Abstract
The reactions of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts with nucleophiles were investigated. Theoretical investigation of reactivity descriptors and structural parameters of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium cation was carried out, and the reactivity centers of the model salts were located. An efficient general method for the synthesis of novel poly-functional derivatives of symmetric triazoles from [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium halides was developed. The structures of the synthesized compounds were confirmed using proton nuclear magnetic resonance (1H NMR), carbon-13 nuclear magnetic resonance (13C NMR), Fourier-transform infrared spectroscopy (FTIR) and single-crystal X-ray analysis.
Introduction
Thiazolo[3,2-b][1,2,4]triazoles and other derivatives of 1,2,4-triazoles show diverse biological activity [1], [2], [3], [4], [5], [6], [7]. The most recent approaches toward the synthesis of the thiazolo[3,2-b][1,2,4]triazole system have been described [8], but only the C-5 substitution in this fused system has been studied as regards the chemical properties [9], [10], [11], [12], [13], [14], [15]. Recently there has been significant interest in developing reactions for thiazolo[3,2-b][1,2,4]triazole and [1,2,4]triazole systems’ functionalizations with the aim of expanding the range of potential bio-active compounds [1], [2], [3], [4], [5], [6], [7]. Herein, we discuss the reactivity of the [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts using quantum chemical calculations followed by experimental studies on the reactions of these salts with O,N-nucleophiles, leading to the formation of new polyfunctional derivatives of 1,2,4-triazole.
Results and discussion
Previously, we reported on the synthetic approach to the [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts via electrophilic heterocyclization [16], [17]. The literature also contains a limited number of sources that describe the chemical properties of azolinium cations; however, these data are contradictory [16], [18], [19], [20], [21], [22]. In order to investigate the reactivity of the [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium cation in more detail, in this work we used a combination of theoretical and experimental approaches.
The bromides 1 and 2 (Figure 1) were used as model salts. The starting geometries of the cations 1 and 2 were pre-optimized using the semiempirical PM7 approach [23], which is one of the best neglect of differential diatomic overlap (NDDO) methods and describes triazole systems well [24]. The re-optimization of the geometries of 1 and 2 with the DFT PBE/6-311G(d,p) level of theory [25] was the next step. This decision was based on our previous studies, where it was found that the PBE functional can model similar C, H, N, S containing organic compounds very well [26], [27]. Table 1 shows the comparison of selected bonds of the optimized geometries of cations 1 and 2 with similar experimental data. Most of the calculated parameters are in good agreement with the experimental parameters and lie within single standard deviations of the experimentally determined values. The only exception is the C33-C34 bond, where the DFT computed length is slightly overestimated as compared with variability range of the experimental bond length (Table 1). This discrepancy can be explained by the fact that none of the presented experimental geometries contain a fragment with the set of bonds chemically and structurally equivalent to the N4+-C33-C34 bonds.

The general view of optimized cations 1 and 2 with numbering. R equals H in cation 1, and R is CH3 in cation 2.
Selected bond lengths of PBE/6-311G(d,p) geometries of cations 1 and 2 and similar experimental data (in Å).
| Bond | Cation 1/Cation 2 | Exp. | Bond | Cation 1/Cation 2 | Exp. | Bond | Cation 1/Cation 2 | Exp. |
|---|---|---|---|---|---|---|---|---|
| N3-N4 | 1.359/1.358 | 1.341(8) [28] | N4-C33 | 1.475/1.490 | 1.470(8) [28] | N5-C7 | 1.420/1.419 | 1.372(9) [28] |
| 1.395(5) [29] | 1.451(6) [29] | 1.349(6) [30] | ||||||
| 1.384(5) [30] | 1.375(3)a | |||||||
| 1.398(3)a | ||||||||
| N3-C7 | 1.328/1.328 | 1.333(8) [28] | N5-C6 | 1.355/1.356 | 1.315(8) [28] | C33-C34 | 1.532/1.545 | 1.509(8) [29] |
| 1.284(6) [29] | 1.377(5) [29] | 1.467(8) [31] | ||||||
| 1.305(6) [30] | 1.377(5) [30] | 1.501(5)a | ||||||
| 1.299(3)a | 1.366(3)a | |||||||
| N4-C6 | 1.338/1.336 | 1.323(9) [28] | S2-C6 | 1.720/1.720 | 1.680(6) [28] | Br1-C34 | 1962/1.963 | 1.961(5) [29] |
| 1.335(5) [29] | 1.738(5) [29] | 1.965(5) [31] | ||||||
| 1.290(6) [30] | 1.727(5) [30] | 1.914(4)a | ||||||
| 1.298(4)a | 1.744(2)a |
aThe dibromide 7 in this work.
The Fukui function [32] and molecular electrostatic potential [33] are well-established reactivity descriptors that can explain the regioselectivity of the reaction [24], [27]. The wavefunction for the production and analysis of reactivity descriptors was generated at the B3LYP/6-311G(d,p) level. It was demonstrated that the B3LYP functional [34] produces relatively good electron densities [35]. Moreover, the predictive power of reactivity descriptors, obtained by this approach and natural population analysis (NPA) partial charges, [36] is satisfactory [37]. The 6-311G(d,p) basis set [38] was chosen as it generates a reliable electrostatic potential [33]. Table 2 presents the computed reactivity indexes of the condensed Fukui function (CFF) for the nucleophilic attack of cations 1 and 2. As can be seen from Table 2, the most reactive atom for both cations is C6. The reactivity order S2>N5>C7 can also be noted. The rearrangement of 1,2,4-triazole system under the action of bases to 1,3,4-thiadiazole system has been described in the literature [31]. Less plausible is the attack at para-carbon in the C19-H29 phenyl ring. Even less probable is the substitution of Br1 atom by nucleophiles. These data strongly suggest breaking of the S2-C6 thiazoline bond under the action of nucleophiles.
CFF reactivity indexes for nucleophilic attack on cations 1 and 2, computed with NPA partial charges at the B3LYP/6-311G(d,p) level.
| Atom | 1 | 2 | Atom | 1 | 2 | Atom | 1 | 2 | Atom | 1 | 2 |
|---|---|---|---|---|---|---|---|---|---|---|---|
| Br1 | 0.069 | 0.067 | C11 | 0.012 | 0.013 | H21 | 0.013 | 0.013 | H31 | 0.020 | 0.019 |
| S2 | 0.103 | 0.098 | H12 | 0.021 | 0.021 | C22 | 0.016 | 0.016 | H32 | 0.017 | 0.015 |
| N3 | 0.039 | 0.041 | C13 | 0.060 | 0.060 | H23 | 0.023 | 0.023 | C33 | −0.004 | −0.001 |
| N4 | 0.018 | 0.016 | H14 | 0.022 | 0.022 | C24 | 0.073 | 0.075 | C34 | −0.013 | −0.012 |
| N5 | 0.053 | 0.051 | C15 | 0.018 | 0.017 | H25 | 0.023 | 0.023 | H35 | 0.008 | 0.008 |
| C6 | 0.133 | 0.128 | H16 | 0.021 | 0.021 | C26 | 0.009 | 0.009 | Xa | 0.020 | −0.004 |
| C7 | 0.076 | 0.079 | C17 | 0.019 | 0.021 | H27 | 0.021 | 0.021 | H37 | 0.003 | 0.008 |
| C8 | −0.012 | −0.011 | H18 | 0.010 | 0.010 | C28 | 0.036 | 0.036 | H38 | – | 0.006 |
| C9 | 0.024 | 0.023 | C19 | −0.009 | −0.007 | H29 | 0.006 | 0.007 | H39 | – | 0.018 |
| H10 | 0.010 | 0.009 | C20 | 0.035 | 0.036 | C30 | 0.006 | 0.005 | H40 | – | −0.001 |
Largest values are shown in bold. aIn the case of cation 1, X=H36=R (see Figure 1), for cation 2, X=C36, which is part of R.
These theoretical results prompted further experimental studies on the behavior of the model salts 1, 2 under the action of nucleophiles (Scheme 1). The reactions between salts 1, 2 and OH− nucleophile in aqueous solutions of NaOH, Na2CO3, Na2SO3 at room temperature were studied. The cleavage of the thiazoline moiety with the formation of poly-functional symmetric triazoles 3, 4 was observed in all cases. Also, as was shown previously [16], the use of more concentrated solutions of reagents and longer reaction times lead to the formation of the elimination product 5 in the case of salt 1. The reaction is believed to proceed through the cleavage of the S2-C6 thiazoline bond by nucleophile action with the subsequent substitution of bromine. Oxidation of the thiol to a disulfide probably occurs in the presence of atmospheric oxygen and gives the corresponding target triazoles 3, 4. It should be noted that performing these reactions under heating or in different solvents leads to a decrease in selectivity [according to the nuclear magnetic resonance (NMR) spectrum of the reaction mixture].

The proton nuclear magnetic resonance (1H NMR), spectrum of compound 3 contains the characteristic signals of exocyclic saturated fragment – the multiplet at 4.79–4.87 ppm for methine proton, the multiplet for oxi-methylene protons at 4.01–4.08 ppm and the doublet of doublets for thiomethylene protons at 3.78 ppm, which (together with the signals of the OH group at 3370 cm−1 and the C=O group at 1695 cm−1 in IR spectrum) confirm the structure. A similar pattern is observed in the case of triazole 4. The 1H NMR spectrum of triazole 4 contains the multiplet for SCH2 at 3.10–3.19 ppm, the multiplet for OCH2 at 4.15–4.21 ppm and a singlet for the methyl group protons at 1.76 ppm. The strong band at 1700 cm−1 (C=O group) is present in the infrared (IR) spectrum of triazole 4; a wide signal at 3400 cm−1 (OH group) is also seen. The different behavior of salts 1 and 2 is observed for the reaction of an N-nucleophile such as morpholine at room temperature with further dilution of the mixture by water (Scheme 1). Specifically, the methyl-substituted salt 2 is transformed into amino-substituted product 6, whereas a similar treatment of the salt 1 gives the elimination product 5, as earlier described by us [16]. In addition to the signals of the CH3, SCH2 and NCH2 groups, the 1H NMR spectrum of triazole 6 also contains the triplets for protons of methylene groups at 3.05 ppm and 3.45 ppm (morpholine moiety). By contrast, there are only signals of the thiopropenyl moiety in the 1H NMR spectrum of triazole 5, namely the singlet for the SCH2 group at 4.20 ppm and the singlets for cis and trans protons of the =CH2 group at 6.00 ppm and 5.57 ppm, respectively. The IR spectra of triazoles 5 and 6 also show a strong band at 1700 cm−1 (C=O group); whereas the signal of the OH group is absent. In the 13C NMR spectra of compounds 3–6, all signals of carbons belong to sp3- and sp2-hybridized carbons. The yields of the triazoles 3–6 are in the range of 64–78%, which indicates regioselectivity of these reactions.
An interesting fact for significantly different stabilities of salts 1 and 2 toward heating can be noted. Thus, salt 1 is stable upon heating in alcohol or acetic acid medium, whereas the heating of salt 2 in ethanol leads to the destruction of the thiazoline ring with the formation of dibromide 7 (Scheme 2). This difference in stability of the N4-C33 bond in cations 1 and 2 is consistent with the following computational results. First, the N4-C33 bond in cation 2 is 0.015 Å longer than that of cation 1 (Table 1). Furthermore, a comparison of the NPA partial charges shows that the N4-C33 bond is more ionic in cation 2 than in 1. Thus, the partial charges of N4/C33 in cation 1 are −0.204/−0.032, whereas in the case of cation 2, the N4/C33 partial charges are −0.216/0.149. The Fukui function is not applicable in the case of Br− attack (Table 2), as it shows the absence of any reactivity of C33 atom. We already faced similar limitations of the CFF [24], [26], [27], and for correct treatment of ionic reactions with the participation of hard nucleophiles the electrostatic forces must be considered. Hence, the electrostatic potential (ESP) is expected to be a useful descriptor for understanding the reaction selectivity between cations 1 and 2 (Figure 2). Thus, in the case of cation 1, the maxima on the ESP surface A, B and C with the corresponding values of 97, 86 and 88 kcal/mol, respectively, are closest to C33. In the case of the SN2 reaction, the attack of a nucleophile must go opposite to the leaving group (the N4 atom), but in the case of cation 1 there is no maximum in the region opposite to N4, and the angles N4-C33-X, where X can be A, B or C, are 71°, 91° and 105°, respectively. Obviously, the reaction is not favorable for cation 1. By contrast, in the case of cation 2, there is a maximum D (83 kcal/mol) located opposite to N4 at the distance of 2.50 Å from C33. The N4-C33-D angle is 141°, which makes possible the attack of bromide anion with cleavage of thiazoline system possible.


The electrostatic potential of cations 1 (A) and 2 (B). The maxima in the N4-C33 region are shown as letters with corresponding values in parentheses (kcal/mol).
The formation of linear dibromide 7 was confirmed by analysis of the NMR spectra and the structure was resolved using single-crystal X-ray diffraction analysis. All calculations relevant to crystal structure determination were performed using the SHELX-97 package [39]. The program OLEX2 1.1 [40] was used for the analysis and visualization of the structure and for rendering the illustrations. As salt 2 is a mixture of R- and S-enantiomers, the resulting dibromide 7 is also obtained in the form of a racemic mixture. Figure 3 presents the structure of dibromide 7 that was obtained by superimposition of R- and S-enantiomers.

Crystal structure of the superimposed R- and S-enantiomers of the molecule 7.
The 1,2,4-triazole-3-thiol moiety is almost planar as in similar triazole systems that we have investigated previously [28], [29], [30]. The angles between the plane of the 1,2,4-triazole-3-thiol fragment and C6-C11/C12-C17 phenyl rings are 67.4°/33.0°, respectively. Intermolecular forces in the crystal are formed by the weak interaction of the N1 atom with the H-CHS group, and by the weak interactions between the 2,3-dibromopropyl fragments through the Br···H-C attraction.
Conclusions
The reactivity of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium cation toward nucleophiles was analyzed using a combination of theoretical and experimental approaches. As a result, an efficient general synthesis of functional derivatives of 1,2,4-triazole was developed.
Experimental
1H NMR (400 MHz) and 13C NMR (100 MHz) spectra were recorded in hexadeutero dimethyl sulfoxide (DMSO-d6) as a solvent and with tetramethylsilane (TMS) as an internal standard on a Varian VXR 400 spectrometer. Melting points were determined on a Stuart SMP30 instrument. IR spectra were recorded on a Shimadzu FTIR Prestige 21 instrument with an attenuated total reflection (ATR) accessory. Elemental analyses were performed on an Elementar Vario MICRO cube analyzer. All crystallographic measurements were performed at room temperature [293(2) K] on a single-crystal diffractometer, Oxford Diffraction Xcalibur.
Computational software
MOPAC2016 was used for the PM7 pre-optimization [41]. The PRIRODA program was used for density functional theory (DFT) calculations [42]. To reduce the computational time, the density fitting technique was used [43]. Calculations of electrostatic potential were performed with Multiwfn 3.3.8 [44]. The JANPA program [45] was used for computation of the NPA charges. Input files preparation and visualization of the structures were made using the programs Avogadro [46], VMD [47] and Jmol (Jmol: an open-source Java viewer for chemical structures in 3D. http://www.jmol.org/).
6-(Bromomethyl)-2,3-diphenyl-5,6-dihydro-3H-[1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium bromide (1)
This compound was synthesized according to the procedure [16]; yield 82% of white powder; mp 195–196°C (lit. [16] mp 195°C).
6-(Bromomethyl)-6-methyl-2,3-diphenyl-5,6-dihydro-3H-[1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium bromide (2)
This compound was synthesized according to the procedure [17]; yield 80% of white powder; mp 284–286°C (lit. [17] mp 285–287°C).
General procedure for reaction of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts 1 and 2 with OH− nucleophile
A solution of salt 1 or 2 (1.0 mmol) in DMF (10 mL) was treated with a solution of a – NaOH, b – Na2CO3 or c – Na2SO3 (2.0 mmol) in water (10 mL). The mixture was stirred for 2 h at room temperature and then treated with water (50 mL). The resultant precipitate of 3 or 4 was filtered, washed with water, dried and crystallized from acetone.
1-2-Hydroxy-1-[3-hydroxy-2-(5-oxo-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-1-yl)propyldisulfanylmethyl]ethyl-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-5-one (3)
This compound was obtained by reaction of 1 with the reagent a–c; yields 58% (a), 64% (b) and 52% (c); white powder; mp 108–110°C; IR: 3370 (OH), 1695 cm−1 (C=O); 1H NMR: δ 7.34–7.76 (m, 10H); 4.79–4.87 (m, 1H, CH), 4.01–4.08 (m, 2H, CH2O), 3.78 (dd, 2H, J=10.4, 3.2 Hz, CH2S); 13C NMR: δ 153.2, 151.0, 135.0, 131.7, 130.9, 129.6, 128.8, 128.2, 125.3, 123.2, 62.6, 59.1, 28.7. Anal. Calcd for C34H32N6O4S2: C, 62.56; H, 4.94; N, 12.87; S, 9.82. Found: C, 62.67; H, 5.11; N, 12.59; S, 9.77.
1-2-Hydroxy-1-[3-hydroxy-2-methyl-2-(5-oxo-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-1-yl)propyldisulfanylmethyl]-1-methylethyl-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-5-one (4)
This compound was obtained from salt 2 with the reagent a–c (see above); yield 72% (a), 78% (b), 63% (c) (white powder); mp 88–90°C; IR: 3400 (OH); 1700 cm−1 (C=O); 1H NMR: δ 7.17–7.47 (m, 10H); 4.15–4.21 (m, 2H, CH2O), 3.10–3.19 (m, 2H, CH2S), 1.76 (s, 3H, CH3); 13C NMR: δ 153.0, 150.6, 134.2, 131.8, 130.2, 129.0, 128.6, 128.2, 126.6, 126.0, 64.8, 63.7, 36.9, 24.8. Anal. Calcd for C36H36N6O4S2: C, 63.51; H, 5.33; N, 12.34; S, 9.42. Found: C, 63.41; H, 5.72; N, 12.27; S, 9.18.
General procedure for the reaction of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts 1 and 2 with morpholine
A solution of salt 1 or 2 (1.0 mmol) and morpholine (2 mL) in DMF (5 mL) was stirred for 2 h at room temperature and then treated with water (100 mL). The resultant precipitate of 5 (from 1) or 6 (from 2) was filtered, washed with water, dried and crystallized from acetone.
1-1-[2-(5-Oxo-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-1-yl)allyldisulfanylmethyl]vinyl-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-5-one (5)
This compound was obtained from 1; yield 70% (lit. [16] yield 76%) of white powder; mp 152–153°C (lit. [16] mp 106–108°C); IR: 1700 cm−1 (C=O); 1H NMR: δ 7.32–7.60 (m, 10H); 6.00 (s, 1H, =CH2-cis), 5.57 (s, 1H, =CH2-trans), 4.20 (s, 2H, SCH2). Anal. Calcd for C34H28N6O2S2: C, 66.21; H, 4.58; N, 13.63; S, 10.40. Found: C, 66.35; H, 4.56; N, 13.54; S, 10.28.
1-1-Methyl-2-[2-methyl-3-morpholino-2-(5-oxo-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-1-yl)propyldisulfanyl]-1-morpholinomethylethyl-3,4-diphenyl-4,5-dihydro-1H-1,2,4-triazol-5-one (6)
This compound was obtained from salt 2; yield 68% (white powder); mp 122°C; IR: 1700 cm−1 (C=O); 1H NMR: δ 7.38–7.80 (m, 10H); 4.26–4.29 (m, 2H, NCH2), 3.45 (t, 4H, J=3.6 Hz, O(CH2)2-morph), 3.38–3.42 (m, 2H, CH2S), 3.05 (t, 4H, J=3.6 Hz, N(CH2)2-morph), 1.74 (s, 3H, CH3); 13C NMR: δ 153.3, 148.6, 129.9, 129.3, 129.1, 129.0, 128.2, 128.1, 127.6, 127.0, 63.7, 63.0, 62.3, 43.3, 36.8, 25.3. Anal. Calcd for C44H50N8O4S2: C, 64.52; H, 6.15; N, 13.68; S, 7.83. Found: C, 64.63; H, 6.12; N, 13.55; S, 7.98.
3-[(2,3-Dibromo-2-methylpropyl)sulfanyl]-4,5-diphenyl-4H- 1,2,4-triazole (7)
A solution of salt 2 (1.0 mmol) in ethanol (40 mL) was heated under reflux for 15 min. After cooling, the resultant precipitate of 7 was filtered, washed with water and crystallized from ethanol as fine colorless needles suitable for X-ray crystal structure analysis; yield 88%; mp 186°C; 1H NMR: δ 7.32–7.62 (m, 10H), 4.09 (s, 2H, CH2Br), 3.90–4.01 (m, 2H, SCH2), 1.87 (s, 3H). Anal. Calcd for C18H17Br2N3S: C, 46.27; H, 3.67; Br, 34.20; N, 8.99; S, 6.86. Found: C, 46.22; H, 3.65; Br, 34.03; N, 8.89; S, 6.91.
For X-ray crystallographic analysis, a single crystal of 7 (from ethanol) was mounted in inert oil and transferred to the cold gas stream of the diffractometer. X-ray crystal data: M 467.23, colorless crystal, crystal size 0.20 mm–0.05 mm–0.04 mm; triclinic, space group P1̅ (no. 2), Z=2, μ(Mo Kα)=4.600, a=5.9160(5) Å, b=10.7456(6) Å, c=14.3759(10) Å, α=84.389(5)°, β=89.460(6)°, γ=84.138(6)°, U=904.76(11) Å3, T=293(2), V=904.76(11) A3, Z=2, ρcalc=1.715 mg mm−3, μ=4.600 mm−1, F(000)=464. The intensity data were collected within the range of 6.54≤θ≤57.1°, using Mo-Kα radiation (λ=0.71078 Å). The intensities of 6273 reflections were collected (3871 unique reflections). The final wR(F2) was 0.1530. Convergence was obtained at RF=0.0504 and wR2=0.1344 for all reflections, RF=0.0760, wR2=0.1530, GOF=0.985.
Supplementary material
X-ray crystallographic data.
Acknowledgments
This study was partially supported by the National Scholarship Program of the Slovak Republic.
References
[1] Zaharia, V.; Silvestru, A.; Palibroda, N.; Mogoşan, C. Heterocycles 28. Synthesis and characterization of some bis and polyhetererocyclic compounds with anti-inflammatory potential. Farmacia2011, 59, 624–635.Search in Google Scholar
[2] Barbuceanu, S. F.; Draghici, C.; Barbuceanu, F.; Bancescu, G.; Saramet, G. Design, synthesis, characterization and antimicrobial evaluation of some heterocyclic condensed systems with bridgehead nitrogen from thiazolotriazole class. Chem. Pharm. Bull.2015, 63, 694–700.10.1248/cpb.c15-00379Search in Google Scholar PubMed
[3] Tozkoparan, B.; Aytaç, S. P.; Gürsoy, S.; Aktay, G. Design and synthesis of some thiazolotriazolyl esters as anti-inflammatory and analgesic agents. Med. Chem. Res.2012, 21, 192–201.10.1007/s00044-010-9508-xSearch in Google Scholar
[4] Lesyk, R.; Vladzimirska, O.; Holota, S.; Zaprutko, L.; Gzella, A. New 5-substituted thiazolo[3,2-b][1,2,4]triazol-6-ones: synthesis and anticancer evaluation. Eur. J. Med. Chem.2007, 42, 641–648.10.1016/j.ejmech.2006.12.006Search in Google Scholar PubMed
[5] Assarzadeh, V. J.; Almasirad, A.; Shafiee, A.; Koopaei, M. N.; Abdollahi, M. Synthesis of new thiazolo[3,2-b][1,2,4]triazole-6(5H)-one derivatives as potent analgesic and anti-inflammatory agents. Med. Chem. Res.2014, 23, 948–957.10.1007/s00044-013-0697-ySearch in Google Scholar
[6] Korol, N. I.; Pantyo, V. V.; Slivka, M. V.; Kryvovyaz, A. O.; Koval, G. M.; Lendel, V. G. The usage of selenium and tellurium-containing derivatives of 1,2,4-triazole-3-thiol as bactericides and fungicides. UA Patent 114460 UA, 2017.Search in Google Scholar
[7] Küçükgüzel, Ş. G.; Çıkla-Süzgün, P. Recent advances bioactive 1,2,4-triazole-3-thiones. Eur. J. Med. Chem.2015, 97, 830–870.10.1016/j.ejmech.2014.11.033Search in Google Scholar PubMed
[8] Korol, N. I.; Slivka, M. V. Recent progress in the synthesis of thiazolo[3,2-b][1,2,4]triazoles. Chem. Heterocycl. Compd.2017, 58, 852–854.10.1007/s10593-017-2136-3Search in Google Scholar
[9] Kazzouli, S. L.; Koubachi, J.; Brahmi, N. E.; Guillaumet, G. Advances in direct C-H arylation of 5,5-, 6,5- and 6,6-fused heterocycles containing heteroatoms (N, O, S). RSC Adv.2015, 5, 15292–15397.10.1039/C4RA15384GSearch in Google Scholar
[10] Le Meur, M.; Bourg, S.; Massip, S.; Marchivie, M.; Jarry, C.; Guillaumet, G.; Routier, S. First palladium-catalyzed direct regioselective C-5 (het)arylation of mono- or disubstituted thiazolo[3,2-b][1,2,4]triazoles. Eur. J. Org. Chem.2014, 2014, 3704–3714.10.1002/ejoc.201402193Search in Google Scholar
[11] Wang, S.-H.; Liu, W.-J.; Zhan, H.-Y.; Zhang, H.; Liang, Y.; Tian, Y. Ligandless microwave-assisted Pd(OAc)2-catalyzed direct arylation of thiazolo[3,2-b]-1,2,4-triazoles. RSC Adv. 2013, 3, 23972–23975.10.1039/c3ra44486dSearch in Google Scholar
[12] Liu, W.; Wang, S.; Lin, J.; Jiang, Y.; Zhang, Q.; Zhong, Y. Palladium-catalyzed direct and regioselective C-5 desulfitative arylation of thiazolo[3,2-b]-1,2,4-triazoles with sodium sulfinates. Synlett2014, 25, 586–590.10.1055/s-0033-1340320Search in Google Scholar
[13] Liu, W.; Wang, S.; Li, Z.; Huang, Y.; Li, S.; Wang, A. Silver-catalyzed direct regioselective phosphonation of thiazolo[3,2-b]-1,2,4-triazoles with dialkyl phosphites. Synlett2015, 26, 2561–2566.10.1002/chin.201613198Search in Google Scholar
[14] Liu, W.; Wang, S.; Zhan, H.; Lin, J.; He, P.; Jiang, Y. Highly regioselective palladium-catalyzed direct alkenylation of thiazolo[3,2-b]-1,2,4-triazoles via CH activation. Tetrahedron Lett. 2014, 55, 3549–3552.10.1016/j.tetlet.2014.04.095Search in Google Scholar
[15] Wang, S.; Liu, W.; Cai, Z.; Li, S.; Liu, J.; Wang, A. Copper-catalyzed regioselective sulfenylation of thiazolo [3,2-b]-1,2,4-triazoles with thiols. Synlett2016, 27, 2264–2268.10.1055/s-0035-1562494Search in Google Scholar
[16] Usenko, R. M.; Slivka, M. V.; Lendel, V. G. Electrophilic heterocyclization of 4,5-disubstituted 3-allylthio-4H-1,2,4-triazoles by the action of halogens. Chem. Heterocycl. Compd.2011, 47, 1029–1036.10.1007/s10593-011-0870-5Search in Google Scholar
[17] Slivka, M. V.; Korol, N. I.; Rusin, I. F.; Lendel, V. G. Synthesis of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium and [1,2,4]triazolo[5,1-b][1,3]thiazin-4-ium salts via regioselective electrophilic cyclization of 3-[(2-alken-1-yl)sulfanyl]-4H-1,2,4-triazoles. Heterocycl. Cоmmun. 2016, 21, 397–401.10.1515/hc-2015-0158Search in Google Scholar
[18] Korotkikh, N. I.; Aslanov, A. F.; Raenko, G. F.; Shvaika, O. P. Halocyclization reactions of recyclization for synthesis of thierane, thioethane and selenethane derivatives of azolones. Russ. J. Org. Chem. 1999, 35, 752–759.Search in Google Scholar
[19] Slivka, M. V.; Khripak, S. M.; Britcun, V. M.; Staninetc, V. I. Synthesis of 3-substituted 1,2,4-triazolo[3,4-b]thiazolium halides. Russ. J. Org. Chem. 2000, 36, 1064–1069.Search in Google Scholar
[20] Korotkikh, N. I.; Shvayka, O. P.; Gorbunov, L. A. Synthesis of 1,3,4-oxadiazole and xanthen-2-one luminophoric epoxide monomers. Chem. Heterocycl. Compd.1999, 35, 358–362.10.1007/BF02259368Search in Google Scholar
[21] Khripak, S. M.; Slivka, M. V.; Zborovskii Yu. L.; Staninets, V. I. The interaction of derivatives of 3-S-substituted 4-bromomethyl-7-p-nitrophenyl-4,5H-1.2.4-triazolo[3,4-b]thiazol-3-ium bromide with secondary amines. Ukrainian Chem. J.2001, 67, 40–43.Search in Google Scholar
[22] Khripak, S. M.; Slivka, M. V.; Slivka, M. V.; Lendel, V. G. Thienooxazolopyrimidinium salts. Reaction of 1-bromomethyl-5-oxo-4-phenyl-1,2,4,5,6,7,8,9-octahydro[1]benzothieno-[3,2-e][1,3]oxazolo[3,2-a] pyrimidin-11-ium bromide with oxygen-centered nucleophiles. Russ J. Org. Chem. 2007, 43, 439–442.10.1134/S1070428007030207Search in Google Scholar
[23] Stewart, J. J. P. Optimization of parameters for semiempirical methods VI: more modifications to the NDDO approximations and re-optimization of parameters. J. Mol. Model.2013, 19, 1–32.10.1007/s00894-012-1667-xSearch in Google Scholar PubMed PubMed Central
[24] Fizer, M.; Sukharev, S.; Slivka, M.; Mariychuk, R.; Lendel, V. Preparation of bisthiourea and 5-amino-4-benzoyl-1,2,4-triazol-3-thione complexes of copper(II), nickel and zinc and their biological evolution. J. Organomet. Chem.2016, 804, 6–12.10.1016/j.jorganchem.2015.12.024Search in Google Scholar
[25] Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett.1996, 77, 3865–3868.10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed
[26] Bevziuk, K.; Chebotarev, A.; Snigur, D.; Bazel, Y.; Fizer, M.; Sidey, V. Spectrophotometric and theoretical studies of the protonation of Allura red AC and Ponceau 4R. J. Mol. Struct.2017, 1144, 216–224.10.1016/j.molstruc.2017.05.001Search in Google Scholar
[27] Fizer, M.; Sidey, V.; Tupys, A.; Ostapiuk, Yu.; Tymoshuk, O.; Bazel, Ya. On the structure of transition metals complexes with the new tridentate dye of thiazole series: Theoretical and experimental studies. J. Mol. Struct.2017, 1149, 669–682.10.1016/j.molstruc.2017.08.037Search in Google Scholar
[28] Slivka, M.; Korol, N.; Pantyo, V.; Baumer, V.; Lendel, V. Regio- and stereoselective synthesis of [1,3]thiazolo[3,2-b][1,2,4]triazol-7-ium salts via electrophilic heterocyclization of 3-S-propargylthio-4H-1,2,4-triazoles and their antimicrobial activity. Heterocycl. Commun.2017, 23, 109–113.10.1515/hc-2016-0233Search in Google Scholar
[29] Fizer, M.; Slivka, M.; Rusanov, E.; Turov, A.; Lendel, V. [1,3]Thiazolo[2′,3′:3,4][1,2,4]triazolo[1,5-a]pyrimidines – a new heterocyclic system accessed via bromocyclization. J. Heterocycl. Chem.2015, 52, 949–952.10.1002/jhet.2073Search in Google Scholar
[30] Fizer, M.; Slivka, M.; Mariychuk, R.; Baumer, V.; Lendel, V. 3-Methylthio-4-phenyl-5-phenylamino-1,2,4-triazole hexabromotellurate: X-ray and computational study. J. Mol. Struct.2018, 1161, 226–236.10.1016/j.molstruc.2018.02.054Search in Google Scholar
[31] Fizer, M.; Slivka, M.; Baumer, V.; Lendel, V. Synthesis of 3-alkyl-5-allylamino-2-benzoylimino-1,3,4-thiadiazoles via Dimroth rearrangement. Heterocycl. Commun.2016, 22, 79–83.10.1515/hc-2015-0279Search in Google Scholar
[32] Ayers, P. W.; Parr, R. G. Variational principles for describing chemical reactions: the Fukui function and chemical hardness revisited. J. Am. Chem. Soc.2000, 122, 2010–2018.10.1021/ja9924039Search in Google Scholar
[33] Riley, K. E.; Tran, K.; Lane, P.; Murray, J. S.; Politzer, P. Comparative analysis of electrostatic potential maxima and minima on molecular surfaces, as determined by three methods and a variety of basis sets. J. Comp. Sci. 2016, 17, 273–284.10.1016/j.jocs.2016.03.010Search in Google Scholar
[34] Becke, A. D. A new mixing of Hartree-Fock and local density functional theories. J. Chem. Phys.1993, 98, 1372–1377.10.1063/1.464304Search in Google Scholar
[35] Medvedev, M. G.; Bushmarinov, I. S.; Sun, J.; Perdew, J. P.; Lyssenko, K. A. Density functional theory is straying from the path toward the exact functional. Science2017, 355, 49–52.10.1126/science.aah5975Search in Google Scholar PubMed
[36] Reed, A. E.; Weinstock, R. B.; Weinhold, F. Natural population analysis. J. Chem. Phys.1985, 83, 735–746.10.1063/1.449486Search in Google Scholar
[37] Cao, J.; Ren, Q.; Chen, F.; Lu, T. Comparative study on the methods for predicting the reactive site of nucleophilic reaction. Sci. China: Chem. 2015, 58, 1845–1852.10.1007/s11426-015-5494-7Search in Google Scholar
[38] Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys.1980, 72, 650–654.10.1063/1.438955Search in Google Scholar
[39] Sheldrick, G. M. A short history of SHELX. Acta Crystallogr.2008, 64A, 112–122.10.1107/S0108767307043930Search in Google Scholar
[40] Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr.2009, 42, 339–341.10.1107/S0021889808042726Search in Google Scholar
[41] Stewart, J. J. P. MOPAC2016 version 16.060W, 2016. Stewart Computational Chemistry, http://OpenMOPAC.net.Search in Google Scholar
[42] Laikov, D. N.; Ustynyuk, A. Yu. PRIRODA-04: a quantum-chemical program suite. New possibilities in the study of molecular systems with the application of parallel computing. Russ. Chem. Bull.2005, 54, 820–826.10.1007/s11172-005-0329-xSearch in Google Scholar
[43] Laikov, D. N. Fast evaluation of density functional exchange-correlation terms using the expansion of the electron density in auxiliary basis sets. Chem. Phys. Lett.1997, 281, 151–156.10.1016/S0009-2614(97)01206-2Search in Google Scholar
[44] Chen, T.; Lu, F. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem.2012, 33, 580–592.10.1002/jcc.22885Search in Google Scholar PubMed
[45] Nikolaienko, T. Y.; Bulavin, L. A.; Hovorun, D. M. JANPA: an open source crossplatform implementation of the natural population analysis on the java platform. Comput. Theor. Chem.2014, 1050, 15–22.10.1016/j.comptc.2014.10.002Search in Google Scholar
[46] Hanwell, M. D.; Curtis, D. E.; Lonie, D. C.; Vandermeersch, T.; Zurek, E.; Hutchison, G. R. Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J. Cheminformatics. 2012, 4, 1–17.10.1186/1758-2946-4-17Search in Google Scholar PubMed PubMed Central
[47] Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph.1996, 14, 33–38.10.1016/0263-7855(96)00018-5Search in Google Scholar
Supplementary Material:
The online version of this article offers supplementary material (https://doi.org/10.1515/hc-2018-0048).
©2018 Walter de Gruyter GmbH, Berlin/Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Frontmatter
- Research Articles
- Synthesis of benzofuro[3,2-b]furo[2,3-d]pyridin-4(5H)-ones, derivatives of a novel heterocyclic system
- Reactions of 3H-furan-2-ones and 2H-chromen-2-ones with pyrazole-3(5)-diazonium salts
- Synthesis of 1,4-oxathian-2-ones by triton B-catalyzed one-pot reaction of epoxides with ethyl mercaptoacetate
- The regioselective catalyst-free synthesis of bis-quinoxalines and bis-pyrido[2,3-b]pyrazines by double condensation of 1,4-phenylene-bis-glyoxal with 1,2-diamines
- [1,3]Thiazolo[3,2-b][1,2,4]triazol-7-ium salts: synthesis, properties and structural studies
- Crystal structure and molecular docking studies of 1,2,4,5-tetraaryl substituted imidazoles
- Design, synthesis and cytotoxicity evaluation of indibulin analogs
- Synthesis of dibenzothiazepine analogues by one-pot S-arylation and intramolecular cyclization of diaryl sulfides and evaluation of antibacterial properties
- Synthesis and preliminary anti-inflammatory evaluation of xanthone derivatives
- Synthesis and antimicrobial evaluation of 3-(4-arylthieno[2,3-d]pyrimidin-2-yl)- 2H-chromen-2-ones
- Corrigendum
- Corrigendum to: Diversity-oriented synthesis of amide derivatives of tricyclic thieno[2,3-d]pyrimidin-4(3H)-ones and evaluation of their influence on melanin synthesis in murine B16 cells
Articles in the same Issue
- Frontmatter
- Research Articles
- Synthesis of benzofuro[3,2-b]furo[2,3-d]pyridin-4(5H)-ones, derivatives of a novel heterocyclic system
- Reactions of 3H-furan-2-ones and 2H-chromen-2-ones with pyrazole-3(5)-diazonium salts
- Synthesis of 1,4-oxathian-2-ones by triton B-catalyzed one-pot reaction of epoxides with ethyl mercaptoacetate
- The regioselective catalyst-free synthesis of bis-quinoxalines and bis-pyrido[2,3-b]pyrazines by double condensation of 1,4-phenylene-bis-glyoxal with 1,2-diamines
- [1,3]Thiazolo[3,2-b][1,2,4]triazol-7-ium salts: synthesis, properties and structural studies
- Crystal structure and molecular docking studies of 1,2,4,5-tetraaryl substituted imidazoles
- Design, synthesis and cytotoxicity evaluation of indibulin analogs
- Synthesis of dibenzothiazepine analogues by one-pot S-arylation and intramolecular cyclization of diaryl sulfides and evaluation of antibacterial properties
- Synthesis and preliminary anti-inflammatory evaluation of xanthone derivatives
- Synthesis and antimicrobial evaluation of 3-(4-arylthieno[2,3-d]pyrimidin-2-yl)- 2H-chromen-2-ones
- Corrigendum
- Corrigendum to: Diversity-oriented synthesis of amide derivatives of tricyclic thieno[2,3-d]pyrimidin-4(3H)-ones and evaluation of their influence on melanin synthesis in murine B16 cells