Next Article in Journal
Super Early Scan of PSMA PET/CT in Evaluating Primary and Metastatic Lesions of Prostate Cancer
Previous Article in Journal
Three New Stigmatellin Derivatives Reveal Biosynthetic Insights of Its Side Chain Decoration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Piperidine-Iodine as Efficient Dual Catalyst for the One-Pot, Three-Component Synthesis of Coumarin-3-Carboxamides

by
Manuel Velasco
1,
Nancy Romero-Ceronio
1,
Rosalía Torralba
2,
Oswaldo Hernández Abreu
1,
Miguel A. Vilchis-Reyes
1,
Erika Alarcón-Matus
1,
Erika M. Ramos-Rivera
1,
David M. Aparicio
3,
Jacqueline Jiménez
4,
Eric Aguilar García
4,
David Cruz Cruz
5,
Clarisa Villegas Gómez
5 and
Cuauhtémoc Alvarado
1,*
1
División Académica de Ciencias Básicas, Universidad Juárez Autónoma de Tabasco, Carretera Cunduacán-Jalpa Km 1, Col. La Esperanza, Cunduacán C.P. 86690, Tabasco, Mexico
2
Complejo Regional Mixteca, Campus Izúcar de Matamoros, Benemérita Universidad Autónoma de Puebla, Carr. Atlixco-Izúcar de Matamoros 141, San Martín Alchichica, Izúcar de Matamoros C.P. 74570, Puebla, Mexico
3
Instituto de Ciencias, Benemérita Universidad Autónoma de Puebla, Edif. IC-9 Complejo de Ciencias C.U. 72570, Puebla, Mexico
4
Facultad de Ciencias Químicas, Benemérita Universidad Autónoma de Puebla, Puebla C.U. 72570, Puebla, Mexico
5
Departamento de Química, División de Ciencias Naturalesy Exactas, Universidad de Guanajuato, Noria Alta S/N 36050, Guanajuato, Mexico
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(14), 4659; https://doi.org/10.3390/molecules27144659
Submission received: 16 June 2022 / Revised: 15 July 2022 / Accepted: 19 July 2022 / Published: 21 July 2022
(This article belongs to the Special Issue Multicomponent Reactions (MCRs))

Abstract

:
A simple and efficient one-pot, three-component synthetic method for the preparation of coumarin-3-carboxamides was carried out by the reaction of salicylaldehyde, aliphatic primary/secondary amines, and diethylmalonate. The protocol employs piperidine-iodine as a dual system catalyst and ethanol, a green solvent. The main advantages of this approach are that it is a metal-free and clean reaction, has low catalyst loading, and requires no tedious workup.

Graphical Abstract

1. Introduction

Today, the development of efficient and environmentally friendly synthesis to obtain complex and highly substituted molecules is one of the most exciting topics for the synthetic chemistry community. In this sense, multicomponent reactions (MCRs) have emerged as a powerful synthetic tool in organic synthesis as an attractive alternative to the conventional multi-step synthesis. These multicomponent strategies provide great molecular diversity in a single step and in a highly efficient manner [1,2,3]. The utility of these processes has been confirmed by the synthesis of a large number of compounds with remarkable biological activity [4,5].
Among the pharmacologically active products, coumarin and its derivatives have attracted attention due to their broad spectrum of applications in medicinal chemistry [6,7]. In particular, coumarin-3-carboxamide has proven to be an important structural core that exhibits diverse biological activities, such as anticancer [8,9], antioxidant [10,11], anti-inflammatory [12], and anticoagulant [13] properties, as well as inhibition against β-secretase (BACE1) [14,15], monoamine oxidase (MAO) [16], acetylcholinesterase [17], and tumorigenesis [18]. In addition, these compounds have found applications as fluorescent probes for Cu2+ and Fe3+ detection [19,20] and molecular sensors for monitoring O2 levels in living cells [21].
Accordingly, several strategies for the synthesis of this important class of compounds have been described in the literature. The traditional approach involves multistep elementary reactions including condensation reactions between activated coumarin derivatives or methyl cyanoacetate with amines (Figure 1a) [22,23]. Alternatively, MCRs using magnetic nanoparticles (NPs) such as Ni–NiO and Fe3O4 have also been applied in the synthesis of a variety of coumarin-3-carboxamides (Figure 1b) [24,25]. Other reported approaches to the synthesis of coumarin-3-carboxamide involve regioselective carboxamidation of coumarins at C-3 with formamides by using the radical initiator tert-butyl peroxybenzoate (TBPB) (Figure 1c) [26], and amidation reactions of coumarin-3-carboxylic acids using tetraalkylthiuram disulfides as amine sources (Figure 1d) [27].
However, most of the above-mentioned protocols suffer from several limitations such as pore functionalization of the substrates, harsh reaction conditions, expensive reagents, long reaction times, laborious workup, and metal-based catalysis. Therefore, from the environmental and economic perspectives, the development of an easy-to-implement protocol for the preparation of coumarin-3-carboxamides would be of great importance in synthetic organic chemistry.
On the other hand, iodine-catalyzed reactions have attracted great research interest because of their simplicity in operation, enhanced reaction rates, and great selectivity. Iodine is also a cheap, non-toxic, and water-tolerant catalyst [28,29,30], which has been explored as a powerful catalyst for MCRs, due to its unique catalytic properties [31,32,33,34].
Based on the use of iodine and the fact that, to the best of our knowledge, no free metal and one-pot syntheses of coumarin-3-carboxamides have been reported, we present here a simple and efficient one-pot, three-component synthetic method for the preparation of coumarin-3-carboxamides using piperidine and molecular iodine as a dual-catalyst system (Figure 1e). The notable advantages of this protocol are (a) transition-metal-free protocol, (b) use of iodine as a catalyst, (c) use of ethanol as a solvent, and (d) workup simplicity.

2. Results

Initially, we explore the multicomponent reaction involving 2-hydroxybenzaldehyde (1), ethanolamine (2), and diethyl malonate (DEM). The reaction was carried out in reflux ethanol in the presence of iodine (10 mol%) and piperidine (10 mol%) over the course of 8 h. As a result, the 3-amidocoumarin 3a was afforded at a 79% yield (Scheme 1).
In view of this success, the optimization of the reaction was performed. Thus, the effects of the solvent, base, and catalyst were studied (Table 1). The solvent effect was studied in the first place, and in a general way, protic solvents produced superior results to aprotic solvents in terms of selectivity (Table 1, entries 1–4 vs. 6–10), since the coumarin-3-carboxamide 3a was the only product detected, though in regular yields (37–79%). Among the protic solvents, the green solvent ethanol [35] was the most effective (Table 1, entry 1), because compound 3a was obtained with the best yield (79%). Otherwise, the use of water and aprotic polar solvents such as CH2Cl2, MeCN, EtOAc, THF, and DMF provided a mixture of coumarin-3-carboxamide 3a, imine 4a, and ethyl coumarin-3-carboxylate 5 (Table 1, entries 5–10). The solvent-free synthesis was found to be interesting (Table 1, entry 11); unfortunately, it required a high temperature (250 °C) and up to 14 h to complete the reaction. In order to establish the role of piperidine, different bases such as Et3N, 1,8-diazabicycloundec-7-ene (DBU), and L-proline were tested, and none of them were effective for the obtention of compound 3a (Table 1, entries 12−14), sadly. Finally, the ratio of base to iodine was investigated (Table 1, entries 15−21). We observed that increasing the amount of iodine from 10 mol% to 20 mol% resulted in a slight decrease in the reaction yield from 79% to 75% (Table 1, entry 1 vs. entry 15). The use of 15 mol% did not affect the product yield (Table 1, entry 16). Fortunately, when using 5 mol% of iodine, the highest yield (85%) for the reaction was found (Table 1, entry 17). However, on reducing the amount of iodine, the product yield decreased (Table 1, entry 18).
Having established the optimal reaction conditions, we subsequently explored the scope for the synthesis of various coumarin-3-carboxamides 3a–n and 7a–e (Scheme 2 and Scheme 3) for this methodology. Accordingly, differently functionalized 2-hydroxy-benzaldehydes 1 were subjected to the multicomponent reaction with diverse aminoalcohols, primary amines 2, and DEM (Scheme 2). Starting with aminoalcohols, the results showed good to very good yields of products 3a, 3d, and 3e (85%, 76%, and 85%, respectively) when ethanolamine, a primary, unhindered amine was used. The yield diminished (3b, 50%) when (R)-(–)-2-phenylglycinol, a hindered amine was used, and surprisingly, the yield obtained using 2-amino-2-(hydroxymethyl)propane-1,3-diol was very good (3c, 88%). This result may be due to the formation of a hydrogen bond between hydroxyl groups of 2-amino-2-(hydroxymethyl)propane-1,3-diol and the carbonyl group of DEM (SI, Scheme S1). Primary amines attached to methylene carbons (i.e., benzylamine and tryptamine) reacted satisfactorily, and the corresponding products 3fk were obtained with regular to very good yields (66–90%). The coumarin-3-carboxamide 3l, derived from aniline (an aromatic amine) was isolated in regular yield (70%). As expected, yields diminished when the primary amine was attached to a methine carbon, due to the increase in steric hindrance. For example, (S)-(–)-α-methylbenzylamine produced 3m at a 50% yield, while isopropylamine delivered 3n in 43% yield. Additionally, the influence of substituents on the aromatic ring of aldehydes was tested. A soft decrease in yield (3d, 76%) was observed when using 3-methoxysalicylaldehyde (with an electron-releasing group). In contrast, a very good yield (3e, 85%) was obtained when using 5-bromosalicylaldehyde (with an electron-withdrawing group).
To further explore the substrate scope, we investigated the multicomponent reaction with secondary amines (including cyclic and aliphatic) 6 (Scheme 3). Unfortunately, yields were from low to regular (22%–54%); for instance, a 23% yield of 7a was obtained when using dimethylamine, and only a 54% yield of compound 7e was isolated when using piperidine, a cyclic amine. Indeed, no reaction occurred when diethylamine, diisopropylamine, or diphenylamine was used (not shown). Again, some unfavorable steric factors may be responsible for both the low yield and not the formation of coumarin-3-carboxamides from long chain secondary amines [24].
Structures of synthesized compounds were confirmed by analytical and spectral data (see SI). In a representative example, the 1H-NMR spectrum of N-(2-hydroxyethyl)-2-oxo-2H-chromene-3-carboxamide (3a) showed one singlet signal at δ = 8.90 ppm, corresponding to H-4 (alkene) of the coumarin nucleus. Signals at 8.00, 7.76, 7.52, and 7.45 ppm were assigned to aromatic protons of the coumarin core. NH and OH protons appear as triplet signals at δ = 8.86 and 4.92 ppm, respectively. Two quadruplet signals at δ = 3.55 and 3.41 ppm were assigned to the protons of the two methylene. The 13C NMR spectrum of 3a exhibited 12 distinct signals in agreement with the suggested structure [36].
In order to show the applicability of this methodology, compound 3g was synthesized at a 5 mmol scale. As a result, compound 3g was afforded at a 87% yield (1.34 g). This result further proved the feasibility to apply this methodology to a larger-scale process (Scheme 4).
To gain insight into the reaction mechanism, control experiments were conducted as shown in Scheme 5. Initially, 2-hydroxybenzaldehyde (1) and ethanolamine (2) were stirred in ethanol at reflux for 2 h, and the imine 4a could be isolated at a 90% yield. Then, imine 4a was further reacted with DEM under standard conditions; this way, the target product 3a could be obtained at a 65% yield (Scheme 5a). Next, the ester 5 was employed in the direct reaction with ethanolamine, and under the standard conditions, a mixture of the 3-amidocoumarin 3a, imine 4a, and N1,N3-bis(2-hydroxyethyl)malonamide 8 (4:60:36) was obtained (Scheme 5b). The obtained result indicates that the formation of 3a from the coumarin 5 is not efficient, mainly due to a competitive reaction involving the attack at the 4-position of coumarin [37] and also due to the fact that imine 4a is the most likely intermediate of the reaction.
Based on the above results, a plausible mechanism is proposed (Scheme 6). Firstly, the condensation between 1 and ethanolamine 2, in the presence of iodine, leads to imine 4a. Secondly, imine 4a undergoes a Mannich-type reaction with DEM in the presence of piperidine to form intermediate I. Thirdly, an intramolecular cyclization produces intermediate chroman-2-one II, which suffers a second intramolecular cyclization to afford intermediate beta-lactam III. Finally, a [1,3]-amino rearrangement involving the β-lactam results in the formation of the target molecule 3a [38]. During the reaction mechanism, iodine is proposed to activate the imine and carbonyl groups as it behaves as a mild Lewis acid [39], thus facilitating the transformations, as long as piperidine acts as a base, which allows for protonic transferences [32].

3. Materials and Methods

3.1. General Information

All chemicals were purchased from Sigma Aldrich (Toluca, Mexico). Melting points were determined on a Stuart SMP10 apparatus by the open capillary technique and are uncorrected. 1H and 13C NMR spectra were recorded at 600 MHz and 150 MHz, respectively, in CDCl3 or DMSO-d6 using a Bruker AscendTM Spectrometer. Chemical shifts are given in ppm and reported to the residual solvent peak (CDCl3: 7.26 ppm for 1H and 77.16 ppm for 13C; DMSO-d6: 2.50 ppm for 1H and 39.51 ppm for 13C). Data are reported as follows: chemical shift (δ), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br s = broad singlet), coupling constant(s) (J, Hz), and integration. Analytical TLC was performed on silica gel 60 F254 plates. IR spectra were obtained using an FT-IR spectrometer, Spectrum One, Perkin Elmer.

3.2. General Procedures and Compound Characterization Data for Coumarin-3-carboxamides 3a–n and 7a–e

To a stirred solution of salicylaldehyde (0.122 g, 1 mmol), primary or secondary amine (1.2 mmol) and diethyl malonate (1.2 mmol) in absolute ethanol (2 mL) was treated with piperidine (10 mol%) and iodine (5 mol%) and refluxed for 8h. After completion, the mixture was filtered and the precipitate was washed with cold ethanol (4 mL) to afford the pure products 3an and 7ae. If necessary, further purification was performed by recrystallization from ethanol. The identity of the known products was confirmed by a comparison of their spectroscopic data and physical properties [20,24,40,41,42,43,44,45,46]. The characterization data of synthesized compounds are given in the Supplementary Materials.

4. Conclusions

We have developed a piperidine/iodine-promoted three-component reaction for the synthesis of various coumarin-3-carboxamides. The simplicity of the synthetic protocol and availability of diverse starting materials make this an attractive strategy for the synthesis of this class of compounds. Further applications of this catalytic system to other multicomponent reactions are now underway.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27144659/s1, characterization data, and 1H-NMR and 13C-NMR of compounds 3a–n and 7a–e.

Author Contributions

Conceptualization, N.R.-C. and O.H.A.; methodology, M.V. and R.T.; software, M.A.V.-R.; validation, E.A.-M. and D.C.C.; formal analysis, O.H.A.; investigation, E.M.R.-R.; resources, C.V.G.; data curation, J.J. and E.A.G.; writing—original draft preparation, M.V. and C.A.; writing—review and editing, D.M.A.; visualization, M.V.; supervision, C.A. and N.R.-C.; project administration, C.A.; funding acquisition, N.R.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universidad Juárez Autónoma de Tabasco, Mexico.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

E.A.G. thanks CONACYT for scholarship 553494 during PhD studies.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Bienaymé, H.; Hulme, C.; Oddon, G.; Schmitt, P. Maximizing Synthetic Efficiency: Multi-Component Transformations Lead the Way. Chem. Eur. J. 2000, 6, 3321–3329. [Google Scholar] [CrossRef]
  2. Ruijter, E.; Scheffelaar, R.; Orru, R.V.A. Multicomponent Reaction Design in the Quest for Molecular Complexity and Diversity. Angew. Chem. Int. Ed. 2011, 50, 6234–6246. [Google Scholar] [CrossRef] [PubMed]
  3. Zhi, S.; Ma, X.; Zhang, W. Consecutive multicomponent reactions for the synthesis of complex molecules. Org. Biomol. Chem. 2019, 17, 7632–7650. [Google Scholar] [CrossRef]
  4. Dömling, A.; Wang, W.; Wang, K. Chemistry and Biology Of Multicomponent Reactions. Chem. Rev. 2012, 112, 3083–3135. [Google Scholar] [CrossRef] [Green Version]
  5. Wang, Z.; Dömling, A. Multicomponent Reactions in Medicinal Chemistry. In Multicomponent Reactions towards Heterocycles: Concepts and Applications; Van der Eycken, E.V., Sharma, U.K., Eds.; John Wiley and Sons: Weinheim, Germany, 2021; pp. 91–137. [Google Scholar]
  6. Murray, R.D.H. Coumarins. Nat. Prod. Rep. 1995, 12, 477–505. [Google Scholar] [CrossRef]
  7. Peng, X.M.; Damu, L.V.; Zhou, C.H. Current developments of coumarin compounds in medicinal chemistry. Curr. Pharm. Des. 2013, 19, 3884–3930. [Google Scholar] [CrossRef] [PubMed]
  8. Reddy, N.S.; Gumireddy, K.; Mallireddigari, M.R.; Cosenza, S.C.; Venkatapuram, P.; Bell, S.C.; Reddy, E.P.; Reddy, M.R. Novel coumarin-3-(N-aryl)carboxamides arrest breast cancer cell growth by inhibiting ErbB-2 and ERK1. Bioorg. Med. Chem. 2005, 13, 3141–3147. [Google Scholar] [CrossRef] [PubMed]
  9. Kempen, I.; Hemmer, M.; Counerotte, S.; Pochet, L.; de Tullio, P.; Foidart, J.M.; Blacher, S.; Noe¨l, A.; Frankenne, F.; Pirott, B. 6- Substituted 2-oxo-2H-1-benzopyran-3-carboxylic acid derivatives in a new approach of the treatment of cancer cell invasion and metastasis. Eur. J. Med. Chem. 2008, 43, 2735–2750. [Google Scholar] [CrossRef]
  10. Melagraki, G.; Afantitis, A.; Igglessi-Markopoulou, O.; Detsi, A.; Koufaki, M.; Kontogiorgis, C.; Hadjipavlou-Litina, D.J. Synthesis and evaluation of the antioxidant and anti-inflammatory activity of novel coumarin-3-aminoamides and their alpha-lipoic acid adducts. Eur. J. Med. Chem. 2009, 44, 3020–3026. [Google Scholar] [CrossRef]
  11. Yang, Y.; Liu, Q.-W.; Shi, Y.; Song, Z.-G.; Jin, Y.-H.; Liu, Z.-Q. Design and synthesis of coumarin-3-acylamino derivatives to scavenge radicals and to protect DNA. Eur. J. Med. Chem. 2014, 84, 1–7. [Google Scholar] [CrossRef]
  12. E Bylov, I.; Vasylyev, M.V.; Bilokin, Y.V. Synthesis and anti-inflammatory activity of N-substituted 2-oxo-2H-1-benzopyran-3-carboxamides and their 2-iminoanalogues. Eur. J. Med. Chem. 1999, 34, 997–1001. [Google Scholar] [CrossRef]
  13. Robert, S.; Bertolla, C.; Masereel, B.; Dogné, J.-M.; Pochet, L. Novel 3-Carboxamide-coumarins as Potent and Selective FXIIa Inhibitors. J. Med. Chem. 2008, 51, 3077–3080. [Google Scholar] [CrossRef] [PubMed]
  14. Edraki, N.; Firuzi, O.; Foroumadi, A.; Miri, R.; Madadkar-Sobhani, A.; Khoshneviszadeh, M.; Shafiee, A. Phenylimino-2 H -chromen-3-carboxamide derivatives as novel small molecule inhibitors of β-secretase (BACE1). Bioorg. Med. Chem. 2013, 21, 2396–2412. [Google Scholar] [CrossRef] [PubMed]
  15. Iraji, A.; Firuzi, O.; Khoshneviszadeh, M.; Tavakkoli, M.; Mahdavi, M.; Nadri, H.; Edraki, N.; Miri, R. Multifunctional iminochromene-2H-carboxamide derivatives containing different aminomethylene triazole with BACE1 inhibitory, neuroprotective and metal chelating properties targeting Alzheimer’s disease. Eur. J. Med. Chem. 2017, 141, 690–702. [Google Scholar] [CrossRef]
  16. Pan, Z.-X.; He, X.; Chen, Y.-Y.; Tang, W.-J.; Shi, J.-B.; Tang, Y.-L.; Song, B.-A.; Li, J.; Liu, X.-H. New 2H-chromene-3-carboxamide derivatives: Design, synthesis and use as inhibitors of hMAO. Eur. J. Med. Chem. 2014, 80, 278–284. [Google Scholar] [CrossRef]
  17. Kamauchi, H.; Noji, M.; Kinoshita, K.; Takanami, T.; Koyama, K. Coumarins with an unprecedented tetracyclic skeleton and coumarin dimers from chemically engineered extracts of a marine-derived fungus. Tetrahedron 2018, 74, 2846–2856. [Google Scholar] [CrossRef]
  18. Endo, S.; Matsunaga, T.; Kuwata, K.; Zhao, H.-T.; El-Kabbani, O.; Kitade, Y.; Hara, A. Chromene-3-carboxamide derivatives discov ered from virtual screening as potent inhibitors of the tumor maker, AKR1B10. Bioorg. Med. Chem. 2010, 18, 2485–2490. [Google Scholar] [CrossRef]
  19. Qian, B.; Váradi, L.; Trinchi, A.; Reichman, S.M.; Bao, L.; Lan, M.; Wei, G.; Cole, I.S. The Design and Synthesis of Fluorescent Coumarin Derivatives and Their Study for Cu2+ Sensing with an Application for Aqueous Soil Extracts. Molecules 2019, 24, 3569. [Google Scholar] [CrossRef] [Green Version]
  20. Yao, J.; Dou, W.; Qin, W.; Liu, W. A new coumarin-based chemosensor for Fe3+ in water. Inorg. Chem. Commun. 2009, 12, 116–118. [Google Scholar] [CrossRef]
  21. Yoshihara, T.; Yamaguchi, Y.; Hosaka, M.; Takeuchi, T.; Tobita, S. Ratiometric Molecular Sensor for Monitoring Oxygen Levels in Living Cells. Angew. Chem. Int. Ed. 2012, 51, 4148–4151. [Google Scholar] [CrossRef]
  22. Chimenti, F.; Secci, D.; Bolasco, A.; Chimenti, P.; Bizzarri, B.; Granese, A.; Carradori, S.; Yáñez, M.; Orallo, F.; Ortuso, F.; et al. Synthesis, Molecular Modeling, and Selective Inhibitory Activity against Human Monoamine Oxidases of 3-Carboxamido-7-Substituted Coumarins. J. Med. Chem. 2009, 52, 1935–1942. [Google Scholar] [CrossRef] [PubMed]
  23. Proenca, F.; Costa, M. A simple and eco-friendly approach for the synthesis of 2-imino and 2-oxo-2H-chromene-3-carboxamides. Green Chem. 2008, 10, 995–998. [Google Scholar] [CrossRef]
  24. Sepay, N.; Guha, C.; Kool, A.; Mallik, A.K. An efficient three-component synthesis of coumarin-3-carbamides by use of Ni–NiO nanoparticles as magnetically separable catalyst. RSC Adv. 2015, 5, 70718–70725. [Google Scholar] [CrossRef]
  25. Payra, S.; Saha, A.; Banerjee, S. Magnetically Recoverable Fe3O4 Nanoparticles for the One-Pot Synthesis of coumarin-3-carboxamide derivatives in aqueous ethanol. Chem. Sel. 2018, 3, 7535–7540. [Google Scholar]
  26. Gupta, M.; Kumar, P.; Bahadur, V.; Kumar, K.; Parmar, V.S.; Singh, B.K. Metal-Free, Regioselective, Dehydrogenative Cross-Coupling between Formamides/Aldehydes and Coumarins by C–H Functionalization. Eur. J. Org. Chem. 2018, 2018, 896–900. [Google Scholar] [CrossRef]
  27. Wang, L.; Ding, S.; Shen, H.; Wang, Y.; Hao, S.; Yin, G.; Qiu, J.; Lin, b.; Wu, Z.; Zhao, M. Generation of Coumarin-3-Carboxamides From Coumarin-3-Carboxylic Acids and Tetraalkylthiuram Disulfides Catalyzed by Copper Salts. Asian J. Org. Chem. 2021, 10, 2544–2548. [Google Scholar] [CrossRef]
  28. Hideo, T.; Shinpei, I. Synthetic use of molecular iodine for organic synthesis. Synlett 2006, 14, 2159–2175. [Google Scholar]
  29. Jereb, M.; Vražič, D.; Zupan, M. Iodine-catalyzed transformation of molecules containing oxygen functional groups. Tetrahedron 2011, 67, 1355–1387. [Google Scholar] [CrossRef]
  30. Parvatkar, P.T.; Parameswaran, P.S.; Tilve, S.G. Recent Developments in the Synthesis of Five- and Six-Membered Heterocycles Using Molecular Iodine. Chem. Eur. J. 2012, 18, 5460–5489. [Google Scholar] [CrossRef]
  31. Ren, Y.-M.; Cai, C.; Yang, R.-C. Molecular iodine-catalyzed multicomponent reactions: An efficient catalyst for organic synthesis. RSC Adv. 2013, 3, 7182–7204. [Google Scholar] [CrossRef]
  32. Alizadeh, A.; Ghanbaripour, R.; Zhu, L.G. An efficient approach to the synthesis of coumarin-bearing 2, 3-dihydro-4 (1H)-quinazolinone derivatives using a piperidine and molecular iodine dual-catalyst system. Synlett 2014, 25, 1596–1600. [Google Scholar] [CrossRef]
  33. Sashidhara, K.V.; Palnati, G.R.; Singh, L.R.; Upadhyay, A.; Avula, S.R.; Kumar, A.; Kant, R. Molecular iodine catalysed one-pot synthesis of chromeno [4, 3-b] quinolin-6-ones under microwave irradiation. Green Chem. 2015, 17, 3766–3770. [Google Scholar] [CrossRef]
  34. Zhang, M.; Yang, W.; Qian, M.; Zhao, T.; Yang, L.; Zhu, C. Iodine-promoted three-component reaction for the synthesis of spirooxindoles. Tetrahedron 2018, 74, 955–961. [Google Scholar] [CrossRef]
  35. Capello, C.; Fischer, U.; Hungerbühler, K. What is a green solvent? A comprehensive framework for the environmental assessment of solvent. Green Chem. 2007, 9, 927–934. [Google Scholar] [CrossRef]
  36. Fonseca, A.; Gaspar, A.; Matos, M.J.; Gomes, L.R.; Low, J.N.; Uriarte, E.; Borges, F. Structural elucidation of a series of 6-methyl-3-carboxamidocoumarins. Org. Magn. Reson. 2017, 55, 373–378. [Google Scholar] [CrossRef]
  37. Mohammed, H.B. Conversion of 3-carbethoxy-4-methyl coumarin derivatives into several new annelated coumarin derivatives. Chin. J. Chem. 2003, 21, 1219–1223. [Google Scholar] [CrossRef]
  38. Hu, S.; Wang, J.; Huang, G.; Zhu, K.; Chen, F. Organocatalytic Asymmetric Domino Oxa-Michael–Mannich-[1, 3]-Amino Rearrangement Reaction of N-Tosylsalicylimines to α, β-Unsaturated Aldehydes by Diarylprolinol Silyl Ethers. J. Org. Chem. 2020, 85, 4011–4018. [Google Scholar] [CrossRef]
  39. Breugst, M.; von der Heiden, D. Mechanisms in iodine catalysis. Chem. Eur. J. 2018, 24, 9187–9199. [Google Scholar] [CrossRef]
  40. Chen, G.F.; Jia, H.M.; Zhang, L.Y.; Hu, J.; Chen, B.H.; Song, Y.L.; Li, J.T.; Bai, G.Y. A highly selective fluorescent sensor for Fe3+ ion based on coumarin derivatives. Res. Chem. Intermed. 2013, 39, 4081–4090. [Google Scholar] [CrossRef]
  41. Areias, F.; Costa, M.; Castro, M.; Brea, J.; Gregori-Puigjané, E.; Proença, F.; Mestres, J.; Loza, M.I. New chromene scaffolds for adenosine A2A receptors: Synthesis, pharmacology and structure–activity relationships. Eur. J. Med. Chem. 2012, 54, 303–310. [Google Scholar] [CrossRef]
  42. Santos-Contreras, R.J.; Martínez-Martínez, F.J.; Mancilla-Margalli, N.A.; Peraza-Campos, A.L.; Morín-Sánchez, L.M.; García-Báez, E.V.; Padilla-Martínez, I.I. Competition between OH⋯O and multiple halogen–dipole interactions on the formation of intramolecular three-centred hydrogen bond in 3-acyl coumarins. CrystEngComm 2009, 11, 1451–1461. [Google Scholar] [CrossRef]
  43. Shi, J.; Lu, W.; Chen, J.; Sun, L.; Yang, S.; Zhou, M.; Xu, L.; Ma, Y.; Yu, L. Synthesis, antiproliferative activities, and DNA binding of coumarin-3-formamido derivatives. Arch. Pharm. 2021, 354, 2000236. [Google Scholar] [CrossRef] [PubMed]
  44. Ghanei-Nasab, S.; Khoobi, M.; Hadizadeh, F.; Marjani, A.; Moradi, A.; Nadri, H.; Emami, S.; Foroumadi, A.; Shafiee, A. Synthesis and anticholinesterase activity of coumarin-3-carboxamides bearing tryptamine moiety. Eur. J. Med. Chem. 2016, 121, 40–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yu, X.; Teng, P.; Zhang, Y.-L.; Xu, Z.-J.; Zhang, M.-Z.; Zhang, W.-H. Design, synthesis and antifungal activity evaluation of coumarin-3-carboxamide derivatives. Fitoterapia 2018, 127, 387–395. [Google Scholar] [CrossRef]
  46. Jang, Y.J.; Syu, S.E.; Chen, Y.J.; Yang, M.C.; Lin, W. Syntheses of furo[3, 4-c] coumarins and related furyl coumarin derivatives via intramolecular Wittig reactions. Org. Biomol. Chem. 2012, 10, 843–847. [Google Scholar] [CrossRef]
Figure 1. General methods for the synthesis of coumarin-3-carboxamides: (a) Traditional approach multistep reactions [22], (b) MCRs using nanoparticles [24,25], (c) C-3 functionalization of coumarins [26], (d) amidation reaction [27], (e) MCRs using piperidine-iodine.
Figure 1. General methods for the synthesis of coumarin-3-carboxamides: (a) Traditional approach multistep reactions [22], (b) MCRs using nanoparticles [24,25], (c) C-3 functionalization of coumarins [26], (d) amidation reaction [27], (e) MCRs using piperidine-iodine.
Molecules 27 04659 g001
Scheme 1. Multicomponent reaction of 2-hydroxybenzaldehyde, ethanolamine, and diethyl malonate.
Scheme 1. Multicomponent reaction of 2-hydroxybenzaldehyde, ethanolamine, and diethyl malonate.
Molecules 27 04659 sch001
Scheme 2. Substrate scope for the multicomponent reaction of salicylaldehyde, primary amines, and DEM.
Scheme 2. Substrate scope for the multicomponent reaction of salicylaldehyde, primary amines, and DEM.
Molecules 27 04659 sch002
Scheme 3. Substrate scope for the multicomponent reaction of salicylaldehyde, secondary amine, and DEM.
Scheme 3. Substrate scope for the multicomponent reaction of salicylaldehyde, secondary amine, and DEM.
Molecules 27 04659 sch003
Scheme 4. Synthesis of 3g at a 5 mmol scale.
Scheme 4. Synthesis of 3g at a 5 mmol scale.
Molecules 27 04659 sch004
Scheme 5. Control experiments.
Scheme 5. Control experiments.
Molecules 27 04659 sch005
Scheme 6. Plausible reaction pathway.
Scheme 6. Plausible reaction pathway.
Molecules 27 04659 sch006
Table 1. Optimization of the reaction conditions a.
Table 1. Optimization of the reaction conditions a.
Molecules 27 04659 i001
EntrySolventBase (mol%)Iodine (mol%)Ratio b
(3a:4a:5)
Yield 3a c
(%)
1EtOHpiperidine 10%10100:0:079
2MeOHpiperidine 10%10100:0:050
3i-PrOHpiperidine 10%10100:0:038
4t-BuOHpiperidine 10%10100:0:037
5H2Opiperidine 10%1035:55:1014
6CH2Cl2piperidine 10%1028:2:708
7MeCNpiperidine 10%1075:0:2532
8EtOAcpiperidine 10%1051:14:3526
9THFpiperidine 10%1066:24:1017
10DMFpiperidine 10%1021:53:2610
11Solvent freepiperidine 10%10100:0:036
12EtOHEt3N 10%10100:0:040
13EtOHDBU 10%10100:0:049
14EtOHL-Proline 10%10100:0:035
15EtOHpiperidine 10%20100:0:075
16EtOHpiperidine 10%15100:0:079
17EtOHpiperidine 10%5100:0:085
18EtOHpiperidine 10%1100:0:060
19EtOHpiperidine 5%5100:0:072
20EtOH-5100:0:063
21EtOHpiperidine 10%-100:0:066
a Reactions were performed at scale of 1 mmol. Relationship salicylaldehyde:ethanolamine:DEM (1.0:1.2:1.2). b Ratios were calculated from the NMR spectrum of the crude reaction. c Isolated yields.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Velasco, M.; Romero-Ceronio, N.; Torralba, R.; Hernández Abreu, O.; Vilchis-Reyes, M.A.; Alarcón-Matus, E.; Ramos-Rivera, E.M.; Aparicio, D.M.; Jiménez, J.; Aguilar García, E.; et al. Piperidine-Iodine as Efficient Dual Catalyst for the One-Pot, Three-Component Synthesis of Coumarin-3-Carboxamides. Molecules 2022, 27, 4659. https://doi.org/10.3390/molecules27144659

AMA Style

Velasco M, Romero-Ceronio N, Torralba R, Hernández Abreu O, Vilchis-Reyes MA, Alarcón-Matus E, Ramos-Rivera EM, Aparicio DM, Jiménez J, Aguilar García E, et al. Piperidine-Iodine as Efficient Dual Catalyst for the One-Pot, Three-Component Synthesis of Coumarin-3-Carboxamides. Molecules. 2022; 27(14):4659. https://doi.org/10.3390/molecules27144659

Chicago/Turabian Style

Velasco, Manuel, Nancy Romero-Ceronio, Rosalía Torralba, Oswaldo Hernández Abreu, Miguel A. Vilchis-Reyes, Erika Alarcón-Matus, Erika M. Ramos-Rivera, David M. Aparicio, Jacqueline Jiménez, Eric Aguilar García, and et al. 2022. "Piperidine-Iodine as Efficient Dual Catalyst for the One-Pot, Three-Component Synthesis of Coumarin-3-Carboxamides" Molecules 27, no. 14: 4659. https://doi.org/10.3390/molecules27144659

Article Metrics

Back to TopTop