Next Article in Journal
Study of Azimuthal Anisotropy of High-pT Charged Particles in Au Au Collisions at √sNN = 200 GeV with RHIC-PHENIX
Previous Article in Journal
Measurements of Anisotropic Flow in Xe–Xe Collisions at √sNN = 5.44 TeV Using the ALICE Detector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis of Imidazo[1,2-a]pyridines via Multicomponent GBBR Using α-isocyanoacetamides †

by
Manuel A. Rentería Gómez
1,
Alejandro Islas-Jácome
2 and
Rocío Gámez-Montaño
1,*
1
Departamento de Química, Universidad de Guanajuato, Noria Alta S/N, Col. Noria Alta, C.P. 36050 Guanajuato, Gto., Mexico
2
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, Del. Iztapalapa, C.P. 09340 Ciudad de México, Mexico
*
Author to whom correspondence should be addressed.
Presented at the 22nd International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2018; Available Online: https://sciforum.net/conference/ecsoc-22.
Proceedings 2019, 9(1), 53; https://doi.org/10.3390/ecsoc-22-05692
Published: 14 November 2018

Abstract

:
Six novel imidazo[1,2-a]pyridines were synthesized by Groebke–Blackburn–Bienaymé reactions (GBBRs) under eco-friendly conditions (10 mol% ammonium chloride catalyst in EtOH at room temperature) with moderate to good yields (76–44%) using 2-isocyano-1-morpholino-3-phenylpropan-1-one. This is the first successful use of this type of α-isocyanoacetamide in a GBBR, as these reactive isonitriles readily undergo ring-chain tautomerization, as reported in other IMCRs (isonitrile-based multicomponent reactions). The product structures contain a peptidomimetic imidazo[1,2-a]pyridine scaffold linked to an α-aminomorpholide and are of interest to medicinal chemists.

1. Introduction

Nitrogen-fused heterocycles are becoming more popular because of their wide range of pharmacological and biological properties. Among them, nitrogen-fused azoles such as imidazo[1,2-a]pyridines have attracted interest over the past decade due to their widespread applications in medicinal chemistry, organometallics, optics, and materials science [1]. These scaffolds are present in many commercially available drugs such as olprinone (cardiotonic agent), miroprofen (analgesic), DS-1 (GABA receptor agonist), zolimidine (peptic ulcers), GSK812397 (HIV infection), and minodronic acid (osteoporosis) [2,3,4]. Imidazo[1,2-a]pyridines are termed as non-benzodiazepine drugs because they possess similar pharmacological properties to benzodiazepines but differ structurally; examples include alpidem (1, anxiety), zolpidem (2, insomnia), saripidem (3, sedative), and necopidem (4, anxiolytic) [5,6]. Additionally, the synthesis of bioimaging probe 5 for benzodiazepine receptors has recently been reported [7]. All of the examples mentioned above contain an amide fragment in their structures (Figure 1).
The most common methodologies for the synthesis of imidazo[1,2-a]pyridines are (i) the condensation of 2-aminopyridines with α-halo carbonyl compounds [8], which suffers from limitations such the scarcity of commercially available α-halo carbonyl compounds and their lachrymatory properties; (ii) copper-catalyzed three-component reactions of 2-aminopyridines, aldehydes, and alkynes [9]; and (iii) Grobke–Blackburn–Bienaymé reactions (GBBRs) between an aldehyde, a 2-aminoazine, and an isocyanide [10,11,12].
In modern synthetic chemistry, there is an urgent need to design and develop new green and efficient methodologies to synthesize complex molecules from simple materials with high atom economy. The isocyanide-based multicomponent reaction (IMCR) is a powerful tool that plays a central role in the synthesis of heterocycles [13]. The GBBR is one of the most common and efficient methodologies to synthesize imidazole analogues and the method of choice to synthesize imidazo[1,2-a]pyridine-3-amines. Normally, this reaction requires a solvent and a catalyst [5,6]. Various GBBR procedures have been reported, using catalysts such as Lewis acids, Bronsted acids, solid supports, organic bases, and inorganic salts [14]. Each of these methodologies has drawbacks such as high temperature, low yields, expensive catalysts, and/or non-green solvents. The design and development of improved GBBR procedures using green solvents and catalysts at room temperature is an underexplored field. There are few GBBR reports available towards imidazo[1,2-a]pyridine-3-amines describing the use of green catalysts [15,16]. For these reasons, it is necessary to increase efforts to develop new, efficient, mild methodologies using green, inexpensive, and readily available catalysts and solvents.
α-isocyanoacetamides present exceptional reactivity because they can undergo intramolecular ring closure. For this reason, they have been extensively explored in certain IMCRs as an Ugi three component reaction [17]. On the other hand, the use of this type of isonitrile in GBBRs is practically unexplored; in fact, there is only one such previous report, by Bienaymé in 1998 (see Scheme 1) [12].
The methodology described here allows the one-pot synthesis of new imidazo[1,2-a]pyridine-3-amines that incorporate a peptidomimetic amide fragment in the isonitrile reactant. To the best of our knowledge, the only other published method for access to this type of compound uses a synthesis strategy of GBBR followed by deprotection and peptide coupling steps (Scheme 1, Valakirev, M.Y. et al.) [7]. Therefore, the methodology described here is attractive due to the use of green reaction conditions and access to the final products in a single stage.

2. Results and Discussion

In order to develop green conditions for the GBBR, we started the synthesis of imidazo[1,2-a]pyridine-3-amine analogue 6a by reacting equimolar amounts of 2-aminopyridine (7), benzaldehyde (8a), and 2-isocyano-1-morpholino-3-phenylpropan-1-one (9). In concordance with our main line of research, green conditions were studied to optimize the reaction. Initially we performed the GBBR under neat conditions at room temperature, generating product 6a in poor yield (10%) after 5 h (Entry 1, Table 1) [18]. When the reaction was performed in water as solvent (Entry 2), only 8% of compound 6a was obtained. Changing the solvent to EtOH (Entry 3) increased the yield to 46%. Seeking a green, inexpensive, and easily available catalyst, we decided to try the reaction with a catalytic amount of NH4Cl at room temperature [19]. This raised the product yield to 72% (Entry 4). The use of iodine and montmorillonite (K-10) as catalysts in GBBR is well-documented [20,21,22,23,24], so we decided to try those catalysts in our methodology. Unfortunately, catalytic iodine or montmorillonite at room temperature resulted in lower yields of 49% and 66%, respectively (Entries 5–6). We then tested phenyl phosphinic acid, which is not a known catalyst for the GBBR, but this catalyst did not result in an improved yield (67%, Entry 7). Performing the NH4Cl-catalyzed reaction at 60 °C lowered the yield of product 6a to 49% (Entry 8, Table 1), which can be attributed to the low stability of this isocyanide in acidic media at elevated temperatures. Indeed, we detected the corresponding oxazole 13, resulting from chain-ring tautomerization of isocyanide 9, as a by-product.
Using our optimized conditions, we synthesized the series of imidazo[1,2-a]pyridines (6a–f) shown in Scheme 2. The versatility of the developed methodology was examined using different benzaldehydes bearing both electron-donating and electron-withdrawing groups (8a–d), and also 9-anthracenecarboxaldehyde (8e) and heptanaldehyde (8f). The respective products 6a–f were obtained in moderate to good yields (44–76%).
A plausible reaction mechanism involves the initial formation of a Schiff base (10a–f) via the condensation of the corresponding aldehyde (8a–f) with 2-aminopyridine (7), which is accompanied by a (nonconcerted) [4+1] cycloaddition between the protonated Schiff base (10a–f) and the isonitrile (9) to give the intermediate 12a–f. A subsequent prototropic shift generates the aromatic, fused imidazo[1,2-a]pyridine (6a–f, Scheme 3).
Figure 2 and Figure 3 show the 1H and 13C NMR spectra for the representative imidazo[1,2-a]pyridine 6a. In the 13C NMR, the carbonyl carbon signal appears at 172.1 ppm, which confirms the formation of the GBBR product and not the formation of the oxazole by an intramolecular ring closure in the isonitrile. All of the other key signals are readily observed in these spectra.

3. Experimental Section

3.1. General Information, Instrumentation, and Chemicals

1H and 13C NMR spectra were acquired on Bruker Avance III spectrometers (500 or 400 MHz). The solvent used was deuterated chloroform (CDCl3). Chemical shifts are reported in parts per million (δ/ppm). The internal reference for 1H NMR spectra is trimethylsilane at 0.0 ppm. The internal reference for 13C NMR spectra is CDCl3 at 77.0 ppm. Coupling constants are reported in Hertz (J/Hz). Multiplicities of the signals are reported using the standard abbreviations: singlet (s), doublet (d), triplet (t), quartet (q), and multiplet (m). NMR spectra were analyzed using the MestreNova software version 10.0.1–14719. IR spectra were acquired on a Perkin Elmer 100 spectrometer using an Attenuated Total Reflectance (ATR) method with neat compounds. The absorbance peaks are reported in reciprocal centimeters (υmax/cm−1). Reaction progress was monitored by Thin-Layer Chromatography (TLC) on precoated silica-gel 60 F254 plates and the spots were visualized under UV light at 254 or 365 nm. Mixtures of hexane with ethyl acetate (EtOAc) were used to run TLC and for measuring retention factors (Rf). Flash column chromatography was performed using silica gel (230–400 mesh) and mixtures of hexane with EtOAc in different proportions (v/v) as the mobile phase. All reagents were purchased from Sigma-Aldrich and were used without further purification. Chemical names and drawings were obtained using the ChemBioDraw Ultra 13.0.2.3020 software package. The purity for all the synthesized products (up to 99%) was assessed by NMR.

3.2. Synthesis and Characterization of the Imidazo[1,2-a]pyridine 6a-f

General procedure (GP): 2-Aminopyridine (7) (1.0 equiv.), the corresponding aldehyde 8a–f (1.0 equiv.), 2-isocyano-1-morpholino-3-phenylpropan-1-one (9), and NH4Cl (10% mol) were placed in a 10-mL sealed vial equipped with a magnetic stirring bar in ethanol [1.0 M]. Then, the mixture was stirred at room temperature (rt) for 12 h. The solvent was removed by rotary evaporation. The residue was purified by flash chromatography using mixtures of hexane–EtOAc (v/v) in different proportions to afford the corresponding imidazo[1,2-a]pyridine 6a–f.

1-Morpholino-3-phenyl-2-((2-phenylimidazo[1,2-a]pyridin-3-yl)amino)propan-1-one (6a)

According to the GP, 2-aminopyridine (26.0 mg, 0.276 mmol), benzaldehyde (29.0 mg, 0.276 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (67.0 mg, 0.276 mmol), and NH4Cl (1.5 mg, 0.027 mmol) were reacted together in EtOH (0.276 mL) to afford the imidazo[1,2-a]pyridine 6a (80.0 mg, 67%, rt) as a white solid; m.p. 142–144 °C; Rf = 0.22 (hexane–EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1631 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 8.06–7.97 (m, 3H), 7.57 (d, J = 9.0 Hz, 1H), 7.45–7.39 (m, 2H), 7.35–7.30 (m, 1H), 7.29–7.22 (m, 3H), 7.18–7.13 (m, 1H), 6.79–6.73 (m, 1H), 4.35–4.23 (m, 1H), 4.12–3.96 (m, 1H), 3.47–3.36 (m, 1H), 3.36–3.21 (m, 3H), 3.12–3.03 (m, 2H), 2.90–2.82 (m, 1H), 2.73–2.59 (m, 2H), 2.59–2.51 (m, 1H); 13C NMR (126 MHz, CDCl3, 25 °C): δ 172.1, 141.2, 136.6, 129.3, 128.6, 127.6, 127.2, 127.0, 124.7, 124.6, 122.6, 117.0, 112.0, 66.3, 65.8, 58.0, 45.6, 42.2, 41.5.

2-((2-(4-Chlorophenyl)imidazo[1,2-a]pyridin-3-yl)amino)-1-morpholino-3-phenylpropan-1-one (6b)

According to the GP, 2-aminopyridine (21.0 mg, 0.220 mmol), 4-chlorobenzaldehyde (31.0 mg, 0.220 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (54.0 mg, 0.220 mmol), and NH4Cl (1.0 mg, 0.022 mmol) were reacted together in EtOH (0.220 mL) to afford the imidazo[1,2-a]pyridine 6b (68.0 mg, 72%, rt) as a white solid; m.p. 81–82 °C; Rf = 0.27 (hexane–EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1618 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 8.03 (d, J = 6.7 Hz, 1H), 7.96 (d, J = 7.9 Hz, 1H), 7.55 (d, J = 8.9 Hz, 1H), 7.36 (d, J = 7.9 Hz, 1H), 7.30–7.25 (m, 3H), 7.19–7.08 (m, 3H), 6.91–6.72 (m, 1H), 4.37–4.17 (m, 1H), 4.08–3.96 (m, 1H), 3.47–3.40 (m, 1H), 3.38–3.29 (m, 2H), 3.28–3.21 (m, 1H), 3.12–3.02 (m, 2H), 2.91–2.85 (m, 2H), 2.75–2.68 (m, 2H), 2.65–2.59 (m, 1H); 13C NMR (126 MHz, CDCl3, 25 °C): δ 172.2, 141.3, 136.6, 134.0, 133.5, 131.1, 129.5, 128.8, 128.7, 128.4, 127.2, 125.0, 124.8, 122.8, 117.1, 112.3, 66.2, 65.6, 57.9, 45.5, 42.0, 41.5.

2-((2-(3,4-Dimethoxyphenyl)imidazo[1,2-a]pyridin-3-yl)amino)-1-morpholino-3-phenylpropan-1-one (6c)

According to the GP, 2-aminopyridine (18.0 mg, 0.192 mmol), 3,4-dimethoxybenzaldehyde (32.0 mg, 0.192 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (47.0 mg, 0.192 mmol), and NH4Cl (1.0 mg, 0.019 mmol) were reacted together in EtOH (0.200 mL) to afford the imidazo[1,2-a]pyridine 6c (81.0 mg, 69%, rt) as a brown gum; Rf = 0.1 (hexane–EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1628 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 8.05 (d, J = 6.7 Hz, 1H), 7.74–7.62 (m, 2H), 7.57 (d, J = 8.3 Hz, 1H), 7.28–7.20 (m, 4H), 7.16–7.08 (m, 2H), 6.91 (d, J = 8.3 Hz, 1H), 6.84–6.79 (m, 1H), 4.47–4.25 (m, 1H), 4.19–3.99 (m, 4H), 3.94 (s, 3H), 3.47–3.42 (m, 1H), 3.36–3.27 (m, 3H), 3.11–3.06 (m, 2H), 2.95–2.90 (m, 1H), 2.75–2.66 (m, 2H), 2.64–2.54 (m, 1H); 13C NMR (126 MHz, CDCl3, 25 °C): δ 171.2, 153.3, 151.6, 142.0, 143.5, 131.1, 129.5, 128.3, 128.7, 128.4, 127.2, 125.0, 124.8, 122.8, 117.1, 112.3, 66.2, 65.6, 57.9, 56.7, 56.3 45.5, 42.0, 41.5.

1-Morpholino-2-((2-(4-nitrophenyl)imidazo[1,2-a]pyridin-3-yl)amino)-3-phenylpropan-1-one (6d)

According to the GP, 2-aminopyridine (19.0 mg, 0.198 mmol), 4-nitrobenzaldehyde (30.0 mg, 0.198 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (49.0 mg, 0.198 mmol), and NH4Cl (1.0 mg, 0.019 mmol) were reacted together in EtOH (0.200 mL) to afford the imidazo[1,2-a]pyridine 6d (72.0 mg, 76%, rt) as a orange solid; m.p. 110–113 °C; Rf = 0.22 (hexane–EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1615 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 8.25–8.13 (m, 4H), 7.33–7.14 (m, 8H), 6.91–6.79 (m, 1H), 4.59–4.36 (m, 1H), 4.13–3.99 (m, 1H), 3.53–3.44 (m, 1H), 3.40–3.24 (m, 3H), 3.17–3.03 (m, 2H), 2.97–2.89 (m, 1H), 2.84–2.63 (m, 3H); 13C NMR (126 MHz, CDCl3, 25 °C): δ 171.3, 142.1, 134.6, 133.0, 132.4, 132.1, 129.9, 128.9, 128.8, 128.2, 127.6, 125.1, 124.3, 122.0, 117.8, 112.4, 66.3, 65.8, 58.0, 45.9, 42.5, 41.9.

2-((2-Hexylimidazo[1,2-a]pyridin-3-yl)amino)-1-morpholino-3-phenylpropan-1-one (6e)

According to the GP, 2-aminopyridine (26.0 mg, 0.280 mmol), heptanaldehyde (32.0 mg, 0.280 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (68.0 mg, 0.280 mmol), and NH4Cl (1.5 mg, 0.028 mmol) were reacted together in EtOH (0.280 mL) to afford the imidazo[1,2-a]pyridine 6e (54.0 mg, 44%, rt) as a brown oil; Rf = 0.22 (hexane –EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1631 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 7.54 (d, J = 8.9 Hz, 1H), 7.42–7.36 (m, 1H), 7.32–7.22 (m, 3H), 7.21–7.16 (m, 3H), 7.14–7.08 (m, 3H), 4.10 (d, J = 9.5 Hz, 1H), 3.79 (q, J = 8.0 Hz, J = 16.2 Hz, 1H), 3.65–3.58 (m, 2H), 3.55–3.44 (m, 2H), 3.27–3.16 (m, 1H), 3.07–3.01 (m, 1H), 2.98–2.86 (m, 4H), 2.59 (t, J = 7 7 Hz, 2H), 1.70–1.62 (m, 1H), 1.62–1.55 (m, 1H), 1.34–1.17 (m, 7H), 0.84–0.79 (m, 3H),; 13C NMR (126 MHz, CDCl3, 25 °C): δ 172.2, 134.0, 133.5, 131.1, 129.5, 128.8, 128.7, 128.4, 127.2, 125.0, 124.8, 122.8, 66.3, 65.8, 58.0, 45.6, 42.2, 38.8, 31.5, 29.4, 29.0, 28.7, 22.7, 22.5, 14.0.

2-((2-(Anthracen-9-yl)imidazo[1,2-a]pyridin-3-yl)amino)-1-morpholino-3-phenylpropan-1-one(6f)

According to the GP, 2-aminopyridine (19.0 mg, 0.203 mmol), 9-anthracenecarboxaldehyde (42.0 mg, 0.203 mmol), 2-isocyano-1-morpholino-3-phenylpropan-1-one (50.0 mg, 0.203 mmol), and NH4Cl (1.0 mg, 0.020 mmol) were reacted together in EtOH (0.200 mL) to afford the imidazo[1,2-a]pyridine 6f (72.0 mg, 67%, rt) as a yellow solid; m.p. 186–187 °C; Rf = 0.22 (hexane–EtOAc = 2:3 v/v); FT-IR (ATR) υmax/cm−1 1640 (C = O); 1H NMR (500 MHz, CDCl3, 25 °C): δ 8.57 (s, 1H), 8.14–8.04 (m, 2H), 8.02–7.93 (m, 1H), 7.88–7.79 (m, 1H), 7.77–7.72 (m, 1H), 7.70–7.64 (m, 1H), 7.55–7.48 (m, 2H), 7.46–7.40 (m, 2H), 7.28–7.20 (m, 1H), 7.14–7.05 (m, 3H), 6.97–6.83 (m, 1H), 6.65–6.56 (m, 2H), 4.31–4.16 (m, 1H), 3.63–3.58 (m, 1H), 3.22–3.04 (m, 4H), 2.74–2.66 (m, 1H), 2.62–2.50 (m, 3H), 2.45–2.32 (m, 1H); 13C NMR (126 MHz, CDCl3, 25 °C): δ 173.0, 141.2, 140.2, 137.6, 135.0, 132.5, 132.1, 129.9, 129.4, 128.7, 127.7, 127.4, 126.2, 124.0, 123.9, 123.5, 118.1, 115.3, 67.1, 66.6, 57.0, 45.9, 42.1, 41.3.

4. Conclusions

In conclusion, we have developed an efficient and mild GBBR-based methodology for the green synthesis of new imidazo[1,2-a]pyridine-3-amines in good yields using α-isocyanoacetamides that incorporate a peptidomimetic amide fragment in the isonitrile component. These results can be interpreted as a reactivity study, the GBBR versus the Ugi three component reaction based on the use of chain ring tautomerizable isonitriles, and as can be seen, iminium ion trapping is the kinetic step favored over the oxazole ring formation. To the best of our knowledge, this is the first example of this reaction using a green, readily available, inexpensive catalyst (NH4Cl) and solvent (EtOH) at room temperature. In addition, the compounds were synthesized in a single step, an improvement over previous multi-stage efforts. This methodology allows the synthesis of imidazo[1,2-a]pyridine-3-amines containing amide substituents that are structurally similar to established drug molecules.

Author Contributions

All authors contributed equally to this work.

Acknowledgments

M.A.R.G. thanks CONACYT for a scholarship (707974/585367). R.G.-M. thanks CONACYT-México (CB-2016-285622) for financial support, Dr. Gustavo Rangel-Porras for providing the montorillonite catalyst K-10, and the Laboratorio Nacional de Caracterización de Propiedades Fisicoquímícas y Estructura Molecular (CONACYT-México, Project: 123732) for the instrumentation time provided. The authors thank David A. Vosburg for helpful comments on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Couty, F.; Evano, G. Comprehensive Heterocyclic Chemistry, 3rd, ed.; Katritzky, A.R., Ramsden, C.A., Scriven, E.F.V., Taylor, R.J.K., Eds.; Elsevier: Oxford, UK, 2008; Volume 11, p. 409. [Google Scholar]
  2. Song, G.; Zhang, Y.; Li, X. Rhodium and Iridium Complexes of Abnormal N-Heterocyclic Carbenes Derived from Imidazo[1,2-a]pyridine. Organometallics 2008, 27, 1936–1943. [Google Scholar] [CrossRef]
  3. John, A.; Shaikh, M.M.; Ghosh, P. Palladium complexes of abnormal N-heterocyclic carbenes as precatalysts for the much preferred Cu-free and amine-free Sonogashira coupling in air in a mixed-aqueous medium. Dalton Trans. 2009, 10581–10591. [Google Scholar] [CrossRef] [PubMed]
  4. Enguehard-Gueiffier, C.; Gueiffier, A. Recent Progress in the Pharmacology of Imidazo[1,2-a]pyridines Mini-Rev. Med. Chem. 2007, 7, 888–899. [Google Scholar] [CrossRef]
  5. Devi, N.; Rawal, R.K.; Singh, V. Diversity-oriented synthesis of fused-imidazole derivatives via Groebkee-Blackburne-Bienayme reaction: A review. Tetrahedron 2015, 71, 183–232. [Google Scholar] [CrossRef]
  6. Volkova, Y.; Gevorgyan, V. Synthesis of functionalyzed imidazo[1,2-a]pyridines via domino A3-coupling/cycloisomerization approach. Chem. Heterocycl. Compd. 2017, 53, 409–412. [Google Scholar] [CrossRef]
  7. Burchak, O.N.; Mugherli, L.; Ostuni, M.; Lacapère, J.J.; Balakirev, M.Y. Combinatorial Discovery of Fluorescent Pharmacophores by Multicomponent Reactions in Droplet Arrays J. Am. Chem. Soc. 2011, 133, 10058–10061. [Google Scholar] [CrossRef]
  8. Pericherla, K.; Kaswan, P.; Pandey, K.; Kumar, A. Recent Developments in the Synthesis of Imidazo[1,2-a]pyridines. Synthesis 2015, 47, 887–912. [Google Scholar] [CrossRef]
  9. Bagdi, A.K.; Santra, S.; Monir, K.; Hajra, A. Synthesis of imidazo[1,2-a]pyridines: A decade update. Chem. Commun. 2015, 51, 1555–1575. [Google Scholar] [CrossRef]
  10. Groebke, K.; Weber, L.; Mehlin, F. Synthesis of Imidazo[1,2-a] annulated Pyridines, Pyrazines and Pyrimidines by a Novel Three-Component Condensation. Synlett 1998, 6, 661–663. [Google Scholar] [CrossRef]
  11. Blackburn, C.; Guan, B.; Fleming, P.; Shiosaki, K.; Tsai, S. Parallel synthesis of 3-aminoimidazo[1,2-a]pyridines and pyrazines by a new three-component condensation. Tetrahedron Lett. 1998, 39, 3635–3638. [Google Scholar] [CrossRef]
  12. Bienaymé, H.; Bouzid, K.A. New Heterocyclic Multicomponent Reaction For the Combinatorial Synthesis of Fused 3-Aminoimidazoles. Angew Chem. 1998, 110, 2349–2352. [Google Scholar] [CrossRef]
  13. Vicente-Garcia, E.; Kielland, N.; Lavilla, R. Functionalization of Heterocycles by MCRs. In Multicomponent Reactions in Organic Synthesis; Zhu, J., Wang, Q., Wang, M.X., Eds.; Wiley-VCH: Weinheim, Germany, 2015; Chapter 6; pp. 159–178. [Google Scholar] [CrossRef]
  14. Shaaban, S.; Abdel-Wahab, B.F. Groebke-Blackburn-Bienaymé multicomponent reaction: Emerging chemistry for drug discovery. Mol. Divers. 2016, 20, 233–254. [Google Scholar] [CrossRef] [PubMed]
  15. Che, C.; Xiang, J.; Wang, G.-X.; Fathi, R.; Quan, J.-M.; Yang, Z. One-Pot Synthesis of Quinoline-Based Tetracycles by a Tandem Three-Component Reaction. J. Comb. Chem. 2007, 9, 982–989. [Google Scholar] [CrossRef] [PubMed]
  16. Shaabani, A.; Soleimani, E.; Maleki, A.; Moghimi-Rad, J. A novel class of extended pi-conjugated systems: One-pot synthesis of bis-3-aminoimidazo[1,2-a]pyridines, pyrimidines and pyrazines. Mol. Divers. 2009, 13, 269–274. [Google Scholar] [CrossRef]
  17. Elders, N.; Ruijter, E.; Nenajdenko, V.G.; Orru, R.V.A. α-Acidic Isocyanides in Multicomponent Chemistry. In Synthesis of Heterocycles via Multicomponent Reactions I. Topics in Heterocyclic Chemistry; Orru, R., Ruijter, E., Eds.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 23. [Google Scholar] [CrossRef]
  18. Claudio-Catalán, M.A.; Pharande, S.G.; Quezada-Soto, A.; Kishore, K.G.; Rentería-Gómez, A.; Padilla-Vaca, F.; Gámez-Montaño, R. Solvent- and Catalyst-Free One-Pot Green Bound-Type Fused Bis- Heterocycles Synthesis via Groebke-Blackburn-Bienaymé Reaction/SNAr/Ring-Chain Azido-Tautomerization Strategy. ACS Omega 2018, 3, 5177–5186. [Google Scholar] [CrossRef]
  19. Kurva, M.; Pharande, S.G.; Quezada-Soto, A.; Gámez-Montaño, R. Ultrasound assisted green synthesis of bound type bis-heterocyclic carbazolyl imidazo[1,2-a]pyridines via Groebke-Blackburn-Bienayme reaction. Tetrahedron Lett. 2018, 59, 1596–1599. [Google Scholar] [CrossRef]
  20. Puttaraju, K.B.; Shivashankar, K. Iodine-catalyzed three component reaction: A novel synthesis of 2-aryl-imidazo[1,2-a]pyridine scaffolds. RSC Adv. 2013, 3, 20883–20890. [Google Scholar] [CrossRef]
  21. Varma, R.S.; Kumar, D. Microwave-accelerated three-component condensation reaction on clay: Solvent-free synthesis of imidazo[1,2-a] annulated pyridines, pyrazines and pyrimidines. Tetrahedron Lett. 1999, 40, 7665. [Google Scholar] [CrossRef]
  22. Bode, M.L.; Gravestock, D.; Moleele, S.S.; van der Westhuyzen, C.W.; Pelly, S.C.; Steenkamp, P.A.; Hoppe, H.C.; Khan, T.; Nkabinde, L.A. Imidazo[1,2-a]pyridin-3-amines as potential HIV-1 non-nucleoside reverse transcriptase inhibitors. Bioorg. Med. Chem. 2011, 19, 4227. [Google Scholar] [CrossRef]
  23. Dahan-Farkas, N.; Langley, C.; Rousseau, A.L.; Yadav, D.B.; Davids, H.; de Koning, C.B. 6-Substituted imidazo[1,2-a]pyridines: Synthesis and biological activity against colon cancer cell lines HT-29 and Caco-2. Eur. J. Med. Chem. 2011, 46, 4573. [Google Scholar] [CrossRef]
  24. Mert-Balci, F.; Conrad, J.; Beifuss, U. Microwave-assisted three-component reaction in conventional solvents and ionic liquids for the synthesis of amino-substituted imidazo[1,2-a]pyridines. Arch. Org. Chem. 2012, 243. [Google Scholar] [CrossRef]
Figure 1. Selected imidazo[1,2-a]pyridines with pharmacological activity and our synthetic targets.
Figure 1. Selected imidazo[1,2-a]pyridines with pharmacological activity and our synthetic targets.
Proceedings 09 00053 g001
Scheme 1. Previous reports and our work.
Scheme 1. Previous reports and our work.
Proceedings 09 00053 sch001
Scheme 2. Sustrate scope.
Scheme 2. Sustrate scope.
Proceedings 09 00053 sch002
Scheme 3. Plausible reaction mechanism involved in the GBBR toward imidazo[1,2-a]pyridines 6a-h.
Scheme 3. Plausible reaction mechanism involved in the GBBR toward imidazo[1,2-a]pyridines 6a-h.
Proceedings 09 00053 sch003
Figure 2. 1H NMR spectrum of imidazo[1,2-a]pyridine 6a.
Figure 2. 1H NMR spectrum of imidazo[1,2-a]pyridine 6a.
Proceedings 09 00053 g002
Figure 3. 13C NMR spectrum of imidazo[1,2-a]pyridine 6a.
Figure 3. 13C NMR spectrum of imidazo[1,2-a]pyridine 6a.
Proceedings 09 00053 g003
Table 1. Screening conditions for synthesis of imidazo[1,2-a]pyridine-3-amine 6a.
Table 1. Screening conditions for synthesis of imidazo[1,2-a]pyridine-3-amine 6a.
Proceedings 09 00053 i001
1------rt10
2H2O---rt8
3EtOH---rt46
4EtOHNH4Clrt72
5EtOHI2rt49
6EtOHK-10rt66
7EtOHPhenyl phosphinic acidrt67
8EtOHNH4Cl6049
a All reactions were carried out using equimolar amounts of 7, 8a, and 9 for 12 h. b [1.0 M] c Isolated yield. rt = room temperature. In all reactions, oxazole 13 was detected as a by-product.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gómez, M.A.R.; Islas-Jácome, A.; Gámez-Montaño, R. Synthesis of Imidazo[1,2-a]pyridines via Multicomponent GBBR Using α-isocyanoacetamides. Proceedings 2019, 9, 53. https://doi.org/10.3390/ecsoc-22-05692

AMA Style

Gómez MAR, Islas-Jácome A, Gámez-Montaño R. Synthesis of Imidazo[1,2-a]pyridines via Multicomponent GBBR Using α-isocyanoacetamides. Proceedings. 2019; 9(1):53. https://doi.org/10.3390/ecsoc-22-05692

Chicago/Turabian Style

Gómez, Manuel A. Rentería, Alejandro Islas-Jácome, and Rocío Gámez-Montaño. 2019. "Synthesis of Imidazo[1,2-a]pyridines via Multicomponent GBBR Using α-isocyanoacetamides" Proceedings 9, no. 1: 53. https://doi.org/10.3390/ecsoc-22-05692

Article Metrics

Back to TopTop