Next Article in Journal
Effect of Curing Agent and Temperature on the Rheological Behavior of Epoxy Resin Systems
Previous Article in Journal
Comprehensive Study of Honey with Protected Denomination of Origin and Contribution to the Enhancement of Legal Specifications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Facile and Efficient Procedure for the Synthesis of New Benzimidazole-2-thione Derivatives

Departamento de Química, Facultad de Ciencias, Universidad Nacional de Colombia, Sede Bogotá, Cra.30 No.45-03, Bogotá 111321, Colombia
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(7), 8578-8586; https://doi.org/10.3390/molecules17078578
Submission received: 4 May 2012 / Revised: 27 June 2012 / Accepted: 9 July 2012 / Published: 17 July 2012
(This article belongs to the Section Organic Chemistry)

Abstract

:
A series of benzimidazole-2-thione derivatives was synthesized using a reaction between the macrocyclic aminal 16H,13H-5:12,7:14-dimethanedibenzo[d,i]-[1,3,6,8] tetraazecine (DMDBTA, 5) and various nucleophiles in the presence of carbon disulfide. A full chemical characterization using IR, 1H-, 13C-NMR and GC-MS analyses of the new compounds is provided. These compounds were separated from the reaction mixture by column chromatography (CC) in highly pure form in 15%–51.4% yield.

Graphical Abstract

1. Introduction

The reaction between carbon disulfide and amines was originally reported by Hoffmann in the 19th century [1]. This reaction with primary and secondary aliphatic amines in ethanol yields reactive dithiocarbamate salt intermediates 1 that can be converted into S-alkyl dithiocarbamates 2, which have several biological activities, via reaction with electrophiles or into dithiocarbamic acids by treatment with mineral acids [2]. In the case of aliphatic diamines, the treatment of the dithiocarbamate intermediate with acid gives cyclic thioureas 3 [3]. In contrast, the reaction with o-phenylenediamine under basic conditions afforded 2-mercaptobenzimidazole (4) [4,5] (Figure 1). In recent years, cyclic thioureas and mercaptobenzimidazoles have had important industrial applications as anticorrosive agents [6,7,8], friction attenuators [9] and heavy metal adsorbents [10]. Moreover, derivatives of cyclic thioureas have important roles in medicinal chemistry owing to their utility as antiseptic [11], antidepressive [12], antitumour [13], and antibacterial agents [14,15,16]. In addition, thioureas have been used as ligands in copper (I)-based complexes [17,18] and in complexes with other metals [19].
Figure 1. The chemical structures of compounds 16.
Figure 1. The chemical structures of compounds 16.
Molecules 17 08578 g001
In continuation of our work with the benzoaminal 6H,13H-5:12,7:14-dimethanedibenzo[d,i]-[1,3,6,8]tetraazecine (DMDBTA, 5) [20,21,22,23,24], we recently became interested in synthesizing the cyclic dithiocarbamate 8H,15H-5:14,9:16-dimethanedibenzo[d,i]-[1,3,6,8,11]thiatetraazatricycledo-decine-6-thione (6). To obtain this product, we planned to use DMDBTA (5) and carbon disulfide in ethanol, following the method reported by Donia et al. [25]. In earlier attempts, we found that the reaction of 5 with carbon disulfide in ethanol did not provide the desired cyclic dithiocarbamate 6. Instead, we found that 5 reacts with carbon disulfide to produce 1,3-bis(ethoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7b) and 1-(ethoxymethyl)-1H-benzimidazole (8b) (Scheme 1).
Scheme 1. General reaction of 5 with nucleophiles in the presence of CS2.
Scheme 1. General reaction of 5 with nucleophiles in the presence of CS2.
Molecules 17 08578 g002
The effects of the solvent on the product yields were examined using 5 as a substrate in various alcohols, including methanol, ethanol, n-propanol, n-butanol and t-butanol. Contrary to our expectations, when 5 was treated with t-butanol, the desired derivatives were not obtained. The efficacy of using other nucleophiles was demonstrated by reacting 5 with CS2 and either hydrogen cyanide or benzotriazole in the aprotic polar non-nucleophilic solvents 1,4-dioxane and acetonitrile, respectively.

2. Results and Discussion

We found the optimum conditions for the synthesis of compounds 7af and 8af. These products were obtained in good yields with high purities. Both the analytical and spectral data (IR, 1H-NMR, 13C-NMR and GC-MS analyses) of all synthesised 1,3-bis(alcoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thiones 7af and 1-(alkoxymethyl)-1,3-dihydro-2H-benzimidazole compounds 8af were in full agreement with the proposed structures.
The IR data for benzimidazole-2-thiones 7a–d clearly indicate the presence of the C=S group, with υ(C=S) vibrations at 1098, 1099, 1104 and 1108. In the 1H-NMR spectra, the N–CH2–O protons of the hemiaminal moiety of 1,3-bis(alcoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione appeared as a singlet at approximately 5.79–5.82 ppm. In the 13C-NMR spectra, the signal of the thiourea functional group was clearly observed at 171.8–169.3 ppm, and these signals are consistent with the signals of analogous molecules [26]. The attributions of the other signals in 13C-NMR spectra were based on the analysis of the HMQC and HMBC spectra. The molecular ions and fragment ions in the mass spectra were also consistent with the assigned structures. The reaction with other nucleophiles in the presence of an inert solvent under the same conditions produces benzimidazole and benzimidazole-2-thione derivatives. When the reaction was carried out with the cyanide anion, 1,3-bis(cyano-1-ylmethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7e) and 1H-benzimidazol-1-yl-acetonitrile (8e) were obtained; in an analogous manner, the reaction with benzotriazole (BtH) afforded 1,3-bis(1H-1,2,3-benzotriazol-1-ylmethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7f) and 1-(1H-benzimidazol-1-ylmethyl)-1H-1,2,3-benzotriazole (8f). The structures of these heterocyclic systems were confirmed by the spectroscopic data.
As shown in Table 1, the results reveal that the yield of the reaction depends on the size of the nucleophile. Consequently, the use of t-butanol under the same reaction conditions afforded complex mixtures, from which we were unable to isolate the expected benzimidazole-2-thione and 1-H-benzimidazole derivatives.
Table 1. Yields of benzimidazole-2-thione (7a–f) and 1-H-benzimidazole (8a–f) derivatives.
Table 1. Yields of benzimidazole-2-thione (7a–f) and 1-H-benzimidazole (8a–f) derivatives.
EntryNuH7 (%)8 (%)t (h)
aMeOH48.551.424
bEtOH30.132.330
cn-PrOH23.324.534
dn-BuOH15.016.740
eHCN25.326.024
fBtH35.236.724
According to these results, in the presence of an adequate nucleophile, the reactions efficiently proceed to provide the ring opening of the benzoaminal. Thus, the first step of the reaction between 5 and CS2 might be fast, and the second step, involving the nucleophilic attack of the dithiocarbamate salt, is slower and consequently should be the rate-limiting step. The onset of the reaction is indicated by the evolution of hydrogen sulphide, whose odour was noticeable during the reaction. Based on these results, we propose a possible pathway for the formation of the products (Scheme 2).
Scheme 2. Proposed pathway for the formation of 7a–f and 8a–f.
Scheme 2. Proposed pathway for the formation of 7a–f and 8a–f.
Molecules 17 08578 g003
We assumed that carbon disulfide first reacts with 5 to form the expected dithiocarbamate salt 9 as a very active intermediate. This intermediate then reacts with one equivalent of the nucleophile, a reaction that includes a proton shift, to produce a second intermediate 10 that is able to form S–H···N intramolecular hydrogen bonds. This intramolecular interaction in this intermediate decreases the relative stability of the intermediate and induces the attack of a second equivalent of nucleophile to give a 1-substituted-3-aryl-2,3,4,5-tetrahydro-1H-1,3,5-benzotriazepine intermediate 11, in which a proton transfer induces an intramolecular rearrangement to give 12. The presence of a positive charge on the 1H-benzimidazole ring of 12 makes this adduct fairly labile, and the central NCH2N moiety in 12 undergoes a regioselective cleavage involving the preferential attack by a third equivalent of nucleophile to give 13 and 14. Then, the benzimidazole-2-thiones7a–f are obtained by cyclisation of the acyclic intermediate 13 with the elimination of a molecule of H2S. Monosubstituted-benzimidazolines 14a–f, the other products formed from 12 under these reaction conditions, smoothly undergo oxidation in air to yield the 1-(alkoxymethyl)-1,3-dihydro-2H-benzimidazole derivatives 8a–f, as observed previously for other benzimidazoline derivatives [27]. Alcohols and cyclic ethers are good solvents for this oxidative process [28,29,30].

3. Experimental

3.1. General

Melting points were determined on an Electrothermal 9100 melting point apparatus and are uncorrected. Chemicals were used without further purification. FT-IR spectra were recorded in potassium bromide pellets using Thermo Nicolet IS10 spectrophotometer. 1H-NMR and 13C-NMR spectra were recorded in CDCl3 using a Bruker Avance AV-400 MHz spectrometer operates at 400 MHz for 1H and 100 MHz for 13C. Elemental analyses (C, H, N) were determined in a Thermo Scientific Flash 2000. Combined GC–MS analysis was performed on a Hewlett–Packard 5973 mass spectrometer at 70 eV coupled to a Hewlett–Packard 6890 gas chromatograph.
General procedure for the reaction of DMDBTA with CS2 in alcohols: Following the general procedure described in the literature [25], carbon disulfide (0.95 mmol, 0.07 mL) was added dropwise over 30 min to a shaking solution of DMDBTA (0.95 mmol) in the desired alcohol (30 mL). This yellow solution was stirred at room temperature in the dark until the DMDBTA had dissolved completely. The reaction was monitored by TLC. The removal of the solvent at reduced pressure (50 mmHg) resulted in the collection of a resinous solid that was then purified via column chromatography on silica gel (elution using benzene:ethyl acetate in a 9:1 mixture).
Procedure for the reaction of DMDBTA with benzotriazole and CS2: A mixture of DMDBTA (0.95 mmol, 0.25 g), 1H-benzotriazole (2.85 mmol, 0.34 g) and carbon disulfide (0.95 mmol, 0.07 mL) was stirred in 1,4-dioxane (30 mL) for 24 h. at room temperature in the dark yielding a resinous solid. The precipitated solid was collected and purified via column chromatography on silica gel (elution using benzene: ethyl acetate in a 9:1 mixture).
Procedure for the reaction of DMDBTA with cyanide anion and CS2: To a solution of DMDBTA (0.95 mmol, 0.25 g) and carbon disulfide (0.95 mmol, 0.07 mL) in acetonitrile (15 mL), an excess of hydrogen cyanide was bubbled in slowly. The reaction mixture was stirred at room temperature in the dark for 24 h. The removal of the solvent at reduced pressure (50 mmHg) resulted in the collection of a resinous product that was then purified via column chromatography on silica gel (elution using benzene:ethyl acetate in a 9:1 mixture).

3.2. Physical and Spectral Data

1,3-Bis(methoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7a): White solid; m.p. 99–101 °C; 1H-NMR d: 3.24 (6H, s, –CH3), 5.79 (4H, s, N-CH2-O), 7.24 (2H, m), 7.35 (2H, m); 13C-NMR d: 52.6, 74.3, 111.5, 121.9, 132.5, 171.8; IR νmax (cm–1): 1,108 (C=S), 1,285 (C-O); EIMS, 70 eV, m/z: 238 (M+); Anal. Calcd. for C11H14N2O2S (238.36): C, 55.44; H, 5.92; N, 11.76; S, 13.46. Found: C, 55.23; H, 5.96; N, 11.46; S, 13.22.
1,3-Bis(ethoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7b): Melting point 103–104 °C (lit. [31] 104–106 °C).
1,3-Bis(propoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7c): White solid; m.p. 120–122 °C; 1H-NMR d: 0.91 (6H, t, J = 6.9 Hz, –CH3), 1.59 (4H, s, J = 6.9 Hz, O-CH2CH2CH3), 3.46 (4H, q, J = 6.9 Hz, O-CH2CH2CH3), 5.82 (4H, s, N–CH2–O), 7.27 (2H, m), 7.39 (2H, m); 13C-NMR d: 10.7, 22.5, 64.9, 73.7, 110.3, 123.8, 132.4, 169.3; IR νmax (cm–1): 1,098 (C=S), 1,289 (C-O); EIMS, 70 eV, m/z: 294 (M+); Anal. Calcd. for C15H22N2O2S (294.40): C, 61.19; H, 7.53; N, 9.52; S, 10.89. Found: C, 61.05; H, 7.66; N, 9.48; S, 10.85.
1,3-Bis(butoxymethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7d): White solid; m.p. 124–126 °C; 1H-NMR d: 0.84 (6H, t, J = 6.7 Hz, –CH3), 1.65–1.68 (8H, m, O-CH2CH2CH2CH3), 3.59 (4H, t, J = 6.7 Hz, O-CH2CH2 CH2CH3), 5.82 (4H, s, N-CH2-O), 7.30 (2H, m), 7.43 (2H, m); 13C-NMR d: 14.3, 18.3, 30.9, 67.4, 72.4, 110.3, 123.6, 131.8, 171.1; IR νmax (cm–1): 1099 (C=S), 1288 (C-O); EIMS, 70 eV, m/z: 322 (M+); Anal. Calcd. for C17H26N2O2S (322.46): C, 63.32; H, 8.13; N, 8.69; S, 9.94. Found: C, 63.24; H, 8.16; N, 8.67; S, 9.91.
1,3-bis(Cyano-1-methyl)-1,3-dihydro-2H-benzimidazole-2-thione (7e): Oily product; 1H-NMR d: 5.32 (4H, s, N-CH2-CN), 7.56 (2H, m), 7.73 (2H, m); 13C-NMR d: 31.9, 111.9, 118.3, 119.4, 134.7, 179.4; IR νmax (cm–1): 2,236 (CN) 1,102 (C=S), 1,285; EIMS, 70 eV, m/z: 228 (M+); Anal. Calcd. for C11H8N4S (228.27): C, 57.88; H, 3.53; N, 24.54; S, 14.05. Found: C, 57.91; H, 3.56; N, 24.48; S, 13.91.
1,3-bis(1H-1,2,3-Benzotriazol-1-ylmethyl)-1,3-dihydro-2H-benzimidazole-2-thione (7f): Yellow Solid; m.p. 142–144 °C; 1H-NMR d: 7.07 (4H, s, N-CH2-Bt), 7.21 (4H, m, benzimidazoline nucleus), 7.35 (2H, t, J = 7.8 Hz, benzotriazole nucleus), 7.52 (2H, t, J = 7.8 Hz, benzotriazole nucleus), 8.08 (2H, d, J = 7.8 Hz, benzotriazole nucleus), 8.43 (2H, d, J = 7.8 Hz, benzotriazole nucleus); 13C-NMR d: 55.3, 110.1, 111.7, 119.9, 124.0, 124.6, 128.5, 130.2, 132.6, 146.2, 169.6. IR νmax (cm−1): 1108 (C=S), 1,271 (C=N); EIMS, 70 eV, m/z: 412 (M+); Anal. Calcd. for C21H16N8S (412.45): C, 61.15; H, 3.91; N, 27.17; S, 7.77. Found: C, 61.23; H, 3.96; N, 27.26; S, 7.55.

4. Conclusions

In conclusion, the reported synthesis is reasonably efficient, direct, and operationally simple. We believe that the methodology presented herein can have wide applications for the development of synthetically useful benzimidazole-2-thiones that were previously inaccessible by other routes.

Acknowledgments

We acknowledge the Dirección de Investigaciones Sede Bogotá (DIB) of Universidad Nacional de Colombia for financial support.

References

  1. Hofmann, A.W. Noch einige Bemerkungen über Senföle. Ber. Dtsch. Chem. Ges. 1875, 8, 105–109. [Google Scholar] [CrossRef]
  2. Wolf-Dieter, R. Reactions of carbon disulfide with Nnucleophiles. J. Sulfur Chem. 2007, 28, 295–339. [Google Scholar] [CrossRef]
  3. Johnson, T.B.; Edens, C.O. Complex Formations between iodine and μ-mercapto-dihydroglyoxalines. J. Am. Chem. Soc. 1942, 64, 2706–2708. [Google Scholar] [CrossRef]
  4. Wang, M.-L.; Liu, B.-L. Reaction of carbon disulfide and o-phenylenediamine catalyzed by tertiary amine in a homogeneous solution. Ind. Eng. Chem. Res. 1995, 34, 3688–3695. [Google Scholar] [CrossRef]
  5. Wang, M.-L.; Liu, B.-L. Homogeneous catalyzed reaction of carbon disulfide and o-phenylenediamine by tetrabutylammonium hydroxide in the presence of potassium hydroxide. J. Chin. Inst. Chem. Eng. 2007, 38, 85–90. [Google Scholar] [CrossRef]
  6. Xue, G.; Huang, X.-Y.; Dong, J.; Zhang, J. The formation of an effective anti-corrosion film on copper surfaces from 2-mercaptobenzimidazole solution. J. Electroanal. Chem. Interfacial Electrochem. 1991, 310, 139–148. [Google Scholar] [CrossRef]
  7. Perrin, F.X.; Pagetti, J. Characterization and mechanism of direct film formation on a Cu electrode through electro-oxidation of 1-mercaptobenzimidazole. Corros. Sci. 1998, 40, 1647–1662. [Google Scholar] [CrossRef]
  8. Zhang, J.; Liu, W.; Xue, Q. Tribological study of a Mannich compound of 2-mercaptobenzimidazole in liquid paraffin. Tribol. Int. 1998, 31, 767–770. [Google Scholar]
  9. Zhang, J.; Liu, W.; Xue, Q. The effect of molecular structure of heterocyclic compounds containing N, O and S on their tribological performance. Wear 1999, 231, 65–70. [Google Scholar] [CrossRef]
  10. Moreira, J.C.; Pavan, L.C.; Gushikem, Y. Adsorption of Cu(II), Zn(II), Cd(II), Hg(II) and Pb(II) from aqueous solutions on a 2-mercaptobenzimidazole-modified silica gel. Microchim. Acta 1990, 102, 107–115. [Google Scholar] [CrossRef]
  11. Dalal, D.S.; Pawar, N.S.; Mahulikar, P.P. Synthesis of 2-mercaptobenzothiazole and of 2-mercaptobenzimidazole derivatives using polymer-supported anions. Org. Prep. Proced. Int. 2005, 37, 539–545. [Google Scholar] [CrossRef]
  12. Bell, S.C.; Wei, P.H. Synthesis of heterocyclic fused thiazole acetic acids 2. J. Med. Chem. 1976, 19, 524–530. [Google Scholar] [CrossRef]
  13. Lin, S.; Lombardo, M.; Malkani, S.; Hale, J.J.; Mills, S.G.; Chapman, K.; Thompson, J.E.; Zhang, W.-X.; Wang, R.; Cubbon, R.M. Novel 1-(2-aminopyrazin-3-yl)methyl-2-thioureas as potent inhibitors of mitogen-activated protein kinase-activated protein kinase 2 (MK-2). Bioorg. Med. Chem. Lett. 2009, 19, 3238–3242. [Google Scholar]
  14. De Almeida, M.V.; Cardoso, S.H.; De Assis, J.V.; De Souza, M.V.N. Synthesis of 2-mercaptobenzothiazole and 2-mercaptobenzimidazole derivatives condensed with carbohydrates as a potential antimicrobial agents. J. Sulfur Chem. 2007, 28, 17–22. [Google Scholar] [CrossRef]
  15. Desai, K.G.; Desai, K.R. Green route for the heterocyclization of 2-mercaptobenzimidazole into β-lactum segment derivatives containing -CONH-bridge with benzimidazole: Screening in vitro antimicrobial activity with various microorganisms. Bioorg. Med. Chem. 2006, 14, 8271–8279. [Google Scholar] [CrossRef]
  16. Proksa, B.; Turdiková, M.; Ertík, M.; Miroslav, K.; Preizingerová, T.; Fuska, J. Effects of derivatives of 2-mercaptobenzimidazole on polyketide biosynthesis in Penicillium frequentans. Biotechnol. Appl. Biochem. 1997, 25, 169–172. [Google Scholar]
  17. Lobana, T.S.; Sharma, R.; Hundal, G.; Butcher, R.J. Synthesis of 1D {Cu63-SC3H6N2)4(μ-SC3H6N2)2(mu-I)2I4}n and 3D {Cu2(μ-SC3H6N2)2(μ-SCN)2}n polymers with 1,3-imidazolidine-2-thione: Bond isomerism in polymers. Inorg. Chem. 2006, 45, 9402–9409. [Google Scholar] [CrossRef]
  18. Lobana, T.S.; Sharma, R.; Sharma, R.; Mehra, S.; Castineiras, A.; Turner, P. Versatility of heterocyclic thioamides in the construction of a trinuclear cluster [Cu3I3(dppe)3 (SC5H4NH)] and Cu(I) linear polymers {Cu6(SC5H4NH)6I6}nx2nCH3CN and {Cu(I)6 (SC3H6N2)6X6}n (X = Br, I). Inorg. Chem. 2005, 44, 1914–1921. [Google Scholar]
  19. Beheshti, A.; Clegg, W.; Dale, S.H.; Hyvadi, R. Synthesis, crystal structures, and spectroscopic characterization of the neutral monomeric tetrahedral [M(Diap)2(OAc)2]H2O complexes (M = Zn, Cd; Diap = 1,3-diazepane-2-thione; OAc = acetate) with N-H…O and O-H…O intra- and intermolecular hydrogen bonding interactions. Inorg. Chim. Acta 2007, 360, 2967–2972. [Google Scholar] [CrossRef]
  20. Rivera, A.; Maldonado, M.; Núñez, M.; Joseph-Nathan, P. Nucleophilic substitution at the aminalic carbon of some macrocyclic polyaminals. Heterocycl. Commun. 2004, 10, 77–80. [Google Scholar] [CrossRef]
  21. Rivera, A.; Maldonado, M. Unexpected behavior of 6H,13H-5:12,7:14-dimethane-dibenzo[d,i][1,3,6,8]tetraazecine (DMDBTA) toward phenols. Tetrahedron Lett. 2006, 47, 7467–7471. [Google Scholar] [CrossRef]
  22. Rivera, A.; Navarro, M.; Rios-Motta, J. Solvent-Free Synthesis of 2-(1H-benzimidazol-1-yl-methyl)-4-substituted-1-hydroxyaryl by the two component Mannich reaction between 6H,13H-5:12,7:14-Dimethanedibenzo[d,i][1,3,6,8]tetraazecine (DMDBTA) and phenols. Heterocycles 2008, 75, 1651–1658. [Google Scholar]
  23. Rivera, A.; Moyano, D.; Maldonado, M.; Ríos-Motta, J.; Reyes, A. FT-IR and DFT studies of the proton affinity of small aminal cages. Spectrochim. Acta A 2009, 74, 588–590. [Google Scholar] [CrossRef]
  24. Rivera, A.; Maldonado, M.; Ríos-Motta, J.; Navarro, M.; González-Salas, D. Novel access to 1-substituted-benzimidazoles via benzotriazole-mediated synthesis. Tetrahedron Lett. 2010, 51, 102–104. [Google Scholar]
  25. Donia, R.A.; Shotton, J.A.; Bentz, L.O.; Smith, G.E.P., Jr. Reactions of mono- and di-amines with carbon disulphide. II. Methylenediamine and imidazolidine-carbon disulphide reactions. J. Org. Chem. 1949, 14, 952–961. [Google Scholar] [CrossRef]
  26. Mohamed, T.A.; Mustafa, A.M.; Zoghaib, W.M.; Afifi, M.S.; Farag, R.S.; Badr, Y. Infrared, Raman and temperature-dependent NMR spectra, vibrational assignments, normal coordinate analysis, and DFT calculations of benzoxazoline-2-thione. Vib. Spectrosc. 2010, 52, 128–136. [Google Scholar] [CrossRef]
  27. Lin, S.; Yang, L. A simple and efficient procedure for the synthesis of benzimidazoles using air as the oxidant. Tetrah edron Lett. 2005, 46, 4315–4319. [Google Scholar] [CrossRef]
  28. Ivanova, N.V.; Sviridov, S.I.; Stepanov, A.E. Parallel solution-phase synthesis of substituted 2-(1,2,4-triazol-3-yl)benzimidazoles. Tetrahedron Lett. 2006, 47, 8025–8027. [Google Scholar] [CrossRef]
  29. Vojtech, M.; Petrušová, M.; Sláviková, E.; Bekešová, S.; Petruš, L. One-pot synthesis of 2-C-glycosylated benzimidazoles from the corresponding methanal dimethyl acetals. Carbohydr. Res. 2007, 342, 119–123. [Google Scholar] [CrossRef]
  30. Gallant, A.J.; Patrick, B.O.; MacLachlan, M.J. Mild and selective reduction of imines: Formation of an unsymmetrical macrocycle. J. Org. Chem. 2004, 69, 8739–8744. [Google Scholar]
  31. Rivera, A.; Mejia-Camacho, A.; Ríos-Motta, J.; Dušek, M.; Fejfarová, K. 1,3-bis(ethoxymethyl)-1H-benzimidazole-2(3H)-thione. Acta Cryst. 2010, E 66, o1135–o1136. [Google Scholar]
  • Sample Availability: Samples of the compounds 7af are available from the authors.

Share and Cite

MDPI and ACS Style

Rivera, A.; Maldonado, M.; Ríos-Motta, J. A Facile and Efficient Procedure for the Synthesis of New Benzimidazole-2-thione Derivatives. Molecules 2012, 17, 8578-8586. https://doi.org/10.3390/molecules17078578

AMA Style

Rivera A, Maldonado M, Ríos-Motta J. A Facile and Efficient Procedure for the Synthesis of New Benzimidazole-2-thione Derivatives. Molecules. 2012; 17(7):8578-8586. https://doi.org/10.3390/molecules17078578

Chicago/Turabian Style

Rivera, Augusto, Mauricio Maldonado, and Jaime Ríos-Motta. 2012. "A Facile and Efficient Procedure for the Synthesis of New Benzimidazole-2-thione Derivatives" Molecules 17, no. 7: 8578-8586. https://doi.org/10.3390/molecules17078578

Article Metrics

Back to TopTop