Next Article in Journal
Insights into the Nature of the Active Sites of Pt-WOx/Al2O3 Catalysts for Glycerol Hydrogenolysis into 1,3-Propanediol
Next Article in Special Issue
Synthesis of Propylene Carbonate by Urea Alcoholysis—Recent Advances
Previous Article in Journal
The Effect of Mass Transfer Rate-Time in Bubbles on Removal of Azoxystrobin in Water by Micro-Sized Jet Array Discharge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Synthesis of 2-(1,2,3-Triazoyl) Quinazolinones through Copper-Catalyzed Multicomponent Reaction

by
Manoela Sacramento
1,
Luís Pedro A. Piúma
1,
José Edmilson R. Nascimento
1,
Roberta Cargnelutti
2,
Raquel G. Jacob
1,
Eder João Lenardão
1 and
Diego Alves
1,*
1
Laboratório de Síntese Orgânica Limpa, LASOL—CCQFA, Universidade Federal de Pelotas, UFPel, P.O. Box 354, Pelotas 96010-900, RS, Brazil
2
Departamento de Química, CCNE, Universidade Federal de Santa Maria, UFSM, Santa Maria 97105-900, RS, Brazil
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(10), 1170; https://doi.org/10.3390/catal11101170
Submission received: 7 September 2021 / Revised: 20 September 2021 / Accepted: 24 September 2021 / Published: 27 September 2021
(This article belongs to the Special Issue Catalysis in Green Chemistry and Organic Synthesis)

Abstract

:
We describe here our results from the copper-catalyzed three component reaction of 2-azidobenzaldehyde, anthranilamide and terminal alkynes, using Et3N as base, and DMSO as solvent. Depending on the temperature and amount of Et3N used in the reactions, 1,2,3-triazolyl-quinazolinones or 1,2,3-triazolyl-dihydroquinazolinone could be obtained. When the reactions were performed at 100 °C using 2 equivalents of Et3N, 1,2,3-triazolyl-dihydroquinazolinone was formed in 82% yield, whereas reactions carried out at 120 °C using 1 equivalent of Et3N provided 1,2,3-triazolyl-quinazolinones in moderate-to-good yields.

1. Introduction

Nitrogen heterocyclic compounds are among the most representative chemical architectures, and they are found in several natural products and commercial drugs. Quinazolinones are important members of the N-heterocycles, and they are building blocks in numerous natural products and widely used in the preparation of synthetic bioactive compounds, especially in the pharmaceutical industry [1]. Quinazolinones derivatives exhibit a wide range of pharmacological activities, including antimicrobial and antitubercular [2], anti-inflammatory [3], antiviral [4], anticancer [5], anti-Alzheimer [6], antimalarial [7], among others [8,9] (Figure 1).
Due to the pharmacological importance of the quinazolinone moiety, in recent years the development of new synthetic methods to access this class of compounds has attracted attention in organic synthesis. Generally, the quinazolinones are synthesized by the condensation of o-aminobenzamides with aldehydes [10], phenylacetic acid [11] and alcohols [12,13]. Other important protocols include carbon-supported acid-catalyzed cascade coupling of isatoic anhydrides with amides and aldehydes [14], oxidative annulation of alcohols with o-aminoarylnitriles [15], and Cu-catalyzed oxidative annulation of anilines, alkylamines and aldehydes [16].
Recently, the use of copper catalysis in the synthesis of quinazolinones has received much attention. An example was reported by Akhlaghinia et al. [17], who described an efficient protocol for the synthesis of 2,3-dihydroquinazolin-4(1H)-ones using Cu(I) functionalized with aminated 3-glycidyloxypropyltrimethoxysilane with thiosemicarbazide [SBA-16/GPTMS-TSC-Cu(I)] as a catalyst in the condensation of 2-aminobenzamide and an aldehyde under solvent-free conditions (Figure 2A). In 2020, Suresh et al. [18] reported the copper-catalyzed oxidative synthesis of quinazolinones in the presence of the reusable catalyst based in Cu3(BTC)2 (BTC = 1,3,5-benzene tricarboxylate) under mild condition, as shown in Figure 2B. There remains, however, a need for in-depth research on the synthesis of highly functionalized quinazolinones from various substrate combinations.
In this sense, substituted 1,2,3-triazoles are an important and useful class of heterocycles and have received considerable attention because of their application in organic synthesis, medicinal chemistry, and materials science [19,20,21,22,23,24]. For instance, the hybrid molecule containing quinoline and 1,2,3-triazole nucleus, 1-(7-chloroquinolin-4-yl)-N-(4-methoxybenzyl)-5-methyl-1H-1,2,3-triazole-4 carboxamide (QTC-4-MeOBnE A), exerts therapeutic effect through multiple pathways involved in AD (Figure 1) [25,26]. Other examples are Rufinamide® [27] and Tazobactam® [28], that having anticonvulsant and antibiotic activities, respectively (Figure 1). Some 1,2,3-triazole-based compounds have shown antibacterial [29,30,31], anticancer [32], antifungal [33,34], antiviral [35], among others [36,37]. Efficient methods to access highly functionalized 1,2,3-triazoles include the copper-catalyzed azide-alkyne cycloaddition (CuAAC) reaction [38,39] and the organocatalyzed [3+2] cycloaddition reactions via the generation of enamines, also known as Ramachary-Bressy-Wang organocatalytic azide-ketone [3+2]-cycloaddition (OrgAKC) [40,41,42]. Both CuAAC and OrgAKC are a powerful synthetic tool, due to simple reaction conditions, broader substrate scope and high atom economy.
Hybrid molecules containing the combination of quinazolinones and 1,2,3-triazoles, leads to compounds with promising pharmacological activities, such as antihistaminic [43], anticancer [44] antitubercular [2], antimicrobial [2], antihypertensive [45], and for the treatment of Alzheimer’s disease [46]. Continuing our studies on the synthesis of new functionalized 1,2,3-triazoles, and due to the pharmacological importance of the 1,2,3-triazoyl-quinazolinones scaffold, we describe herein the synthesis of a range of new 2-(1,2,3-triazoyl)quinazolin-4(3H)-ones by a copper-catalyzed multicomponent reaction of 2-azidobenzaldehyde 1, anthranilamide 2 and terminal alkynes 3 (Figure 2C).

2. Results

The original objective of this work was the synthesis of ten-membered heterocycles fused to a triazole ring (4), as a continuation of our previous works on the synthesis of medium-sized rings [47]. Thus, aiming to prepare the bicyclic triazole 4, a mixture of 2-azidobenzaldehyde 1 (0.5 mmol), anthranilamide 2 (0.5 mmol) and phenylacetylene 3a (0.5 mmol) in the presence of CuI (10 mol%) as a catalyst and Et3N (2 equivalents) as a base, in DMSO (1.5 mL) and under N2 atmosphere was stirred at 100 °C (Figure 3). However, after 16 h of reaction, the only isolated product was the triazoyl-2,3-dihydroquinazolinone 5a, which was isolated in 82% yield. The structure of 5a was confirmed after fully characterization by high- and low-resolution mass spectrometry, 1H and 13C NMR analysis (data are described in Supplementary Materials).
After confirming the identity of compound 5a, and checking in the literature the importance and versatility of the quinazolinone nucleus, our efforts were aimed at optimizing the reaction described in Figure 3. For this purpose, freshly prepared 2-azidobenzaldehyde 1 (0.5 mmol), anthranilamide 2 (0.5 mmol) and phenylacetylene 3a (0.5 mmol) were used as standard substrates, and different copper(I) and copper(II) salts, as well as a diversity of solvents and temperatures were evaluated (Table 1).
When the same conditions of Figure 3 (Table 1, entry 1) were performed using other copper(I) salts, such as CuBr and CuCl, it was observed a significant decrease in the yield of product 5a, from 82% to 21% and 27%, respectively (Table 1, entry 1 vs. entries 2–3). When copper(II) salts like CuBr2, CuCl2, and Cu(OAc)2·H2O were tested, only the latter afforded the expected product 5a, in 38% yield (Table 1, entries 4–6 ). By increasing the amount of the best catalyst (CuI) to 20 mol%, the yield of 5a was the same, whereas when we decrease the catalyst charge to 5 mol%, a decrease in the yield to 31% was observed (Table 1, entry 7–8). By fixing 10 mol% of CuI as catalyst, the following studies were intended to verify which would be the best solvent for the reaction. When 1,4-dioxane was used, the product 5a was obtained in 36% yield (Table 1, entry 9), while no product was observed when N,N-dimethylformamide (DMF) was used as solvent (Table 1, entry 10). Polyethylene glycol (PEG-400) and toluene were also tested, affording the product 5a in 58% and 61% yield, respectively (Table 1, entries 11,12). These outcomes indicate that DMSO is the best solvent for the synthesis of 5a. When a large excess of phenylacetylene 3a was used (2.0 equiv), compound 5a was obtained in 65% yield, a lower yield than that observed using an equivalent amount (Table 1, entry 13 vs. entry 1). A decrease in the reaction performance was observed at 80 °C instead of 100 °C, with product 5a being obtained in 60% yield (Table 1, entry 14). Notably, when higher temperatures were evaluated (120 and 140 °C), the formation of the aromatic quinazolin-4(3H)-one 6a was detected (Table 1, entries 15–16). The overall yield refers to the mixture of compounds 5a and 6a as this cannot be separated by column chromatography. The mixture of products could be observed in the 1H NMR spectrum of the crude reaction.
Our next aim was then establishing an optimal condition to selectively lead to the formation of the aromatic quinazolinone 6a. Firstly, the amount of base was reduced in order to verify its influence in the oxidation stage; using equivalent amount of Et3N at 120 °C, the aromatic product 6a was exclusively obtained in 87% yield (Table 1, entry 17). However, both the selectivity and the yield were diminished when the temperature was reduced to 100 °C (66% yield; 5a:6a ratio: 0.2:1) (Table 1, entry 18 vs. entry 17). Finally, the role of the catalyst was verified, the formation of products 5a and 6a was not observed in the absence of CuI (Table 1, entry 19). Compound 6a, once isolated, was characterized by high- and low-resolution mass spectrometry, 1H and 13C NMR, and X-ray diffraction analysis (data are described in Supplementary Materials). Selected bond lengths (Å) and angles (°) for compound 6a are summarized in Table S2 (Supplementary Materials).
Thus, with the best condition to prepare the triazoylquinazolin-4(3H)-one 6a in hand (Table 1, entry 16), this methodology was expanded to the reaction of 2-azidobenzaldehyde 1 and anthranilamide 2 with a variety of alkyl and aryl terminal alkynes 3a-n (Figure 4).
When analyzing the results shown in Figure 4, it is observed that the developed protocol tolerates a range of substituents in the alkynes counterpart. For instance, the variation of the groups at the para-position of the pendent phenyl ring, did not influence in the reactivity, and both electron-donating and electron-withdrawing substituents afforded the expected product 6 in around the same average yield. For instance, the electron-rich compounds 6b (R = 4-Me), 6c (R = 4-Et), 6d (R = 4-tBu), and 6e (R = 4-OMe) were obtained respectively in 54%, 67%, 51% and 66% yield, while the electron-poor one 6g (R = 4-Cl) was isolated in 60% yield (Figure 4). The presence of an ortho-substituent, however, caused a decrease in the yield, probably due a slightly steric hindrance compared to the para-substituted analogs, and compound 6f (R = 2-OMe) was obtained in 42% yield. A good result was obtained starting from 2-ethynylnaphthalene (3h), with the desired product 6h being obtained in 63% yield, while the aliphatic 1-hexyne (3i) afforded 6i in 57% yield after reaction with 1 and 2. The effect of the presence of different functional groups in the alkyne evaluated. Propagyl alcohols 3j and 3k satisfactorily reacted to afford the respective products 6j and 6k in 41% and 52%, respectively. In the case of 1-ethynylcyclopentanol (3k), an inseparable mixture of aromatic (6j) and non-aromatic (5j) products was obtained. The method was suitable also to ethyl propiolate (3l); however the expected product 5l was obtained in only 20% yield, with the formation of several by-products. Interestingly, alkynyl selenide 3m and alkynyl sulfide 3n could be converted to the respective selenium- and thio- derivatives 5m and 5n in 53% and 55% yield, respectively.
Thus, based on our results, and on studies described in the literature [38,47,48,49,50,51], a plausible mechanism for the multicomponent synthesis of quinazolinones functionalized with 1,2,3-triazoles 6 was proposed (Figure 5). Firstly, the triazole nucleus is formed, through the reaction between the terminal alkyne 3, Et3N and CuI, generating the copper acetylide intermediate A. Then, intermediate A reacts with 2-azidobenzaldehyde (1), through the complexation of the nucleophilic nitrogen from azide with the metal, forming of intermediate B. The β-vinylidene attack of copper(I) acetylide to the terminal electrophilic nitrogen of the azide occurs, affording the unstable six-membered ring C, which undergoes a ring contraction to give the copper triazoyl D [38]. After a protolysis step, the intermediate E is formed and the copper catalyst is regenerated in the catalytic cycle. In the sequence, a condensation reaction occurs by the nucleophilic attack of the nitrogen atom of the amino group of the anthranilamide (2) to the carbonyl of intermediate E, eliminating a water molecule and affording the intermediate F. Then, an intramolecular cyclization reaction occurs after a nucleophilic attack from the amide nitrogen to the imine carbon, forming the triazoyl-2,3-dihydroquinazolinone 5, which is oxidized in the reaction media [52,53,54] to the more stable aromatic quinazolin-4(3H)-one 6 (Figure 5).

3. Materials and Methods

3.1. General Information

The reactions were monitored by TLC carried out on Merck silica gel (60 F254) by using UV light as visualizing agent and 5% vanillin in 10% H2SO4 and heat as developing agents. Baker silica gel (particle size 0.040–0.063 mm) was used for flash chromatography. Hydrogen nuclear magnetic resonance spectra (1H NMR) were obtained at 400 MHz on Bruker DPX 400 spectrometer. Spectra were recorded in DMSO-d6 solutions. Chemical shifts are reported in ppm, referenced to tetramethylsilane (TMS) as the external reference. Coupling constants (J) are reported in Hertz. Abbreviations to denote the multiplicity of a particular signal are s (singlet), d (doublet), dd (double doublet), t (triplet), q (quartet), quin (quintet), sex (sextet) and m (multiplet). Carbon-13 nuclear magnetic resonance spectra (13C NMR) were obtained at 100 MHz on Bruker DPX 400 spectrometer. Chemical shifts are reported in ppm, referenced to the solvent peak of DMSO-d6. Low-resolution mass spectra were obtained with a Shimadzu GC-MS-QP2010 mass spectrometer. High resolution mass spectra (HRMS) were recorded on a Bruker Micro TOF-QII spectrometer 10416. The collection of X-ray diffraction data of compound 6a was performed on a Bruker D8 Quest diffractometer equipped with a Photon 100 detector, Incoatec microfocus Montel optic X-ray tube with Cu-Kα radiation (1.54178 Ǻ). Structure solutions and refinements were done through direct methods, with the SHELX program package [55,56]. Hydrogen atoms were included in the refinement at calculated positions. The graphical representation of the crystal structure was performed using the DIAMOND program (version 4.6.0) [57]. The crystallographic information file (CIF) for the compound 6a was deposited at the Cambridge Crystallographic Data Centre (CCDC) under identification number 2106971. Crystal data and more details of the data collection and refinement of 6a is provided in Table S1 (Supplementary Materials).

3.2. General Procedure for the Synthesis of Triazoylquinazolin-4(3H)-ones 6a–n

In a test tube was added 2-azidobenzaldehyde (1, 0.5 mmol), anthranilamide (2, 0.5 mmol), CuI (10 mol%). Then, DMSO (1.5 mL) was added, followed by triethylamine (0.5 mmol) and the terminal alkyne 3a–n (0.5 mmol). The system was heated to 120 °C under a nitrogen atmosphere and magnetic stirring for 16 h. Afterwards, the organic phase was received in water (20 mL), extracted with dichloromethane (3 × 10 mL), dried with MgSO4, and concentrated on a rotary evaporator followed by a vacuum pump. Finally, a purification by column chromatography of silica gel as a stationary phase, and a mixture of hexanes/ethyl acetate as eluent (50/50) was performed, providing the desired products 6a–n (20–87%).

4. Conclusions

In this work, an efficient method was developed for the synthesis of a series of unprecedented quinazolinones functionalized with 1,2,3-triazoles in modrate to good yields. The reaction involves a 1,3-dipolar cycloaddition between terminal alkynes and 2-azidobenzaldehyde, to obtain the triazole nucleus, followed by a cyclocondensation with anthranilamide to form the quinazolinone nucleus, which is oxidized in situ to the aromatic product. Fourteen new compounds (6a–n) were obtained in yields from 20 to 87%, presenting in their structure different substituents with a diversity of functional groups, including aryl, naphthyl, alkyl, ester, alcohol, as well as sulfide and selenide.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4344/11/10/1170/s1, 1H and 13C NMR spectra of all compounds. Table S1. Crystallographic and structure refinement data for compound 6a, Figure S1. Ellipsoid representations (50% probability) of 6a, Figure S2. Molecular structure of 6a determined by single-crystal X-ray crystallography, Table S2. Selected bond lengths (Å) and angles (°) of both independent molecules present in the compound 6a.

Author Contributions

Conceptualization, J.E.R.N., R.G.J., E.J.L. and D.A.; methodology, J.E.R.N., L.P.A.P., M.S. and R.C.; investigation, J.E.R.N., L.P.A.P. and M.S.; resources, E.J.L. and D.A.; data curation, J.E.R.N., R.C., E.J.L. and D.A., writing—original draft preparation, J.E.R.N. and M.S.; writing—review and editing, R.C., E.J.L. and D.A.; visualization, R.C., R.G.J., E.J.L. and D.A.; supervision, R.G.J., E.J.L. and D.A.; funding acquisition, R.G.J., E.J.L. and D.A. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financed in party by the Coordenação de Aperfeiçoamento de Pessoal de Nivel Superior—Brasil (CAPES)—Finance Code 001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful for the financial support and scholarships from the Brazilian agencies CNPq and FAPERGS. CNPq is also acknowledged for the fellowship for R.G.J., E.J.L. and D.A. In addition, we are grateful to Thiago Barcellos from University of Caxias do Sul (UCS), Brazil, for providing the HRMS analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Auti, P.S.; George, G.; Paul, A.T. Recent advances in the pharmacological diversification of quinazoline/quinazolinone hybrids. RSC Adv. 2020, 10, 41353–41392. [Google Scholar] [CrossRef]
  2. Pandey, S.K.; Yadava, U.; Upadhyay, A.; Sharma, M.L. Synthesis, biological evaluation and molecular docking studies of novel quinazolinones as antitubercular and antimicrobial agents. Bioorg. Chem. 2021, 108, 104611. [Google Scholar] [CrossRef]
  3. Abdel-Aziz, A.A.-M.; Abou-Zeid, L.A.; ElTahir, K.E.H.; Ayyad, R.R.; El-Sayed, M.A.-A.; El-Azab, A.S. Synthesis, anti-inflammatory, analgesic, COX-1/2 inhibitory activities and molecular docking studies of substituted 2-mercapto-4(3H)-quinazolinones. Eur. J. Med. Chem. 2016, 121, 410–421. [Google Scholar] [CrossRef]
  4. Brown, E.C.; Kong, T.; McNulty, J.; D’Aiuto, L.; Williamson, K.; McClain, L.; Piazza, P.; Nimgaonkar, L.V. Discovery of potent antiviral (HSV-1) quinazolinones and initial structure-activity relationship studies. Bioorg. Med. Chem. Lett. 2017, 27, 4601–4605. [Google Scholar] [CrossRef]
  5. Gatadi, S.; Pulivendala, G.; Gour, J.; Malasala, S.; Bujji, S.; Parupalli, R.; Shaikh, M.; Godugu, C.; Nanduri, S. Synthesis and evaluation of new 4(3H)-quinazolinone derivatives as potential anticancer agentes. J. Mol. Struct. 2020, 1200, 127097. [Google Scholar] [CrossRef]
  6. Decker, M. Novel inhibitors of acetyl- and butyrylcholinesterase derived from the alkaloids dehydroevodiamine and rutaecarpine. Eur. J. Med. Chem. 2005, 40, 305–313. [Google Scholar] [CrossRef] [PubMed]
  7. Zhu, S.; Chandrashekar, G.; Meng, L.; Robinson, K.; Chatterji, D. Febrifugine analogue compounds: Synthesis and antimalarial evaluation. Bioorg. Med. Chem. 2012, 20, 927–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Alagarsamy, V.; Chitra, K.; Saravanan, G.; Solomon, V.R.; Sulthana, M.T.; Narendhar, B. An overview of quinazolines: Pharmacological significance and recent developments. Eur. J. Med. Chem. 2018, 151, 628–685. [Google Scholar] [CrossRef] [PubMed]
  9. He, D.; Wang, M.; Zhao, S.; Shu, Y.; Zeng, H.; Xiao, C.; Lu, C.; Liu, Y. Pharmaceutical prospects of naturally occurring quinazolinone and its derivatives. Fitoterapia 2017, 119, 136–149. [Google Scholar] [CrossRef]
  10. Gupta, R.; Arora, G.; Yadav, Y.; Dixit, R.; Srivastavab, A.; Sharma, R.K. A magnetically retrievable copper ionic liquid nanocatalyst for cyclooxidative synthesis of 2-phenylquinazolin-4(3H)-ones. Dalton Trans. 2021, 50, 890–898. [Google Scholar] [CrossRef]
  11. To, T.A.; Vo, Y.H.; Nguyen, H.T.T.; Ha, P.T.M.; Doan, S.H.; Doan, T.L.H.; Li, S.; Le, H.V.; Tu, T.N.; Phan, N.T.S. Iron-catalyzed one-pot sequential transformations: Synthesis of quinazolinones via oxidative Csp3-H bond activation using a new metal-organic framework as catalyst. J. Catal. 2019, 370, 11–20. [Google Scholar] [CrossRef]
  12. Kerdphon, S.; Sanghong, P.; Chatwichien, J.; Choommongkol, V.; Rithchumpon, P.; Singh, T.; Meepowpan, P. Commercial copper-catalyzed aerobic oxidative synthesis of quinazolinones from 2-aminobenzamide and methanol. Eur. J. Org. Chem. 2020, 2730–2734. [Google Scholar] [CrossRef]
  13. Wang, Y.; Meng, X.; Chen, G.; Zhao, P. Direct synthesis of quinazolinones by heterogeneous Cu(OH)X/OMS-2 catalyst under oxygen. Catal. Commun. 2018, 104, 106–111. [Google Scholar] [CrossRef]
  14. Zhang, X.; Luo, C.; Chen, X.; Ma, W.; Li, B.; Lin, Z.; Chen, X.; Li, Y.; Xie, F. Direct synthesis of quinazolinones via the carbon-supported acid-catalyzed cascade reaction of isatoic anhydrides with amides and aldehydes. Tetrahedron Lett. 2021, 66, 152835. [Google Scholar] [CrossRef]
  15. Wang, Q.; Lv, M.; Liu, J.; Li, Y.; Xu, Q.; Xu, Z.; Cao, H. Efficient synthesis of quinazolinones by transition-metal-free direct aerobic oxidative cascade annulation of alcohols with o-aminoarylnitriles. ChemSusChem 2019, 12, 3043–3048. [Google Scholar] [CrossRef] [PubMed]
  16. Liang, Y.; Tan, Z.; Jiang, H.; Zhu, Z.; Zhang, M. Copper-catalyzed oxidative multicomponent annulation reaction for direct synthesis of quinazolinones via an imine-protection strategy. Org. Lett. 2019, 21, 4725–4728. [Google Scholar] [CrossRef]
  17. Erfan, M.A.; Akhlaghinia, B.; Ghodsinia, S.S.E. An efficient green protocol for synthesis of 2,3-dihydroquinazolin-4(1H)-ones using SBA-16/GPTMS-TSC-CuI under solvent-free conditions. ChemistrySelect 2020, 5, 2306–2316. [Google Scholar] [CrossRef]
  18. Latha, G.; Devarajan, N.; Suresh, P. Framework copper catalyzed oxidative synthesis of quinazolinones: A benign approach using Cu3(BTC)2 MOF as an efficient and reusable catalyst. ChemistrySelect 2020, 5, 10041–10047. [Google Scholar] [CrossRef]
  19. Rani, A.; Singh, G.; Singh, A.; Maqbool, U.; Kaur, G.; Singh, J. CuAAC-ensembled 1,2,3-triazole-linked isosteres as pharmacophores in drug discovery: Review. RSC Adv. 2020, 10, 5610–5635. [Google Scholar] [CrossRef]
  20. Jain, A.; Piplani, P. Exploring the chemistry and therapeutic potential of triazoles: A comprehensive literature review. Mini Rev. Med. Chem. 2019, 19, 1298–1368. [Google Scholar] [CrossRef]
  21. Mantoani, S.P.; Andrade, P.; Chierrito, T.P.C.; Figueredo, A.S.; Carvalho, I. Potential triazole-based molecules for the treatment of neglected diseases. Curr. Med. Chem. 2019, 26, 4403–4434. [Google Scholar] [CrossRef]
  22. Asif, M. Pharmacological activities of triazole analogues as antibacterial, antifungal, antiviral agents. Pharm. Sci. Asia 2017, 44, 59–74. [Google Scholar] [CrossRef]
  23. Bonandi, E.; Christodoulou, M.S.; Fumagalli, G.; Perdicchia, D.; Rastelli, G.; Passarella, D. The 1,2,3-triazole ring as a bioisostere in medicinal chemistry. Drug Discov. Today 2017, 22, 1572–1581. [Google Scholar] [CrossRef]
  24. Massarotti, A.; Aprile, S.; Mercalli, V.; Del Grosso, E.; Grosa, G.; Sorba, G.; Tron, G.C. Are 1,4- and 1,5-disubstituted 1,2,3-triazoles good pharmacophoric groups? ChemMedChem 2014, 9, 2497–2508. [Google Scholar] [CrossRef]
  25. Fronza, M.G.; Baldinotti, R.; Fetter, J.; Sacramento, M.; Sousa, F.S.S.; Seixas, F.K.; Collares, T.; Alves, D.; Pratico, D.; Savegnago, L. QTC-4-MeOBnE Rescues Scopolamine-Induced Memory Deficits in Mice by Targeting Oxidative Stress, Neuronal Plasticity, and Apoptosis. ACS Chem. Neurosci. 2020, 11, 1259–1269. [Google Scholar] [CrossRef]
  26. Fronza, M.G.; Baldinotti, R.; Sacramento, M.; Gutierres, J.; Carvalho, F.B.; Fernandes, M.C.; Sousa, F.S.S.; Seixas, F.K.; Collares, T.; Alves, D.; et al. Effect of QTC-4-MeOBnE Treatment on Memory, Neurodegeneration, and Neurogenesis in a Streptozotocin-Induced Mouse Model of Alzheimer’s Disease. ACS Chem. Neurosci. 2021, 12, 109–122. [Google Scholar] [CrossRef]
  27. Cheng-Hakimian, A.; Anderson, G.D.; Miller, J.W. Rufinamide: Pharmacology, clinical trials, and role in clinical practice. Int. J. Clin. Pract. 2006, 60, 1497–1501. [Google Scholar] [CrossRef] [PubMed]
  28. Sanders, W.E., Jr.; Sanders, C.C. Piperacillin/Tazobactam: A critical review of the evolving clinical literature. Clin. Infect. Dis. 1996, 22, 107–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Aufort, M.; Herscovici, J.; Bouhours, P.; Moreau, N.; Girard, C. Synthesis and antibiotic activity of a small molecules library of 1,2,3-triazole derivatives. Bioorg. Med. Chem. Lett. 2008, 18, 1195–1198. [Google Scholar] [CrossRef] [PubMed]
  30. Leaver, D.J.; Dawson, R.M.; White, J.M.; Polyzos, A.; Hughes, A.B. Synthesis of 1,2,3-triazole linked galactopyranosides and evaluation of cholera toxin inhibition. Org. Biomol. Chem. 2011, 9, 8465–8474. [Google Scholar] [CrossRef]
  31. Zhang, B. Comprehensive review on the anti-bacterial activity of 1,2,3-triazole hybrids. Eur. J. Med. Chem. 2019, 168, 357–372. [Google Scholar] [CrossRef]
  32. Xu, Z.; Zhao, S.-J.; Liu, Y. 1,2,3-Triazole-containing hybrids as potential anticancer agents: Current developments, action mechanisms and structure-activity relationships. Eur. J. Med. Chem. 2019, 183, 111700. [Google Scholar] [CrossRef]
  33. Lima-Neto, R.G.; Cavalcante, N.N.M.; Srivastava, R.M.; Junior, F.J.B.M.; Wanderley, A.G.; Neves, R.P.; Anjos, A.P. Synthesis of 1,2,3-triazole derivatives and in vitro antifungal evaluation on Candida strains. Molecules 2012, 17, 5882–5892. [Google Scholar] [CrossRef] [Green Version]
  34. Shalini, K.; Kumar, N.; Drabu, S. Advances in synthetic approach to and antifungal activity of triazoles. Beilstein J. Org. Chem. 2011, 7, 668–677. [Google Scholar] [CrossRef]
  35. Silva, F.C.; Souza, M.C.B.V.; Frugulhetti, I.I.P.; Castro, H.C.; Souza, S.L.O.; Souza, T.M.L.; Rodrigues, D.Q.; Souza, A.M.T.; Abreu, P.A.; Passamani, F.; et al. Synthesis, HIV-RT inhibitory activity and SAR of 1-benzyl-1H-1,2,3-triazole derivatives of carbohydrates. Eur. J. Med. Chem. 2009, 44, 373–383. [Google Scholar] [CrossRef]
  36. Raj, R.; Singh, P.; Singh, P.; Gut, J.; Rosenthal, P.J.; Kumar, V. Azide-alkyne cycloaddition en route to 1H-1,2,3-triazole-tethered 7-chloroquinoline-isatin chimeras: Synthesis and antimalarial evaluation. Eur. J. Med. Chem. 2013, 62, 590–596. [Google Scholar] [CrossRef] [Green Version]
  37. Im, C.; Maiti, S.N.; Micetich, R.G.; Daneshtalab, M.; Atchison, K.; Phillips, O.A.; Kunugita, C.J. Synthesis and β-lactamase inhibitory activity of 6-[(1-heteroarylthioethyl-1, 2, 3-triazol-4-yl)-methylene]penam sulfones. J. Antibiot. 1994, 47, 1030–1040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Rostovtsev, V.V.; Green, L.G.; Fokin, V.V.; Sharpless, K.B. A stepwise Huisgen cycloaddition process: Copper(I)-catalyzed regioselective “ligation” of azides and terminal alkynes. Angew. Chem. Int. Ed. 2002, 41, 2596–2599. [Google Scholar] [CrossRef]
  39. Tornøe, C.W.; Christensen, C.; Meldal, M. Peptidotriazoles on solid phase: [1,2,3]-triazoles by regiospecific copper(I)-catalyzed 1,3-dipolar cycloadditions of terminal alkynes to azides. J. Org. Chem. 2002, 67, 3057–3064. [Google Scholar] [CrossRef] [PubMed]
  40. Ramachary, D.B.; Ramakumar, K.; Narayana, V.V. Amino Acid-Catalyzed Cascade [3+2]-Cycloaddition/Hydrolysis Reactions Based on the Push–Pull Dienamine Platform: Synthesis of Highly Functionalized NH-1,2,3-Triazoles. Chem. Eur. J. 2008, 14, 9143–9147. [Google Scholar] [CrossRef]
  41. Belkheira, M.; El Abed, D.; Pons, J.M.; Bressy, C. Organocatalytic Synthesis of 1,2,3-Triazoles from Unactivated Ketones and Arylazides. Chem. Eur. J. 2011, 17, 12917–12921. [Google Scholar] [CrossRef]
  42. Danence, L.J.T.; Gao, Y.; Li, M.; Huang, Y.; Wang, J. Organocatalytic Enamide–Azide Cycloaddition Reactions: Regiospecific Synthesis of 1,4,5-Trisubstituted-1,2,3-Triazoles. Chem. Eur. J. 2011, 17, 3584–3587. [Google Scholar] [CrossRef] [PubMed]
  43. Alagarsamy, V.; Solomon, V.R.; Murugan, M. Synthesis and pharmacological investigation of novel 4-benzyl-1-substituted-4H-[1,2,4]triazolo[4,3-a]quinazolin-5-ones as new class of H1-antihistaminic agents. Bioorg. Med. Chem. 2007, 15, 4009–4015. [Google Scholar] [CrossRef] [PubMed]
  44. Driowya, M.; Leclercq, J.; Verones, V.; Barczyk, A.; Lecoeur, M.; Renault, N.; Flouquet, N.; Ghinet, A.; Berthelot, P.; Lebegue, N. Synthesis of triazoloquinazolinone based compounds as tubulin polymerization inhibitors and vascular disrupting agents. Eur. J. Med. Chem. 2016, 115, 393–405. [Google Scholar] [CrossRef] [PubMed]
  45. Ram, V.J.; Srimal, R.C.; Kushwaha, D.S.; Mishra, L. Chemotherapeutic agents. XIX. Synthesis of [1,2,4]-Triazoloquinazolinones and related compounds as antihypertensive agents. J. Prakt. Chem. 1990, 332, 629–639. [Google Scholar] [CrossRef]
  46. Bertelli, L.; Biagi, G.; Giorgi, I.; Livi, O.; Manera, C.; Scartoni, V.; Lucacchini, A.; Giannaccini, G.; Barili, P.L. Substituted 1,2,3-triazolo[1,5-a]quinazolines: Synthesis and binding to benzodiazepine and adenosine receptors. Eur. J. Med. Chem. 2000, 35, 333–341. [Google Scholar] [CrossRef]
  47. Peringer, F.; Nascimento, J.E.R.; Abib, P.B.; Barcellos, T.; Eycken, E.V.V.; Perin, G.; Jacob, R.G.; Alves, D. Copper-catalyzed multicomponent reactions: Synthesis of fused 1,2,3-triazolo-1,3,6-triazonines. Eur. J. Org. Chem. 2017, 2017, 2579. [Google Scholar] [CrossRef]
  48. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide–alkynecycloaddition (CuAAC) and beyond: New reactivity of copper(i) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef]
  49. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  50. Li, L.; Zang, Z. Development and Applications of the Copper-Catalyzed Azide-Alkyne Cycloaddition (CuAAC) as a Bioorthogonal Reaction. Molecules 2016, 21, 1393. [Google Scholar] [CrossRef]
  51. Rammah, M.M.; Gati, W.; Mtiraoui, H.; Rammah, M.E.B.; Ciamala, K.; Knorr, M.; Rousselin, Y.; Kubicki, M.M. Synthesis of Isoxazole and 1,2,3-Triazole Isoindole Derivatives via Silver- and Copper-Catalyzed 1,3-Dipolar Cycloaddition Reaction. Molecules 2016, 21, 307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Guo, S.; Li, Y.; Tao, L.; Zhang, W.; Fan, X. Rapid assembly of quinazolinone scaffold via copper-catalyzed tandem reaction of 2-bromobenzamides with aldehydes and aqueous ammonia: Application to the synthesis of the alkaloid tryptanthrin. RSC Adv. 2014, 4, 59289–59296. [Google Scholar] [CrossRef]
  53. Hu, B.-Q.; Wang, L.-X.; Yang, L.; Xiang, J.-F.; Tang, Y.-L. Copper-Catalyzed Intramolecular C–C Bond Cleavage To Construct 2-Substituted Quinazolinones. Eur. J. Org. Chem. 2015, 4504–4509. [Google Scholar] [CrossRef]
  54. Upadhyaya, K.; Thakur, R.K.; Shukla, S.K.; Tripathi, R.P. One-Pot Copper(I)-Catalyzed Ligand/Base-Free Tandem Cyclooxidative Synthesis of Quinazolinones. J. Org. Chem. 2016, 81, 5046–5055. [Google Scholar] [CrossRef] [PubMed]
  55. Sheldrick, G.M. Crystal structure refinement with SHELXL Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  56. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  57. Brandenburg, K.; Putz, H. Diamond—Crystal and Molecular Structure Visualization; Crystal Impact: Bonn, Germany, 2014. [Google Scholar]
Figure 1. Quinazolinones and 1,2,3-triazoles with pharmacological activities.
Figure 1. Quinazolinones and 1,2,3-triazoles with pharmacological activities.
Catalysts 11 01170 g001
Figure 2. Recent examples of the copper-catalyzed synthesis of quinazolinones and this work.
Figure 2. Recent examples of the copper-catalyzed synthesis of quinazolinones and this work.
Catalysts 11 01170 g002
Figure 3. Synthesis of triazoylquinazolinone 5a.
Figure 3. Synthesis of triazoylquinazolinone 5a.
Catalysts 11 01170 g003
Figure 4. Reaction scope to obtain triazoylquinazolin-4(3H)-ones 6a-n. Reactions performed using 2-azidobenzaldehyde 1 (0.5 mmol), anthranylamide 2 (0.5 mmol), terminal alkynes 3a-n (0.5 mmol), CuI (10 mol%), and Et3N (0.5 mmol) in DMSO (1.5 mL) at 120 °C under N2 atmosphere for 16 h. Yields of isolated product after column chromatography. a Obtained as a mixture of 5j and 6j.
Figure 4. Reaction scope to obtain triazoylquinazolin-4(3H)-ones 6a-n. Reactions performed using 2-azidobenzaldehyde 1 (0.5 mmol), anthranylamide 2 (0.5 mmol), terminal alkynes 3a-n (0.5 mmol), CuI (10 mol%), and Et3N (0.5 mmol) in DMSO (1.5 mL) at 120 °C under N2 atmosphere for 16 h. Yields of isolated product after column chromatography. a Obtained as a mixture of 5j and 6j.
Catalysts 11 01170 g004
Figure 5. Proposed mechanism for the formation of triazoylquinazolin-4(3H)-one 6.
Figure 5. Proposed mechanism for the formation of triazoylquinazolin-4(3H)-one 6.
Catalysts 11 01170 g005
Table 1. Optimization of the reaction conditions a.
Table 1. Optimization of the reaction conditions a.
Catalysts 11 01170 i001
EntryEt3N
(mmol)
Catalyst
(10 mol%)
SolventTemp. (°C)Yield (%) bRatio c
5a:6a
11CuIDMSO100821:0
21CuBrDMSO100211:0
31CuClDMSO100271:0
41CuBr2DMSO100n.f.--
51CuCl2DMSO100n.f.--
61Cu(OAc)2·H2ODMSO100381:0
7 d1CuIDMSO100811:0
8 e1CuIDMSO100311:0
91CuI1,4-dioxane100361:0
101CuIDMF100NF1:0
111CuIPEG-400100581:0
121CuIToluene100611:0
13 f1CuIDMSO100651:0
141CuIDMSO80601:0
151CuIDMSO120881:0.5
161CuIDMSO140901:0.7
170.5CuIDMSO120870:1
180.5CuIDMSO100660.2:1
190.5--DMSO120NF--
a Reaction using 2-azidobenzaldehyde 1 (0.5 mmol), anthranilamide 2 (0.5 mmol) and phenylacetylene 3a (0.5 mmol) in the presence of copper salt as catalyst, Et3N as base, in 1.5 mL of solvent fo 16 h. b Yield of 5a + 6a. c Determined by 1H NMR. d Used 20 mol% of CuI. e Used 5 mol% of CuI. f Reaction using 2.0 equiv of phenylacetylene 3a. NF = not formed.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sacramento, M.; Piúma, L.P.A.; Nascimento, J.E.R.; Cargnelutti, R.; Jacob, R.G.; Lenardão, E.J.; Alves, D. Selective Synthesis of 2-(1,2,3-Triazoyl) Quinazolinones through Copper-Catalyzed Multicomponent Reaction. Catalysts 2021, 11, 1170. https://doi.org/10.3390/catal11101170

AMA Style

Sacramento M, Piúma LPA, Nascimento JER, Cargnelutti R, Jacob RG, Lenardão EJ, Alves D. Selective Synthesis of 2-(1,2,3-Triazoyl) Quinazolinones through Copper-Catalyzed Multicomponent Reaction. Catalysts. 2021; 11(10):1170. https://doi.org/10.3390/catal11101170

Chicago/Turabian Style

Sacramento, Manoela, Luís Pedro A. Piúma, José Edmilson R. Nascimento, Roberta Cargnelutti, Raquel G. Jacob, Eder João Lenardão, and Diego Alves. 2021. "Selective Synthesis of 2-(1,2,3-Triazoyl) Quinazolinones through Copper-Catalyzed Multicomponent Reaction" Catalysts 11, no. 10: 1170. https://doi.org/10.3390/catal11101170

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop