Next Article in Journal
1H NMR Spin-Lattice Relaxometry of Cement Pastes with Polycarboxylate Superplasticizers
Previous Article in Journal
Sustainability in the Aerospace, Naval, and Automotive Supply Chain 4.0: Descriptive Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Luminescent 4H-1,2,4-Triazole Derivatives: Synthesis, Molecular Structure and Photophysical Properties

by
Monika Olesiejuk
1,*,
Agnieszka Kudelko
1 and
Marcin Świątkowski
2
1
Department of Chemical Organic Technology and Petrochemistry, The Silesian University of Technology, Krzywoustego 4, PL-44100 Gliwice, Poland
2
Department of X-ray Crystallography and Crystal Chemistry, Institute of General and Ecological Chemistry, Lodz University of Technology, Żeromskiego 116, PL-90924 Łódź, Poland
*
Author to whom correspondence should be addressed.
Materials 2020, 13(24), 5627; https://doi.org/10.3390/ma13245627
Submission received: 15 November 2020 / Revised: 4 December 2020 / Accepted: 8 December 2020 / Published: 10 December 2020

Abstract

:
An alternative approach to the Suzuki cross-coupling reaction is used to synthesize a series of new luminophores based on 4-alkyl-4H-1,2,4-triazole cores conjugated via 1,4-phenylene linker to fused-bicyclic and tricyclic aromatic, or heteroaromatic arrangements. The described methodology allows one to conduct the coupling reaction with the use of commercially available boronic acids in the presence of conventional solvents or ionic liquids and produced excellent yields. It was found that the use of ultrasounds or microwaves significantly accelerates the reaction. The obtained compounds exhibited high luminescent properties and a large quantum yield of emitted photons. The X-ray molecular structures of three highly conjugated 4H-1,2,4-triazole representatives are also presented.

Graphical Abstract

1. Introduction

In recent decades, organic compounds containing both donor and acceptor moieties connected through π-conjugated linkages have been widely investigated due to the possibility of their application in optoelectronic devices [1,2,3], such as organic light-emitting diodes (OLEDs), photovoltaic cells and organic field-effect transistors (OFETs). However, a potential luminophore should contain not only an extended π-conjugated system, but also other features including the proper electron–hole transporting properties, a high external quantum efficiency, as well as thermal and chemical stabilities [4,5]. Heterocyclic arrangements with highly electronegative nitrogen atoms perfectly fit the requirements for such systems. The presence of nitrogen in the aromatic ring considerably affects the electron distribution within the molecule, and also improves the intramolecular electron transport.
1,2,4-Triazoles belong to the vast family of the five-membered heterocyclic arrangements and are especially interesting due to their high nitrogen content. This scaffold occurs in two tautomeric forms with the 1H-1,2,4-triazole system predominating over the 4H-1,2,4-triazole form [6,7]. The derivatives of 4H-1,2,4-triazole exhibit interesting properties and, thus, have applications in diverse fields such as optoelectronics (emission properties) [8,9,10,11,12,13,14], materials science (corrosion inhibition) [15,16,17,18,19,20], medicine [21,22,23,24,25,26,27,28,29,30,31,32] and agriculture [33,34,35,36,37,38] (biological properties). Due to the numerous applications, the 1,2,4-triazole core has been the subject of several recent literature reviews commenting on the synthesis of these moieties [7,39,40,41,42,43]. The most popular methods for the construction of trisubstituted 4H-1,2,4-triazole derivatives are the cyclocondensation of N,N’-dichloromethylidenehydrazine using an appropriate amine, and the cyclodehydration of N-acylamidrazone derivatives from a large variety of precursors depending on the nature of their substituents [7,30,40]. An interesting alternative seems to be the use of the one-pot method in the synthesis of 4H-1,2,4-triazole derivatives [44,45,46].
One of the most effective reactions to form a C–C bond are cross-coupling reactions. Among them, the methodology using a variety of commercially available organoboron compounds is particularly interesting. The Suzuki cross-coupling reaction is typically completed with various combinations of catalysts, bases or solvents, and their mutual integration affects the efficiency and selectivity of the reaction [47,48,49]. Suzuki cross-coupling reaction is well utilized by chemists due to the possibility of using alternative solvents such as ionic liquids (IL), or ultrasound and microwave irradiation as accelerating factors. A combination of these two aspects, such as microwave or ultrasound-assisted reactions using IL, are necessary to improve the environmental impact of chemistry. In the case of alternative solvents, the important topic worth mentioning is the low vapor pressure of the ionic liquid, and the possibility of its regeneration and reuse, which eliminates classic volatile organic solvents. It is possible to reduce energy expenditure by heating chemical reactions using microwaves or by reducing the number of reagents applied when the ionic liquid acts as a catalyst [50,51,52,53,54].
In our previous work, we analyzed the derivatives of five-membered heterocyclic systems such as oxadiazole, thiadiazole, and triazole, with potential applications in optoelectronics [55,56,57,58]. We showed that 1,2,4-triazole derivatives have very good emission properties. Here, we present a synthesis proceeding in an environmentally friendly manner of more extended π-conjugated systems with triazole core hoping for their equally high quantum yields of fluorescence.

2. Materials and Methods

2.1. General Information

The melting points of all compounds were measured on a Stuart SMP3 melting point apparatus (Stuart, Staffordshire, UK). The 1H-NMR and 13C-NMR spectra were performed in CDCl3 solution using tetramethylsilane (TMS) as the internal standard on an Agilent 400-NMR spectrometer (Agilent Technologies, Waldbronn, Germany). X-ray diffraction data were collected on the Synergy Dualflex Pilatus diffractometer (Rigaku, Tokyo, Japan) (for the detailed description of the crystal structure determination, see Supplementary Materials, Section 3). All FT-IR spectra were registered between 4000 and 650 cm−1 using an FT-IR Nicolet 6700 apparatus with a Smart iTR accessory (Thermo Fischer Scientific, Wesel, Germany). UV–Vis spectra were recorded on a Jasco V-660 spectrophotometer (Jasco Corporation, Tokyo, Japan) at room temperature in CH2Cl2 (c = 5 × 10−6 mol/dm3). Fluorescence spectra were recorded at room temperature in CH2Cl2 (c = 5 × 10−6 mol/dm3) using a Jasco F-6300 fluorescence spectrophotometer (Jasco Corporation, Tokyo, Japan). High-resolution mass spectra measurements were completed on a Waters ACQUITY UPLC/Xevo G2QT instrument (Waters Corporation, Milford, MA, USA). Thin-layer chromatography was performed on silica gel 60 F254 Thin-Layer Chromatography (TLC) plates (Merck KGaA, Darmstadt, Germany) using CHCl3/EtOAc (5:1 v/v) as the mobile phase. All reagents from commercial sources were used without additional purification. An aqueous solution of IL: choline hydroxide solution 46 wt.% in H2O (Choline–OH) was used throughout. Ultrasound, for sonication, was generated using an EMAG Technologies Emmi 40HC operating at a frequency 38 kHz, 230 V/50 Hz with bath dimensions of 300 mm × 155 mm × 100 mm (Emag Polska, Juszczyn, Poland). The experiments using microwave radiation were performed in a CEM Discover microwave-enhanced synthesis system (CEM Corporation, Matthews, NC, USA) equipped with a personal computer.

2.2. Synthesis of Compounds

2.2.1. The Synthesis of Suzuki Cross-Coupling Precursors

4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazoles (5a–d) was completed according to the literature procedure [58] using a four-step reaction methodology starting with commercially available 4-bromobenzoic acid (1) (Scheme 1).

2.2.2. A General Procedure for Conventional Suzuki Cross-Coupling Reactions

Synthesis of the 4-alkyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazoles (7–10a).
4-Alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole (5a–d, 1.00 mmol), 4-(N,N-diphenylamino) phenylboronic acid (6a, 2.50 mmol), palladium catalyst Pd(PPh3)4 (0.058 g, 0.05 mmol), a phase transfer catalyst NBu4Br (0.032 g, 0.10 mmol), and base K2CO3 (1.382 g, 10.00 mmol) were added to a combination of toluene (9 mL), H2O (6 mL) and EtOH (3 mL). The mixture was heated under reflux in an oil bath (130 °C) for 7 h (TLC). After cooling, the mixture was transferred to a separating funnel and chloroform (50 mL) was added. After additional extraction with chloroform (2 × 10 mL), the combined organic layers were filtered through a silica gel plug (10 mL). The silica layer was then rinsed with CHCl3/EtOAc (5:1 v/v). The filtrate was dried with MgSO4 and concentrated on the rotary evaporator. The product was precipitated using EtOAc (5 mL), filtered, washed with fresh EtOAc, and air-dried to give pure 4-alkyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazoles (7–10a).

2.2.3. A General Procedure for IL Alternative Approach for Suzuki Cross-Coupling Reactions

Synthesis of the 4-alkyl-3,5-bis(4-arylphenyl)-4H-1,2,4-triazoles 7a–k, 8a, 9a–k, 10a.
4-Alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole (5a–d, 1.00 mmol), an appropriate boronic acid (6a–k, 2.50 mmol), and Pd(PPh3)4 (0.058 g, 0.05 mmol) were added to a mixture of aqueous choline hydroxide solution 46 wt.% (Choline–OH, 10 mL) and toluene (1 mL). The reaction monitored by TLC was carried out using conventional heating—an oil bath (130 °C) or ultrasonication—an ultrasonic bath (80 °C), or microwave irradiation (150 W) in 3–5 cycles of 90 s with 2 min intervals at a temperature 50–100 °C. After cooling, mixture was transferred to a separating funnel and chloroform (50 mL) was added. After the additional extraction with chloroform (2 × 10 mL), the combined chloroform layers were filtered through a silica gel plug (10 mL). The silica layer was then rinsed with CHCl3/EtOAc (5:1 v/v). The filtrate was dried with MgSO4 and concentrated on the rotary evaporator. The product was precipitated using EtOAc (5 mL), filtered, washed with fresh EtOAc, and air-dried to give corresponding 4-alkyl-3,5-bis(4-arylphenyl)-4H-1,2,4-triazoles 7a–k, 8a, 9a–k, 10a.

2.3. Characterization of Compounds

4-Ethyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazole (7a). Yellow solid in 97% yield, 0.357 g, m.p. 268–270 °C; UV (CH2Cl2) λmax 351.0 nm (ε⋅10−3 59.5 cm−1M−1); IR (ATR) ν: 3033, 1731, 1588, 1509, 1483, 1327, 1272, 1219, 1181, 1075, 1028, 821, 772, 749, 694, 666, 621 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.16 (t, J = 7.2 Hz, 3H, CH3), 4.23 (q, J = 7.2 Hz, 2H, CH2), 7.06 (t, J = 7.2 Hz, 4H, ArH), 7.13–7.18 (m, 12H, ArH), 7.28 (t, J = 7.2 Hz, 8H, ArH), 7.54 (d, J = 8.4 Hz, 4H, ArH), 7.71–7.75 (m, 8H, ArH); 13C-NMR (100 MHz, CDCl3): δ 15.8, 40.0, 123.2, 123.6, 124.6, 126.0, 126.9, 127.8, 129.3, 129.4, 133.5, 142.2, 147.5, 147.9, 155.1; HRMS m/z calcd for (C52H41N5 + H+): 736.3440; found: 736.3441.
4-Ethyl-3,5-bis[4-(naphthalen-1-yl)phenyl]-4H-1,2,4-triazole (7b). Beige solid in 88% yield, 0.221 g, m.p. 308–310 °C; UV (CH2Cl2) λmax 297.0 nm (ε⋅10−3 36.4 cm−1M−1); IR (ATR) ν: 3055, 2980, 1591, 1506, 1478, 1425, 1395, 1340, 1250, 1108, 1018, 967, 951, 829, 802, 793, 780, 754, 729 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.29 (t, J = 7.2 Hz, 3H, CH3), 4.36 (q, J = 7.2 Hz, 2H, CH2), 7.47–7.59 (m, 8H, ArH), 7.69 (d, J = 8.4 Hz, 4H, ArH), 7.86 (d, J = 8.4 Hz, 4H, ArH), 7.91–7.96 (m, 6H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.1, 40.1, 125.4, 125.7, 126.0, 126.4, 126.7, 127.0, 128.2, 128.4, 128.9, 130.7, 131.4, 133.8, 139.2, 142.7, 155.3; HRMS m/z calcd for (C36H27N3 + H+): 502.2283; found: 502.2281.
4-Ethyl-3,5-bis[4-(naphthalen-2-yl)phenyl]-4H-1,2,4-triazole (7c). White solid in 99% yield, 0.248 g, m.p. 292–295 °C; UV (CH2Cl2) λmax 268.0 nm (ε⋅10−3 57.3 cm−1M−1) and 305.0 (45.0); IR (ATR) ν: 3051, 1473, 1423, 1401, 1360, 1211, 1080, 1017, 969, 956, 898, 857, 815, 752, 719 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.22 (t, J = 7.2 Hz, 3H, CH3), 4.29 (q, J = 7.2 Hz, 2H, CH2), 7.50–7.57 (m, 4H, ArH), 7.81 (dd, J = 8.4 and 1.6 Hz, 2H, ArH), 7.84 (d, J = 8.8 Hz, 4H, ArH), 7.88–7.95 (m, 8H, ArH), 7.97 (d, J = 8.4 Hz, 2H, ArH), 8.13 (d, J = 1.6 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 125.2, 126.1, 126.3, 126.5, 126.7, 127.7, 127.9, 128.3, 128.7, 129.4, 132.9, 133.6, 137.4, 142.8, 155.2; HRMS m/z calcd for (C36H27N3 + H+): 502.2283; found: 502.2287.
4-Ethyl-3,5-bis[4-(quinolin-3-yl)phenyl]-4H-1,2,4-triazole (7d). White solid in 79% yield, 0.198 g, m.p. 296–298 °C; UV (CH2Cl2) λmax 269.0 nm (ε⋅10−3 52.8 cm−1M−1) and 305.0 (42.6); IR (ATR) ν: 3069, 3002, 2964, 1494, 1477, 1468, 1453, 1379, 1358, 1543, 1202, 1081, 1019, 963, 952, 931, 912, 868, 856, 841, 795, 781, 767, 761, 744, 738, 726 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.23 (t, J = 7.2 Hz, 3H, CH3), 4.30 (q, J = 7.2 Hz, 2H, CH2), 7.62 (t, J = 8.0 Hz, 2H, ArH), 7.77 (t, J = 8.0 Hz, 2H, ArH), 7.88–7.94 (m, 10H, ArH), 8.18 (d, J = 8.0 Hz, 2H, ArH), 8.40 (d, J = 2.4 Hz, 2H, ArH), 9.25 (d, J = 2.4 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 127.3, 127.4, 127.8, 127.9, 128.1, 129.3, 129.7, 129.8, 132.7, 133.6, 139.7, 147.7, 149.5, 155.0; HRMS m/z calcd for (C34H25N5 + H+): 504.2188; found: 504.2188.
4-Ethyl-3,5-bis[4-(quinolin-6-yl)phenyl]-4H-1,2,4-triazole (7e). Pearl solid in 94% yield, 0.236 g, m.p. 302–305 °C; UV (CH2Cl2) λmax 269.0 nm (ε⋅10−3 56.4 cm−1M−1) and 306.0 (42.0); IR (ATR) ν: 3052, 1593, 1570, 1499, 1474, 1424, 1347, 1326, 1121, 1014, 971, 892, 868, 856, 831, 797, 781, 754 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.22 (t, J = 7.2 Hz, 3H, CH3), 4.30 (q, J = 7.2 Hz, 2H, CH2), 7.47 (dd, J = 8.4 and 4.4 Hz, 2H, ArH), 7.85 (d, J = 8.4 Hz, 4H, ArH), 7.91 (d, J = 8.4 Hz, 4H, ArH), 8.05 (dd, J = 8.4 and 2.0 Hz, 2H, ArH), 8.09 (d, J = 2.0 Hz, 2H, ArH), 8.22–8.26 (m, 4H, ArH), 8.96 (dd, J = 4.4 and 2.0 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 121.7, 125.8, 127.0, 128.0, 128.5, 128.9, 129.5, 130.2, 136.3, 138.1, 142.0, 147.9, 150.8, 155.1; HRMS m/z calcd for (C34H25N5 + H+): 504.2188; found: 504.2189.
3,5-Bis[4-(dibenzothiophen-4-yl)phenyl]-4-ethyl-4H-1,2,4-triazole (7f). Creamy solid in 96% yield, 0.296 g, m.p. 281–283 °C; UV (CH2Cl2) λmax 240.0 nm (ε⋅10−3 85.0 cm−1M−1) and 290.0 (48.5); IR (ATR) ν: 3062, 1470, 1440, 1425, 1391, 1376, 1360, 1246, 1112, 1016, 863, 836, 800, 753, 743, 730, 718, 704 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.28 (t, J = 7.2 Hz, 3H, CH3), 4.35 (q, J = 7.2 Hz, 2H, CH2), 7.48–7.51 (m, 4H, ArH), 7.56 (d, J = 7.6 Hz, 2H, ArH), 7.61 (t, J = 7.6 Hz, 2H, ArH), 7.86 (d, J = 7.6 Hz, 2H, ArH), 7.90 (d, J = 8.4 Hz, 4H, ArH), 7.95 (d, J = 8.4 Hz, 4H, ArH), 8.20–8.23 (m, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 121.0, 121.8, 122.7, 124.5, 125.2, 127.0, 127.3, 128.9, 129.4, 135.7, 135.9, 136.5, 138.5, 139.4, 142.4, 152.2, 155.1; HRMS m/z calcd for (C40H27N3S2 + H+): 614.1725; found: 614.1727.
3,5-Bis[4-(dibenzofuran-4-yl)phenyl]-4-ethyl-4H-1,2,4-triazole (7g). Creamy solid in 92% yield, 0.267 g, m.p. 255–256 °C; UV (CH2Cl2) λmax 286.0 nm (ε⋅10−3 70.5 cm−1M−1) and 315.0 (41.0); IR (ATR) ν: 3055, 2949, 1586, 1479, 1450, 1430, 1412, 1395, 1360, 1259, 1189, 1101, 1058, 1014, 968, 874, 838, 801, 791, 741, 729 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.27 (t, J = 7.2 Hz, 3H, CH3), 4.35 (q, J = 7.2 Hz, 2H, CH2), 7.39 (t, J = 8.0 Hz, 2H, ArH), 7.46–7.52 (m, 4H, ArH), 7.65 (d, J = 8.0 Hz, 2H, ArH), 7.70 (d, J = 8.0 Hz, 2H, ArH), 7.91 (d, J = 8.4 Hz, 4H, ArH), 7.99–8.03 (m, 4H, ArH), 8.14 (d, J = 8.4 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.1, 40.1, 111.9, 120.3, 120.8, 123.0, 123.4, 124.1, 124.7, 125.2, 126.8, 127.0, 127.4, 129.2, 129.3, 138.2, 153.3, 155.2, 156.2; HRMS m/z calcd for (C40H27N3O2 + H+): 582.2182; found: 582.2183.
4-Ethyl-3,5-bis[4-(9-methyl-9H-carbazol-3-yl)phenyl]-4H-1,2,4-triazole (7h). Creamy solid in 70% yield, 0.107 g, m.p. 313–315 °C; UV (CH2Cl2) λmax 300.0 nm (ε⋅10−3 69.9 cm−1M−1) and 321.0 (46.0); IR (ATR) ν: 3054, 2990, 2935, 1598, 1483, 1466, 1424, 1359, 1337, 1323, 1294, 1254, 1221, 1120, 1012, 844, 825, 804, 764, 750, 739, 719 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.22 (t, J = 7.2 Hz, 3H, CH3), 3.92 (s, 6H, CH3), 4.29 (q, J = 7.2 Hz, 2H, CH2), 7.29 (t, J = 7.6 Hz, 2H, ArH), 7.45 (d, J = 7.6 Hz, 2H, ArH), 7.50–7.55 (m, 4H, ArH), 7.80–7.83 (m, 6H, ArH), 7.91 (d, J = 8.4 Hz, 4H, ArH), 8.18 (d, J = 7.6 Hz, 2H, ArH), 8.40 (s, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 15.9, 29.3, 40.0, 108.7, 108.9, 118.9, 119.2, 120.4, 122.9, 123.4, 125.1, 125.7, 126.1, 127.6, 129.4, 131.1, 140.9, 141.5, 143.8, 155.3; HRMS m/z calcd for (C42H33N5 + H+): 608.2814; found: 608.2817.
4-Ethyl-3,5-bis[4-(9-ethyl-9H-carbazol-3-yl)phenyl]-4H-1,2,4-triazole (7i). Creamy solid in 76% yield, 0.121 g, m.p. 232–234 °C; UV (CH2Cl2) λmax 243.0 nm (ε⋅10−3 54.6 cm−1M−1), 301.0 (76.5) and 322.0 (49.9); IR (ATR) ν: 3047, 2977, 1597, 1459, 1379, 1346, 1334, 1295, 1254, 1233, 1196, 1154, 1125, 1086, 848, 806, 769, 747, 729, 695 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.22 (t, J = 7.2 Hz, 3H, CH3), 1.49 (t, J = 7.2 Hz, 6H, CH3), 4.29 (q, J = 7.2 Hz, 2H, CH2), 4.43 (q, J = 7.2 Hz, 4H, CH2), 7.28 (t, J = 7.6 Hz, 2H, ArH), 7.45 (d, J = 7.6 Hz, 2H, ArH), 7.49–7.53 (m, 4H, ArH), 7.79 (d, J = 7.6 Hz, 2H, ArH), 7.83 (d, J = 8.4 Hz, 4H, ArH), 7.90 (d, J = 8.4 Hz, 4H, ArH), 8.18 (d, J = 7.6 Hz, 2H, ArH), 8.40 (s, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.8, 15.9, 37.7, 40.0, 108.7, 108.9, 119.0, 119.1, 120.5, 123.0, 123.6, 125.1, 125.7, 126.0, 127.6, 129.3, 131.0, 139.8, 140.5, 143.8, 155.3; HRMS m/z calcd for (C44H37N5 + H+): 636.3127; found: 636.3124.
3,5-Bis[4′-(9H-carbazol-9-yl)biphenyl-4-yl]-4-ethyl-4H-1,2,4-triazole (7j). Grey solid in 77% yield, 0.141 g, m.p. 277–280 °C; UV (CH2Cl2) λmax 293.0 nm (ε⋅10−3 50.0 cm−1M−1) and 322.0 (41.7); IR (ATR) ν: 3055, 1740, 1599, 1521, 1477, 1450, 1414, 1367, 1334, 1318, 1303, 1227, 1184, 1170, 1119, 1004, 969, 825, 742, 719 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.25 (t, J = 7.2 Hz, 3H, CH3), 4.32 (q, J = 7.2 Hz, 2H, CH2), 7.32 (t, J = 7.6 Hz, 4H, ArH), 7.45 (t, J = 7.6 Hz, 4H, ArH), 7.51 (d, J = 7.6 Hz, 4H, ArH), 7.72 (d, J = 8.4 Hz, 4H, ArH), 7.86–7.92 (m, 12H, ArH), 8.19 (d, J = 7.6 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 109.8, 120.1, 120.4, 123.5, 126.0, 126.9, 127.5, 127.6, 128.6, 129.5, 137.6, 139.0, 140.8, 142.0, 155.1; HRMS m/z calcd for (C52H37N5 + H+): 732.3127; found: 732.3126.
4-Ethyl-3,5-bis[4-(thiantren-1-yl)phenyl]-4H-1,2,4-triazole (7k). Grey solid in 87% yield, 0.585 g, m.p. 198–200 °C; UV (CH2Cl2) λmax 263.0 nm (ε⋅10−3 53.9 cm−1M−1); IR (ATR) ν: 3051, 1475, 1441, 1425, 1396, 1190, 1107, 1017, 967, 853, 842, 787, 748, 728, 686, 664, 653, 604 cm−1; 1H-NMR (400 MHz, CDCl3): δ 1.30 (t, J = 7.2 Hz, 3H, CH3), 4.35 (q, J = 7.2 Hz, 2H, CH2), 7.20 (t, J = 7.6 Hz, 2H, ArH), 7.26 (t, J = 7.6 Hz, 2H, ArH), 7.30–7.36 (m, 4H, ArH), 7.39 (d, J = 7.6 Hz, 2H, ArH), 7.51 (d, J = 7.6 Hz, 2H, ArH), 7.57 (d, J = 7.6 Hz, 2H, ArH), 7.60 (d, J = 8.4 Hz, 4H, ArH), 7.84 (d, J = 8.4 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 16.0, 40.1, 127.2, 127.7, 127.9, 128.3, 128.6, 128.6, 128.7, 128.9, 129.1, 130.1, 135.0, 135.6, 135.9, 136.2, 141.5, 142.0, 155.2; HRMS m/z calcd for (C40H27N3S4 + H+): 678.1166; found: 678.1166.
3,5-Bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4-propyl-4H-1,2,4-triazole (8a). Yellow solid in 93% yield, 0.347 g, m.p. 189–192 °C; UV (CH2Cl2) λmax 352.0 nm (ε⋅10−3 51.8 cm−1M−1); IR (ATR) ν: 3033, 2958, 1735, 1587, 1509, 1486, 1327, 1272, 1241, 1179, 1046, 1004, 862, 843, 819, 773, 753, 737, 693 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.66 (t, J = 7.2 Hz, 3H, CH3), 1.49 (sext, J = 7.2 Hz, 2H, CH2), 4.13 (t, J = 7.2 Hz, 2H, CH2), 7.06 (t, J = 7.2 Hz, 4H, ArH), 7.14–7.18 (m, 12H, ArH), 7.28 (t, J = 7.2 Hz, 8H, ArH), 7.55 (d, J = 8.4 Hz, 4H, ArH), 7.71–7.75 (m, 8H, ArH); 13C-NMR (100 MHz, CDCl3): δ 10.7, 23.4, 46.6, 123.2, 123.6, 124.6, 126.1, 126.9, 127.7, 129.3, 129.4, 133.5, 142.1, 147.5, 147.8, 155.5; HRMS m/z calcd for (C53H43N5 + H+): 750.3597; found: 750.3588.
4-Butyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazole (9a). Creamy solid in 99% yield, 0.382 g, m.p. 208–209 °C; UV (CH2Cl2) λmax 351.0 nm (ε⋅10−3 52.5 cm−1M−1); IR (ATR) ν: 3034, 2959, 2864, 1735, 1586, 1487, 1326, 1273, 1239, 1178, 1046, 1004, 974, 861, 843, 821, 754, 737, 693 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.68 (t, J = 7.6 Hz, 3H, CH3), 1.06 (sext, J = 7.6 Hz, 2H, CH2), 1.43 (quin, J = 7.6 Hz, 2H, CH2), 4.17 (t, J = 7.6 Hz, 2H, CH2), 7.06 (t, J = 7.2 Hz, 4H, ArH), 7.14–7.18 (m, 12H, ArH), 7.29 (t, J = 7.2 Hz, 8H, ArH), 7.54 (d, J = 8.4 Hz, 4H, ArH), 7.71–7.75 (m, 8H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.2, 19.3, 32.0, 44.8, 123.2, 123.6, 124.6, 126.1, 126.9, 127.7, 129.3, 129.4, 133.5, 142.1, 147.5, 147.9, 155.4; HRMS m/z calcd for (C54H45N5 + H+): 764.3753; found: 764.3749.
4-Butyl-3,5-bis[4-(naphthalen-1-yl)phenyl]-4H-1,2,4-triazole (9b). Grey solid in 62% yield, 0.165 g, m.p. 217–218 °C; UV (CH2Cl2) λmax 296.0 nm (ε⋅10−3 40.6 cm−1M−1); IR (ATR) ν: 3058, 2928, 2869, 1472, 1429, 1393, 1336, 1183, 1020, 965, 952, 858, 845, 831, 799, 791, 774, 752, 724 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.77 (t, J = 7.6 Hz, 3H, CH3), 1.17 (sext, J = 7.6 Hz, 2H, CH2), 1.57 (quin, J = 7.6 Hz, 2H, CH2), 4.30 (t, J = 7.6 Hz, 2H, CH2), 7.47–7.59 (m, 8H, ArH), 7.69 (d, J = 8.4 Hz, 4H, ArH), 7.86 (d, J = 8.4 Hz, 4H, ArH), 7.90–7.96 (m, 6H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.2, 44.8, 125.4, 125.7, 126.0, 126.4, 126.8, 127.0, 128.2, 128.4, 128.9, 130.7, 131.4, 133.8, 139.2, 142.7, 155.5; HRMS m/z calcd for (C38H31N3 + H+): 530.2596; found: 530.2596.
4-Butyl-3,5-bis[4-(naphthalen-2-yl)phenyl]-4H-1,2,4-triazole (9c). Creamy solid in 89% yield, 0.236 g, m.p. 331–334 °C; UV (CH2Cl2) λmax 269.0 nm (ε⋅10−3 73.8 cm−1M−1) and 305.0 (58.9); IR (ATR) ν: 3050, 2954, 2872, 1597, 1489, 1465, 1402, 1368, 1211, 1016, 972, 951, 895, 841, 813, 751, 717 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.72 (t, J = 7.6 Hz, 3H, CH3), 1.10 (sext, J = 7.6 Hz, 2H, CH2), 1.50 (quin, J = 7.6 Hz, 2H, CH2), 4.24 (t, J = 7.6 Hz, 2H, CH2), 7.51–7.57 (m, 4H, ArH), 7.81–7.85 (m, 6H, ArH), 7.89–7.95 (m, 8H, ArH), 7.97 (d, J = 8.8 Hz, 2H, ArH), 8.14 (d, J = 1.2 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.1, 44.9, 125.2, 126.1, 126.3, 126.5, 126.8, 127.7, 127.9, 128.3, 128.7, 129.4, 132.9, 133.6, 137.4, 142.7, 155.4; HRMS m/z calcd for (C38H31N3 + H+): 530.2596; found: 530.2598.
4-Butyl-3,5-bis[4-(quinolin-3-yl)phenyl]-4H-1,2,4-triazole (9d). Grey solid in 99% yield, 0.263 g, m.p. 310–313 °C; UV (CH2Cl2) λmax 270.0 nm (ε⋅10−3 37.5 cm−1M−1) and 304.0 (30.4); IR (ATR) ν: 3060, 2961, 2918, 2862, 1574, 1489, 1471, 1436, 1363, 1208, 1018, 972, 951, 906, 861, 844, 833, 783, 745, 724 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.73 (t, J = 7.6 Hz, 3H, CH3), 1.11 (sext, J = 7.6 Hz, 2H, CH2), 1.50 (quin, J = 7.6 Hz, 2H, CH2), 4.26 (t, J = 7.6 Hz, 2H, CH2), 7.63 (t, J = 8.0 Hz, 2H, ArH), 7.77 (t, J = 8.0 Hz, 2H, ArH), 7.88–7.95 (m, 10H, ArH), 8.18 (d, J = 8.0 Hz, 2H, ArH), 8.41 (d, J = 2.4 Hz, 2H, ArH), 9.26 (d, J = 2.4 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.2, 45.0, 127.3, 127.5, 127.8, 127.9, 128.1, 129.3, 129.7, 129.8, 132.7, 133.6, 139.6, 147.7, 149.5, 155.3; HRMS m/z calcd for (C36H29N5 + H+): 532.2501; found: 532.2507.
4-Butyl-3,5-bis[4-(quinolin-6-yl)phenyl]-4H-1,2,4-triazole (9e). White solid in 91% yield, 0.242 g, m.p. 268–270 °C; UV (CH2Cl2) λmax 271.0 nm (ε⋅10−3 52.9 cm−1M−1) and 306.0 (39.8); IR (ATR) ν: 2954, 2872, 1593, 1571, 1481, 1351, 1127, 975, 886, 870, 827, 792, 776, 743, 716 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.72 (t, J = 7.6 Hz, 3H, CH3), 1.11 (sext, J = 7.6 Hz, 2H, CH2), 1.50 (quin, J = 7.6 Hz, 2H, CH2), 4.25 (t, J = 7.6 Hz, 2H, CH2), 7.47 (dd, J = 8.4 and 4.4 Hz, 2H, ArH), 7.86 (d, J = 8.4 Hz, 4H, ArH), 7.92 (d, J = 8.4 Hz, 4H, ArH), 8.05 (dd, J = 8.4 and 2.0 Hz, 2H, ArH), 8.10 (d, J = 2.0 Hz, 2H, ArH), 8.22–8.27 (m, 4H, ArH), 8.96 (dd, J = 4.4 and 2.0 Hz, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.1, 44.9, 121.7, 125.8, 127.2, 127.9, 128.5, 128.9, 129.5, 130.2, 136.3, 138.1, 141.9, 147.9, 150.8, 155.4; HRMS m/z calcd for (C36H29N5 + H+): 532.2501; found: 532.2504.
3,5-Bis[4-(dibenzothiophen-4-yl)phenyl]-4-butyl-4H-1,2,4-triazole (9f). Creamy solid in 93% yield, 0.299 g, m.p. 256–258 °C; UV (CH2Cl2) λmax 241.0 nm (ε⋅10−3 40.5 cm−1M−1) and 290.0 (22.4); IR (ATR) ν: 3056, 2959, 1472, 1441, 1377, 1249, 1101, 1048, 1020, 973, 846, 826, 804, 796, 755, 742, 727, 704 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.76 (t, J = 7.6 Hz, 3H, CH3), 1.16 (sext, J = 7.6 Hz, 2H, CH2), 1.55 (quin, J = 7.6 Hz, 2H, CH2), 4.29 (t, J = 7.6 Hz, 2H, CH2), 7.48–7.51 (m, 4H, ArH), 7.56 (d, J = 7.6 Hz, 2H, ArH), 7.61 (t, J = 7.6 Hz, 2H, ArH), 7.86 (d, J = 7.6 Hz, 2H, ArH), 7.89 (d, J = 8.4 Hz, 4H, ArH), 7.95 (d, J = 8.4 Hz, 4H, ArH), 8.20–8.23 (m, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.2, 44.9, 121.0, 121.8, 122.7, 124.5, 125.2, 127.0, 127.4, 128.9, 129.4, 135.7, 135.9, 136.5, 138.5, 139.4, 142.4, 152.2, 155.4; HRMS m/z calcd for (C42H31N3S2 + H+): 642.2038; found: 642.2034.
3,5-Bis[4-(dibenzofuran-4-yl)phenyl]-4-butyl-4H-1,2,4-triazole (9g). Creamy solid in 88% yield, 0.268 g, m.p. 228–230 °C; UV (CH2Cl2) λmax 286.0 nm (ε⋅10−3 57.0 cm−1M−1) and 317.0 (30.6); IR (ATR) ν: 3049, 2960, 1737, 1471, 1450, 1430, 1411, 1386, 1237, 1220, 1192, 1174, 1063, 1014, 975, 854, 841, 812, 792, 772, 741 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.75 (t, J = 7.6 Hz, 3H, CH3), 1.15 (sext, J = 7.6 Hz, 2H, CH2), 1.55 (quin, J = 7.6 Hz, 2H, CH2), 4.30 (t, J = 7.6 Hz, 2H, CH2), 7.39 (t, J = 8.0 Hz, 2H, ArH), 7.46–7.53 (m, 4H, ArH), 7.65 (d, J = 8.0 Hz, 2H, ArH), 7.71 (d, J = 8.0 Hz, 2H, ArH), 7.91 (d, J = 8.4 Hz, 4H, ArH), 7.99–8.03 (m, 4H, ArH), 8.14 (d, J = 8.4 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.2, 45.0, 111.9, 120.3, 120.8, 122.9, 123.3, 124.1, 124.7, 125.2, 126.8, 127.1, 127.4, 129.2, 129.3, 138.1, 153.4, 155.5, 156.2; HRMS m/z calcd for (C42H31N3O2 + H+): 610.2495; found: 610.2491.
4-Butyl-3,5-bis[4-(9-methyl-9H-carbazol-3-yl)phenyl]-4H-1,2,4-triazole (9h). Creamy solid in 97% yield, 0.308 g, m.p. 267–268 °C; UV (CH2Cl2) λmax 300.0 nm (ε⋅10−3 76.8 cm−1M−1) and 323.0 (51.0); IR (ATR) ν: 3047, 2957, 2934, 1597, 1483, 1466, 1424, 1360, 1335, 1323, 1294, 1254, 1221, 1154, 1121, 1012, 843, 825, 813, 799, 772, 749, 739, 721 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.73 (t, J = 7.6 Hz, 3H, CH3), 1.11 (sext, J = 7.6 Hz, 2H, CH2), 1.50 (quin, J = 7.6 Hz, 2H, CH2), 3.91 (s, 6H, CH3), 4.24 (t, J = 7.6 Hz, 2H, CH2), 7.28 (t, J = 7.6 Hz, 2H, ArH), 7.44 (d, J = 7.6 Hz, 2H, ArH), 7.49–7.54 (m, 4H, ArH), 7.80–7.82 (m, 6H, ArH), 7.90 (d, J = 8.4 Hz, 4H, ArH), 8.18 (d, J = 7.6 Hz, 2H, ArH), 8.40 (s, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 29.3, 32.1, 44.9, 108.7, 108.9, 118.9, 119.2, 120.4, 122.9, 123.4, 125.1, 125.8, 126.1, 127.6, 129.3, 131.1, 140.9, 141.5, 143.7, 155.6; HRMS m/z calcd for (C44H37N5 + H+): 636.3127; found: 636.3123.
4-Butyl-3,5-bis[4-(9-ethyl-9H-carbazol-3-yl)phenyl]-4H-1,2,4-triazole (9i). Creamy solid in 98% yield, 0.326 g, m.p. 269–270 °C; UV (CH2Cl2) λmax 243.0 nm (ε⋅10−3 51.1 cm−1M−1), 301.0 (70.7) and 323.0 (47.1); IR (ATR) ν: 3049, 2973, 1740, 1598, 1460, 1379, 1346, 1325, 1295, 1252, 1233, 1197, 1154, 1125, 1087, 848, 804, 769, 747, 730, 694 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.72 (t, J = 7.6 Hz, 3H, CH3), 1.11 (sext, J = 7.6 Hz, 2H, CH2), 1.46–1.53 (m, 8H, CH2 + CH3), 4.24 (t, J = 7.6 Hz, 2H, CH2), 4.42 (q, J = 7.2 Hz, 4H, CH2), 7.28 (t, J = 7.6 Hz, 2H, ArH), 7.45 (d, J = 7.6 Hz, 2H, ArH), 7.49–7.53 (m, 4H, ArH), 7.78–7.82 (m, 6H, ArH), 7.90 (d, J = 8.4 Hz, 4H, ArH), 8.18 (d, J = 7.6 Hz, 2H, ArH), 8.40 (s, 2H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 13.8, 19.4, 32.1, 37.7, 44.9, 108.7, 108.8, 119.0, 119.1, 120.5, 123.0, 123.6, 125.1, 125.8, 126.0, 127.6, 129.3, 131.0, 139.8, 140.4, 143.7, 155.6; HRMS m/z calcd for (C46H41N5 + H+): 664.3440; found: 664.3439.
4-Butyl-3,5-bis[4′-(9H-carbazol-9-yl)biphenyl-4-yl]-4H-1,2,4-triazole (9j). Grey solid in 91% yield, 0.344 g, m.p. 320–323 °C; UV (CH2Cl2) λmax 293.0 nm (ε⋅10−3 50.5 cm−1M−1) and 320.0 (42.6); IR (ATR) ν: 3054, 2954, 1599, 1521, 1490, 1478, 1451, 1413, 1367, 1334, 1317, 1303, 1226, 1185, 1171, 1004, 974, 826, 771, 743, 719 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.74 (t, J = 7.6 Hz, 3H, CH3), 1.13 (sext, J = 7.6 Hz, 2H, CH2), 1.52 (quin, J = 7.6 Hz, 2H, CH2), 4.26 (t, J = 7.6 Hz, 2H, CH2), 7.32 (t, J = 7.6 Hz, 4H, ArH), 7.45 (t, J = 7.6 Hz, 4H, ArH), 7.50 (d, J = 7.6 Hz, 4H, ArH), 7.71 (d, J = 8.4 Hz, 4H, ArH), 7.85–7.89 (m, 8H, ArH), 7.92 (d, J = 8.4 Hz, 4H, ArH), 8.18 (d, J = 7.6 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.4, 32.1, 44.9, 109.8, 120.1, 120.4, 123.5, 126.0, 127.0, 127.5, 127.6, 128.6, 129.5, 137.6, 139.0, 140.8, 141.9, 155.4; HRMS m/z calcd for (C54H41N5 + H+): 760.3440; found: 760.3438.
4-Butyl-3,5-bis[4-(thiantren-1-yl)phenyl]-4H-1,2,4-triazole (9k). Creamy solid in 86% yield, 0.305 g, m.p. 264–266 °C; UV (CH2Cl2) λmax 263.0 nm (ε⋅10−3 66.7 cm−1M−1); IR (ATR) ν: 3048, 2952, 2870, 1474, 1441, 1397, 1247, 1188, 1109, 1018, 973, 843, 807, 788, 748, 728, 686 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.78 (t, J = 7.6 Hz, 3H, CH3), 1.20 (sext, J = 7.6 Hz, 2H, CH2), 1.57 (quin, J = 7.6 Hz, 2H, CH2), 4.29 (t, J = 7.6 Hz, 2H, CH2), 7.21 (t, J = 7.6 Hz, 2H, ArH), 7.26 (t, J = 7.6 Hz, 2H, ArH), 7.31–7.34 (m, 4H, ArH), 7.37 (d, J = 7.6 Hz, 2H, ArH), 7.52 (d, J = 7.6 Hz, 2H, ArH), 7.57 (d, J = 7.6 Hz, 2H, ArH), 7.60 (d, J = 8.4 Hz, 4H, ArH), 7.83 (d, J = 8.4 Hz, 4H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.3, 19.3, 32.1, 44.8, 127.2, 127.7, 127.9, 128.3, 128.6, 128.7, 128.8, 128.9, 129.1, 130.1, 135.1, 135.7, 135.9, 136.2, 141.5, 141.9, 155.4; HRMS m/z calcd for (C42H31N3S4 + H+): 706.1479; found: 706.1474.
4-Hexyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazole (10a). Yellow solid in 61% yield, 0.240 g, m.p. 244–246 °C; UV (CH2Cl2) λmax 352.0 nm (ε⋅10−3 68.0 cm−1M−1); IR (ATR) ν: 3033, 2925, 2855, 1586, 1508, 1484, 1411, 1328, 1278, 1203, 1173, 1121, 1076, 1025, 1005, 968, 822, 748, 729, 691 cm−1; 1H-NMR (400 MHz, CDCl3): δ 0.73 (t, J = 7.2 Hz, 3H, CH3), 0.99–1.02 (m, 4H, 2 × CH2), 1.08 (quin, J = 7.2 Hz, 2H, CH2), 1.44 (quin, J = 7.2 Hz, 2H, CH2), 4.16 (t, J = 7.2 Hz, 2H, CH2), 7.06 (t, J = 7.2 Hz, 4H, ArH), 7.13–7.18 (m, 12H, ArH), 7.29 (t, J = 7.2 Hz, 8H, ArH), 7.55 (d, J = 8.4 Hz, 4H, ArH), 7.71–7.75 (m, 8H, ArH); 13C-NMR (100 MHz, CDCl3): δ 13.8, 22.2, 25.7, 29.9, 30.7, 45.0, 123.2, 123.6, 124.6, 126.1, 126.9, 127.8, 129.3, 129.4, 133.5, 142.1, 147.5, 147.9, 155.5; HRMS m/z calcd for (C56H49N5 + H+): 792.4066; found: 792.4061.

3. Results and Discussion

Four basic 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole (5a–d) substrates were synthesized starting from 4-bromobenzoic acid (1), using our previously elaborated methodologies [58] (Scheme 1). For each reaction, triazole precursors (5a–d) containing substituents, differing in their alkyl chain length at the nitrogen atom (position 4), were coupled with 4-(N,N-diphenylamino)phenylboronic acid (6a) in a Suzuki reaction. Transformations were completed using a two-phase solvent system with conventional heating in an oil bath over a sufficiently long period (7 h, TLC) using Pd(PPh3)4 as a catalyst, and potassium carbonate as a base which gives the desired products in high yields (Scheme 2).
Three (8a10a) of the final products form single crystals in the solid-state (Figure 1), enabling the determination of their molecular structures (Table S1). Compounds 8a and 9a crystallize as ethanol solvates (Figure S1), while 10a forms a solventless compound. This is caused by the presence of an n-hexyl substituent on the triazole ring of 10a. The n-hexyl terminal methyl and methylene atoms (-CH2–CH3) are located in the space occupied by ethanol molecules in 8a and 9a (Figures S2–S4). The shorter n-butyl and n-propyl substituents cannot fill the crystal net space and leave empty voids with volumes of 153–216 Å3, which are accessible for solvent molecules. Due to discrepancy between the volume of the solvent molecules (~97 Å3 for one ethanol molecule) and voids, these molecules are disordered in the crystal net (Figures S1 and S5). This affects the neighboring parts of alkyl substituents, which are also disordered in 9a and 10a due to the relative freedom of their packing (Figure S1). Compounds 8a and 9a in the solid-state are effectively isostructural, whilst 10a adopts a different packing pattern. In all compounds (8a–10a), the molecules extend in one direction of the crystal net, but their mutual shifts are different. In 8a and 9a, the subsequent molecules are shifted to approximately half of the preceding molecule length, i.e., the terminal N,N-diphenylamino group of one molecule is located near the central triazole ring of the neighboring molecule (Figures S2 and S3). In 10a, the terminal N,N-diphenylamino groups of the neighboring molecules are adjacent (Figure S4). The overall conformation of molecules 8a–10a is similar, such that the molecules extend along axes going through the N(triazole)–C(phenylene) bonds due to the presence of 1,4-phenylene moieties. The mutual arrangement of the ring groups is different in all compounds (8a–10a), including clearly visible differences in the isostructural 8a and 9a. The dihedral angles between the least-squares plane of the diphenylene moiety rings vary from 12.3(1)° to 31.95(6)°, while the angles between the N,N-diphenylamino moiety rings differ from 52.80(5)° to 66.52(6)° (Table S3). The C–N and N–N bonds with the triazole ring possess similar lengths (Table S2) and respective values are between the length of single and double bonds C–N and N–N, [59] what proves the electron delocalization within the ring. The C(triazole)-C(phenylene) bonds exhibit shortening observed typically in exocyclic bonds connected to aromatic rings [59], which results in the redirecting of the electron density toward the C–C bond. The complete hydrogen bonding scheme of compounds 8a and 9a could not be resolved due to the disorder in the solvent molecules’ positions. Nevertheless, the studied molecules possess only two classical hydrogen bonds acceptors and non-classical hydrogen bonds donors [60]. The N1 and N2 atoms participate in weak C–H···N hydrogen bonds (Table S4). The isostructural forms of 8a and 9a have the same π···π stacking interaction motifs due to the similarity of their crystal packing [61]. One of the phenyl rings from each terminal N,N-diphenylamino substituent interacts with their symmetry generated equivalent rings. Additionally, one of these rings interacts with the symmetry generated ring of the N,N-diphenylamino moiety located on the opposite end of the molecule (Table S5). Consequently, one of the four rings belonging to N,N-diphenylamino substituents interacts with two rings, meaning two interact with one ring, and one does not form any π···π stacking interactions. In 10a, only two different π···π stacking interactions exist. The first one is formed between the neighboring triazole rings and the second one between one at the phenyl rings of each terminal N,N-diphenylamino moiety, i.e., similarly to 8a and 9a, the phenyl ring of one terminal moiety interacts with the symmetry generated ring of the oppositely located terminal moiety (Table S5).
The final products were tested for their luminescent properties, displaying strong fluorescent properties (Table 1, Figure 2). The compounds have a relatively large Stokes shift (Δλ = ~90 nm) and a near-unity quantum yield (Φ > 0.98). From these results, an alternative approach to the Suzuki reaction was investigated (Table 2).
The use of ionic liquids in the Suzuki cross-coupling reaction opens up the possibility for the additional acceleration of the reaction with the use of ultrasound or microwaves. These variants (Table 2, entries: 3–6, 9–12, 15–18 and 21–24) have a shorter reaction time compared to conventional heating (Table 2, entries: 1, 2, 7, 8, 13, 14, 19 and 20) in an oil bath while maintaining a high coupling efficiency. The most beneficial aspects from an economic point of view is the use of a small addition of a classical solvent (toluene, 1 mL) and conducting the reaction in a microwave reactor, where the complete conversion of substrates takes place after only 6 min (TLC), and the products are obtained in high yields (Table 2, entries: 6, 12, 18 and 24, 62–97%).
The highest yields were obtained for derivatives with two and four carbon alkyl chains, the subsequent reactions were carried out for these two series of triazole precursors, 5a and 5c (Scheme 3), and then their emission properties were tested (Table 3). Generally, compounds containing a 4H-1,2,4-triazole core substituted with an ethyl group at the position 4 were synthesized with a slightly smaller range of yields (70–99% for 7ak, Table 1 and Table 3) than their counterparts with the butyl substituent (62–99% for 9ak, Table 1 and Table 3).
Three-dimensional fluorescence characteristics were determined for all the final products from the outlined reactions. The spectra of compounds 7a–7e, 8a, 9a–9e and 10a possess a single fluorescence maximum, while compounds 7f–7k and 9f–9k have two maxima (Figure 2, Figure 3 and Figure 4). Almost all of the synthesized compounds emit strong fluorescence upon irradiation with UV light. The only exceptions are the compounds 7k and 9k containing the thianthrene substituent (Table 3). This is due to the presence of two sulfur atoms in the thianthrene moiety. Typically, the presence of one sulfur atom in a molecule causes the severe quenching of fluorescence [65,66]. In each molecule of compound 7k and 9k, four sulfur atoms are present, thus, the almost total quenching of fluorescence is observed (Figure 3 and Figure 4). Similarly, the existence of two sulfur atoms per one molecule in 7f and 9f distinctly decreases the number of emitted photons in comparison to the other studied compounds without a sulfur atom in the structure. The change in the aliphatic substituent length at the triazole nitrogen N atom does not affect the position and shape of the three-dimensional emission maxima. The effect of increased emission wavelengths for aliphatic substituents with even number of carbon atoms (in comparison to these with odd number of carbon atoms) is not observed in the studied case [58]. This suggests that the presence of substituents containing a single ring at the ends of the 4-alkyl-3,5-bis(phenyl)-4H-1,2,4-triazole core is necessary for the aforementioned even/odd number of carbon atoms and its influence on fluorescence. In the studied compounds, the terminal substituents consists of two up to four rings (in the subsequent or fused arrangement) and this completely diminishes the effect of the aliphatic substituent parity at the triazole nitrogen N atom.
The number of rings in terminal substituents also affects the excitation and emission wavelengths (Figure S8). There is a decrease in the excitation wavelength for the first maximum on the spectrum (accompanied by an increase in the excitation wavelength for the second maximum on the spectrum) with an increase in the number of rings (Figure S8). The difference between the excitation and emission wavelengths at the maxima falls by approximately 70–84 nm (for compounds possessing two fluorescence maxima, at least one fulfills this relation), thus the energy gap between the respective orbitals responsible for the fluorescence is very similar in all studied compounds. The spectra of the compounds with rings separated by a single bond (7a–10a) or with no more than two fused rings (7b–7e and 9b–9e) possess a single maximum in the three-dimensional fluorescence spectrum. The compounds with three fused rings (7f–7k and 9f–9k) possess two more or less shaped maxima at the same emission wavelength and two different excitation wavelengths. Manifestations of these maxima were faintly but unambiguously visible in the compounds with almost completely quenched fluorescence (7k and 9k). All the single maxima possess a considerable degree of asymmetry, with the emission drastically fading with a decrease in the emission wavelengths and/or increase in the excitation wavelengths. In two other directions of the spectra, the emission decreases slowly. This originates from the electronic transitions of the excited state (S1) to the different energy sub-levels (vibrational energy levels 0, 1, 2, …) of the ground state (S0). The asymmetricity increases with the number of electrons in the delocalized system (7b7c9b9c < 7d7e9d9e < 7a8a9a10a), due to the increased number of vibrational energy levels. The presence of two fluorescence maxima in the spectra of compounds 7f–7k and 9f–9k is caused by the presence of ring systems possessing two different heteroatomic rings (central and terminal ones). The intensity of both maxima is similar for compounds 7g–7j and 9g–9j. For compounds 7f and 9f, the first (at lower excitation wavelengths) maxima is stronger than the second one (existing at larger excitation wavelengths), while for compounds 7j and 9j the mutual relation of maxima intensities is reversed. This effect is caused by the presence of a thiophene moiety in the thianthrene ring system of 7f and 9f and an occurrence of an additional phenylene moiety in 7j and 9j (in comparison to the other studied compounds). The conjugation of the triazole ring with appropriately selected substituents containing together more than six aromatic rings (7a–10a, 7g–7j, and 9g–9j excluding 7f, 7k, 9f, and 9k, containing sulfur atoms) leads to reach a nearly 1.00 value of quantum yield (Φ). Such properties are not achieved by triazoles conjugated with four aromatic rings [58]. Their Φ, although high, reaches values maximally around 0.9. The triazoles which are not bonded directly to aryl substituents are characterized by a very low Φ (below 0.01) [67]. To improve the fluorescence properties of triazoles conjugated with a lower number of aromatic rings, the coordination compounds with d10 transition metals are synthesized [68]. Even then, the zinc compounds of triazoles with one aromatic ring achieve Φ only up to 0.075 [69]. The Φ of the studied compounds correlate well with the fluorescence intensities (larger fluorescence leads to larger Φ, Figure S9), whereas there is no discernible relationship between Φ and absorption (Figure S10). This suggests that the mechanism of fluorescence is similar in all compounds and the relatively large differences in the absorption values derive from the variations in the amount of electromagnetic energy (photons) transformed directly into internal energy. The above described effects (presence/lack of a correlation) exist also in 4-alkyl-3,5-bis(phenyl)-4H-1,2,4-triazole substituted at the ends by a single ring [58] and there is no such effects in the di(1,3,4-oxadiazol-2-yl)-1,2,4,5-tetrazine and 3,6-di(1,3,4-thiadiazol-2-yl)-1,2,4,5-tetrazine derivatives [70]. In the case of the compounds exhibiting a single fluorescence maximum, the n→π* absorption transition is the main source of the excited states leading to the subsequent emission of the fluorescence photons. The origin of fluorescence in the compounds exhibiting two fluorescence maxima are n→π* (for higher λ m a x e x ) and ππ* (for lower λ m a x e x ) absorption transitions [68]. The involved n-type orbital originates from the nitrogen atom of the triazole ring (N1 or N2) and the source of the π-type orbital is also the triazole ring [71]. The lowest unoccupied molecular orbital in similarly conjugated 1,2,4-triazole compounds is antibonding delocalized orbital (π*), which most often is the benzene ring directly attached to the triazole ring [72] or fused π* orbital of triazole and the neighboring benzene [73]. The presence of a single fluorescence maximum or two fluorescence maxima of the same emission wavelength indicates that π*→π transitions are responsible for the fluorescence in the studied compounds [74]. Noteworthy is the fact that upon ultraviolet irradiation, the compounds 7a–10a emit a strong deep-blue fluorescence light visible by the naked eye. The blue-violet color emitted light is clearly visible for compound 7c–7e, 7h–7j, 9c–9e, 9h–9j.

4. Conclusions

Four alternative approaches using palladium-catalyzed Suzuki cross-coupling reactions to synthesize a series of highly-conjugated 4H-1,2,4-triazole derivatives were presented. The reactions of the intermediate 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazoles and boronic acids were conducted using conventional heating in a two-phase solvent system (toluene/EtOH/H2O) or using an ionic liquid (choline–OH). For the ionic liquid approach, it was possible to use ultrasound or microwave assistance, which significantly reduces the reaction time. Additionally, this type of methodology removes the need for large amounts of organic solvents and for the addition of auxiliary reagents such as a base or a phase transfer catalyst. Generally, 4-alkyl-4H-1,2,4-triazole derivatives conjugated to different fused-bicyclic and fused-tricyclic systems via a 1,4-phenylene linker were obtained in excellent yields. Almost all compounds are effective fluorophores and emit visible light upon irradiation by ultraviolet electromagnetic waves. The exceptions are the products possessing sulfur atoms within terminal moieties, which cause a large decrease in the fluorescence. The studied compounds produce fluorescence with large quantum yields (up to almost 100%), which depend on a structure of the terminal substituents. The obtained compounds, due to their good emission properties and relatively good solubility in typical organic solvents, seem to be of particular interest for optoelectronic purposes.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/13/24/5627/s1, copy of 1H NMR, 13C NMR and HRMS spectra; X-ray crystallography data; absorption spectrometry data. Figure S1: Complete asymmetric units of the structures of compounds 8a, 9a, and 10a, with atom numbering scheme, plotted with 50% probability of displacement ellipsoids of non-hydrogen atoms. Hydrogen atoms are plotted as spheres of arbitrary radii, Figure S2: Solvent accessible voids within crystal structure of 8a, Figure S3: Solvent accessible voids within crystal structure of 9a, Figure S4: The part of crystal packing in 10a, Figure S5: Disorder model of solvent in 8a, Figure S6: UV-Vis spectra of 7a–10a and 7b–7k, Figure S7: UV-Vis spectra of 9b–9k, Figure S8: Positions of global maxima for studied compounds (divided into groups containing 2 or 3 rings at the ends of 4-alkyl-3,5-bis(phenyl)-4H-1,2,4-triazole core). “1st max.” and “2nd max.” indicate the first (for smaller excitation wavelengths) and the second (for larger excitation wavelengths) maxima on three-dimensional fluorescence spectra, Figure S9: Quantum yield of studied compounds as a function of fluorescence intensity at global and local maximum, Figure S10: Quantum yield of studied compounds in relation to absorption at global and local maximum of fluorescence, Table S1: Crystal data and structure refinement details for 8a, 9a, and 10a. The structures of 8a and 9a were refined twice: against measured and squeezed data, Table S2: Selected structural data of 8a, 9a, and 10a, Table S3: Dihedral angles (°) between ring least squares planes in 8a, 9a, and 10a. Each ring is indicated by one atom, which belongs solely to this ring, Table S4: Non-classic hydrogen bonds and the first level graph motifs in the studied compounds (Å, °), Table S5: Stacking interactions in the studied compounds. Each ring is indicated by one atom, which belongs solely to this ring. The α is a dihedral angle between planes I and J, β is an angle between Cg(I)-Cg(J) vector and normal to plane I, dp is a perpendicular distance of Cg(I) on ring J plane.

Author Contributions

Conceptualization, M.O. and A.K.; methodology, M.O. and A.K.; formal analysis, M.O. and M.Ś.; investigation, M.O.; data curation, M.Ś.; writing—original draft preparation, M.O. and M.Ś.; writing—review and editing, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Polish National Science Centre grant UMO-2016/23/N/ST5/02036.

Acknowledgments

The authors would like to thank the late Rafał Kruszyński for his contribution to this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kraft, A.; Grimsdale, A.C.; Holmes, A.B. Electroluminescent Conjugated Polymers—Seeing Polymers in a New Light. Angew. Chem. Int. Ed. 1998, 37, 402–428. [Google Scholar] [CrossRef]
  2. Segura, J.L. The chemistry of electroluminescent organic materials. Acta Polym. 1998, 49, 319–344. [Google Scholar] [CrossRef]
  3. Mitschke, U.; Bäuerle, P. The electroluminescence of organic materials. J. Mater. Chem. 2000, 10, 1471–1507. [Google Scholar] [CrossRef]
  4. Bera, M.K.; Pal, P.; Malik, S. Solid-state emissive organic chromophores: Design, strategy and building blocks. J. Mater. Chem. C 2020, 8, 788–802. [Google Scholar] [CrossRef]
  5. Zhu, M.; Yang, C. Blue fluorescent emitters: Design tactics and applications in organic light-emitting diodes. Chem. Soc. Rev. 2013, 42, 4963–4976. [Google Scholar] [CrossRef] [PubMed]
  6. Katritzky, A.R.; Ramsden, C.A.; Joule, J.A.; Zhdankin, V.V. Structure of Five-membered Rings with Two or More Heteroatoms. In Handbook of Heterocyclic Chemistry; Elsevier: Oxford, UK, 2010; Volume 2, pp. 139–209. [Google Scholar]
  7. Curtis, A.D.M.; Jennings, N. 1,2,4-Triazoles. In Comprehensive Heterocyclic Chemistry III; Katritzky, A.R., Ramsden, C.A., Scriven, E.F.V., Taylor, R.J.K., Eds.; Elsevier: Oxford, UK, 2008; Volume 5, pp. 159–209. [Google Scholar]
  8. Liu, H.; Wang, L.; Wu, Y.; Liao, Q. Luminescence emission-modulated based on specific two-photon compound of triazole-conjugated pyrene derivative. RSC Adv. 2017, 7, 19002–19006. [Google Scholar] [CrossRef] [Green Version]
  9. Li, W.; Yan, L.; Zhou, H.; You, W. A General Approach toward Electron Deficient Triazole Units to Construct Conjugated Polymers for Solar Cells. Chem. Mater. 2015, 27, 6470–6476. [Google Scholar] [CrossRef]
  10. Wu, C.-S.; Lee, S.-L.; Chen, Y. Bipolar copoly(aryl ether) containing distyrylbenzene, triphenylamine, and 1,2,4-triazole moieties: Synthesis and optoelectronic properties. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 3099–3108. [Google Scholar] [CrossRef]
  11. Tao, Y.; Wang, Q.; Ao, L.; Zhong, C.; Yang, C.; Qin, J.; Ma, D. Highly Efficient Phosphorescent Organic Light-Emitting Diodes Hosted by 1,2,4-Triazole-Cored Triphenylamine Derivatives: Relationship between Structure and Optoelectronic Properties. J. Phys. Chem. C 2010, 114, 601–609. [Google Scholar] [CrossRef]
  12. Meng, X.; Chen, M.; Bai, R.; He, L. Cationic Iridium Complexes with 3,4,5-Triphenyl-4 H-1,2,4-Triazole Type Cyclometalating Ligands: Synthesis, Characterizations, and Their Use in Light-Emitting Electrochemical Cells. Inorg. Chem. 2020, 59, 9605–9617. [Google Scholar] [CrossRef]
  13. Demirbaş, Ü.; Özçifçi, Z.; Akçay, H.T.; Menteşe, E. Novel phthalocyanines bearing 1,2,4 triazole substituents: Synthesis, characterization, photophysical and photochemical properties. Polyhedron 2020, 181, 114470. [Google Scholar] [CrossRef]
  14. Abdurahman, A.; Wang, L.; Zhang, Z.; Feng, Y.; Zhao, Y.; Zhang, M. Novel triazole-based AIE materials: Dual-functional, highly sensitive and selective fluorescence probe. Dye. Pigment. 2020, 174, 108050. [Google Scholar] [CrossRef]
  15. Yan, Y.; Lin, X.; Zhang, L.; Zhou, H.; Wu, L.; Cai, L. Electrochemical and quantum-chemical study on newly synthesized triazoles as corrosion inhibitors of mild steel in 1 M HCl. Res. Chem. Intermed. 2017, 43, 3145–3162. [Google Scholar] [CrossRef]
  16. Jiang, L.; Lan, Y.; He, Y.; Li, Y.; Li, Y.; Luo, J. 1,2,4-Triazole as a corrosion inhibitor in copper chemical mechanical polishing. Thin Solid Films 2014, 556, 395–404. [Google Scholar] [CrossRef]
  17. Sherif, E.-S.M.; Erasmus, R.M.; Comins, J.D. Effects of 3-amino-1,2,4-triazole on the inhibition of copper corrosion in acidic chloride solutions. J. Colloid Interface Sci. 2007, 311, 144–151. [Google Scholar] [CrossRef]
  18. Wang, L. Inhibition of mild steel corrosion in phosphoric acid solution by triazole derivatives. Corros. Sci. 2006, 48, 608–616. [Google Scholar] [CrossRef]
  19. Shet, N.; Nazareth, R.; Suchetan, P.A. Corrosion inhibition of 316 stainless steel in 2M HCl by 4-{[4-(dimethylamino)benzylidene]amino}-5-methyl-4H-1,2,4-triazole-3-thiol. Chem. Data Collect. 2019, 20, 100209. [Google Scholar] [CrossRef]
  20. He, P.; Wu, B.; Shao, S.; Teng, T.; Wang, P.; Qu, X.-P. Characterization of 1,2,4-Triazole as Corrosion Inhibitor for Chemical Mechanical Polishing of Cobalt in H2O2 Based Acid Slurry. ECS J. Solid State Sci. Technol. 2019, 8, P3075–P3084. [Google Scholar] [CrossRef]
  21. Zhang, J.; Wang, S.; Ba, Y.; Xu, Z. 1,2,4-Triazole-quinoline/quinolone hybrids as potential anti-bacterial agents. Eur. J. Med. Chem. 2019, 174, 1–8. [Google Scholar] [CrossRef]
  22. Gao, F.; Wang, T.; Xiao, J.; Huang, G. Antibacterial activity study of 1,2,4-triazole derivatives. Eur. J. Med. Chem. 2019, 173, 274–281. [Google Scholar] [CrossRef]
  23. Venugopala, K.N.; Kandeel, M.; Pillay, M.; Deb, P.K.; Abdallah, H.H.; Mahomoodally, M.F.; Chopra, D. Anti-Tubercular Properties of 4-Amino-5-(4-Fluoro-3-Phenoxyphenyl)-4H-1,2,4-Triazole-3-Thiol and Its Schiff Bases: Computational Input and Molecular Dynamics. Antibiotics 2020, 9, 559. [Google Scholar] [CrossRef] [PubMed]
  24. Mehta, C.C.; Patel, A.; Bhatt, H.G. New molecular insights into dual inhibitors of tankyrase as Wnt signaling antagonists: 3D-QSAR studies on 4H-1,2,4-triazole derivatives for the design of novel anticancer agents. Struct. Chem. 2020, 31, 2371–2389. [Google Scholar] [CrossRef]
  25. Saadaoui, I.; Krichen, F.; Ben Salah, B.; Ben Mansour, R.; Miled, N.; Bougatef, A.; Kossentini, M. Design, synthesis and biological evaluation of Schiff bases of 4-amino-1,2,4-triazole derivatives as potent angiotensin converting enzyme inhibitors and antioxidant activities. J. Mol. Struct. 2019, 1180, 344–354. [Google Scholar] [CrossRef]
  26. Fan, Y.-L.; Ke, X.; Liu, M. Coumarin-triazole Hybrids and Their Biological Activities. J. Heterocycl. Chem. 2018, 55, 791–802. [Google Scholar] [CrossRef]
  27. Kaur, P.; Chawla, A. 1,2,4-Triazole: A Review Of Pharmacological Activities. Parminder Kaur Anshul Chawla. Int. Res. J. Pharm 2017, 8, 10–29. [Google Scholar] [CrossRef]
  28. Zhang, S.; Xu, Z.; Gao, C.; Ren, Q.-C.; Chang, L.; Lv, Z.-S.; Feng, L.-S. Triazole derivatives and their anti-tubercular activity. Eur. J. Med. Chem. 2017, 138, 501–513. [Google Scholar] [CrossRef] [PubMed]
  29. Asif, M. Biological Potentials of Biological Active Triazole Derivatives: A Short Review. Org. Chem. Curr. Res. 2016, 5, 1000173. [Google Scholar] [CrossRef]
  30. Maddila, S.; Pagadala, R.; Jonnalagadda, S. 1,2,4-Triazoles: A Review of Synthetic Approaches and the Biological Activity. Lett. Org. Chem. 2013, 10, 693–714. [Google Scholar] [CrossRef]
  31. Zhou, C.-H.; Wang, Y. Recent Researches in Triazole Compounds as Medicinal Drugs. Curr. Med. Chem. 2012, 19, 239–280. [Google Scholar] [CrossRef]
  32. Sever, B.; Altıntop, M.D.; Demir, Y.; Pekdoğan, M.; Akalın Çiftçi, G.; Beydemir, Ş.; Özdemir, A. An extensive research on aldose reductase inhibitory effects of new 4H-1,2,4-triazole derivatives. J. Mol. Struct. 2020, 129446. [Google Scholar] [CrossRef]
  33. Hull, J.W.; Romer, D.R.; Adaway, T.J.; Podhorez, D.E. Development of Manufacturing Processes for a New Family of 2,6-Dihaloaryl 1,2,4-Triazole Insecticides. Org. Process Res. Dev. 2009, 13, 1125–1129. [Google Scholar] [CrossRef]
  34. Zhang, Z.; Gao, B.; He, Z.; Li, L.; Shi, H.; Wang, M. Enantioselective metabolism of four chiral triazole fungicides in rat liver microsomes. Chemosphere 2019, 224, 77–84. [Google Scholar] [CrossRef] [PubMed]
  35. Li, J.; Ren, G.-Y.; Zhang, Y.; Yang, M.-Y.; Ma, H.-X. Two Cu(II) complexes of 1,2,4-triazole fungicides with enhanced antifungal activities. Polyhedron 2019, 157, 163–169. [Google Scholar] [CrossRef]
  36. Verweij, P.E.; Kema, G.H.; Zwaan, B.; Melchers, W.J. Triazole fungicides and the selection of resistance to medical triazoles in the opportunistic mould Aspergillus fumigatus. Pest Manag. Sci. 2013, 69, 165–170. [Google Scholar] [CrossRef] [PubMed]
  37. Ribas e Ribas, A.D.; Spolti, P.; Del Ponte, E.M.; Donato, K.Z.; Schrekker, H.; Fuentefria, A.M. Is the emergence of fungal resistance to medical triazoles related to their use in the agroecosystems? A mini review. Braz. J. Microbiol. 2016, 47, 793–799. [Google Scholar] [CrossRef] [Green Version]
  38. Lohmann, J.K.; Craig, I.R.; Brahm, L.; Fehr, M.; Weber, A.; Seet, M.; Mueller, B.; Wiebe, C.; Winter, C.H.; Grote, T.; et al. European Patent Application Substituted [1,2,4]triazoles as Agricultural Fungicides; European Patent Organisation: Dallas, TX, USA, 2020; pp. 1–79. [Google Scholar]
  39. Shalini, K.; Kumar, N.; Drabu, S.; Sharma, P.K. Advances in synthetic approach to and antifungal activity of triazoles. Beilstein J. Org. Chem. 2011, 7, 668–677. [Google Scholar] [CrossRef]
  40. Moulin, A.; Bibian, M.; Blayo, A.-L.; El Habnouni, S.; Martinez, J.; Fehrentz, J.-A. Synthesis of 3,4,5-Trisubstituted-1,2,4-triazoles. Chem. Rev. 2010, 110, 1809–1827. [Google Scholar] [CrossRef]
  41. Mustafa, S.; Nair, V.; Chittoor, J.; Krishnapillai, S. Synthesis of 1,2,4-Triazoles and Thiazoles from Thiosemicarbazide and its Derivatives. Mini Rev. Org. Chem. 2004, 1, 375–385. [Google Scholar] [CrossRef]
  42. Ibrahim, M.A.; El-Gohary, N.M. Heterocyclization with Some Heterocyclic Diamines: Synthetic Approaches for Nitrogen Bridgehead Heterocyclic Systems. Heterocycles 2014, 89, 1125–1157. [Google Scholar] [CrossRef]
  43. Al-Masoudi, I.A.; Al-Soud, Y.A.; Al-Salihi, N.J.; Al-Masoudi, N.A. 1,2,4-Triazoles: Synthetic approaches and pharmacological importance. (Review). Chem. Heterocycl. Compd. 2006, 42, 1377–1403. [Google Scholar] [CrossRef]
  44. Jilloju, P.C.; Vinaykumar, A.; Shyam, P.; Vedula, R.R. One-pot, Multicomponent Cascade Reaction for the Synthesis of Various Aralkyl/alkylthio-3,5-dimethyl-1H-pyrazolyl-4H-1,2,4-triazol-4-amine and Their Docking Studies. J. Heterocycl. Chem. 2019, 56, 1012–1019. [Google Scholar] [CrossRef]
  45. Nikpassand, M.; Farshami, M.J. One-Pot Synthesis of Novel 3-Pyrazolyl-4H-1,2,4-triazoles Using Amino Glucose-Functionalized Silica-Coated NiFe2O4 Nanoparticles as a Magnetically Separable Catalyst. J. Clust. Sci. 2020, 1–8. [Google Scholar] [CrossRef]
  46. Jilloju, P.C.; Srikanth, M.; Kumar, S.V.; Vedula, R.R. One-Pot, Multi-Component Synthesis of Substituted 2-(6-Phenyl-7H-[1,2,4]Triazolo[3,4-b][1,3,4]Thiadiazin-3-yl)-2,3-Dihydrophthalazine-1,4-Diones. Polycycl. Aromat. Compd. 2020, 1–12. [Google Scholar] [CrossRef]
  47. de Meijere, A.; Diederich, F. Metal-Catalyzed Cross-Coupling Reactions; de Meijere, A., Diederich, F., Eds.; Wiley-VCH: Weinheim, Germany, 2004; Volume 1. [Google Scholar]
  48. Miyaura, N. Metal-Catalyzed Cross-Coupling Reactions of Organoboron Compounds with Organic Halides. In Metal-Catalyzed Cross-Coupling Reactions; de Meijere, A., Diederich, F., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp. 41–123. [Google Scholar]
  49. Peruzzini, M.; Gonsalvi, L. Phosphorus Compounds. Advanced Tools in Catalysis and Material Sciences. In Catalysis by Metal Complexes; Springer: Berlin/Heidelberg, Germany, 2011; Volume 37. [Google Scholar]
  50. Mandal, B.; Ghosh, S.; Basu, B. Task-Specific Properties and Prospects of Ionic Liquids in Cross-Coupling Reactions. Top. Curr. Chem. 2019, 377, 1–43. [Google Scholar] [CrossRef] [PubMed]
  51. Yousaf, M.; Zahoor, A.F.; Akhtar, R.; Ahmad, M.; Naheed, S. Development of green methodologies for Heck, Chan–Lam, Stille and Suzuki cross-coupling reactions. Mol. Divers. 2019, 28, 1–19. [Google Scholar] [CrossRef] [PubMed]
  52. Boruah, P.R.; Gehlot, P.S.; Kumar, A.; Sarma, D. Palladium immobilized on the surface of MMT K 10 with the aid of [BMIM][BF4]: An efficient catalyst for Suzuki-Miyaura cross-coupling reactions. Mol. Catal. 2018, 461, 54–59. [Google Scholar] [CrossRef]
  53. Hooshmand, S.E.; Heidari, B.; Sedghi, R.; Varma, R.S. Recent advances in the Suzuki-Miyaura cross-coupling reaction using efficient catalysts in eco-friendly media. Green Chem. 2019, 21, 381–405. [Google Scholar] [CrossRef]
  54. Massaro, M.; Riela, S.; Lazzara, G.; Gruttadauria, M.; Milioto, S.; Noto, R. Green conditions for the Suzuki reaction using microwave irradiation and a new HNT-supported ionic liquid-like phase (HNT-SILLP) catalyst. Appl. Organomet. Chem. 2014, 28, 234–238. [Google Scholar] [CrossRef] [Green Version]
  55. Kudelko, A.; Wróblowska, M.; Jarosz, T.; Katarzyna, Ł. Synthesis, spectral characteristics and electrochemistry of symmetrically substituted hybrids derived from 2,5-bis(4-bromophenyl)-1,3,4-oxadiazole under Suzuki cross-coupling reaction. Arkivoc 2015, 2015, 287–302. [Google Scholar] [CrossRef]
  56. Wróblowska, M.; Kudelko, A.; Łapkowski, M. Efficient Synthesis of Conjugated 1,3,4-Thiadiazole Hybrids through Palladium-Catalyzed Cross-Coupling of 2,5-Bis(4-bromophenyl)-1,3,4-thiadiazole with Boronic Acids. Synlett 2015, 26, 2127–2130. [Google Scholar] [CrossRef]
  57. Wróblowska, M.; Kudelko, A.; Kuźnik, N.; Łaba, K.; Łapkowski, M. Synthesis of Extended 1,3,4-Oxadiazole and 1,3,4-Thiadiazole Derivatives in the Suzuki Cross-coupling Reactions. J. Heterocycl. Chem. 2017, 54, 1550–1557. [Google Scholar] [CrossRef]
  58. Olesiejuk, M.; Kudelko, A.; Swiatkowski, M.; Kruszynski, R. Synthesis of 4-Alkyl-4H-1,2,4-triazole Derivatives by Suzuki Cross-coupling Reactions and Their Luminescence Properties. Molecules 2019, 24, 652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Prince, E. (Ed.) International Tables for Crystallography, Volume C: Mathematical, Physical and Chemical Tables; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2004. [Google Scholar]
  60. Jeffrey, G.A.; Saenger, W. Hydrogen Bonding in Biological Structures; Springer: New York, NY, USA, 1991. [Google Scholar]
  61. Kruszynski, R.; Sierański, T. Can Stacking Interactions Exist beyond the Commonly Accepted Limits? Cryst. Growth Des. 2016, 16, 587–595. [Google Scholar] [CrossRef]
  62. Melhuish, W.H. Quantum efficiencies of fluorescence of organic substan ces: Effect of solvent and concentration of the fluorescent solute. J. Phys. Chem. 1961, 65, 229–235. [Google Scholar] [CrossRef]
  63. Birks, J.B.; Dyson, D.J. The relations between the fluorescence and absorption properties of organic molecules. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1963, 275, 135–148. [Google Scholar]
  64. Brouwer, A.M. Standards for photoluminescence quantum yield measurements in solution (IUPAC technical report). Pure Appl. Chem. 2011, 83, 2213–2228. [Google Scholar] [CrossRef] [Green Version]
  65. Shamsipur, M.; Chaichi, M.J. A study of quenching effect of sulfur-containing amino acids l-cysteine and l-methionine on peroxyoxalate chemiluminescence of 7-amino-4-trifluoromethylcumarin. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 61, 1227–1231. [Google Scholar] [CrossRef]
  66. D’Auria, S.; Staiano, M.; Kuznetsova, I.M.; Turoverov, K.K. The Combined Use of Fluorescence Spectroscopy and X-Ray Crystallography Greatly Contributes to Elucidating Structure and Dynamics of Proteins. In Reviews in Fluorescence; Geddes, C.D., Lakowicz, J.R., Eds.; Springer: Boston, MA, USA, 2005; pp. 25–61. [Google Scholar]
  67. Kayumova, R.R.; Ostakhov, S.S.; Mamykin, A.V.; Muslukhov, R.R.; Iskhakova, G.F.; Ivanov, S.P.; Meshcheryakova, S.A.; Klen, E.E.; Khaliullin, F.A.; Kazakov, V.P. Structure and luminescence of thietane-containing 1,2,4-triazoles. Russ. J. Gen. Chem. 2011, 81, 1203–1210. [Google Scholar] [CrossRef]
  68. Liu, K.; Shi, W.; Cheng, P. The coordination chemistry of Zn(II), Cd(II) and Hg(II) complexes with 1,2,4-triazole derivatives. Dalton Trans. 2011, 40, 8475–8490. [Google Scholar] [CrossRef]
  69. Yunusova, S.N.; Novikov, A.S.; Khoroshilova, O.V.; Kolesnikov, I.E.; Demakova, M.Y.; Bolotin, D.S. Solid-state fluorescent 1,2,4-triazole zinc(II) complexes: Self-organization via bifurcated (N[sbnd]H)2⋯Cl contacts. Inorg. Chim. Acta 2020, 510, 119660. [Google Scholar] [CrossRef]
  70. Kędzia, A.; Kudelko, A.; Świątkowski, M.; Kruszyński, R. Microwave-promoted synthesis of highly luminescent s-tetrazine-1,3,4-oxadiazole and s-tetrazine-1,3,4-thiadiazole hybrids. Dye. Pigment. 2020, 172, 107865. [Google Scholar] [CrossRef]
  71. Meng, S.; Duan, A.; Xue, J.; Zheng, X.; Zhao, Y. UV-Vis, Fluorescence, and Resonance Raman Spectroscopic and Density Functional Theoretical Studies on 3-Amino-1,2,4-triazole: Microsolvation and Solvent-Dependent Nonadiabatic Excited State Decay in Solution. J. Phys. Chem. A 2018, 122, 8530–8538. [Google Scholar] [CrossRef] [PubMed]
  72. Gusev, A.; Braga, E.; Baluda, Y.; Kiskin, M.; Kryukova, M.; Karaush-Karmazin, N.; Baryshnikov, G.; Kuklin, A.; Minaev, B.; Ågren, H.; et al. Structure and tuneable luminescence in polymeric zinc compounds based on 3-(3-pyridyl)-5-(4-pyridyl)-1,2,4-triazole. Polyhedron 2020, 191, 114768. [Google Scholar] [CrossRef]
  73. Jin, X.D.; Li, B.; Gao, H.; Zhang, X.; Liu, W.Y. Synthesis, crystal structure, fluorescent property and DFT calculations of a new Zn(II) complex based on 3-(2-pyridyl)-5-(4-pyridyl)-1H-1,2,4-triazole. Chin. J. Struct. Chem. 2016, 35, 1129–1136. [Google Scholar] [CrossRef]
  74. Yao, Y.G.; Yang, J.X.; Zhang, X.; Qin, Y.Y. N-donor auxiliary ligand influence on the coordination mode variations of v-shaped triazole dicarboxylic acid ligand affording seven new luminescent Zn(II) compounds with variable structural motifs. Cryst. Growth Des. 2020, 20, 6366–6381. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the 4-alkyl-4H-1,2,4-triazole precursors.
Scheme 1. Synthesis of the 4-alkyl-4H-1,2,4-triazole precursors.
Materials 13 05627 sch001
Scheme 2. The conventional Suzuki cross-coupling reaction for 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole derivatives. Reagents and conditions: aryl dibromide 5a–d (1.00 mmol), 4-(N,N-diphenylamino)phenylboronic acid (6a) (2.50 mmol), Pd(PPh3)4 (0.05 mmol), NBu4Br (0.10 mmol), K2CO3 (10 mmol), toluene/H2O/EtOH (10:6:3 mL), oil bath 130 °C, 7 h.
Scheme 2. The conventional Suzuki cross-coupling reaction for 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole derivatives. Reagents and conditions: aryl dibromide 5a–d (1.00 mmol), 4-(N,N-diphenylamino)phenylboronic acid (6a) (2.50 mmol), Pd(PPh3)4 (0.05 mmol), NBu4Br (0.10 mmol), K2CO3 (10 mmol), toluene/H2O/EtOH (10:6:3 mL), oil bath 130 °C, 7 h.
Materials 13 05627 sch002
Figure 1. The molecular structures of compounds 8a10a, with an atom numbering scheme, plotted with a 50% probability of the displacement ellipsoids of non-hydrogen atoms. Hydrogen atoms are plotted as spheres with arbitrary radii. The disordered solvent molecules of 8a and 9a, as well as a minor contribution from the distorted alkyl chains of 9a and 10a, are omitted for clarity.
Figure 1. The molecular structures of compounds 8a10a, with an atom numbering scheme, plotted with a 50% probability of the displacement ellipsoids of non-hydrogen atoms. Hydrogen atoms are plotted as spheres with arbitrary radii. The disordered solvent molecules of 8a and 9a, as well as a minor contribution from the distorted alkyl chains of 9a and 10a, are omitted for clarity.
Materials 13 05627 g001
Scheme 3. The Suzuki coupling reaction of 4H-1,2,4-triazole precursors 5a and 5c with selected boronic acids 6b–k. Reagents and conditions: aryl dibromide 5a,c (1.00 mmol), arylboronic acid 6b–k (2.50 mmol), Pd(PPh3)4 (0.05 mmol), choline–OH/toluene (10:1 mL), and microwave irradiation, 5–7 min.
Scheme 3. The Suzuki coupling reaction of 4H-1,2,4-triazole precursors 5a and 5c with selected boronic acids 6b–k. Reagents and conditions: aryl dibromide 5a,c (1.00 mmol), arylboronic acid 6b–k (2.50 mmol), Pd(PPh3)4 (0.05 mmol), choline–OH/toluene (10:1 mL), and microwave irradiation, 5–7 min.
Materials 13 05627 sch003
Figure 2. Three-dimensional fluorescence spectra of compounds 7a10a. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Figure 2. Three-dimensional fluorescence spectra of compounds 7a10a. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Materials 13 05627 g002
Figure 3. Three-dimensional fluorescence spectra of compounds 7b7k. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Figure 3. Three-dimensional fluorescence spectra of compounds 7b7k. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Materials 13 05627 g003
Figure 4. Three-dimensional fluorescence spectra of compounds 9b9k. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Figure 4. Three-dimensional fluorescence spectra of compounds 9b9k. The color scale represents the flux of emitted photons. The number above the color scales indicates the maximum relative value of emission intensity represented by color scale (and indicated in the respective figure). The unit of measurement in each spectrum represents the same number of emitted photons per second, i.e., all values of intensity presented in Figure 2, Figure 3 and Figure 4 can be directly compared. The green area (upper left triangles in each spectrum) indicates the non-fluorescent region in which λex is close to or larger than λem.
Materials 13 05627 g004
Table 1. 4-Alkyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazoles (7–10a) prepared using the Suzuki cross-coupling reaction a.
Table 1. 4-Alkyl-3,5-bis[4′-(N,N-diphenylamino)biphenyl-4-yl]-4H-1,2,4-triazoles (7–10a) prepared using the Suzuki cross-coupling reaction a.
ProductYield a (%) Absorption   Maximum   λ m a x a b s   ( nm ) Excitation   Wavelength   λ m a x e x   ( nm ) Emission   Wavelength   λ m a x e m   ( nm ) Stokes Shift b Δ (nm)Quantum Yield c,d,e Φ
7a9735135244392>0.98 c,d,e
8a9335235244290>0.98 c,d,e
9a9935135144392>0.98 c,d,e
10a6135235244391>0.98 c,d,e
a isolated yield. b stokes shift (Δ) from the equation Δ   =   λ m a x e m λ m a x a b s . Wavelength determined from the 3D emission spectrum for C = 5.0 × 10−6 M CH2Cl2 solution (nm). c quinine sulfate was used as a standard [62]. d trans,trans-1,4-diphenyl-1,3-butadiene was used as a standard [63]. e exact value cannot be determined due to nonlinearity of standard/sample dependence in the Φ region of 0.97–1.00 [64].
Table 2. An alternative ionic liquids (IL) Suzuki cross-coupling reaction of the 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole moiety (7a–10a).
Table 2. An alternative ionic liquids (IL) Suzuki cross-coupling reaction of the 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole moiety (7a–10a).
Materials 13 05627 i001
EntrySubstrateApproachReaction TimeAdditional Solvent (1 mL)Yield a (%)
15aConventional heating
(oil bath 130 °C)
24 h--
2Toluene85
3Ultrasounds10 h-5
4Toluene89
5Microwaves6 min-8
6Toluene94
75bConventional heating
(oil bath 130 °C)
24 h--
8Toluene80
9Ultrasounds10 h-3
10Toluene87
11Microwaves6 min-10
12Toluene89
135cConventional heating
(oil bath 130 °C)
24 h--
14Toluene85
15Ultrasounds10 h-3
16Toluene93
17Microwaves6 min-9
18Toluene97
195dConventional heating
(oil bath 130 °C)
24 vh--
20Toluene47
21Ultrasounds10 h--
22Toluene56
23Microwaves6 min-5
24Toluene62
a yield with respect to the 4-alkyl-3,5-bis(4-bromophenyl)-4H-1,2,4-triazole (5a–d). Reagents and conditions: aryl dibromide 5a–d (1.00 mmol), boronic acid 6a (2.50 mmol), Pd(PPh3)4 (0.05 mmol), choline–OH (10 mL).
Table 3. 4-Alkyl-3,5-bis(4-arylphenyl)-4H-1,2,4-triazoles 7b–k and 9b–k prepared using the Suzuki cross-coupling reaction a.
Table 3. 4-Alkyl-3,5-bis(4-arylphenyl)-4H-1,2,4-triazoles 7b–k and 9b–k prepared using the Suzuki cross-coupling reaction a.
ProductYield a (%) Absorption   Maximum   λ m a x a b s   ( nm ) Excitation   Wavelength   λ m a x e x   ( nm ) Emission   Wavelength   λ m a x e m   ( nm ) Stokes Shift b Δ (nm)Quantum Yield c,d,e Φ
7b88297304382850.75 c/0.74 d
7c9930531339489>0.98 c,d,e
7d79305312395900.75 c/0.74 d
7e94306312394880.55 c/0.54 d
7f96290291371810.11 c/0.11 d
335335374390.18 c/0.18 d
7g922862903881020.63 c/0.62 d
315316388730.95 c/0.93 d
7h703003044031030.77 c/0.76 d
32132240180>0.98 c,d,e
7i763013044041030.78 c/0.77 d
32232240381>0.98 c,d,e
7j772932934091160.77 c/0.76 d
32232840886>0.98 c,d,e
7k87263----
9b62296304382860.64 c/0.62 d
9c89305313393880.95 c/0.93 d
9d99304311395910.40 c/0.39 d
9e91306313393870.60 c/0.58 d
9f93290291373830.14 c/0.14 d
335335376410.27 c/0.26 d
9g882862923871010.76 c/0.75 d
31731838871>0.98 c,d,e
9h973003044021020.80 c/0.78 d
32332340279>0.98 c,d,e
9i983013044031020.78 c/0.77 d
32332340279>0.98 c,d,e
9j912932954091160.81 c/0.79 d
32032740888>0.98 c,d,e
9k86263----
a isolated yield. b stokes shift (Δ) from the equation Δ   =   λ m a x e m λ m a x a b s . Wavelength determined from the 3D emission spectrum for C = 5.0 × 10−6 M CH2Cl2 solution (nm). c quinine sulfate was used as a standard [62]. d trans,trans-1,4-diphenyl-1,3-butadiene was used as a standard [63]. e exact value cannot be determined due to the nonlinearity of standard/sample dependence in the Φ region 0.97–1.00 [64].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Olesiejuk, M.; Kudelko, A.; Świątkowski, M. Highly Luminescent 4H-1,2,4-Triazole Derivatives: Synthesis, Molecular Structure and Photophysical Properties. Materials 2020, 13, 5627. https://doi.org/10.3390/ma13245627

AMA Style

Olesiejuk M, Kudelko A, Świątkowski M. Highly Luminescent 4H-1,2,4-Triazole Derivatives: Synthesis, Molecular Structure and Photophysical Properties. Materials. 2020; 13(24):5627. https://doi.org/10.3390/ma13245627

Chicago/Turabian Style

Olesiejuk, Monika, Agnieszka Kudelko, and Marcin Świątkowski. 2020. "Highly Luminescent 4H-1,2,4-Triazole Derivatives: Synthesis, Molecular Structure and Photophysical Properties" Materials 13, no. 24: 5627. https://doi.org/10.3390/ma13245627

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