Bis-arylsulfenyl- and bis-arylselanyl-benzo-2,1,3-thiadiazoles: synthesis and photophysical characterization

Renata A. Balagueza, Vanessa G. Ricordia, Rodrigo C. Duarteb, Josene M. Toldoc, Cristtofer M. Santosc, Paulo H. Schneiderd, Paulo F. B. Gonçalvesc, Fabiano S. Rodembusch*b and Diego Alves*a
aLaboratório de Síntese Orgânica Limpa, Universidade Federal de Pelotas – UFPel, PO Box 354 – CEP 96010-900, Pelotas, RS, Brazil. E-mail: diego.alves@ufpel.edu.br; Fax: +55 53 32757533; Tel: +55 53 32757533
bGrupo de Pesquisa em Fotoquímica Orgânica Aplicada, Universidade Federal do Rio Grande do Sul – Instituto de Química, Avenida Bento Gonçalves 9500, CEP 91501-970, Porto Alegre, RS, Brazil. E-mail: fabiano.rodembuschs@ufrgs.br; Fax: +55 51 33087204; Tel: +55 51 33087204
cGrupo de Química Teórica e Computacional, Universidade Federal do Rio Grande do Sul – Instituto de Química, Avenida Bento Gonçalves, 9500, CP 15003, CEP 91501-970, Porto Alegre, RS, Brazil
dUniversidade Federal do Rio Grande do Sul, Instituto de Química, Departamento de Química Orgânica, Av. Bento Gonçalves, 9500, Agronomia, CEP 91501-970, PO Box 15003, Porto Alegre, RS, Brazil. E-mail: paulos@iq.ufrgs.br; Tel: +55 51 33089636

Received 15th February 2016 , Accepted 3rd May 2016

First published on 5th May 2016


Abstract

Bis-arylsulfenyl- and bis-arylselanyl-benzo-2,1,3-thiadiazoles were synthesized in good yields by copper-catalysed cross-coupling reaction of arylthiols or diaryl diselenides with the commercially available 4,7-dibromobenzo[c][1,2,5]thiadiazole. The arylsulfenyl derivatives present absorptions in the visible region (∼420 nm) with molar absorptivity coefficient and radiative rate constant values ascribed to spin and symmetry allowed π–π* electronic transitions, with almost complete absence of solvatochromic effect. An emission located in the cyan green to green region (514–570 nm), with a large Stokes shift (90–146 nm) was observed, probably associated to the charge transfer character of the S1 state. Theoretical calculations were also performed in order to study the geometry, charge distribution and photophysical properties of the molecules in their ground and excited electronic states. TD-DFT calculations were performed using the PBE1PBE and CAM-B3LYP functionals with cc-pVDZ basis set for geometrical optimisations in the S0 and S1 states and jun-cc-pVTZ basis set to obtain vertical transition energies and electronic properties. Solvent effects were included by IEF-PCM formalism using solvents with different dielectric constants. The computationally predicted transition energies calculated with CAM-B3LYP are in good agreement with the experimental results. No substantial solvatochromic effect was found in the absorption maxima, but in the emission from S1 state a redshift was observed on increasing the solvent polarity. This fact, combined with higher dipole moment in the first excited state and some spatial separation of HOMO and LUMO orbitals could indicate an intramolecular charge transfer character of the S1 state.


Introduction

Organyl sulfides, represent special scaffolds usually found in various naturally occurring and biologically active compounds1 as well as being employed as important intermediates in organic synthesis.2 Thus, the development of efficient and new methodologies for the synthesis of these compounds has emerged in organic synthesis. Traditional methods for the formation of C–S bonds frequently require harsh reaction conditions2 and recently considerable attention has been devoted to transition metal-catalysed protocols for this purpose.3 In this sense, copper-catalysed reactions using specific ligands and additives are probably the most important method for C–S bond formation. Copper-catalysed protocols have been widely studied using organic disulfides or thiols, as sulfur source, and a plethora of organyl sulfides, including sulfenyl heterocycles, was synthesized in high yields.4

In the context of heterocyclic compounds, benzo-2,1,3-thiadiazole (BTD) derivatives comprise an interesting class of molecules and possess interesting electro-optical properties.5 Their charge transport capability make them attractive candidates for organic light-emitting diodes (OLEDs).6 In addition, BTD derivatives have received much attention in recent years because of their use as fungicides, herbicides and antibacterials,7 and were employed as bioprobes for the analyses of numerous cell types.8

Recently, a range of arylsulfonyl-BTDs 6 was synthesized and used as pyruvate kinase M2 (PKM2) modulators for the treatment of cancer.9 In this study the authors described the synthesis of sulfenyl-benzo-2,1,3-thiadiazoles derivatives via sequential palladium-catalysed reactions and subsequent oxidation of sulfides to arylsulfonyl-BTDs 6 using H2O2 and AcOH (eqn (1), Scheme 1).9 Sulfenyl-benzo[1,2-c:4,5-c]bis([1,2,5]thiadiazole) derivatives 7 were synthesized by Yamashita and co-workers by reaction of benzo[1,2-c:4,5-c]bis([1,2,5]thiadiazole) with thiophenol in DMF at 80 °C under argon atmosphere (eqn (2), Scheme 1).10


image file: c6ra04157d-s1.tif
Scheme 1 Previous works on synthesis of sulfur-containing benzo-2,1,3-thiadiazoles.

To the best of our knowledge, however, the direct synthesis and photophysical characterization of functionalized bis-arylsulfenyl-benzo-2,1,3-thiadiazoles has not been explored. In this context and in continuation of our interest in the synthesis of heterocycles bearing organochalcogen moieties, here we describe the direct copper-catalysed synthesis of arylsulfenyl-benzo-2,1,3-thiadiazoles by reaction of arylthiols with 4,7-dibromobenzo[c][1,2,5]thiadiazole 1. In addition, we dedicate our efforts to understand the photophysics of the synthesized molecules comparing experimental and theoretical properties predicted with TD-DFT calculations.

Experimental

General information

The reactions were monitored by TLC carried out on Merck silica gel (60 F254) by using UV light as visualizing agent and 5% vanillin in 10% H2SO4 and heat as developing agents. Baker silica gel (particle size 0.040–0.063 mm) was used for flash chromatography. Proton nuclear magnetic resonance spectra (1H NMR) were obtained at 300 MHz on Bruker DPX 300 spectrometer. Spectra were recorded in CDCl3 solutions. Chemical shifts are reported in ppm, referenced to tetramethylsilane (TMS) as the external reference. Coupling constants (J) are reported in hertz. Abbreviations to denote the multiplicity of a particular signal are s (singlet), d (doublet), dd (doublet of doublet) and m (multiplet). Carbon-13 nuclear magnetic resonance spectra (13C NMR) were obtained at 75 MHz on Bruker DPX 300 spectrometer. Chemical shifts are reported in ppm, referenced to the solvent peak of CDCl3. Low-resolution mass spectra were obtained with a Shimadzu GC-MS-QP2010 mass spectrometer. High resolution mass spectra (HRMS) were recorded on a Bruker Micro TOF-QII spectrometer 10416. The solvents and reagents were used as received or purified using standard procedures. Spectroscopic grade solvents (Merck) were used for fluorescence and UV-Vis measurements. UV-Vis absorption spectra in solution were performed on a Shimadzu UV-2450 spectrophotometer at a concentration range of 10−4 to 10−5 M. Steady state fluorescence spectra were taken with a Shimadzu spectrofluorometer model RF-5301PC. The maximum absorption wavelength was used as excitation wavelength for fluorescence measurements. The quantum yield of fluorescence (ΦF) was measured at 25 °C using spectroscopic grade solvents within solutions with absorbance intensity lower than 0.05 (optical dilute methodology). Coumarin 343 (Aldrich) in ethanol was used as fluorescence quantum yield standard (ΦF = 0.63).11 All measurements were performed at room temperature (25 °C).

General procedure for the synthesis of bis-arylsulfenyl-benzo-2,1,3-tiadiazoles 3a–f

To a 5 mL round-bottomed flask containing the 7-dibromobenzo[c][1,2,5]thiadiazole 1 (0.5 mmol), the appropriated arylthiol 2a–f (1.0 mmol), CuO nanoparticles (CuO NPs) (20 mol%), KOH (2.0 mmol) and DMSO (1.5 mL) were added. The homogeneous reaction mixture was stirred at 80 °C for 24 hours under N2 atmosphere. After this time, the solution was cooled to room temperature, diluted with ethyl acetate (20 mL), and washed with water (3 × 20 mL). The organic phase was separated, dried over MgSO4 and concentrated under vacuum. The obtained products 3a–f were purified by chromatography on neutral alumina using a mixture of ethyl acetate/hexane (10[thin space (1/6-em)]:[thin space (1/6-em)]90) as the eluent.
4,7-Bis((4-methoxyphenyl)thio)benzo[c][1,2,5]thiadiazole (3a). Yield: 0.179 g (87%); orange oil. 1H NMR (CDCl3, 300 MHz): δ 7.34 (d, J = 8.8 Hz, 4H); 6.77 (d, J = 8.8 Hz, 4H); 6.57 (s, 2H); 3.67 (s, 6H). 13C NMR (CDCl3, 75 MHz): δ 160.32, 152.21, 136.36, 129.66, 125.52, 120.67, 115.11, 55.18. MS (relative intensity) m/z: 412 (59), 273 (100), 207 (23), 139 (28), 96 (16), 77 (11). HRMS calcd for C20H17N2O2S3 [M + H]+ 413.0446. Found: 413.0422.
4,7-Bis(phenylthio)benzo[c][1,2,5]thiadiazole (3b). Yield: 0.151 g (86%); yellow solid; mp 125–126 °C. 1H NMR (CDCl3, 300 MHz): δ 7.42–7.39 (m, 4H); 7.27–7.25 (m, 6H); 6.84 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 152.89, 133.47, 131.41, 129.56, 128.91, 128.60, 127.64. MS (relative intensity) m/z: 352 (60), 243 (93), 207 (48), 109 (40), 77 (100), 51 (88). HRMS calcd for C18H13N2S3 [M + H]+ 353.0235. Found: 353.0238.
4,7-Bis(p-tolylthio)benzo[c][1,2,5]thiadiazole (3c). Yield: 0.158 g (83%); orange oil. 1H NMR (CDCl3, 300 MHz): δ 7.34 (d, J = 8.1 Hz, 4H); 7.12 (d, J = 8.1 Hz, 4H); 6.76 (s, 2H); 2.29 (s, 6H). RMN 13C (CDCl3, 75 MHz): δ 152.77, 139.13, 134.14, 130.45, 129.35, 127.48, 126.84, 21.22. MS (relative intensity) m/z: 380 (14), 281 (25), 257 (25), 207 (100), 73 (20), 40 (41). HRMS calcd for C20H17N2S3 [M + H]+ 381.0548. Found: 381.0548.
4,7-Bis((4-chlorophenyl)thio)benzo[c][1,2,5]thiadiazole (3d). Yield: 0.168 g (80%); orange solid; mp 110–112 °C. 1H NMR (CDCl3, 300 MHz): δ 7.34 (d, J = 8.5 Hz, 4H); 7.25 (d, J = 8.5 Hz, 4H); 6.91 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 152.89, 134.94, 134.58, 130.07, 129.81, 128.75, 127.97. MS (relative intensity) m/z: 420 (15), 277 (100), 242 (39), 207 (17), 108 (32), 75 (26). HRMS calcd for C18H11Cl2N2S3 [M + H]+ 420.9455. Found: 420.9451.
4,7-Bis((4-fluorophenyl)thio)benzo[c][1,2,5]thiadiazole (3e). Yield: 0.173 g (89%); yellow solid; mp 132–134 °C. 1H NMR (CDCl3, 300 MHz): δ 7.44 (dd, J = 8.8 and 5.2 Hz, 4H); 7.00 (t, J = 8.8 Hz, 4H); 6.77 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 163.14 (d, J = 249.7 Hz), 152.65, 136.17 (d, J = 8.4 Hz), 129.27, 126.93, 126.26 (d, J = 3.7 Hz), 116.87 (d, J = 22.0 Hz). MS (relative intensity) m/z: 388 (23), 281 (27), 261 (36), 207 (100), 73 (23). HRMS calcd for C18H11F2N2S3 [M + H]+ 389.0046. Found: 389.0043.
4,7-Bis(naphthalen-1-ylthio)benzo[c][1,2,5]thiadiazole (3f). The yield: 0.203 g (90%); yellow solid; mp 104–106 °C. 1H NMR (CDCl3, 300 MHz): δ 7.92 (m, 2H); 7.71–7.63 (m, 6H); 7.41–7.37 (m, 6H); 6.85 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 152.92, 133.70, 132.91, 132.81, 129.93, 129.26, 128.88, 128.59, 127.82, 127.65, 127.54, 126.83, 126.65. MS (relative intensity) m/z: 452 (2), 368 (11), 111 (25), 83 (23), 55 (100), 43 (93). HRMS calcd for C26H17N2S3 [M + H]+ 453.0548. Found: 453.0547.

General procedure for the synthesis of bis-arylselanyl-benzo-2,1,3-tiadiazoles 5a–d

To a 3 mL round-bottomed flask containing 7-dibromobenzo[c][1,2,5]thiadiazole 1 (0.5 mmol), an appropriated diaryl diselenide 4a–d (0.5 mmol), CuI (20 mol%), 1,10-phenanthroline (1,10-phen) (20 mol%), KOH (2.0 mmol) and DMSO (1.5 mL) were added. The homogeneous reaction mixture was stirred at 110 °C for 24 hours under N2 atmosphere. After this time, the solution was cooled to room temperature, diluted with ethyl acetate (20 mL), and washed with water (3 × 20 mL). The organic phase was separated, dried over MgSO4 and concentrated under vacuum. The obtained products 5a–d were purified by chromatography on neutral alumina using a mixture of ethyl acetate/hexane (10[thin space (1/6-em)]:[thin space (1/6-em)]90) as the eluent.
4,7-Bis(phenylselanyl)benzo[c][1,2,5]thiadiazole (5a). Yield: 0.175 g (78%); dark yellow solid; mp 129–131 °C. 1H NMR (CDCl3, 300 MHz): δ 7.54–7.50 (m, 4H); 7.25–7.22 (m, 6H); 6.86 (s, 2H). RMN 13C (CDCl3, 75 MHz): δ 153.47, 135.73, 135.36, 129.90, 129.48, 127.05, 125.08. MS (relative intensity) m/z: 448 (9), 291 (25), 207 (14), 77 (79), 51 (51), 40 (100). HRMS calcd for C18H13N2SSe2 [M + H]+ 448.9126. Found: 448.9107.
4,7-Bis(p-tolylselanyl)benzo[c][1,2,5]thiadiazole (5b). Yield: 0.164 g (69%); yellow solid; mp 101–103 °C. 1H NMR (CDCl3, 300 MHz): δ 7.53 (d, J = 8.1 Hz, 4H); 7.15 (d, J = 8.1 Hz, 4H); 6.89 (s, 2H); 2.36 (s, 6H). 13C NMR (CDCl3, 75 MHz): δ 153.50, 139.12, 136.04, 130.56, 129.43, 125.41, 123.22, 21.24. MS (relative intensity) m/z: 476 (5), 305 (12), 111 (17), 81 (57), 69 (100), 43 (88). HRMS calcd for C20H17N2SSe2 [M + H]+ 476.9439. Found: 476.9392.
4,7-Bis((4-chlorophenyl)selanyl)benzo[c][1,2,5]thiadiazole (5c). Yield: 0.188 g (73%); orange solid; mp 87–89 °C. 1H NMR (CDCl3, 300 MHz): δ 7.47 (d, J = 8.4 Hz, 4H); 7.23 (d, J = 8.4 Hz, 4H); 6.93 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 153.26, 136.53, 134.96, 130.00, 129.67, 125.15, 124.73. MS (relative intensity) m/z: 516 (8), 325 (25), 290 (5), 81 (63), 69 (100), 41 (33). HRMS calcd for C18H11Cl2N2SSe2 [M + H]+ 516.8339. Found: 516.8318.
4,7-Bis((4-fluorophenyl)selanyl)benzo[c][1,2,5]thiadiazole (5d). Yield: 0.189 g (78%); orange solid; mp 151–153 °C. 1H NMR (CDCl3, 300 MHz): δ 7.62 (dd, J = 8.8 and 5.4 Hz, 4H); 7.03 (t, J = 8.8 Hz, 4H); 6.89 (s, 2H). 13C NMR (CDCl3, 75 MHz): δ 163.39 (d, J = 249.8 Hz), 141.03, 138.06 (d, J = 8.1 Hz), 129.72, 125.48, 121.87 (d, J = 3.6 Hz), 117.03 (d, J = 22.0 Hz). MS (relative intensity) m/z: 484 (39), 324 (40), 309 (100), 229 (30), 83 (35), 69 (24). HRMS calcd for C18H11F2N2SSe2 [M + H]+ 484.8937. Found: 484.8906.

Theoretical calculations

DFT and TDDFT calculations were used to provide information about the electronic structure of 3a–f and 5a–d. Among the functionals, PBE1PBE12 has provided good agreement with experimental results and has been one of the most used functionals to calculate 2,1,3-benzothiadiazole (BTD) derivatives.13 Thus this functional was initially choose to calculate the electronic spectrum of the selected structures. The ground electronic state and first excited state equilibrium geometries were obtained by optimising the geometry of each species by using PBE1PBE functional together with cc-pVDZ basis set. Subsequently, a single-point energy and population evaluation was performed using the same functional, but changing the basis function to the jun-cc-pVTZ basis set to obtain the vertical absorptions and emissions.14 The jun-basis set, named calendar basis set, is recommended by Truhlar et al. as a better default option than aug-basis sets because of its performance versus cost. The calendar basis set are constructed by removing diffuse functions from the aug basis sets. The Jun-cc-pV*Z basis sets remove the diffuse function from H and He and also removes the highest angular momentum diffuse function from all other atoms, from aug-cc-pV*Z. All equilibrium geometries were confirmed by vibrational analysis and have no imaginary frequencies. Solvent effects were included in all cases by the Integral Equations Formalism of the Polarisable Continuum Model (IEF-PCM).15 The solvents considered at PBE1PBE level were hexane, toluene, dichloromethane, ethanol and N,N-dimethylformamide. In addition, the ground and first excited state were also optimised and have their electronic transitions calculated using CAM-B3LYP16 functional with the same basis set previously mentioned. With this functional, the solvent effects were considered using hexane, toluene, 1,4-dioxane and dichloromethane as solvents. The main difference between CAM-B3LYP and PBE1PBE is the long-range corrections considered by the first one, which provides a more accurate description of charge-transfer excited states.16,17 Thus, this functional was chosen because of its good results in molecules with donor–acceptor design containing the benzothiadiazole group.18,19 The ground and excited state electrostatic potential surfaces and dipole moment were obtained by population analysis using ChelpG formalism (charges from electrostatic potential using a grid based method)20 and self-consistent field density and CI density respectively, at the corresponding optimised geometries. The calculations were carried out using Gaussian 09 Package.21

Results and discussion

Synthesis

Initially, 4,7-dibromobenzo[c][1,2,5]thiadiazole 1 and 4-methoxythiophenol 2a were selected as a model system to optimize the reaction conditions. Thus, in a first attempt, a mixture of substrates 1 (0.5 mmol) and 2a (1.0 mmol) were stirred in DMSO at 110 °C in N2 atmosphere for 24 h, using CuO NPs (5 mol%) as a catalyst and KOH (2.0 eq.) as base.22 Under these reaction conditions, only traces of the desired product 3a were observed (Table 1, entry 1). When the same reaction was performed using 4.0 eq. of KOH, the desired product 3a was obtained, however in low yield (Table 1, entry 2). Fortunately, when the amount of catalyst was increased from 5 to 20 mol%, moderate to good yields of product 3a were obtained (Table 1, entries 3 and 4). A good result was achieved when we performed this reaction at 80 °C using 20 mol% of CuI and 4.0 eq. of KOH, with the product 3a being obtained in 86% yield (Table 1, entry 5). When the reaction was performed using 4.0 eq. of KOH without CuO NPs, the desired product 3a was not formed (Table 1, entry 6). A decrease in the yield of 3a was observed when this reaction was performed using the catalytic system CuI (20 mol%)23 and 1,10-phen (20 mol%) in reactions carried out at 80 and 110 °C (Table 1, entries 7 and 8).
Table 1 Optimization of reaction conditionsa

image file: c6ra04157d-u1.tif

Entry KOH (eq.) CuO NPs (mol%) Temperature (°C) Yield of 3ab (%)
a Reactions are performed in the presence of compound 1 (0.5 mmol) and 2a (1.0 mmol), in DMSO (1.5 mL), for 24 h under N2 atmosphere.b Yields are given for isolated product.c Reaction was performed using CuI (20 mol%) and 1,10-phenanthroline (20 mol%) as catalytic system.
1 2.0 5 110 Traces
2 4.0 5 110 25
3 4.0 10 110 43
4 4.0 20 110 75
5 4.0 20 80 87
6 4.0 80
7 4.0 20c 80 46
8 4.0 20c 110 55


In an optimised reaction, 4,7-dibromobenzo[c][1,2,5] thiadiazole 1 (0.5 mmol), 4-methoxythiophenol 2a (1.0 mmol), CuO NPs (20 mol%) and KOH (4.0 eq.) were dissolved in DMSO (1.5 mL) and after that the mixture was stirred at 80 °C for 24 h under N2 atmosphere, affording the correspondent 4,7-bis((4-methoxyphenyl)thio)benzo[c][1,2,5]thiadiazole 3a in 87% yield. Then, the possibility to synthesize the selenium analogues of compound 3a was also studied. Interest in the chemistry and application of different selenium functionalized compounds as potential pharmaceuticals,24 new materials,25 fluorescent molecules26 and ionic liquids27 has expanded over the last years. Thus, a mixture of 4,7-dibromobenzo[c][1,2,5]thiadiazole 1 (0.5 mmol), diphenyl diselenide 4a (0.5 mmol), CuO NPs (20 mol%), KOH (4.0 eq.) and DMSO (1.5 mL) was stirred at 80 °C for 24 h under N2 atmosphere, and the desired product 5a was obtained only in 33% yield (Scheme 2, Condition 1). However, when the reaction of compounds 1 and 4a was performed using the catalytic system CuI (20 mol%) and 1,10-phen (20 mol%), KOH (4.0 eq.) as base, DMSO (1.5 mL) as solvent at 110 °C in N2 atmosphere for 24 h, selenylation product 5a was achieved in 78% yield (Scheme 2, Condition 2).


image file: c6ra04157d-s2.tif
Scheme 2 Synthesis of compound 5a. Condition 1: CuO NPs (20 mol%), KOH (4.0 eq.), DMSO (1.5 mL), 80 °C, N2, 24 h (33%). Condition 2: CuI (20 mol%), 1,10-phen (20 mol%), KOH (4.0 eq.), DMSO (1.5 mL), 110 °C, N2, 24 h (78%).

To extend the scope of our synthetic methodology, the possibility to perform these reactions of benzo[c][1,2,5]thiadiazole 1 with arylthiols 2a–f or diaryl diselenides 4a–d was investigated under our two developed methodologies and the results are presented in Table 2.

Table 2 Synthesis of different 4,7-bis(arylthio)- and 4,7-bis(arylselanyl)- benzo[c][1,2,5]thiadiazolesa

image file: c6ra04157d-u2.tif

Entry Product (yield)b Entry Product (yield)b
a Condition A: reactions are performed with compound 1 (0.5 mmol), arylthiols 2a–f (1.0 mmol), CuO NPs (20 mol%) and KOH (4 eq.) in DMSO (1.5 mL) at 80 °C for 24 h under N2 atmosphere. Condition B: reactions are performed with compound 1 (0.5 mmol), diaryl diselenides 4a–d (0.5 mmol), CuI (20 mol%), 1,10-phen (20 mol%) and KOH (4 eq.) in DMSO (1.5 mL) at 110 °C for 24 h under N2 atmosphere.b Yields are given for isolated product.
1 image file: c6ra04157d-u3.tif 6 image file: c6ra04157d-u4.tif
2 image file: c6ra04157d-u5.tif 7 image file: c6ra04157d-u6.tif
3 image file: c6ra04157d-u7.tif 8 image file: c6ra04157d-u8.tif
4 image file: c6ra04157d-u9.tif 9 image file: c6ra04157d-u10.tif
5 image file: c6ra04157d-u11.tif 10 image file: c6ra04157d-u12.tif
image file: c6ra04157d-u13.tif


Firstly, we employed a range of arylthiols 2a–f and our results reveal that the reactions are not sensitive to the electronic effect of the aromatic ring in the arylthiols (Table 2, entries 1–6). For example, arylthiols containing electron-donating (OMe, Me), electron-withdrawing (Cl, F) and electron-neutral group at the aromatic ring, gave good yields of the desired sulfenyl-benzo[c][1,2,5]thiadiazoles 3a–e (Table 2, entries 1–5). In addition, reaction performed with α-naphthylmercaptan 2f furnished the respective product 3f in 90% yield (Table 2, entry 6). It is worth mentioning that it was also evaluated the reactivity of benzo[c][1,2,5]thiadiazole 1 with other diaryl diselenides 4a–d under the Condition 2 described in Scheme 1. Different diaryl diselenides containing EDG and EWG were reacted with compound 1, affording the respective selanyl-benzo[c][1,2,5]thiadiazoles 5a–d in moderated to good yields (Table 2, entries 7–10). The possible mechanism for these cross-coupling reactions using CuO NPs or CuI involving aryl halides and thiols/dichalcogenides has been well explored and described in some studies.3a,22,23

Photophysical characterisation

The absorption spectra of the bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d) are shown in Fig. 1 using hexane and dichloromethane as solvents. The additional data in 1,4-dioxane and toluene presented similar results and are presented in the ESI. The relevant data from UV-Vis absorption spectroscopy are presented in Table 3, where the solvents are presented in order of increasing dipole moment. The UV-V is spectra allowed the obtention of the pure radiactive lifetimes (τ0) and the radiactive rate constant (k0e) using the Strickler–Berg relations:28
 
image file: c6ra04157d-t1.tif(1)

image file: c6ra04157d-f1.tif
Fig. 1 UV-Vis absorption spectra in solution of the bis-arylsulfenyl- (3b–e) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d) in hexane (top) and dichloromethane (bottom).
Table 3 Relevant photophysical data of the UV-Vis spectra of the bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d), where λabs is the absorption maxima (nm), ε is the molar absorptivity (M−1 cm−1), fe is the calculated oscillator strength, k0e is the calculated radiactive rate constant (108 s−1) and τ0 is the inherent emission lifetime (ns)
Dye Solvent λabs ε fe k0e τ0
3a Hexane 427 3183 0.064 0.35 28.41
Toluene 435 3479 0.071 0.38 26.53
1,4-Dioxane 424 3173 0.067 0.37 27.00
Dichloromethane 432 2418 0.051 0.27 36.39
3b Hexane 416 6445 0.131 0.76 13.16
Toluene 420 5698 0.117 0.66 15.10
1,4-Dioxane 415 6101 0.131 0.76 13.16
Dichloromethane 418 5113 0.098 0.56 17.80
3c Hexane 423 6662 0.133 0.74 13.47
Toluene 426 6927 0.140 0.77 12.98
1,4-Dioxane 423 6445 0.136 0.76 13.14
Dichloromethane 425 5303 0.109 0.60 16.53
3d Hexane 412 4406 0.090 0.53 18.87
Toluene 417 4878 0.101 0.58 17.23
1,4-Dioxane 417 5540 0.118 0.68 14.75
Dichloromethane 414 4023 0.083 0.48 20.63
3e Hexane 415 4457 0.089 0.52 19.39
Toluene 421 4746 0.094 0.53 18.91
1,4-Dioxane 417 4511 0.095 0.55 18.34
Dichloromethane 416 3479 0.073 0.42 23.73
3f Hexane 422 9063 0.188 1.06 9.46
Toluene 430 9754 0.205 1.11 9.04
1,4-Dioxane 422 9500 0.203 1.14 8.75
Dichloromethane 420 8500 0.186 1.05 9.48
5a Hexane 416 1646 0.032 0.19 53.38
Toluene 429 1703 0.037 0.20 50.24
1,4-Dioxane 423 1537 0.032 0.18 55.44
Dichloromethane 421 1323 0.029 0.16 61.25
5b Hexane 421 7586 0.149 0.84 11.93
Toluene 434 7511 0.142 0.75 13.25
1,4-Dioxane 419 7308 0.151 0.86 11.64
Dichloromethane 420 5942 0.124 0.70 14.22
5c Hexane 422 3296 0.063 0.35 28.34
Toluene 428 3608 0.067 0.37 27.24
1,4-Dioxane 422 3533 0.071 0.40 24.96
Dichloromethane 423 2740 0.057 0.32 31.48
5d Hexane 427 5167 0.100 0.55 18.27
Toluene 429 5033 0.092 0.50 19.92
1,4-Dioxane 424 4743 0.098 0.54 18.39
Dichloromethane 424 3877 0.081 0.45 22.19


In this equation [v with combining macron]0 is the wavenumber (energy in 1/λ units) of the maximum of the absorption band and the integral image file: c6ra04157d-t2.tif is the area under the absorption curve from a plot of the molar absorptivity coefficient ε (M−1 cm−1) vs. wavenumber [v with combining macron] (cm−1), related to a single electron oscillator and the pure radiative lifetime τ0 is defined as 1/k0e.29 The oscillator strength (fe) can also be calculated from eqn (2):

 
image file: c6ra04157d-t3.tif(2)

The bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d) present absorption maxima located around 420 and 424 nm, respectively. Additionally, the intense absorption bands observed around 310 nm can be associated to the absorption of the benzo-2,1,3-tiadiazole core. The molar absorptivity coefficient ε values, as well as the calculated radiative rate constant (k0e) for all compounds indicate that spin and symmetry allowed electronic transitions, which could be related to 1ππ* transitions. From Fig. 1 it can be observed that any significative change takes place on the absorption maxima location associated with the different organic groups present in the benzo-2,1,3-thiadiazoles moiety. The same behavior was observed when sulfur atom was changed to selenium. In this way, changes on the substituent in the benzo-2,1,3-tiadiazoles core from bis-arylsulfenyl- to bis-arylselanyl do not affect the photophysics in the ground state of these compounds.

Additionally, a very small solvatochromic effect could be observed in the ground state (Δλabs from 3 to 10 nm) for the bis-arylsulfenyl compounds 3a–f, indicating an almost absent intramolecular charge transfer state (see ESI).

The bis-arylselanyl derivatives 5a–d present more significative solvatochromism in the ground state (Δλabs from 5 to 15 nm), probably due to the better electron delocalization provided by the selenium atom. It could also be observed that changes in the chalcogen atom – sulfur to selenium – shifts the absorption maxima to longer wavelengths, also indicating that the electrons are in the ground state held tighter in the molecular structure of the sulfur derivative than in its selenium analogs.30

The studied compounds presented non constant inherent emission lifetimes τ0. In a general way, the inherent lifetimes directly excited into S0 → S1 electronic transition increases to the arylselanyl derivatives in despite of the sulfur analogues, probably indicating to the selenium derivatives a reduction in the non-radiative relaxation due to more restricted molecular motion. These results indicate an efficient non-radiative energy transfer to compounds 5a–d, which can be related to lower calculated fluorescence quantum yields (Table 4).

Table 4 Relevant photophysical data of the steady-state fluorescence spectra of the bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d), where λem is the emission maxima (nm), ΔλST is the Stokes shift (nm) and ΦF is the fluorescence quantum yield (%)
Dye Solvent λem ΔλST ΦF
3a Hexane 526 99 0.369
Toluene 554 119 0.323
1,4-Dioxane 558 134 0.264
Dichloromethane 570 138 0.199
3b Hexane 515 99 0.260
Toluene 541 121 0.055
1,4-Dioxane 554 139 0.023
Dichloromethane 561 146 0.058
3c Hexane 521 98 0.367
Toluene 547 121 0.340
1,4-Dioxane 551 128 0.207
Dichloromethane 566 141 0.213
3d Hexane 516 104 0.329
Toluene 535 118 0.297
1,4-Dioxane 542 125 0.203
Dichloromethane 555 141 0.167
3e Hexane 512 97 0.373
Toluene 537 116 0.338
1,4-Dioxane 545 128 0.243
Dichloromethane 557 141 0.242
3f Hexane 519 97 0.131
Toluene 546 116 0.250
1,4-Dioxane 553 131 0.147
Dichloromethane 564 144 0.129
5a Hexane 517 101 0.096
Toluene 546 116 0.057
1,4-Dioxane 552 129 0.036
Dichloromethane 564 143 0.072
5b Hexane 514 93 0.101
Toluene 550 116 0.060
1,4-Dioxane 550 131 0.031
Dichloromethane 559 139 0.061
5c Hexane 512 90 0.099
Toluene 542 114 0.079
1,4-Dioxane 550 128 0.054
Dichloromethane 560 137 0.068
5d Hexane 518 91 0.094
Toluene 544 115 0.070
1,4-Dioxane 554 130 0.031
Dichloromethane 562 138 0.089


The normalised fluorescence emission spectra of the bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d) are presented in Fig. 2. The emission curves were obtained exciting the compounds at the absorption maxima wavelength. The relevant data from fluorescence emissions are summarised in Table 4, where the solvents are presented in order of increasing dipole moment.


image file: c6ra04157d-f2.tif
Fig. 2 Fluorescence emission spectra in solution of the bis-arylsulfenyl- (3b–e) and bis-arylselanyl-benzo-2,1,3-tiadiazoles (5a–d) in hexane (top) and dichloromethane (bottom). The absorption maxima were used as excitation wavelengths.

The bis-arylsulfenyl- and bis-arylselanyl-benzo-2,1,3-tiadiazoles present emission in the cyan green to green regions (∼540 nm). As already observed in the ground state, the bis-arylsulfenyl derivatives 3a–f present comparable location to the emission maxima, which indicates that the different organic moieties in the benzo-2,1,3-tiadiazole core do not play a fundamental role on the excited state of these compounds. A similar behavior was observed to the compounds (5a–d). However, a significative solvatochromic effect was observed in the excited state (Δλem from 39 to 45 nm) for the bis-arylsulfenyl compounds 3a–f. The bis-arylselanyl derivatives 5a–d also presented a significative solvatochromism in the excited state (Δλem from 44 to 48 nm), where increasing the dipole moment of the solvent (hexane to dichloromethane), the emission maxima redshifts (Fig. 3).


image file: c6ra04157d-f3.tif
Fig. 3 Normalised fluorescence emission spectra in solution of the derivatives 3b (left) and 5a (right). The absorption maxima were used as excitation wavelengths.

This behavior can be understood when a relaxed excited state is more polar than the ground state (i.e., μe > μg). In this case, the stronger the interaction between the solute and the solvent, the lower the energy of the excited state, the larger the redshift of the emission band.31,32 As already observed to chalcogen compounds, this results can probably also be attributed to the nature of the 1ππ* electronic transition, indicating that these compounds allow better electronic delocalization in the excited state.32 The large Stokes shift, as well as the solvathochromic effect in the excited state indicates that these compounds presents charge separation, such as an intramolecular charge transfer character in the excited state (ICT state). Moreover, the bis-arylsulfenyl derivatives 3a–f presented higher fluorescence quantum yields if compared to the selanyl analogues 5a–d, indicating that the presence of the selenium atom seems to be an efficient nonradiactive deactivation channel in these structures.

Fig. 4 depicts the compound 3f in solution under UV radiation, where a positive solvatochromism from green to yellow region, could be observed by varying the polarity of the solvent. A similar behavior could be observed with the other derivatives.


image file: c6ra04157d-f4.tif
Fig. 4 Picture of the solutions of compound 3f in hexane (left), toluene (middle) and dichloromethane (right) under UV radiation (365 nm).

Theoretical calculations

The main structural characteristic of the bis-arylsulphenyl- and bis-aryselanyl-benzo-2,1,3-thiadiazoles is the loss of planarity. The planar structures are not stable, which means they are not a minimum at the surface potential energy. The lowest energy conformations present the groups attached to the 4- and 7-positions of BTD core almost perpendicular to the core π system. The conformational analysis for the 3b structure can be seen in Fig. 5. The conformation A has a lower energy than all remaining conformations found to the 3b molecule (B to D) at PBE1PBE/jun-cc-pVTZ//PBE1PBE/cc-pVDZ level of theory, in the gas phase. Considering the A conformation found to 3b, the remaining structures were optimised (3a–f and 5a–f).
image file: c6ra04157d-f5.tif
Fig. 5 Conformational analysis for the 3b structure at PBE1PBE/jun-cc-pVTZ//PBE1PBE/cc-pVDZ level of theory, in the gas phase. Energies (E) in Hartree and ΔE (energy difference in relation to the more stable conformation, A) in kcal mol−1.

The most stable conformations for the ground and excited state obtained using CAM-B3LYP and hexane as solvent are represented in Fig. 6. The most relevant calculated structural parameters obtained with both functionals and all solvents are provided in the ESI together with the optimised structures at PBE1PBE level (Table ESI1). Due to the fact that PBE1PBE did not describe well the systems (see discussion below), all geometric parameters are discussed based on CAM-B3LYP functional.


image file: c6ra04157d-f6.tif
Fig. 6 Molecular geometries for 3a–f (left) and 5a–d (right) series at ground state (on the left) and first excited state (on the right) calculated at CAM-B3LYP/cc-pVDZ level in hexane.

In the ground state, all structures presents similar geometries, with symmetry near to C2v, independent of the R substituent attached to phenyl ring or if it is S or Se. Due to the symmetry, some angles and bond lengths are the same in both sides of the molecule and are not explicitly presented (Table ESI1). Both S0 and S1 states have a similar geometry, indicating a high stability of those structures. This reflects a small geometrical reorganization in both transitions. The main structural difference between S0 and S1 states is the change in the dihedral angle between phenyl rings and the benzothiadiazole core (d1). In the ground state, d1 is close to 91 degrees, except in the 3f structure where the dihedral angle is close to 82 degrees. Moving to the excited state, this dihedral angle becomes to be equal to 66–70 degrees for 3a–d and 82–87 degrees for 5a–c. Although in the ground state there is no significant difference between the geometric parameters obtained with both functionals, using PBE1PBE the values obtained for d1 in the first excited state are smaller (50–57 degrees in the structures 3a–e, ∼78 in the 3f and 57–61 in the 5a–d). Another important change from S0 to S1 is the bond length between N–S2 (r3), that increase in the excited state, while the C7–S17 (r1) decreases. In fact, in the ground state all the structural parameters are very close to each other, but in excited state the dependence with R group attached to the phenyl rings and sulfur/selenium are reflected mainly on the dihedral angle d1. Moreover, the change in the dihedral angle d1 is the main structural difference calculated by the two functionals. No significant structural changes were observed when the solvent is changed.

The lack of planarity in 2,1,3-benzothiadiazole derivative may also play a role in the long wavelength absorption bands. Decreasing conjugation is known for increasing the HOMO–LUMO gap of conjugated molecules.18 The calculated absorption and emission wavelengths are shown in Table 5 together with their respective oscillator strength and dipole moments.

Table 5 Calculated photophysical data of the bis-arylsulfenyl- (3a–f) and bis-arylselanyl-benzo-2,1,3-thiadiazoles (5a–d) with CAM-B3LYP/jun-cc-pVTZ//CAM-B3LYP/cc-pVDZ. The λabs is the absorption maxima (nm), λem is the emission maxima, fe is the oscillator strength and μ is the dipole moment (D) for molecules in their respective S0 and S1 electronic states
Dye Solvent S0 S1
λabs fe μ λem fe μ
3a Hexane 413.43 0.187 7.7 547.71 0.170 13.0
1,4-Dioxane 413.97 0.190 7.9 550.25 0.177 13.3
Toluene 414.60 0.195 8.0 551.42 0.180 13.4
CH2Cl2 414.03 0.187 8.9 564.03 0.217 15.1
3b Hexane 410.96 0.184 5.6 541.05 0.166 11.3
1,4-Dioxane 411.58 0.187 5.8 544.10 0.173 11.3
Toluene 412.23 0.192 5.8 544.74 0.176 11.4
CH2Cl2 411.83 0.184 6.6 558.72 0.216 12.9
3c Hexane 413.11 0.186 6.8 545.19 0.168 12.3
1,4-Dioxane 413.68 0.189 6.9 547.80 0.175 12.6
Toluene 414.32 0.194 7.0 548.91 0.178 12.7
CH2Cl2 413.82 0.186 7.9 562.85 0.216 14.3
3d Hexane 406.24 0.193 1.8 534.79 0.180 7.3
1,4-Dioxane 406.86 0.196 1.9 537.14 0.187 7.5
Toluene 407.50 0.201 1.9 538.16 0.189 7.6
CH2Cl2 407.55 0.192 2.3 551.84 0.226 8.7
3e Hexane 407.27 0.186 2.3 530.07 0.159 7.9
1,4-Dioxane 407.93 0.189 2.3 532.67 0.166 8.2
Toluene 408.59 0.194 2.4 533.79 0.169 8.3
CH2Cl2 408.71 0.185 2.9 549.02 0.208 9.4
3f Hexane 411.69 0.190 5.5 536.81 0.161 10.9
1,4-Dioxane 411.91 0.195 5.6 539.20 0.168 11.6
Toluene 412.49 0.200 5.7 540.22 0.171 11.7
CH2Cl2 411.63 0.193 6.5 553.04 0.210 13.3
5a Hexane 419.06 0.188 5.3 543.59 0.164 10.7
1,4-Dioxane 419.43 0.192 5.5 545.82 0.172 11.0
Toluene 420.01 0.197 5.5 546.78 0.175 10.6
CH2Cl2 418.69 0.189 6.2 559.55 0.220 12.5
5b Hexane 420.96 0.190 6.5 546.93 0.165 12.0
1,4-Dioxane 421.89 0.198 6.7 550.11 0.176 12.4
Toluene 421.31 0.193 6.7 549.15 0.173 12.3
CH2Cl2 420.45 0.190 7.6 562.50 0.22 13.9
5c Hexane 414.53 0.196 1.4 536.27 0.171 6.9
1,4-Dioxane 414.91 0.199 1.4 538.49 0.178 7.1
Toluene 415.49 0.204 1.4 539.45 0.182 7.2
CH2Cl2 414.67 0.196 1.8 552.55 0.225 8.2
5d Hexane 415.46 0.191 1.9 537.60 0.166 7.3
1,4-Dioxane 415.87 0.194 1.9 539.82 0.173 7.5
Toluene 416.47 0.199 2.0 540.88 0.177 7.6
CH2Cl2 415.70 0.190 2.4 554.42 0.221 8.7


It is possible to see that, in contrast to the other works, the PBE1PBE functional did not describe well λabs and λem of these systems, probably because the chalcogen atoms. This functional predicted a higher λabs and an even higher λem compared to the experimental data, mainly for the 5a–d structures (see Table ESI2). On the other hand, the theoretical results calculated using CAM-B3LYP functional are very close to the experimental data, mainly in the ground state, validating the rationality of our method and basis set chosen. With this functional, the calculated absorption maxima obtained for the bis-aryl-sulfenyl-benzo-2,1,3-thiadiazoles (3a–f) is located around 410 nm and for the bis-aryl-selenyl-benzo-2,1,3-thiadiazoles (5a–d) the absorption maxima is located around 417 nm. The weak oscillator strengths presented are also associated with the lack of planarity since the more planar conformation, the higher oscillator strength.

No significant change takes place on the λabs related to different organic groups attached to the benzo-2,1,3-thiadiazole core. The same behavior was observed when sulfur atom was changed by selenium, in accordance with experimental data, although with Se the absorption maxima is shifted to longer wavelength. In addition, a quite similar absorption wavelength was obtained for all solvents, indicating the absence of the solvatochromic effect.

The emission maxima wavelength showed for 3a–e compounds are 530–564 nm, 536–553 nm for 3f and 536–562 nm for 5a–d. In the same way as in the ground state, a small influence of the substituent group attached to the benzene ring on the emission spectra was found. However, the presence of the halogen atom at this position slightly decreases both the emission and absorption wavelength. This could be explained by the inductive effect, i.e. simple electrostatic effects which arise from changes in electron distribution. It is known that the substitution of fluorine in positions close to a chromophore group, but not conjugated with it, leads to short wavelength shifts of the absorption bands of the more highly excited states due to the inductive effect of the fluorine atom.33 The relative magnitude of this effect differs in the ground and excited states and depends upon the degree of excitation. When there are a large overlap between the upper orbitals and the substituent group, energy changes in the excited orbitals could be produced, which make the band appear to behave anomalously on substitution.33

In contrast to the almost absent solvent dependence in absorption spectra, the fluorescence emission spectra displays a clear solvent dependence, with a bathochromic shift in the emission wavelength maxima when solvent polarity is increased (from hexane to dichloromethane around 18 nm in 3a–e and around 16 nm in 3f and 5a–d). This reflects the larger change in dipole moment from S0 to S1. Since the molecules are more polar in the excited state, more polar solvents will stabilize more the S1 than S0, resulting in emission at lower energies, as the solvent polarity is increased. In addition, the oscillator strengths also are increased from hexane to dichloromethane in excited state. Absorption spectrum is less sensitive to solvent polarity because the molecule is exposed to the same local environmental in the ground and excited states.34 The representative HOMO and LUMO molecular orbitals are shown in Fig. 7 for 3b and 5a structures.


image file: c6ra04157d-f7.tif
Fig. 7 Representative HOMO and LUMO molecular orbitals obtained for 3b (above) and 5a (below) at CAM-B3LYP level using hexane as solvent.

No significant differences were observed between HOMO and LUMO surface contour for the remaining structures or even when solvent is changed. Both HOMO and LUMO are π type, but the LUMO orbital is mostly centered on the BTD ring as a whole, whereas the HOMO is mainly located on the phenyl portion of the BTD ring (donor) and extend the conjugation over the chalcogen atom attached to it. This feature is important to understand de ICT contribution to the excited state. The vertical absorption from S0 → S1 is a 1ππ* electronic transition from HOMO to LUMO orbitals, according to the molar absorptivity coefficient values and radiative rate constant obtained experimentally. Although changing the chalcogen atom also does not affect appreciably the HOMO and LUMO, one interesting feature of these plots is the small lobes centred on sulfur atom in the LUMO that is not present changing the chalcogen to Se.

The large Stokes shift and dipole moment in the first excited state together with the solvatochromic effect in emission energies and a observed spatial separation of the HOMO and LUMO orbitals are strong evidences to ICT character of the S1 electronic state.35,36 The intensity and position of the intramolecular charge transfer bands depends on the nature of the structure and substituents influencing the donor and acceptor character. According to the literature, planar π-conjugated moiety linked with distorted π-substituents associated with well separated HOMO and LUMO distributions are a clear indication of an efficient stabilizing ICT.37

In the systems studied by us, although there is a π-conjugation of the chalcogen atoms with the donor part of the molecule, the π-extension is not so efficient since the RC6H4 unit is not conjugated with the 2,1,3-benzothiadiazole ring. In addition, since the LUMO orbital is delocalized over all 2,1,3-benzothiadiazole core, the charge separation is affected. This feature could decreases the band intensity related to an intramolecular charge transfer. Further details about the less energetic electronic transitions and molecular orbitals involved on it can be found in the ESI (Tables ESI3, ESI4 and Fig. ESI1–4).

Additionally, one more intense band than the first one is presented in all theoretical calculations. This band is observed around 278 nm, 281 nm and 273 nm, for 3a–e, 3f and 5a–d structures, respective and could be related to the local excitation (LE) 1ππ* at the 2,1,3-benzothiadoazole core. This band involves a transition between an occupied molecular orbital of the lower energy than HOMO, with π symmetry and fully spread over at 2,1,3-benzothiadiazole core, to LUMO. Increasing the solvent polarity, this band becomes less intense.

In Fig. 8, the electronic changes to which structures 3a–f and 5a–d are subjected when excitation occurs can be observed.


image file: c6ra04157d-f8.tif
Fig. 8 Electrostatic potential surfaces of 3a–f and 5a–d structures in the ground electronic state (S0) and first electronic excited state (S1). The highest electron density potential is represented in red and the lowest electron density is represented in blue.

In each case, significant displacement of charge and increase in the magnitude of the dipole moment takes place upon excitation (see Table 5). The dipole moment is considerably higher in excited states than in ground states. With the increase in dielectric constant of solvent, the dipole moment also increase, as result of the higher stabilization of molecules when solvent is more polar, particularly at the excited state. The main effect of the halogen atom (bonded at para position) on the dipole moment is its significant decrease with respect to the other structures in S0 and S1. Additionally, the sulfur analogues present the higher dipole moment in the ground and excited state. It can be seen that upon excitation to S1 state, the electrostatic potential over the thiophene unit increases, particularly over the sulfur atom, indicating an increase of the local electronic density. At the same time, it can be seen a decrease in electronic density over the side benzene ring when –Cl or –F are attached to it. No significant differences can be seen changing the chalcogen atom or changing the solvent.

Conclusions

In summary, new classes of benzo-2,1,3-thiadiazoles derivatives containing an organochalcogen moiety in their structures were synthesized and photophysically characterized. These compounds were synthesized in good yields by copper-catalysed cross-coupling reaction of arylthiols or diaryl diselenides with 4,7-dibromobenzo[c][1,2,5]thiadiazole. The obtained compounds present absorptions in the visible region with molar absorptivity coefficient and radiative rate constant values ascribed to spin and symmetry allowed π–π* electronic transitions. An emission located in the cyan green to green regions with a large Stokes shift (90–146 nm) was observed, probably due to intramolecular charge transfer state. DFT and TD-DFT calculations were performed and showed to be in full accordance with experimental data, at CAM-B3LYP level. The non-planarity of the 3a–f and 5a–d molecules increases the HOMO/LUMO gap, decreases the oscillator strength and shifts the absorption spectra to the smaller wavelength. The lowest excited state can be ascribed to be an intramolecular charge-transfer, but its intensity is decreased by the non-planar architecture of these systems, leading to a less efficient HOMO/LUMO separation. The vertical electronic transitions of all the molecules in lower energy absorption bands are a π → π* type. Changing the chalcogen atom from S to Se, there are no significant differences in the absorption and emission spectra, but λabs are slightly redshifts while λem are slightly blueshifts. Increasing the solvent polarity does not affect the maximum absorption wavelength but redshift the maximum emission wavelength.

Acknowledgements

We are grateful for the financial support and scholarships from the Brazilian agencies CNPq (400150/2014-0 and 447595/2014-8), FAPERGS (PRONEM 11/2024-9) and CAPES and Instituto Nacional de Inovação em Diagnósticos para a Saúde Pública (INDI-Saúde). Theoretical calculations carried out with the support of Centro Nacional de Supercomputação (CESUP), Universidade Federal do Rio Grande do Sul (UFRGS).

Notes and references

  1. (a) A. Gangjee, Y. Zeng, T. Talreja, J. J. McGuire, R. L. Kisliuk and S. F. Queener, J. Med. Chem., 2007, 50, 3046 CrossRef CAS PubMed; (b) W. Hu, Z. Guo, F. Chu, A. Bai, X. Yi, G. Cheng and J. Li, Bioorg. Med. Chem. Lett., 2003, 13, 1153 CrossRef; (c) P. Caboni, R. E. Sammelson and J. E. Casida, J. Agric. Food Chem., 2003, 24, 7055 CrossRef PubMed; (d) Y. Wang, S. Chackalamannil, Z. Hu, J. W. Clader, W. Greenlee, W. Billard, H. I. I. I. Binch, G. Crosby, V. Ruperto, R. A. Duffy, R. McQuade and J. Lachowicz, Bioorg. Med. Chem. Lett., 2000, 10, 2247 CrossRef CAS PubMed; (e) S. F. Nielson, E. O. Nielson, G. M. Olsen, T. Liljefors and D. Peters, J. Med. Chem., 2000, 43, 2217 CrossRef; (f) G. De Martino, M. C. Edler, G. La Regina, A. Cosuccia, M. C. Barbera, D. Barrow, R. I. Nicholson, G. Chiosis, A. Brancale, E. Hamel, M. Artico and R. Silvestri, J. Med. Chem., 2006, 49, 947 CrossRef CAS PubMed.
  2. (a) P. Metzner and A. Thuillier, in Sulfur Reagents in Organic Synthesis, ed. A. R. Katritzky, O. MethCohn and C. W. Rees, Academic Press, San Diego, 1994 Search PubMed; (b) T. Kondo and T. Mitsudo, Chem. Rev., 2000, 100, 3205 CrossRef CAS PubMed; (c) S. V. Ley and A. W. Thomas, Angew. Chem., Int. Ed., 2003, 42, 5400 CrossRef CAS PubMed.
  3. (a) I. P. Beletskaya and V. P. Ananikov, Chem. Rev., 2011, 111, 1596 CrossRef CAS PubMed; (b) V. P. Ananikov, S. S. Zalesskiy and I. P. Beletskaya, Curr. Org. Synth., 2011, 8, 2 CrossRef CAS.
  4. Some examples: (a) C. G. Bates, R. K. Gujadhur and D. Venkataraman, Org. Lett., 2002, 4, 2803 CrossRef CAS PubMed; (b) F. Y. Kwong and S. L. Buchwald, Org. Lett., 2002, 4, 3517 CrossRef CAS PubMed; (c) C. G. Bates, P. Saejueng, M. Q. Doherty and D. Venkataraman, Org. Lett., 2004, 6, 5005 CrossRef CAS PubMed; (d) W. Deng, Y. Zou, Y. F. Wang, L. Liu and Q. X. Guo, Synlett, 2004, 1254 CAS; (e) S. Fukuzawa, E. Shimizu, Y. Atsuumi, M. Haga and K. Ogata, Tetrahedron Lett., 2009, 50, 2374 CrossRef CAS; (f) D. Zhu, L. Xu, F. Wu and B. S. Wan, Tetrahedron Lett., 2006, 47, 5781 CrossRef CAS; (g) L. Rout, P. Saha, S. Jammi and T. Punniyamurthy, Eur. J. Org. Chem., 2008, 640 CrossRef CAS; (h) E. Sperotto, G. P. M. van Klink, J. G. de Vries and G. van Koten, J. Org. Chem., 2008, 73, 5625 CrossRef CAS PubMed; (i) L. Rout, T. K. Sen and T. Punniyamurthy, Angew. Chem., Int. Ed., 2007, 46, 5583 CrossRef CAS PubMed.
  5. (a) B. A. D. Neto, A. A. M. Lapis, E. N. da Silva Júnior and J. Dupont, Eur. J. Org. Chem., 2013, 228 CrossRef CAS; (b) S. Kato, T. Matsumoto, T. Ishi-i, T. Thiemann, M. Shigeiwa, H. Gorohmaru, S. Maeda, Y. Yamashita and S. Mataka, Chem. Commun., 2004, 2342 RSC; (c) X. Zhang, H. Gorohmaru, M. Kadowaki, T. Kobayashi, T. Ishi-i, T. Thiemann and S. Mataka, J. Mater. Chem., 2004, 14, 1901 RSC; (d) M. Velusamy, K. R. J. Thomas, J. T. Lin, Y. C. Hsu and K. C. Ho, Org. Lett., 2005, 7, 1899 CrossRef CAS PubMed; (e) M. Akhtaruzzaman, N. Kamata, J. Nishida, S. Ando, H. Tada, M. Tomura and Y. Yamashita, Chem. Commun., 2005, 3183 RSC.
  6. (a) K. R. J. Thomas, M. Velusamy, J. T. Lin, S. S. Sun, Y. T. Tao and C. H. Chuen, Chem. Commun., 2004, 2328 RSC; (b) K. R. J. Thomas, J. T. Lin, M. Velusamy, Y. T. Tao and C. H. Chuen, Adv. Funct. Mater., 2004, 14, 83 CrossRef CAS.
  7. (a) S. Mataka, K. Takahashi, T. Imura and M. Tashiro, J. Heterocycl. Chem., 1982, 19, 1481 CrossRef CAS; (b) F. Gozzo, J. Agric. Food Chem., 2003, 51, 4487 CrossRef CAS PubMed; (c) T. Balasankar, M. Gopalakrishnan and S. Nagarajan, Eur. J. Med. Chem., 2005, 40, 728 CrossRef CAS PubMed.
  8. B. A. D. Neto, P. H. P. R. Carvalho and J. R. Correa, Acc. Chem. Res., 2015, 48, 1560 CrossRef CAS PubMed.
  9. G. Kukreja, S. Phukan, J. Kodam, D. M. More, M. V. Uravane, V. P. Palle and R. K. Kamboj, Patent WO2013005157 A1, 2013.
  10. Y. Yamashita, K. One, M. Tomura and S. Tanaka, Tetrahedron, 1997, 53, 10169 CrossRef CAS.
  11. C. Wurth, M. Grabolle, J. Pauli, M. Spieles and U. Resch-Genger, Nat. Protoc., 2013, 8, 1535 CrossRef PubMed.
  12. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158 CrossRef CAS.
  13. (a) T. O. Lopes, D. A. D. S. Filho, A. A. M. Lapis, H. C. B. de Oliveira and B. A. D. Neto, J. Phys. Org. Chem., 2014, 27, 303 CrossRef CAS; (b) F. Pop, A. Amacher, N. Avarvari, J. Ding, L. M. L. Daku, A. Hauser, M. Koch, J. Hauser, S. X. Liu and S. Decurtins, Chem.–Eur. J., 2013, 19, 250 CrossRef PubMed; (c) B. A. D. Neto, J. R. Corrêa, P. H. P. R. Carvalho, D. C. B. D. Santos, B. C. Guido, C. C. Gatto, H. C. B. de Oliveira, M. Fasciotti, M. N. Eberlin and E. N. da Silva Jr, J. Braz. Chem. Soc., 2012, 23, 770 CrossRef CAS; (d) E. H. G. Cruz, P. H. P. R. Carvalho, J. R. Corrêa, D. A. C. Silva, E. B. T. Diogo, J. D. D. S. Filho, B. C. Cavalcanti, C. Pessoa, H. C. B. de Oliveira, B. C. Guido, D. A. D. S. Filho, B. A. D. Neto and E. N. D. S. Júnior, New J. Chem., 2014, 38, 2569 RSC.
  14. E. Papajak, J. Zheng, H. R. Leverentz and D. G. Truhlar, J. Chem. Theory Comput., 2011, 7, 3027 CrossRef CAS PubMed.
  15. (a) S. Miertuš and J. Tomasi, Chem. Phys., 1982, 65, 239 CrossRef; (b) S. Miertuš, E. Scrocco and J. Tomasi, Chem. Phys., 1981, 55, 117 CrossRef; (c) B. Mennucci, E. Cancès and J. Tomasi, J. Phys. Chem. B, 1997, 101, 10506 CrossRef CAS.
  16. T. Yanai, D. Tew and N. Handy, Chem. Phys. Lett., 2004, 393, 51 CrossRef CAS.
  17. M. J. Peach, P. Benfield, T. Helgaker and D. J. Tozer, J. Chem. Phys., 2008, 128, 044118 CrossRef PubMed.
  18. G. Gibson, T. M. McCormick and D. S. Seferos, J. Phys. Chem. C, 2013, 117, 16606 CAS.
  19. J.-Z. Zhang, J. Zhang, H.-B. Li, Y. Wu, H.-L. Xu, M. Zhang, Y. Geng and Z.-M. Su, J. Power Sources, 2014, 267, 300 CrossRef CAS.
  20. C. M. Breneman and K. B. Wiberg, J. Comput. Chem., 1990, 11, 361 CrossRef CAS.
  21. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2013 Search PubMed.
  22. S. Jammi, S. Sakthivel, L. Rout, T. Mukherjee, S. Mandal, R. Mitra, P. Saha and T. Punniyamurthy, J. Org. Chem., 2009, 74, 1971 CrossRef CAS PubMed.
  23. N. Taniguchi, J. Org. Chem., 2007, 72, 1241 CrossRef CAS PubMed.
  24. (a) C. W. Nogueira and J. B. T. Rocha, in The Chemistry of Organic Selenium and Tellurium Compounds, Patai's Chemistry of Functional Groups, ed. Z. Rappoport, John Wiley & Sons, Chichester, 2011, vol. 3 Search PubMed; (b) M. Carland and T. Fenner, The use of selenium-based drugs in medicine, in Metallotherapeutic drugs and metal-based diagnostic agents, ed. M. Gielen and E. R. T. Ticking, Wiley, Chichester, UK, 2005 Search PubMed.
  25. (a) K. D. Volkova, V. B. Kovalska, M. Y. Losytskyy, A. Bento, L. V. Reis, P. F. Santos, P. Almeida and S. M. Yarmoluk, J. Fluoresc., 2008, 18, 877 CrossRef CAS PubMed; (b) P. F. Santos, L. V. Reis, P. Almeida and D. E. Lynch, CrystEngComm, 2011, 13, 1333 RSC; (c) N. R. Conley, A. Dragulescu-Andrasi, J. Rao and W. E. Moerner, Angew. Chem., Int. Ed., 2012, 51, 1 CrossRef PubMed; (d) A. Patra, Y. H. Wijsboom, G. Leitus and M. Bendikov, Chem. Mater., 2011, 23, 896 CrossRef CAS; (e) B. Kim, H. R. Yeom, M. H. Yun, J. Y. Kim and C. Yang, Macromolecules, 2012, 45, 8658 CrossRef CAS; (f) D. Gao, J. Hollinger and D. S. Seferos, ACS Nano, 2012, 6, 7114 CrossRef CAS PubMed.
  26. (a) D. S. Rampon, F. S. Rodembusch, J. M. F. M. Schneider, I. H. Bechtold, P. F. B. Gonçalves, A. Merlo and P. H. Schneider, J. Mater. Chem., 2010, 20, 715 RSC; (b) I. Samb, J. Bell, P. Y. Toullec, V. Michelet and I. Leray, Org. Lett., 2011, 13, 1182 CrossRef CAS PubMed; (c) S. Goswami, A. Hazra, R. Chakrabarty and H. K. Fun, Org. Lett., 2009, 11, 4350 CrossRef CAS PubMed; (d) B. Tang, Y. Xing, P. Li, N. Zhang, F. Yu and G. Yang, J. Am. Chem. Soc., 2007, 129, 11666 CrossRef CAS PubMed.
  27. (a) H. S. Kim, Y. J. Kim, H. Lee, K. Y. Park, C. Lee and C. S. Chin, Angew. Chem., Int. Ed., 2002, 41, 4300 CrossRef CAS; (b) E. J. Lenardão, S. Feijó, J. O. Thurow, G. Perin, R. G. Jacob and C. C. Silveira, Tetrahedron Lett., 2009, 50, 5215 CrossRef; (c) E. J. Lenardão, E. L. Borges, S. R. Mendes, G. Perin and R. G. Jacob, Tetrahedron Lett., 2008, 49, 1919 CrossRef; (d) S. Thurow, V. A. Pereira, D. M. Martinez, D. Alves, G. Perin, R. G. Jacob and E. J. Lenardão, Tetrahedron Lett., 2011, 52, 640 CrossRef CAS; (e) E. E. Alberto, L. L. Rossato, S. H. Alves, D. Alves and A. L. Braga, Org. Biomol. Chem., 2011, 9, 1001 RSC.
  28. S. J. Strickler and R. A. Berg, J. Phys. Chem., 1962, 37, 814 CrossRef CAS.
  29. N. J. Turro, J. C. Scaiano and V. Ramamurthy, in Principles of Molecular Photochemistry: An Introduction, University Science Books, 1st edn, 2008 Search PubMed.
  30. (a) D. S. Rampon, F. S. Santos, R. R. Descalzo, J. M. Toldo, P. F. B. Gonçalves, P. H. Schneider and F. S. Rodembusch, J. Phys. Org. Chem., 2014, 27, 336 CrossRef; (b) D. S. Rampon, F. S. Rodembusch, P. F. B. Gonçalves, R. V. Lourega, A. A. Merlo and P. H. Schneider, J. Braz. Chem. Soc., 2010, 21, 2100 CrossRef CAS.
  31. Z. L. Gong, L. W. Zheng and B. X. Zhao, J. Lumin., 2012, 132, 318 CrossRef CAS.
  32. C. Reichardt, Solvents and S. Effects, in Organic Chemistry, Wiley-VCH, Marburg, 3rd edn, 2004, pp. 352–353 Search PubMed.
  33. R. Bralsford, P. V. Harris and W. C. Price, Proc. R. Soc. London, Ser. A, 1960, 258, 459 CrossRef CAS.
  34. J. R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer, New York, 3rd edn, 2006 Search PubMed.
  35. J. Pina, J. S. de Melo, D. Breusov and U. Scherf, Phys. Chem. Chem. Phys., 2013, 15, 15204 RSC.
  36. A. P. Kulkarni, P. T. Wu, T. W. Kwon and S. A. Jenekhe, J. Phys. Chem. B, 2005, 109, 19584 CrossRef CAS PubMed.
  37. A. Granzhan, H. Ihmels and G. Viola, J. Am. Chem. Soc., 2007, 129, 1254 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: Spectroscopic characterization of the compounds and additional results of the theoretical calculations. See DOI: 10.1039/c6ra04157d

This journal is © The Royal Society of Chemistry 2016