) 50%/50%, (6) 70%/30%. The irradiation starts at t = 10 s.
Polymers 14 01998 g007
Figure 8. Free radical photopolymerization for the photocomposite synthesis upon near-UV irradiation at 395 nm (4 W·cm−2) using Cu2/Iod/NPG (0.1%/1%/1% w/w/w) in TA [(1)–(4)], (5)) Cu2/Iod/NPG (0.1%/1%/1% w/w/w) in TMPTA, (6) Cu1/Iod/NPG (0.1%/1%/1% w/w/w) in TMPTA and (7) Cu1/Iod/NPG (0.1%/1%/1% w/w/w) in TA.
Figure 8. Free radical photopolymerization for the photocomposite synthesis upon near-UV irradiation at 395 nm (4 W·cm−2) using Cu2/Iod/NPG (0.1%/1%/1% w/w/w) in TA [(1)–(4)], (5)) Cu2/Iod/NPG (0.1%/1%/1% w/w/w) in TMPTA, (6) Cu1/Iod/NPG (0.1%/1%/1% w/w/w) in TMPTA and (7) Cu1/Iod/NPG (0.1%/1%/1% w/w/w) in TA.
Polymers 14 01998 g008
Figure 9. 3D patterns produced by free radical photopolymerization of acrylate functions (TA or TMPTA) using a laser diode at 405 nm and their characterization by numerical optical microscopy: (A) Cu2/Iod/TMA (0.057%/1%/0.46% w/w/w) in TA, (B) Cu1/Iod/TMA (0.058%/0.506%/0.44% w/w/w) in TMPTA; (C) Cu2/Iod/TMA (0.048%/0.65%/0.259% w/w/w) in TA and (D) Cu1/Iod/TMA (0.05%/0.5%/0.305% w/w/w).
Figure 9. 3D patterns produced by free radical photopolymerization of acrylate functions (TA or TMPTA) using a laser diode at 405 nm and their characterization by numerical optical microscopy: (A) Cu2/Iod/TMA (0.057%/1%/0.46% w/w/w) in TA, (B) Cu1/Iod/TMA (0.058%/0.506%/0.44% w/w/w) in TMPTA; (C) Cu2/Iod/TMA (0.048%/0.65%/0.259% w/w/w) in TA and (D) Cu1/Iod/TMA (0.05%/0.5%/0.305% w/w/w).
Polymers 14 01998 g009
Figure 10. Photolysis experiments of (A) Cu1 with Iod salt (10−2 M), (B) Cu2 with Iod salt (10−2 M) and (C) Cu2/Iod/NPG (10−2 M) upon exposure to LED at 405 nm. (D) Consumption percentage of Cu2: (1) alone, (2) with Iod, (3) with Iod/NPG.
Figure 10. Photolysis experiments of (A) Cu1 with Iod salt (10−2 M), (B) Cu2 with Iod salt (10−2 M) and (C) Cu2/Iod/NPG (10−2 M) upon exposure to LED at 405 nm. (D) Consumption percentage of Cu2: (1) alone, (2) with Iod, (3) with Iod/NPG.
Polymers 14 01998 g010
Figure 11. (A) Photolysis of Cu2 with Iod (10−2 M) upon irradiation @375 nm. (B) Consumption of Cu2 with irradiation @375 nm and 405 nm.
Figure 11. (A) Photolysis of Cu2 with Iod (10−2 M) upon irradiation @375 nm. (B) Consumption of Cu2 with irradiation @375 nm and 405 nm.
Polymers 14 01998 g011
Figure 12. (A) Fluorescence quenching of Cu2 by Iod salt, (B) ES1 determination for Cu2, (C) Stern–Volmer coefficient determination of Cu2 quenched by Iod and (D) oxidation potential determination of Cu2.
Figure 12. (A) Fluorescence quenching of Cu2 by Iod salt, (B) ES1 determination for Cu2, (C) Stern–Volmer coefficient determination of Cu2 quenched by Iod and (D) oxidation potential determination of Cu2.
Polymers 14 01998 g012
Scheme 3. Proposed chemical mechanisms.
Scheme 3. Proposed chemical mechanisms.
Polymers 14 01998 sch003
Table 1. Maximum absorption wavelengths (λmax), extinction coefficients at λmax, and molar extinction coefficients for the investigated copper complexes at different emission wavelengths for different LEDs used.
Table 1. Maximum absorption wavelengths (λmax), extinction coefficients at λmax, and molar extinction coefficients for the investigated copper complexes at different emission wavelengths for different LEDs used.
λmax
(nm)
εmax
(M−1∙cm−1)
ε375nm
(M−1∙cm−1)
ε405nm
(M−1∙cm−1)
ε455nm
(M−1∙cm−1)
Cu144514,0007150846013,600
Cu244112,7005760795011,740
Table 2. Final reactive functions conversion (FC) for TA monomer using two or three-component PISs upon exposure at different wavelengths (LED at 405 and 455 nm).
Table 2. Final reactive functions conversion (FC) for TA monomer using two or three-component PISs upon exposure at different wavelengths (LED at 405 and 455 nm).
At 405 nmAt 455 nm
Thick SampleThin SampleThick SampleThin Sample
Cu1/Iod56% a62% b36% a28% b64% a64% b26% a26% b
Cu2/Iod64% a67% b45% a33% b70% a87% b48% a46% b
Cu1/Iod/amine57% c80% d58% c65% d69% c85% d59% c65% d
Cu2/Iod/amine82% c83% d65% c74% d71% c87% d64% c65% d
a: Cu/Iod (0.1%/1% w/w), b: Cu/Iod (0.2%/1% w/w), c: Cu/Iod/EDB (0.1%/1%/1% w/w/w), d: Cu/Iod/NPG (0.1%/1%/1% w/w/w).
Table 3. Final reactive function conversions for EPOX monomer using two and three-component PISs upon irradiation at 405 nm during the polymerization of thin samples.
Table 3. Final reactive function conversions for EPOX monomer using two and three-component PISs upon irradiation at 405 nm during the polymerization of thin samples.
Cu/Iod (0.1%/1% w/w)Cu/Iod/NVK (0.1%/2%/3% w/w/w)
Cu127%49%
Cu216%45%
Table 4. Final conversions of acrylate and epoxides functions for the IPN synthesis of TA/EPOX blend using Cu1 (or Cu2)/Iod/NPG (0.1%/1%/1%/ w/w/w) as PIS upon exposure to visible light at 405 nm.
Table 4. Final conversions of acrylate and epoxides functions for the IPN synthesis of TA/EPOX blend using Cu1 (or Cu2)/Iod/NPG (0.1%/1%/1%/ w/w/w) as PIS upon exposure to visible light at 405 nm.
IPN Synthesis of TA/EPOX Blend Performed in Thick Sample at 405 nmIPN Synthesis of TA/EPOX Blend Performed in Thin Sample at 405 nm
30%/70%50%/50%30%/70%30%/70%50%/50%70%/30%
Cu190%/25%90%/15%93%/27%88%/25%87%/31%84%/47%
Cu299%/30%98%/20%96%/38%92%/22%90%/35%80%/32%
Table 5. Final conversions of acrylate and epoxides functions for the IPN synthesis of TA/EPOX blend using Cu1 (or Cu2)/Iod/NPG (0.1%/1%/1%/ w/w/w) as PIS upon exposure to visible light at 455 nm.
Table 5. Final conversions of acrylate and epoxides functions for the IPN synthesis of TA/EPOX blend using Cu1 (or Cu2)/Iod/NPG (0.1%/1%/1%/ w/w/w) as PIS upon exposure to visible light at 455 nm.
IPN Synthesis of TA/EPOX Blend Performed in Thick Sample @455 nmIPN Synthesis of TA/EPOX Blend Performed in Thin Sample @455 nm
30%/70%50%/50%30%/70%30%/70%50%/50%70%/30%
Cu190%/30%90%/23%90%/41%90%/15%90%/25%90%/15%
Cu2100%/38%98%/55%99%/51%98%/22%99%/30%98%/22%
Table 6. Photocomposite synthesis results using TA as acrylic monomer (or TMPTA) and number of passes to reach the tack-free character on the surfaces.
Table 6. Photocomposite synthesis results using TA as acrylic monomer (or TMPTA) and number of passes to reach the tack-free character on the surfaces.
ThicknessNumber of Passes to Reach Tack-Free Character on the SurfaceNumber of Passes to Reach Tack-Free Character on the Bottom
Cu2/Iod/NPG1.6 mm11
Cu2/Iod/NPG2.9 mm12
Cu2/Iod/NPG4.2 mm16
Cu2/Iod/NPG (TMPTA)7.1 mm125
Cu2/Iod/NPG6.5 mm130
Cu1/Iod/NPG5.25 mm140
Cu1/Iod/NPG (TMPTA)5.1 mm145
Table 7. The tensile strength of IPN materials as a result of hybrid polymerization of the TA/EPOX mixture using Cu/Iod/NPG (0.1%/1%/1% w/w/w) as a photoinitiating system.
Table 7. The tensile strength of IPN materials as a result of hybrid polymerization of the TA/EPOX mixture using Cu/Iod/NPG (0.1%/1%/1% w/w/w) as a photoinitiating system.
Tensile Strength [MPa]0.1%PA/Iod/NPG
TA/EPOX (30% 70%)
@ 395 nm
0.1% PA/Iod/NPG
TA/EPOX (50% 50%)
@ 395 nm
0.1% PA/Iod/NPG
TA/EPOX (70% 30%)
@ 395 nm
Cu16.57.334.3
Cu 27.226.337.2
Table 8. Parameters characterizing the chemical mechanisms between Cu1 (or Cu2) and Iod. For Iod, a reduction potential of −0.7 eV was used for the ∆Get calculations.
Table 8. Parameters characterizing the chemical mechanisms between Cu1 (or Cu2) and Iod. For Iod, a reduction potential of −0.7 eV was used for the ∆Get calculations.
Eox
(V)
ES1
(eV)
ΔG(Cu/Iod)
(eV)
ΚSV(Cu/Iod)Φ(Cu/Iod)Eox
(V)
Cu11.072.34−0.6213.580.55Cu1
Cu20.972.41−0.7461.830.74Cu2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rahal, M.; Noirbent, G.; Graff, B.; Toufaily, J.; Hamieh, T.; Gigmes, D.; Dumur, F.; Lalevée, J. Novel Copper Complexes as Visible Light Photoinitiators for the Synthesis of Interpenetrating Polymer Networks (IPNs). Polymers 2022, 14, 1998. https://doi.org/10.3390/polym14101998

AMA Style

Rahal M, Noirbent G, Graff B, Toufaily J, Hamieh T, Gigmes D, Dumur F, Lalevée J. Novel Copper Complexes as Visible Light Photoinitiators for the Synthesis of Interpenetrating Polymer Networks (IPNs). Polymers. 2022; 14(10):1998. https://doi.org/10.3390/polym14101998

Chicago/Turabian Style

Rahal, Mahmoud, Guillaume Noirbent, Bernadette Graff, Joumana Toufaily, Tayssir Hamieh, Didier Gigmes, Frédéric Dumur, and Jacques Lalevée. 2022. "Novel Copper Complexes as Visible Light Photoinitiators for the Synthesis of Interpenetrating Polymer Networks (IPNs)" Polymers 14, no. 10: 1998. https://doi.org/10.3390/polym14101998

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