An efficient, green synthesis of novel regioselective and stereoselective indan-1,3-dione grafted spirooxindolopyrrolizidine linked 1,2,3-triazoles via a one-pot five-component condensation using PEG-400

M. Rajeswari, Jayant Sindhu, Harjinder Singh and Jitender M. Khurana*
Department of Chemistry, University of Delhi, New Delhi – 110 007, India. E-mail: jmkhurana@chemistry.du.ac.in; Fax: +91 11 27666605; Tel: +91 11 27667725

Received 26th February 2015 , Accepted 8th April 2015

First published on 8th April 2015


Abstract

An efficient synthesis of highly diversified novel functionalized indan-1,3-dione grafted spirooxindolopyrrolizidine linked 1,2,3-triazole conjugates via a one-pot, five-component condensation of indan-1,3-diones, aldehydes, sarcosine, N-propargylated isatin and azides using Cu(I) as a catalyst in PEG-400 as the reaction medium is reported. The reaction proceeds in a highly regio- and stereoselective manner involving a catalyst free Knoevenagel condensation followed by two successive 1,3-dipolar cycloaddition reactions. This protocol is suitable for aromatic, heteroaromatic and aliphatic aldehydes. In situ generation of azomethine ylides and their selectivity towards exocyclic double bonds result in highly functionalized molecular hybrids. All the compounds are obtained in high yield (6a–6s) and were characterized by spectroscopic methods.


Introduction

Designing highly efficient protocols for accessing biologically active compounds possessing structural diversity from simple starting compounds is one of the major challenges in organic synthesis.1 Multi-component reaction (MCR) protocols offer remarkable advantages such as high selectivity, operational simplicity, reduction in the number of work-ups, high yields and structural diversity of drug-like compounds.2 The 1,3-dipolar cycloaddition reaction of azomethine ylides with olefinic dipolarophiles and the Cu(I) catalyzed [3 + 2] cycloaddition of azides with triple bonds constitute facile approaches for the construction of five-membered nitrogen containing heterocycles.3 Azomethine ylides can be generated in situ for the construction of highly functionalized spirooxindolopyrrolizidines in an efficient manner.4 Spirooxindolopyrrolizidines exhibit a wide range of biological activities such as anticonvulsive,5 antileukaemic,6 anesthetic,7 antiviral8 and antibacterial activity.9 In addition the spirooxindolopyrrolizidine ring system also occurs in alkaloids,10 for example, (−)-horsfiline10a and spirotryprostatin A10b (Fig. 1).
image file: c5ra03505h-f1.tif
Fig. 1 Some representative compounds containing the spirocyclic oxindole, indanone and 1,2,3-triazole moieties.

1,2,3-Triazoles are privileged structures associated with biological activities such as anti-HIV,11 antimicrobial,12 antiviral,13 antiproliferative,14 insecticidal,15 and fungicidal activity.16 Fluconazole is a well-known antifungal drug consisting of two 1,2,3-triazole moieties (Fig. 1). 1,2,3-Triazoles can be readily constructed from alkynes and azides by a Cu(I) catalyzed 1,3-dipolar addition. Indanone-fused heterocycles (Fig. 1) have also attracted the attention of chemists and pharmacologists17 due to their role as topoisomerase-I inhibitors.18,19

Therefore, in continuation of our work on the synthesis of potentially bioactive heterocyclic compounds with diverse applications through hybridization,20,21 we decided to link spirooxindolopyrrolizidines, indanones and 1,2,3-triazoles in a single matrix through a one-pot five-component reaction using PEG-400 as an efficient and green reaction media.

Results and discussion

The present manuscript reports a new, diversity oriented and highly efficient green protocol for the synthesis of novel functionalized indan-1,3-dione grafted spirooxindolopyrrozolidine linked 1,2,3-triazole conjugates via a one-pot, five-component reaction by using Cu(I) as the catalyst in PEG-400 as an efficient and green reaction medium (Scheme 1).
image file: c5ra03505h-s1.tif
Scheme 1 Synthesis of 1-N-methyl-spiro[2.3′]-1′-N-((1-(4-fluorophenyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(4-chlorophenyl)-pyrrolidine.

The optimized reaction conditions for the above synthesis were identified by attempting the reaction of N-propargylated isatin (1.0 mmol) (1), indane-1,3-dione (1.0 mmol) (2), 4-cholorobenzaldehyde (1.0 mmol) (3), 4-fluorophenyl azide (1.0 mmol) (4) and sarcosine (1.0 mmol) (5) under different conditions. Initially the reaction was attempted in ethanol (10 mL) in the presence of aq. CuSO4·5H2O (10 mol%) and sodium ascorbate (20 mol%) (in a 50 mL round-bottomed flask) maintained at 80 °C in an oil-bath. The reaction was incomplete even after 2 h as indicated by TLC (ethyl acetate[thin space (1/6-em)]:[thin space (1/6-em)]petroleum ether, 30[thin space (1/6-em)]:[thin space (1/6-em)]70, v/v) (Table 1, entry 1). The reaction was quenched and worked up. After flash chromatography, a solid was obtained which was identified as 1-N-methyl-spiro[2.3′]-1′-N-((1-(4-fluorophenyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(4-chlorophenyl)-pyrrolidine (6a) (55% yield) by 1H NMR and 13C NMR spectroscopy, mass spectrometry, IR spectroscopy and X-ray crystallography. Spectroscopic studies revealed the formation of only one isomer though other isomeric products are possible.

Table 1 Optimization of reaction conditions for the synthesis of 1-N-methyl-spiro[2.3′]-1′-N-((1-(4-fluorophenyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(4-chlorophenyl)-pyrrolidine
Entry Solvent Catalystc (mol%) Temp (°C) Time (min) Yield (%)
a Incomplete reaction.b Reaction performed under ultrasonic irradiation.c aq. CuSO4·5H2O (10 mol%) and aq. sodium ascorbate (20 mol%) were added in entries 1–13 after 35–40 min.
1 EtOH 80 120 55a
2 MeOH 80 120 60a
3 CH3CN 80 100 65a
4 THF 80 100 60a
5 CH3COOH 80 90 70a
6 H2O 80 120 35a
7 PEG-400 80 45 85
8 PEG-600 80 45 82
9 PEG-400 100 45 80
10 PEG-400 60 70 75
11 PEG-400 40b 120 70
12 PEG-400 p-TSA (20) 80 45 82
13 PEG-400 L-Proline (20) 80 45 80


The above reaction was then attempted in different reaction media under otherwise identical conditions (Table 1, entries 2–8). Reactions carried out in methanol, acetonitrile, THF, AcOH and water were not complete and gave inferior yields of 6a after work-up (entries 2–6). The same reaction attempted in PEG-400 and PEG-600 at 80 °C was complete in 45 min and yielded 85% and 82% of the desired product (6a) respectively (Table 1, entries 7–8). The five-component reaction was then attempted at different temperatures, under ultrasonic irradiation and in the presence of catalysts in PEG-400 (Table 1, entries 9–13).

It can be inferred from Table 1 that the above one-pot five-component reaction in PEG-400 using aq. CuSO4·5H2O (10 mol%) and aq. sodium ascorbate (20 mol%) as catalyst at 80 °C gave the highest yield of 6a (85%) (Table 1, entry 7).

The structure of 6a was elucidated using one and two-dimensional NMR spectroscopy, IR spectroscopy and HRMS. The HMBC and COSY correlations are useful in the signal assignments of 6a, and various characteristic signals are shown in Fig. 2. The 1H NMR spectrum of 6a revealed one sharp singlet at δ 2.0 due to the N-methyl protons. The benzylic proton on the C4 carbon of the pyrrolidine ring exhibits a multiplet at δ 5.05–5.01. The two protons on the C5 carbon of pyrrolidine exhibit multiplets at δ 4.01–3.97 and 3.61–3.56. The two protons on the N–CH2 carbon appear as a multiplet at δ 4.97–4.91. One proton, on the C5 carbon of the triazole, appears at δ 8.63 which confirms the formation of the triazole ring. Aromatic protons appeared as a multiplet in the region of δ 7.92–6.71.


image file: c5ra03505h-f2.tif
Fig. 2 HMBC and COSY correlations useful in the signal assignments of 6a and various characteristic 1H and 13C NMR peaks.

The regiochemistry of 6a formed in the reaction was confirmed by 1H NMR. The regioisomer 6a should exhibit a multiplet for the benzylic proton on the C4 carbon of the pyrrolidine ring, whereas the other possible regioisomer 8 (see Fig. 4) would show a singlet. The 1H NMR of the product showed a multiplet at δ 5.05–5.01 rather than a singlet thus suggesting the formation of regioisomer 6a. Furthermore, the off-resonance decoupled 13C NMR of the product exhibited signals at δ 76.8 and 69.6 which correspond to the C3 spiro carbon and the C2 carbon of the pyrrolidine ring of 6a. The signals at δ 197.3 and 196.3 for the product 6a correspond to the keto carbonyls of indan-1,3-dione. The resonance at δ 173.4 is due to the oxindole carbonyl carbon. The signal at δ 34.6 is due to N–CH3 carbon and peaks at δ 45.3 and 55.9 are due to the C4 and C5 carbons of the pyrrolidine ring. The mass spectrum of 6a showed a molecular ion peak at m/z 618.1705 (M+ + 1). The formation of only one regioisomer i.e. 6a was also confirmed by single crystal X-ray structural analysis (Fig. 3).


image file: c5ra03505h-f3.tif
Fig. 3 Single crystal X-ray structure of 6a.

The generality of the above protocol was confirmed by carrying out the reactions of N-propargylated isatin (1), indane-1,3-dione (2) and sarcosine (5) with aromatic/aliphatic azides and aromatic/aliphatic aldehydes. All the reactions proceeded smoothly to yield a diverse library of 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidines (6a–6s) in high yields under the optimized protocol (Scheme 2). The results are summarized in Table 2.


image file: c5ra03505h-s2.tif
Scheme 2 Synthesis of 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidine.
Table 2 Synthesis of 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidine
Product code R1 R2 Time (min) Yield (%)
6a 4-ClC6H4 4-FC6H4 45 85
6b 4-ClC6H4 4-(CH3)C6H4 50 80
6c 4-BrC6H4 4-FC6H4 50 83
6d 4-FC6H4 4-(CH3)C6H4 40 82
6e 4-BrC6H4 4-(CH3)C6H4 50 81
6f 4-FC6H4 4-FC6H4 45 84
6g 4-ClC6H4 4-(NO2)C6H4 50 82
6h 4-BrC6H4 4-(NO2)C6H4 45 86
6i 4-(CH3)C6H4 4-(NO2)C6H4 40 79
6j 4-(NO2)C6H4 4-(OCH3)C6H4 45 84
6k 4-FC6H4 4-(NO2)C6H4 40 82
6l 4-(CF3)C6H4 4-(NO2)C6H4 40 87
6m 4-(CF3)C6H4 7-Chloroquinoline 50 80
6n 4-(CF3)C6H4 4-FC6H4 45 85
6o 4-(CH3)C6H4 7-Chloroquinoline 50 74
6p Furfuraldehyde n-Butyl 50 80
6q Piperonal 4-FC6H4 50 86
6r Isobutyl 7-Chloroquinoline 60 78
6s Isobutyl 4-(OCH3)C6H4 65 74


The proposed pathway for the formation of 6 is given in Fig. 4. The pathway consists of two sequential steps. The first step involves formation of intermediate 7 by Cu(I) catalyzed [3 + 2] azide-alkyne cycloaddition. The Cu(I) is generated in situ by the reduction of Cu(II) to Cu(I) by sodium ascorbate.21 In the second part, the azomethine ylide, generated in situ via decarboxylative condensation of sarcosine with intermediate 7, undergoes a [3 + 2] cycloaddition reaction with the Knoevenagel condensation product of indan-1,3-dione and aldehyde, resulting in the formation of product 6. The [3 + 2] dipolar cycloaddition reaction between the azomethine ylide and the exocyclic double bond can proceed through two paths i.e. path (a) and path (b). However, in the case of path (b), there are secondary orbital interactions between the carbonyl group of indan-1,3-dione with the carbonyl group of isatin in the transition state, which results in the formation of only 6. The formation of intermediate 7 was confirmed by CO-TLC with an authentic sample of 7. The intermediacy of 7 was confirmed by an independent reaction of preformed 7 with the Knoevenagel product of indan-1,3-dione and sarcosine in PEG-400 which led to the formation of 6.


image file: c5ra03505h-f4.tif
Fig. 4 Plausible mechanism for the regio- and stereo-selective formation of 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidines.

The role of Cu(I) in catalyzing only the first part of the pathway was also confirmed by an independent reaction of 7 with the Knoevenagel product and sarcosine. The reaction was attempted both in the presence and absence of CuSO4·5H2O (10 mol%) and sodium ascorbate (20 mol%). The reactions resulted in the formation of 6a in 85 and 82% yield, respectively, in 45 min, which suggests that Cu(I) has no effect on the dipolar cycloaddition reaction of the azomethine ylide and the double bond. The formation of regioisomer 8, as shown in Fig. 4, has already been ruled out based on 1H and 2D NMR spectroscopy and X-ray crystallography.

Conclusion

In conclusion, we have reported an efficient multicomponent methodology for the synthesis of indan-1,3-dione grafted spirooxindolopyrrolizidine linked 1,2,3-triazole hybrids, namely 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidines (6a–6s), by the reaction of N-propargylated isatin (1), indane-1,3-dione (2), aldehydes (3), azides (4) and sarcosine (5) using Cu(I) as the catalyst in PEG-400 at 80 °C. The products could be obtained in high yields using a simple work-up.

Experimental

Chemistry

Silica gel 60 F254 (precoated aluminium plates) from Merck was used to monitor reaction progress. Melting points were determined on Buchi melting point 545 apparatus and are uncorrected. IR (KBr) spectra were recorded on a Perkin Elmer FTIR spectrophotometer, and the values are expressed as νmax cm−1. The 1H and 13C spectra were recorded on Jeol JNM ECX-400P at 400 MHz and 100 MHz, respectively. Chemical shift values are recorded on a δ scale, and the coupling constants (J) are in Hertz. Mass spectra were recorded on a Bruker Micro TOF Q – II. The aryl azides and propargylated isatin were prepared from aromatic amines and isatin respectively by a reported procedure.22

General procedure for the synthesis of 1-N-methyl-spiro[2.3′]-1′-N-((1-(aryl/alkyl)-1H-1,2,3-triazol-4-yl)methyl)oxindole-spiro[3.2′′]-indan-1,3-dione-(aryl/alkyl)-pyrrolidine (6a–6s)

An equimolar mixture of N-propargylated isatin (1) (1.0 mmol), indan-1,3-dione (2) (1.0 mmol), aldehydes (3) (1.0 mmol), azides (4) (1.0 mmol) and sarcosine (5) (1.0 mmol) was dissolved in PEG-400 (10 mL) in a 50 mL round-bottomed flask. An aqueous solution of CuSO4·5H2O (10 mol%) followed by an aqueous solution of sodium ascorbate (20 mol%) were then added to the reaction mixture. The reaction contents were stirred magnetically in a pre-heated oil-bath maintained at 80 °C for 45–70 min (Table 2). The progress of the reaction was monitored by TLC (eluent: ethyl acetate[thin space (1/6-em)]:[thin space (1/6-em)]petroleum ether, 30[thin space (1/6-em)]:[thin space (1/6-em)]70 v/v). After completion of the reaction, the reaction mixture was allowed to cool to room temperature and was quenched with water (∼5 mL). The precipitate formed was collected by filtration at the pump and washed with water. The crude material was purified by flash chromatography over silica gel (230–400 mesh) to afford pure products. The products were characterized by IR spectroscopy, 1H NMR and 13C NMR spectroscopy and mass spectrometry.

References

  1. (a) S. L. Schreiber, Science, 2000, 287, 1964 CrossRef CAS; (b) J. P. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005, p. 1499 Search PubMed.
  2. (a) M. D. Burke and S. L. Schreiber, Angew. Chem., Int. Ed., 2004, 43, 46 CrossRef PubMed; (b) D. S. Tan, Nat. Chem. Biol., 2005, 1, 74 CrossRef CAS PubMed; (c) R. J. Spandl, A. Bender and R. D. Spring, Org. Biomol. Chem., 2008, 6, 1149 RSC.
  3. (a) R. Huisgen, in 1,3-Dipolar cycloaddition chemistry, ed. A. Padwa, Wiley, New York, 1984, vol. 1, pp. 1–176 Search PubMed; (b) R. Grigg and V. Sridharan, Advances in cycloaddition, Jai Press, London, 1993, vol. 3, pp.161–180 Search PubMed.
  4. J. Sindhu, H. Singh and J. M. Khurana, Mol. Diversity, 2014, 18, 345–355 CrossRef CAS PubMed.
  5. (a) H. Jiang, J. Zhao, X. Han and S. Zhu, Tetrahedron, 2006, 62, 11008–11011 CrossRef CAS PubMed; (b) E. Coutouli-Argyropoulou, P. Lianis, M. Mitakou, A. Giannoulis and J. Nowak, Tetrahedron, 2006, 62, 1494–1501 CrossRef CAS PubMed; (c) P. J. S. Gomes, C. M. Nunes, A. A. C. C. Pais, T. M. V. D. Pinho e Melo and L. G. Arnaut, Tetrahedron Lett., 2006, 47, 5475–5479 CrossRef CAS PubMed.
  6. M. A. Abou-Gharbia and P. H. Doukas, Heterocycles, 1979, 12, 637 CrossRef CAS.
  7. M. J. Kornett and A. P. Thio, J. Med. Chem., 1976, 19, 892 CrossRef.
  8. K. Lundahl, J. Schut, J. L. M. A. Schlatmann, G. B. Paerels and A. Peters, J. Med. Chem., 1972, 15, 129 CrossRef CAS.
  9. (a) M. S. Chande, R. S. Verma, P. A. Barve and R. R. Khanwelkar, Eur. J. Med. Chem., 2005, 40, 1143–1148 CrossRef CAS PubMed; (b) A. Dandia, M. Sati, K. Arya, R. Sharma and A. Loupy, Chem. Pharm. Bull., 2003, 51, 1137–1141 CrossRef CAS; (c) R. R. Kumar, S. Perumal, P. Senthilkumar, P. Yogeeswari and D. Sriram, J. Med. Chem., 2008, 51, 5731–5735 CrossRef CAS PubMed; (d) R. R. Kumar, S. Perumal, P. Senthil kumar, P. Yogeeswari and D. Sriram, Tetrahedron, 2008, 64, 2962–2971 CrossRef CAS PubMed.
  10. (a) G. Cravotto, G. B. Giovenzana, T. Pilati, M. Sisti and G. Palmisano, J. Org. Chem., 2001, 66, 8447–8453 CrossRef CAS PubMed; (b) T. Onishi, P. R. Sebahar and R. M. Williams, Org. Lett., 2003, 5, 3135–3137 CrossRef CAS PubMed; (c) R. Grigg, E. L. Millington and M. Thornton-Pett, Tetrahedron Lett., 2002, 43, 2605–2608 CrossRef CAS.
  11. S. Velazquez, R. Alvarez, C. Perez, F. Gago and M. J. Camarasa, Antiviral Chem. Chemother., 1998, 9, 481–489 CrossRef CAS PubMed.
  12. (a) M. J. Genin, D. A. Allwine, D. J. Anderson, M. R. Barbachyn, D. E. Emmert, S. A. Garmon, D. R. Graber, K. C. Grega, J. B. Hester, D. K. Hutchinson, J. Morris, R. J. Reischer, C. W. Ford, G. E. Zurenko, J. C. Hamel, R. D. Schaadt, D. Stapert and B. H. Yagi, J. Med. Chem., 2000, 43, 953–970 CrossRef CAS PubMed; (b) M. Kume, T. Kubota, Y. Kimura, H. Nakashimizu, K. Motokawa and M. Nakano, J. Antibiot., 1993, 46, 177–192 CrossRef CAS.
  13. A. K. Jordão, V. F. Ferreira, T. M. Souza, G. G. Faria, V. Machado, J. L. Abrantes, M. C. Souza and A. C. Cunha, Bioorg. Med. Chem., 2011, 19, 1860–1865 CrossRef PubMed.
  14. S. G. Agalave, S. R. Maujan and V. S. Pore, Chem.–Asian J., 2011, 6, 2696–2718 CrossRef CAS PubMed.
  15. I. K. Boddy, G. G. Briggs, R. P. Harrison, T. H. Jones, M. J. O’Mahony, I. D. Marlow, B. G. Roberts, R. J. Willis and R. Bardsley, J. Pestic. Sci., 1996, 48, 189–196 CrossRef CAS.
  16. K. H. Buechel, H. Gold, P. E. Frohberger and H. Kaspers, German Pat. 2407305, 1975Chem. Abstr., 1975, 83, 206.
  17. (a) Y. J. Duan, J. L. Liu and C. L. Wang, Chin. J. Org. Chem., 2010, 30, 988–996 CAS; (b) N. M. Evdokimov, S. V. Slambrouck, P. Heffeter, L. Tu, B. L. Calvé, D. Lamoral-Theys, C. J. Hooten, P. Y. Uglinskii, S. Rogelj, R. Kiss, W. F. A. Steelant, W. Berger, J.-J. Yang, C. G. Bologa, A. Kornienko and I. V. Magedov, J. Med. Chem., 2011, 54, 2012–2021 CrossRef CAS PubMed.
  18. T. Utsugi, K. Aoyagi, T. Asano, S. Okazaki, Y. Aoyagi, M. Sano, K. Wierzba and Y. Yamada, J. Cancer Res., 1997, 88, 992–1002 CAS.
  19. S. Antony, M. Jayaraman, G. Laco, G. Kohlhagen, K. W. Kohn, M. Cushman and Y. Pommier, Cancer Res., 2003, 63, 7428–7435 CAS.
  20. (a) H. Singh, J. Sindhu and J. M. Khurana, J. Iran. Chem. Soc., 2013, 10, 883–888 CrossRef CAS; (b) J. Sindhu, H. Singh, J. M. Khurana, C. Sharma and K. R. Aneja, Aust. J. Chem., 2013, 66, 710–717 CrossRef CAS; (c) H. Singh, J. Sindhu, J. M. Khurana, C. Sharma and K. R. Aneja, Aust. J. Chem., 2013, 66, 1088–1096 CrossRef CAS; (d) H. Singh, J. Sindhu and J. M. Khurana, RSC Adv., 2013, 3, 22360–22366 RSC; (e) H. Singh, J. Sindhu, J. M. Khurana, C. Sharma and K. R. Aneja, RSC Adv., 2014, 4, 5915–5926 RSC; (f) H. Singh, J. Sindhu and J. M. Khurana, Sens. Actuators, B, 2014, 192, 536–542 CrossRef CAS PubMed.
  21. F. Himo, T. Lovell, R. Hilgraf, V. V. Rostovtsev, L. Noodleman, K. B. Sharpless and V. V. Fokin, J. Am. Chem. Soc., 2004, 127, 210–216 CrossRef PubMed.
  22. (a) N. D. Obushak, N. T. Pokhodylo, N. I. Pidlypnyi and V. S. Matiichuk, Russ. J. Org. Chem., 2008, 44, 1522–1527,  DOI:10.1134/S1070428008100217; (b) H. Singh, J. Sindhu, J. M. Khurana, C. Sharma and K. R. Aneja, Eur. J. Med. Chem., 2014, 77, 145–154 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. CCDC 996871. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra03505h

This journal is © The Royal Society of Chemistry 2015