Next Article in Journal
Nanomaterials for Direct Air Capture of CO2: Current State of the Art, Challenges and Future Perspectives
Previous Article in Journal
Recent Advances on Biomass-Derived Carbon Materials-Based Electrochemical Sensors
Previous Article in Special Issue
Flavonoids and Furanocoumarins Involved in Drug Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pd/Ligand-Free Synthesis of 2-Alkynylated Pyrano[4,3-d]imidazol-4-ones via One-Pot Cu-Mediated Tandem Sonogashira Coupling/Regioselective 6-endo-dig Oxacyclization Reaction

1
Laboratoire de Physico-Chimie des Matériaux et des Electrolytes Pour l’Energie (PCM2E), EA 6299, Faculté des Sciences et Techniques, Avenue Monge, Faculté des Sciences, Université de Tours, Parc de Grandmont, 37200 Tours, France
2
Laboratoire de Chimie Analytique et Electrochimie (LCAE), Faculté des Sciences de Tunis, Université de Tunis El Manar, Campus Universitaire, El-Manar, Tunis 2092, Tunisia
3
Laboratoire de Chimie Moléculaire et Substances Naturelles, Faculté des Sciences, Université Moulay Ismail, B.P. 11201, Zitoune, Meknès 50050, Morocco
4
Laboratoire de Synthèse et Isolement de Molécules BioActives, SIMBA EA 7502, Faculté de Pharmacie, Université de Tours, 31 Avenue Monge, 37200 Tours, France
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(14), 3045; https://doi.org/10.3390/molecules30143045
Submission received: 14 June 2025 / Revised: 10 July 2025 / Accepted: 14 July 2025 / Published: 21 July 2025

Abstract

Herein, we report a one-pot palladium- and ligand-free tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization reaction of 2,4-diiodo-1-methyl-imidazole-5-carboxylic acid with terminal alkynes mediated by Copper(I). This impressive approach offers a straightforward, practical, and efficient tandem procedure for accessing 2-alkynylated pyrano[4,3-d]imidazol-4-one in moderate to good yields with an exclusive 6-endo-dig oxacyclization. Notably, this cost-effective methodology demonstrates broad substrate compatibility with various commercially available aliphatic and (hetero)aromatic terminal alkynes. Furthermore, DFT studies were performed to elucidate the origin of this regioselective 6-endo-dig oxacyclization reaction.

Graphical Abstract

1. Introduction

2H-Pyran-2-ones, otherwise known as 2-pyrones or 2-pyranones (α-pyrones), are six-membered heterocyclic esters embedding an oxygen atom and a carbonyl group. They are frequently encountered in a large number of natural products, pharmaceutical compounds, and other functional molecules, displaying a broad spectrum of important bioactive properties [1,2]. Particularly, fused 2H-pyran-2-ones with a five-membered heteroaryl ring, which are undoubtedly of great value because of their own structural diversity and biological properties, have garnered widespread attention from the chemical research community [3,4,5]. Pyrano[4,3-d]imidazol-4-one derivatives are typical examples of these types of compounds, which have also demonstrated notable bioactivities, including the inhibition of bacterial DNA gyrase and topoisomerase IV (Figure 1a) [6], antagonists of the angiotensin II receptor with different affinities for the AT1 and AT2 subtypes (Figure 1b–d) [7,8,9], as well as partial agonists of peroxisome proliferator-activated receptor-γ (PPARγ) (Figure 1c) [9].
Given the diverse biological properties of 2H-pyran-2-ones derivatives, the development of practical and highly efficient synthetic methods is still in high demand. Although significant efforts have been devoted to the synthesis of these heterocycles, only a few synthetic methods for the preparation of 2H-pyran-2-ones fused heterocyclic systems have been reported. Among them, the one-pot tandem Sonogashira coupling/oxacyclization reaction from heteroaromatic β-iodo/bromo-α,β-unsaturated carboxylic acid derivatives and terminal alkynes has emerged as a general approach to obtain these valuable compounds. However, this transformation generally requires the use of expensive palladium catalysts accompanied by copper as a cocatalyst, including ligand effects in some cases, which significantly limits the usefulness of these conditions (Figure 2a) [10,11,12]. In the last decade, our group has developed a new versatile, direct, and elegant method under palladium- and ligand-free conditions, using exclusively copper(I) as a cheaper catalyst. Moreover, this protocol provides novel, facile, rapid, and efficient access to structurally diverse bi- and tri-heterocyclic 2H-pyran-2-ones based on thiophene [13,14], indole [14], and imidazo[1,2-a]pyridine [15] rings in a highly atom-economical manner (Figure 2b).
It is widespread recognition that the Sonogashira cross-coupling reaction is routinely conducted with a palladium-based catalyst and a copper(I) salt as a cocatalyst [16]. Although numerous different copper-free, palladium-based catalytic systems have been recently established for this reaction [17,18], efficient palladium- and ligand-free copper-based catalytic systems have remained relatively underexplored (Figure 2b), even though they are obviously be much more interesting in terms of environmental and economic points of view than palladium-based ones [19].
In light of the diverse biological properties of 2H-pyran-2-ones derivatives and fascinated by the great features of Sonogashira cross-coupling catalyzed exclusively by copper, which are cheaper, more abundant, and more environmentally benign than those catalyzed by Pd complexes, herein we report an unprecedented one-pot copper-mediated tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization reaction of 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 with terminal alkynes (Figure 2b).
Contrary to previous proposals, our approach offers notable improvements in terms of environmental impact, atom economy, efficiency, and product selectivity when compared to traditional and alternative literature-reported methods. Additionally, this strategy avoided the use of precious metal and ligand, with a drastic reduction in catalyst toxicity.
Our methodology offers a potential simplified approach to the synthesis of alkynylated isocoumarins, which are valuable building blocks for the preparation of an important class of antibiotics such as Rubromycin-type compounds [20,21,22] while circumventing multiple synthetic steps and eliminating the need for palladium catalysts. In addition to the synthetic utility of our methodology, the imidazole fused 2H-pyran-2-one structural motif present in the synthesized compounds is of considerable interest in medicinal chemistry, due to the significant biological activities exhibited by 2H-pyran-2-ones fused with five-membered heteroaryl rings, including antileishmanial, antifungal, and anticancer properties [3,4,5]. Future work will focus on the biological evaluation of these compounds to fully explore their therapeutic potential, as well as on expanding the scope of this methodology for the synthesis of natural products, pharmaceutically relevant compounds, and a broader range of heterocyclic scaffolds.

2. Results and Discussion

At the outset of our studies, 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4, a five-membered N-heterocycle bearing a β-iodo-α,β-unsaturated carboxylic acid, was readily prepared in three steps on a multigram scale, as depicted in Scheme 1.
First, the intermediate 2 was synthesized in 93% yield via the bisiodination reaction of commercially available ethyl imidazole-4-carboxylate 1 using N-iodosuccinimide (NIS). The subsequent reaction of 2 with methyl iodide leads to regioselective methylation of the N1 position, yielding the desired product 3 in 60% yield. The synthesis was completed by the saponification reaction of ethyl ester 3 in the presence of sodium hydroxide, providing the corresponding carboxylic acid, the key intermediate 4, in 55% yield.
Since the main goal of this study relies on the one-pot Cu-mediated tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization reaction, 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 and phenylacetylene were selected as the model substrates to optimize the reaction conditions. In order to find the optimal conditions, the coupling partners were subjected to an extensive screening of several reaction parameters, such as catalyst and base, in DMF as a commonly used solvent for this reaction. Detailed information is summarized in Table 1.
We began our investigations with reaction conditions based on the work of Pal [10] and their co-workers using a Pd/Cu co-catalytic system (Table 1, entry 1). The reaction did not go to full conversion since only a mixture of desired product 5a and the starting material 4 was observed in this case in 47:53 ratio. Furthermore, increasing the amount of copper iodide to 1.0 equivalent improved the desired product ratio 5a (5a/4 70:30) but did not achieve the complete conversion of starting material 4 (Table 1, entry 2). Interestingly, the use of a mild inorganic base (K2CO3) instead of an organic base (Et3N) resulted in a full conversion of 4 after 3 h, providing the target product 5a in 44% isolated yield (Table 1, entry 3). Encouraged by this promising result, we then examined whether both palladium and copper were essential to the success and efficiency of this reaction. Therefore, in the absence of palladium, the reaction afforded the desired product 5a in just 27% yield after 24h, indicating that the reaction can proceed without palladium despite the low yield (Table 1, entry 4).
Gratifyingly, when the amount of CuI was increased to 2.0 equiv., the isolated yield of 5a was improved to 51% after 3 h (Table 1, entry 5). Thereafter, the isolated yield of 5a was slightly decreased to 47% when 2.5 equiv. of CuI was used as the catalyst loading (Table 1, entry 6). Unsurprisingly, when the alkyne loading was reduced to 2.05 and 1.05, no change was noted in the selectivity of the reaction, but the isolated yields decreased drastically to 24% and 12%, respectively (Table 1, entries 7–8). Subsequently, the influence of temperature on the reaction outcome was evaluated, but no further improvement in reaction yield was observed (Table 1, entries 9–10). Finally, switching from K2CO3 to various bases such as Na2CO3, K3PO4, Cs2CO3, or Et3N did not give better reaction results (Table 1, entries 11–14).
After successfully establishing the optimal reaction conditions (Table 1, entry 5), we evaluated the reaction scope and functional group tolerance against the catalytic system. Various commercially available aliphatic and (hetero)aromatic terminal alkynes reacted with 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4, giving the desired products 5ak in 41–78% yields (Scheme 2).
The structure of 5a was determined through comprehensive NMR and HRMS spectroscopic analyses and unambiguously confirmed by the X-ray crystallography of the single crystal (CCDC 2463986) (see Scheme 2) [23].
Reactions of 4 with phenyl-substituted acetylenes containing a methyl (Me) group at the para and meta positions proceeded smoothly under the well-established reaction conditions, providing the corresponding desired products 5b and 5c in 70% and 68% yields, respectively. It is noteworthy that the reaction of the phenyl substituent bearing an electron-donating group such as methoxy and tert-butyl at the para position worked efficiently in the current protocol, generating the expected products 5d and 5e in 78% and 44% yields, respectively. To our delight, the chloro-containing group at ortho and meta positions was a compatible functional group for this transformation, giving the desired products 5f and 5g in 58% and 65% yields, respectively. More importantly, heteroaromatic substrates such as 2-ethynylthiophene and 3-ethynylthiophene were also suitable substrates for this methodology and afforded the desired pyrano[4,3-d]imidazol-4-ones 5h and 5i in 55% and 41% yields, respectively. Interestingly, phenethylacetylene and cyclopropylacetylene, two aliphatic alkynes, were successfully reacted with 4, leading the corresponding targeted products 5j and 5k in 49% and 52% yields, respectively.
In the course of our investigation on the mechanistic evidence of this transformation, an additional experiment was performed. Under the same standard conditions, the reaction carried out with the intermediate 3 bearing an ester group instead of 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4, exclusively affording 2,6-dialkynylated imidazole 6 instead of 5a in 51% yield, as shown in Scheme 3.
This result indicates that, unlike the carboxylic acid function, the ester group of 3 was unable to participate in the regioselective 6-endo-dig oxacyclization step under the optimized conditions (Table 1, entry 5).
Based on similar studies reported in the literature [24,25,26] and the experiment performed before, a plausible mechanism of this transformation is proposed as shown in Scheme 4.
In the presence of the base K2CO3, the carboxylic acid 4 leads to the formation of potassium carboxylate A. Then, the nitrogen (NX) atom of imidazole, the nitrogen-based intra-ligand, coordinates the copper to form intermediate B. This species coordinates alkyne, and the deprotonated form of alkyne reacts with the Cu(I)–N(imidazole) ligand complex to provide the Cu(I)–acetylide species D. Next, the oxidative addition of heteroaryl iodide to intermediate D generates a four-coordinated Cu (III) complex E. Finally, the reductive elimination from E affords the desired dialkynylated intermediate F. Cu(I)I could coordinate with the electron-rich alkynyl bond of species F to give intermediate G. Afterwards, the o-alkynyl imidazoate salt G underwent intramolecular ring closure via oxyanion attacks at the C-β position of the acetylenic bond via 6-endo-dig cyclization to give intermediate H, which is further converted into 5 upon protodemetalation.

Theoretical Calculation (Computational Studies)

To validate our mechanistic hypothesis for intramolecular 6-endo-dig oxacyclization through β-addition, we conducted density functional theory (DFT) calculations to analyze the natural population of each carbon in the alkyne moiety, which appears to be critical in determining the resulting cyclization (Figure 3 and Supplementary Materials) [27,28,29,30,31,32,33].
The calculated NPA revealed that the partial positive charge is located on the β-carbon of alkyne leading to the 6-endo-dig oxacyclization, which is in excellent agreement with the experiment results, regardless of the nature of the alkyne substituent.

3. Materials and Methods

3.1. General Information

All reactions were performed under an inert atmosphere of argon in oven-dried glassware equipped with a magnetic stir bar. Solvents for reactions were obtained from Thermo Fisher Scientific (Illkirch-Graffenstaden, France) in extra dry quality and stored under argon over activated 3 Å sieves. All reagents were purchased from Fluorochem (Cork, Ireland) and used as received without additional purification. Reactions were monitored by thin-layer chromatography (TLC) analysis using silica gel 60 F254 plates. Column chromatography was performed using silica gel 60 (230–400.13 mesh, 0.040–0.063 mm). Eluents were distilled by the standard methods before each use. All new compounds were characterized by NMR spectroscopy (1H, and 13C), and high-resolution mass spectroscopy (HRMS). NMR spectra were recorded at 300 MHz for 1H, and 75 MHz for 13C with a Bruker® (Wissembourg, France) 300 MHz NMR spectrometer. Proton and carbon magnetic resonance spectra (1H NMR and 13C NMR) were recorded using tetramethylsilane (TMS) as an external standard and CDCl3 (7.28 ppm for 1H NMR and 77.04 ppm for 13C NMR). Data for NMR are reported as follows: chemical shift (δ ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, sep = septet, m = multiplet, and br = broad resonance), and coupling constants J are reported in Hertz (Hz). All NMR spectra were processed in MestReNova. HRMS experiments were performed on a hybrid tandem quadrupole/time-of-flight (Q-TOF) instrument, equipped with a pneumatically assisted electrospray (Z-spray) ion source (Micromass, Manchester, UK) operated in the positive mode. The melting points (Mp [°C]) of samples were measured using open capillary tubes and recorded on a StuartTM (Paris, France) melting point apparatus SMP3.
  • Ethyl-2,4-diiodo-1H-imidazol-5-carboxylate (2)
Compound 2 was prepared following a slightly modified procedure based on literature methods. Ethyl-1H-imidazole-5-carboxylate 1 (1.0 g, 7.1 mmol) was dissolved in dry acetonitrile (50 mL), and N-iodosuccinimide (NIS) (4.01 g, 17.8 mmol, 2.5 equiv.) was added to the solution. The reaction mixture was stirred at 80 °C for 3 h under an argon atmosphere. The progress of the reaction was monitored by TLC. Upon completion, the solvent was removed under reduced pressure. 1H NMR (300 MHz, CDCl3): δ = 8.28 (s, 1H), 4.40 (q, J = 7.1 Hz, 2H), 1.41 (t, J = 7.1 Hz, 3H). 13C NMR (75 MHz, CDCl3): δ = 177.6, 158.9, 91.2, 87.6, 59.7, 14.2.
  • Ethyl-2,4-diiodo-1-methyl-1H-imidazol-5-carboxylate (3)
Compound 3 was synthesized from ethyl-2,4-diiodo-1H-imidazole-5-carboxylate 2 (1.0 g, 2.55 mmol) and potassium tert-butoxide (0.31 g, 2.80 mmol), which were dissolved in dry DMF in a round-bottom flask under anhydrous conditions. The reaction mixture was cooled to 0 °C, and iodomethane was added dropwise. Stirring was continued for 24 h at room temperature under an argon atmosphere. After completion of the reaction (monitored by TLC), the solvent was removed under reduced pressure. The crude product was purified directly by column chromatography on silica gel using a mixture of petroleum ether and ethyl acetate as eluent, affording the desired product 3. 1H NMR (300 MHz, CDCl3): δ = 4.37 (q, J = 7.1 Hz, 2H), 3.93 (s, 3H), 1.43 (t, J = 7.1 Hz, 3H). 13C NMR (75 MHz, CDCl3): δ = 177.6, 158.9, 89.3, 61.8, 29.6, 14.2. HRMS (ESI): m/z calcd for C7H9I2N2O2 [M + H]+ 406.8747; found: 406.8741.
  • 2,4-Diiodo-1-methyl-1H-imidazole-5-carboxylic acid (4)
Compound 4 was obtained from ethyl-2,4-diiodo-1-methyl-1H-imidazole-5-carboxylate 3 (1.0 g, 2.45 mmol) and sodium hydroxide (0.14 g, 3.68 mmol), which were added to a round-bottom flask and dissolved in ethanol. The reaction mixture was stirred at 80 °C for 1 h. After completion of the reaction (monitored by TLC), the solvent was removed under reduced pressure. The residue was diluted with water, cooled to 0 °C, and acidified by the slow addition of 2 M HCl until the pH reached 1. The resulting precipitate was collected by vacuum filtration to afford compound 4. 1H NMR (300 MHz, (CD3)2SO)): δ = 3.82 (s, 1H). 13C NMR (75 MHz, (CD3)2SO)): δ = 150.9, 122.7, 94.8, 78.6, 40.6, 40.3, 40.1, 39.8, 39.5, 39.2, 38.9, 30.6. HRMS (ESI): m/z calcd for C5H5I2N2O2 [M + H]+ 378.8434; found: 378.8429.

3.2. General Procedure for the Synthesis of 3-Methyl-6-aryl-2-alcynylpyrano[3,4-d]imidazole-4(3H)-one (5ak)

An oven dried 25 mL Schlenk tube equipped with a magnetic stir bar was charged with 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 5 (100 mg, 1 equiv.), CuI (2 equiv.), and K2CO3 (2 equiv.) solubilized in 4 mL of dry DMF. The vial was sealed with a septum-lined pierceable cap, evacuated, and back-filled with argon ( × 3). The terminal alkyne (3 equiv.) was dissolved in anhydrous DMF (2 mL) and added to the reaction mixture via a syringe. The reaction vessel was placed in a pre-heated oil bath and stirred at 130 °C for 3 h under an argon atmosphere. After completion of the reaction (monitored by TLC), the mixture was cooled to room temperature and filtered through a thin pad of celite eluting with ethyl acetate (25 mL), and the combined filtrate was concentrated in vacuo. The residue was directly purified by column chromatography on silica gel using a mixture of petroleum ether/EtOAc as eluent to give the desired products 5ak.
  • 3-Methyl-6-phenyl-2-(phenylethynyl)pyrano[3,4-d]imidazol-4(3H)-one (5a)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5a as a white solid (43.8 mg, 51%). Mp 223–225 °C. 1H NMR (300 MHz, CDCl3): δ = 7.87–7.84 (m, 2H), 7.66–7.65 (m, 2H), 7.47–7.42 (m, 6H), 7.12 (s, 1H), 4.16 (s, 3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 156.6, 155.1, 150.0, 140.3, 132.2, 132.0, 130.3, 129.9, 128.9, 128.7, 125.4, 120.5, 117.3, 98.3, 97.2, 77.3, 33.2. HRMS (ESI): m/z calcd for C21H15N2O2 [M + H]+ 327.1128; found: 327.1120.
  • 3-Methyl-6-(4-methylphenyl)-2-[(4-methylphenyl)ethynyl]pyrano[3,4-d]imidazol-4(3H)-one (5b)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5b as a white solid (65.1 mg, 70%). Mp 179–181 °C. 1H NMR (300 MHz, CDCl3): δ = 7.78–7.75 (d, J = 8.2 Hz, 2H), 7.56–7.53 (d, J = 8.1 Hz, 2H), 7.29–7.23 (m, 4H), 7.08 (s, 1H), 4.17 (s,3H), 2.43 (s, 6H) ppm. 13C NMR (75 MHz, CDCl3): δ = 156.5, 155.0, 150.0, 140.2, 132.2, 132.0, 130.2, 129.9, 128.8, 128.7, 125.3, 120.4, 117.2, 98.2, 97.2, 77.2, 33.2. HRMS (ESI): m/z calcd for C23H19N2O2 [M + H]+ 355.1441; found: 355.1433.
  • 3-Methyl-6-(3-methylphenyl)-2-[(3-methylphenyl)ethynyl]pyrano[3,4-d]imidazol-4(3H)-one (5c)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5c as a white solid (63.7 mg, 68%). Mp 186–188 °C. 1H NMR (300 MHz, CDCl3): δ = 7.68–7.62 (m, 2H), 7.45–7.43 (m, 2H), 7.37–7.22 (m, 4H), 7.09 (s, 1H), 4.15 (s,3H), 2.42 (s, 3H), 2.39 (s, 3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 156.8, 155.1, 150.1, 140.3, 138.6, 138.5, 132.7, 132.0, 131.2, 130.7, 129.3, 128.7, 128.6, 126.0, 122.5, 120.3, 117.2, 98.2, 97.5, 33.2, 21.5, 21.2. HRMS (ESI): m/z calcd for C23H19N2O2 [M + H]+ 355.1441; found: 355.1434.
  • 6-(4-Methoxyphenyl)-2-[(4-methoxyphenyl)ethynyl]-3-methylpyrano[3,4-d]imidazol-4(3H)-one (5d)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5d as a white solid (79.5 mg, 78%). Mp 253–255 °C. 1H NMR (300 MHz, CDCl3): δ = 7.81–7.78 (d, J = 8.8 Hz, 2H), 7.59–7.56 (d, J = 8.7 Hz, 2H), 6.99–6.94 (m, 4H), 6.91 (s, 1H), 4.14 (s,3H), 3.86 (s, 6H) ppm. 13C NMR (75 MHz, CDCl3): δ = 161.1, 161.0, 156.6, 155.2, 150.4, 140.6, 133.9, 133.2, 126.9, 124.7, 116.6, 114.4, 114.2, 112.4, 97.6, 96.7, 76.5, 55.4, 33.1. HRMS (ESI): m/z calcd for C23H19N2O4 [M + H]+ 387.1339; found: 387.1332.
  • 6-(4-(Tert-butyl)phenyl)-2-((4-(tert-butyl)phenyl)ethynyl)-3-methylpyrano[3,4-d]imidazol-4(3H)-one (5e)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (27.8 mg, 44%). Mp 261–263 °C. 1H NMR (300 MHz, CDCl3): δ = 7.80 (d, J = 8.6 Hz, 2H), 7.58 (d, J = 8.4 Hz, 2H), 7.49–7.43 (m, 4H), 7.12 (s, 1H), 4.16 (s,3H), 1.35 (s, 9H), 1.34 (s, 9H) ppm. 13C NMR (75 MHz, CDCl3): δ = 161.1, 161.0, 156.6, 155.2, 150.4, 140.6, 133.9, 133.2, 126.9, 124.7, 116.6, 114.4, 114.2, 112.4, 97.6, 96.7, 76.5, 55.4, 33.1. HRMS (ESI): m/z calcd for C29H31N2O2: 439.2380; found: 439.2376.
  • 6-(2-Chlorophenyl)-2-[(2-chlorophenyl)ethynyl]-3-methylpyrano[3,4-d]imidazol-4(3H)-one (5f)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5f as a white solid (74.8 mg, 58%). Mp 179–181 °C. 1H NMR (300 MHz, CDCl3): δ = 7.70–7.66 (m, 2H), 7.51–7.48 (m, 2H), 7.41–7.33 (m, 4H), 7.15 (s, 1H), 4.22 (s,3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 155.3, 154.5, 149.6, 140.1, 136.7, 134.1, 132.6, 131.8, 131.4, 130.9, 130.9, 130.8, 129.8, 127.1, 127.0, 120.9, 117.8, 104.2, 93.8, 82.1, 33.6. HRMS (ESI): m/z calcd for C21H13Cl2N2O2: 395.0348; found: 395.0342.
  • 6-(3-Chlorophenyl)-2-[(3-chlorophenyl)ethynyl]-3-methylpyrano[3,4-d]imidazol-4(3H)-one (5g)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (22.7 mg, 65%). Mp 225–227 °C. 1H NMR (300 MHz, CDCl3): δ = 7.85–7.52 (m, 4H), 7.47–7.34 (m, 4H), 7.12 (s, 1H), 4.17 (s,3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 155.2, 154.7, 149.8, 140.1, 135.3, 134.8, 133.8, 132.1, 130.8, 130.5, 130.3, 130.1, 130.1, 125.6, 123.5, 122.2, 117.8, 99.2, 95.7, 78.2, 33.5. HRMS (ESI): m/z calcd for C21H13Cl2N2O2: 395.0348; found: 395.0342.
  • 3-Methyl-6-(thiophen-2-yl)-2-(thiophen-2-ylethynyl)pyrano[3,4-d]imidazol-4(3H)-one (5h)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (40 mg, 55%). Mp 197–199 °C. 1H NMR (300 MHz, CDCl3): δ = 7.85–7.74 (m, 2H), 7.43–7.36 (m, 3H), 7.29–7.27 (m, 1H), 6.94 (s, 1H), 4.13 (s, 3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 155.0, 153.5, 150.1, 140.4, 134.4, 132.2, 129.9, 127.1, 126.4, 124.6, 124.2, 119.7, 117.1, 97.9, 92.8, 77.2, 33.4. HRMS (ESI): m/z calcd for C17H12N2O2S2: 339.0256; found: 339.0249.
  • 3-Methyl-6-(thiophen-3-yl)-2-(thiophen-3-ylethynyl)pyrano[3,4-d]imidazol-4(3H)-one (5i)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (22.2 mg, 41%). Mp 218–220 °C. 1H NMR (300 MHz, CDCl3): δ = 7.85 (dd, J = 3.0, 1.4 Hz, 1H), 7.75 (dd, J = 3.0, 1.1 Hz, 1H), 7.42 (dd, J = 5.1, 1.4 Hz, 1H), 7.40–7.37 (m, 2H), 7.29 (dd, J = 5.0, 1.2 Hz, 1H), 6.94 (s, 1H), 4.13 (s, 3H) ppm. 13C NMR (75 MHz, CDCl3): δ = 155.1, 153.4, 150.2, 140.4, 134.4, 132.2, 129.9, 127.1, 126.4, 124.6, 124.2, 119.7, 117.1, 98.0, 92.6, 77.2, 33.4. HRMS (ESI): m/z calcd for C17H11N2O2S2: 339.0256; found: 339.0252.
  • 3-Methyl-6-phenethyl-2-(4-phenylbut-1-yn-1-yl)pyrano[3,4-d]imidazol-4(3H)-one (5j)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (28.1 mg, 49%). Mp 165–167 °C. 1H NMR (300 MHz, CDCl3): δ = 7.29–7.21 (m, 8H), 7.21–7.19 (m, 2H), 6.35 (s, 1H), 3.85 (s, 3H), 3.00–2.96 (m, 4H), 2.86–2.82 (m, 4H) ppm. 13C NMR (75 MHz, CDCl3): δ = 159.8, 155.9, 149.6, 140.2, 140.0, 139.7, 128.7, 128.6, 128.5, 128.4, 126.8, 126.4, 116.2, 100.3, 98.7, 70.2, 35.6, 34.1, 33.5, 32.9, 29.8, 21.6. HRMS (ESI): m/z calcd for C25H23N2O2: 383.1754; found: 383.1749.
  • 6-Cyclopropyl-2-(cyclopropylethynyl)-3-methylpyrano[3,4-d]imidazol-4(3H)-one (5k)
The crude product was purified by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), followed by recrystallization from Et2O to afford compound 5e as a white solid (21.4 mg, 52%). Mp 132–134 °C. 1H NMR (300 MHz, CDCl3): δ = 6.41 (s, 1H), 3.96 (s, 3H), 1.87–1.77 (m, 2H), 1.58–1.49 (m, 1H), 1.04–0.94 (m, 8H) ppm. 13C NMR (75 MHz, CDCl3): δ = 161.4, 155.6, 150.2, 140.4, 102.7, 98.0, 64.5, 33.0, 29.8, 14.2, 9.4, 7.5, 0.2. HRMS (ESI): m/z calcd for C15H15N2O2: 255.1128; found: 255.1123.
  • Ethyl 1-methyl-2,4-bis(phenylethynyl)-1H-imidazole-5-carboxylate (6)
Compound 6 was synthesized according to the General Procedure for the Synthesis of 3-methyl-6-aryl-2-alkynylpyrano[3,4-d]imidazole-4(3H)-ones (5ak), described above. After purification by column chromatography on silica gel (PE/EtOAC, 9.5:0.5), compound 6 was obtained as a white solid (36 mg, 41%). 1H NMR (300 MHz, CDCl3): δ = 7.61–7.55 (m, 4H), 7.43–7.34 (m, 6H), 4.42 (q, J = 6.6 Hz, 1H), 4.05 (s, 1H), 1.45 (t, J = 6.8 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ = 159.8, 136.2, 132.2, 131.9, 131.2, 129.9, 128.7, 128.5, 123.0, 121.1, 95.1, 93.0, 83.0, 61.3, 34.6, 14.5.

4. Conclusions

In conclusion, we have developed a novel approach for one-pot copper-mediated tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization reaction of 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid and terminal alkynes. The following unprecedented method demonstrates broad compatibility with structurally diverse al-iphatic, aromatic, and heteroaromatic alkynes. Notably, 5-exo-dig heterocycles were not detected in all experiments, demonstrating the high regioselectivity of the intramolecular 6-endo-dig oxacyclization step. Furthermore, theoretical calculations (computational results) indicated that the favorable reaction position is the β-carbon of alkyne, which is in close agreement with the experimental results.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30143045/s1, S1. Copies of 1H and 13C and NMR spectra, S2. HRMS spectra, S3. Details of DFT calculations, S4. References.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data set presented in this study is available in this article.

Acknowledgments

We thank the “Departement d’Analyses Chimiques et Medicales” (Tours, France) for the chemical analyses.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kishore, D.R.; Sreenivasulu, C.; Satyanarayana, G. 2H-Pyran-2-ones’ Synthesis: State-of-the-Art Methods and Applications. Asian J. Org. Chem. 2025, 14, e202400726. [Google Scholar] [CrossRef]
  2. Gogoi, N.; Parhi, R.; Tripathi, R.K.P.; Pachuau, L.; Kaishap, P.P. Recent advances in synthesis of isocoumarins: An overview. Tetrahedron 2024, 150, 133740. [Google Scholar] [CrossRef]
  3. Ram, V.J.; Goel, A.; Shukla, P.K.; Kapil, A. Synthesis of thiophenes and thieno[3,2-c]pyran-4-ones as antileishmanial and antifungal agents. Bioorg. Med. Chem. Lett. 1997, 7, 3101. [Google Scholar] [CrossRef]
  4. Thabet, H.K.; Althagbi, H.I.; Al Zahrani, N.A.; Helal, M.H.; Gouda, M.A. Recent Advances in the Chemistry of Thienopyranone Heterocycles: Synthesis, Reactivity, and Application. Mini-Rev. Org. Chem. 2025, 1875–6298. [Google Scholar] [CrossRef]
  5. Nakhi, A.; Adepu, R.; Rambabu, D.; Kishore, R.; Vanaja, G.R.; Kalle, A.M.; Pal, M. Thieno[3,2-c]pyran-4-one based novel small molecules: Their synthesis, crystal structure analysis and in vitro evaluation as potential anticancer agents. Bioorg. Med. Chem. Lett. 2012, 22, 4418–4427. [Google Scholar] [CrossRef]
  6. Ombrato, R.; Garofalo, B.; Mangano, G.; Capezzone de Joannon, A.; Corso, G.; Cavarishia, C.; Furlotti, G.; Iacoangeli, T. Antibacterial Compounds Having Broad Spectrum of Activity. US10221144B2, 5 March 2019. [Google Scholar]
  7. Mederski, W.W.K.R.; Dorsch, D.; Osswald, M.; Beier, N.; Lues, I.; Minck, K.-O.; Schelling, P.; Ladstetter, B.J. 4,5-Dihydro-4-oxo-3H-imidazo[4,5-c]pyridines: Potent arylacetic acid-derived AT1 antagonists with improved affinity for the AT2 receptor. Bioorg. Med. Chem. Lett. 1995, 5, 2665–2670. [Google Scholar] [CrossRef]
  8. Mederski, W.W.K.R.; Dorsch, D.; Osswald, M.; Schwartz, H.; Beier, N.; Christadler, M.; Minck, K.O.; Schelling, P.; Schmitges, C.J. Novel 4,5-dihydro-4-oxo-3H-imidazo[4,5-c]pyridines. Potent angiotensin II receptor antagonists with high affinity for both the AT1 and AT2 subtypes. Eur. J. Med. Chem. 1997, 32, 479–491. [Google Scholar] [CrossRef]
  9. Casimiro-Garcia, A.; Heemstra, R.J.; Bigge, C.F.; Chen, J.; Ciske, F.A.; Davis, J.A.; Ellis, T.; Esmaeil, N.; Flyn, D.; Han, S.; et al. PowellDesign, synthesis, and evaluation of imidazo[4,5-c]pyridin-4-one derivatives with dual activity at angiotensin II type 1 receptor and peroxisome proliferator-activated receptor-γ. Bioorg. Med. Chem. Lett. 2013, 23, 767–772. [Google Scholar] [CrossRef]
  10. Raju, S.; Batchu, V.R.; Swamy, N.K.; Dev, R.V.; Babu, J.M.; Kumar, P.R.; Mukkanti, K.; Pal, M. Palladium-mediated synthesis of 5-substituted 4-alkynylthieno[2,3-c]pyran-7-ones. Tetrahedron Lett. 2006, 47, 83–88. [Google Scholar] [CrossRef]
  11. Gorja, D.R.; Batchu, V.R.; Ettam, A.; Pal, M. Pd/C-mediated synthesis of α-pyrone fused with a five-membered nitrogen heteroaryl ring: A new route to pyrano[4,3-c]pyrazol-4(1H)-ones. Beilstein J. Org. Chem. 2009, 5, 64. [Google Scholar] [CrossRef]
  12. Roy, P.; Ghorai, B.K. One-pot synthesis of pyrano[4,3-b]quinolinones from 2-alkynyl-3-formylquinolines via oxidative 6-endo-dig ring closure. Tetrahedron Lett. 2012, 53, 235–238. [Google Scholar] [CrossRef]
  13. Rao, M.S.; Haritha, M.; Chandrasekhar, N.; Rao, M.V.B.; Pal, M. Pd/ligand-free synthesis of thienopyranones via Cu-catalyzed coupling-cyclization in PEG 400 under ultrasound. Tetrahedron Lett. 2014, 55, 1660–1663. [Google Scholar] [CrossRef]
  14. Ngi, S.I.; Guilloteau, V.; Abarbri, M.; Thibonnet, J. Regioselective Copper-Mediated Synthesis of Thieno[2,3-c]pyrane-7-one, Indolo[2,3-c]pyrane-1-one, and Indolo[3,2-c]pyrane-1-one. J. Org. Chem. 2011, 76, 8347–8354. [Google Scholar] [CrossRef]
  15. Bahlaouan, Z.; Abarbri, M.; Duchêne, A.; Thibonnet, J.; Henry, N.; Enguehard-Gueiffier, C.; Gueiffier, A. Copper(I)-mediated preparation of new pyrano[3′,4′:4,5]imidazo[1,2-a]pyridin-1-one compounds under mild palladium-free conditions. Org. Biomol. Chem. 2011, 9, 1212–1218. [Google Scholar] [CrossRef] [PubMed]
  16. Kanwal, I.; Mujahid, A.; Rasool, N.; Rizwan, K.; Malik, A.; Ahmad, G.; Shah, S.A.A.; Rashid, U.; Nadiah, M.N. Palladium and Copper Catalyzed Sonogashira cross Coupling an Excellent Methodology for C-C Bond Formation over 17 Years: A Review. Catalysts 2020, 10, 443. [Google Scholar] [CrossRef]
  17. Mohajer, F.; Heravi, M.M.; Zadsirjan, V.; Poormohammad, N. Copper-free Sonogashira cross-coupling reactions: An overview. RSC Adv. 2021, 11, 6885–6925. [Google Scholar] [CrossRef]
  18. Gazvoda, M.; Virant, M.; Pinter, B.; Košmrlj, J. Mechanism of copper-free Sonogashira reaction operates through palladium-palladium transmetallation. Nat. Commun. 2018, 9, 4814. [Google Scholar] [CrossRef]
  19. Thomas, A.M.; Sujathaa, A.; Anilkumar, G. Recent advances and perspectives in copper-catalyzed Sonogashira coupling reactions. RSC Adv. 2014, 4, 21688–21698. [Google Scholar] [CrossRef]
  20. Brasholz, M.; Luan, X.; Reissig, H.U. Towards the Rubromycins: An Efficient Synthesis of a Suitable Isocoumarin Precursor, its Lactam Analogue, and Palladium-Catalyzed Couplings. Synthesis 2005, 20, 3571–3580. [Google Scholar] [CrossRef]
  21. Brasholz, M.; Reissig, H.U. An Expedient and Short Synthesis of a 6-Iodo Isocoumarin Building Block for the Rubromycin Family and its First Palladium-Catalyzed Coupling. Synlett 2004, 15, 2736–2738. [Google Scholar] [CrossRef]
  22. Waters, S.P.; Kozlowski, M.C. Synthesis of the isocoumarin portion of the rubromycins. Tetrahedron Lett. 2001, 42, 3567–3570. [Google Scholar] [CrossRef]
  23. CCDC 2463986 [For Compound 5a] Contains the Supplementary Crystallographic Data for this Paper. These Data Can Be Obtained Free of Charge from the Cambridge Crystallographic Data Centre. Available online: https://www.ccdc.cam.ac.uk/ (accessed on 13 June 2025).
  24. Monnier, F.; Turtaut, F.; Duroure, L.; Taillefer, M. Copper-Catalyzed Sonogashira-Type Reactions Under Mild Palladium-Free Conditions. Org. Lett. 2008, 10, 3203–3206. [Google Scholar] [CrossRef] [PubMed]
  25. Kotovshchikov, Y.N.; Latyshev, G.V.; Lukashev, N.V.; Beletskaya, I.P. Alkynylation of steroids via Pd-free Sonogashira coupling. Org. Biomol. Chem. 2015, 13, 5542–5555. [Google Scholar] [CrossRef] [PubMed]
  26. Das, S.; Paul, S.; Mandal, K.K.; Anoop, A.; Nanda, S. Asymmetric Total Synthesis of Naturally Occurring (R)-2′-Methoxydihydroartemidin, (R)-(E)-3′-Hydroxyartemidin, and Its Structural Congeners: Method Optimization and Mechanistic Analysis. J. Org. Chem. 2024, 89, 15764–15776. [Google Scholar] [CrossRef] [PubMed]
  27. O’Boyle, N.M.; Banck, M.; James, C.A.; Morley, C.; Vandermeersch, T.; Hutchison, G.R. Open Babel: An open chemical toolbox. J. Cheminf. 2011, 3, 33. [Google Scholar] [CrossRef]
  28. Bannwarth, C.; Caldeweyher, E.; Ehlert, S.; Hansen, A.; Pracht, P.; Seibert, J.; Spicher, S.; Grimme, S. Extended tight-binding quantum chemistry methods. WIREs Comput. Mol. Sci. 2020, 11, e01493. [Google Scholar] [CrossRef]
  29. Neese, F. The ORCA program system. WIREs Comput Mol Sci 2012, 2, 73–78. [Google Scholar] [CrossRef]
  30. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  31. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  32. Raghavachari, K. Perspective on “Density functional thermochemistry. III. The role of exact exchange”. Theor. Chem. Acc. 2000, 103, 361–363. [Google Scholar] [CrossRef]
  33. Arnim, H.; Dmitrij, R. Development of new auxiliary basis functions of the Karlsruhe segmented contracted basis sets including diffuse basis functions (def2-SVPD, def2-TZVPPD, and def2-QVPPD) for RI-MP2 and RI-CC calculations. Phys. Chem. Chem. Phys. 2015, 17, 1010–1017. [Google Scholar]
Figure 1. Some examples of bioactive compounds containing the pyrano[4,3-d]imidazol-4-one skeleton and derivatives. (a) DNA gyrase and topo IV inhibitor; (bd) antagonists of the angiotensin II.
Figure 1. Some examples of bioactive compounds containing the pyrano[4,3-d]imidazol-4-one skeleton and derivatives. (a) DNA gyrase and topo IV inhibitor; (bd) antagonists of the angiotensin II.
Molecules 30 03045 g001
Figure 2. Previous and present work on Sonogashira/6-endo-dig cyclization of different heteroaromatic β-iodo-α,β-unsaturated carboxylic acid. (a) Pd/Cu co-catalyst process developed by M. Pal et al. (2006) [10], M. Pal et al. (2009) [11] and by B. K. Ghorai et al. (2012) [12]; (b) Cu catalyst process developed by our group (2011) [14,15] and by M. Pal et al. (2014) [13].
Figure 2. Previous and present work on Sonogashira/6-endo-dig cyclization of different heteroaromatic β-iodo-α,β-unsaturated carboxylic acid. (a) Pd/Cu co-catalyst process developed by M. Pal et al. (2006) [10], M. Pal et al. (2009) [11] and by B. K. Ghorai et al. (2012) [12]; (b) Cu catalyst process developed by our group (2011) [14,15] and by M. Pal et al. (2014) [13].
Molecules 30 03045 g002
Scheme 1. Synthesis of 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 from ethyl 1H-imidazole-5-carboxylate 1.
Scheme 1. Synthesis of 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 from ethyl 1H-imidazole-5-carboxylate 1.
Molecules 30 03045 sch001
Scheme 2. Substrate scope for the synthesis of pyrano[4,3-d]imidazol-4-one derivatives 5ak.
Scheme 2. Substrate scope for the synthesis of pyrano[4,3-d]imidazol-4-one derivatives 5ak.
Molecules 30 03045 sch002
Scheme 3. Copper-mediated Sonogashira reaction at C-2 and C-4 positions between 3 and phenylacetylene.
Scheme 3. Copper-mediated Sonogashira reaction at C-2 and C-4 positions between 3 and phenylacetylene.
Molecules 30 03045 sch003
Scheme 4. Plausible mechanism of the Cu-mediated tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization.
Scheme 4. Plausible mechanism of the Cu-mediated tandem Sonogashira coupling/regioselective 6-endo-dig oxacyclization.
Molecules 30 03045 sch004
Figure 3. Plausible natural population analysis charge on the carbons Cα and Cβ of the alkyne substituent.
Figure 3. Plausible natural population analysis charge on the carbons Cα and Cβ of the alkyne substituent.
Molecules 30 03045 g003
Table 1. Optimization of the reaction conditions between 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 and phenylacetylene.
Table 1. Optimization of the reaction conditions between 2,4-diiodo-1-methyl-1H-imidazole-5-carboxylic acid 4 and phenylacetylene.
Molecules 30 03045 i001
EntryBaseCatalyst (equiv.)Alkyne (equiv.)T (°C)Time (h)Ratio (%) a5a/4
1Et3NCuI/PdCl2(PPh3)2 (0.2/0.1)3130347/53
2Et3NCuI/PdCl2(PPh3)2 (1/0.1)3130370/30
3K2CO3CuI/PdCl2(PPh3)2 (1/0.1)31303100/0 (44%) b
4K2CO3CuI (1)313024100/0 (27%) b
5K2CO3CuI (2)31303100/0 (51%) b
6K2CO3CuI (2.5)31303100/0 (47%) b
7K2CO3CuI (2)2.051303100/0 (24%) b
8K2CO3CuI (2)1.051303100/0 (12%) b
9K2CO3CuI (2)31503100/0 (44%) b
10K2CO3CuI (2)310024100/0 (12%) b
11Na2CO3CuI (2)31303100/0 (9%) b
12K3PO4CuI (2)3130355/45
13Cs2CO3CuI (2)3130352/48
14Et3NCuI (2)3130355/45
a The ratio of mixture (5a/4) was calculated from the crude 1H NMR spectrum. b Isolated yield.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ayachi, A.; Tikad, A.; Lazeran, V.; Allouchi, H.; Bletry, M.; Besbes, R.; Abarbri, M.; Jismy, B. Pd/Ligand-Free Synthesis of 2-Alkynylated Pyrano[4,3-d]imidazol-4-ones via One-Pot Cu-Mediated Tandem Sonogashira Coupling/Regioselective 6-endo-dig Oxacyclization Reaction. Molecules 2025, 30, 3045. https://doi.org/10.3390/molecules30143045

AMA Style

Ayachi A, Tikad A, Lazeran V, Allouchi H, Bletry M, Besbes R, Abarbri M, Jismy B. Pd/Ligand-Free Synthesis of 2-Alkynylated Pyrano[4,3-d]imidazol-4-ones via One-Pot Cu-Mediated Tandem Sonogashira Coupling/Regioselective 6-endo-dig Oxacyclization Reaction. Molecules. 2025; 30(14):3045. https://doi.org/10.3390/molecules30143045

Chicago/Turabian Style

Ayachi, Abir, Abdellatif Tikad, Vincent Lazeran, Hassan Allouchi, Marc Bletry, Rafâa Besbes, Mohamed Abarbri, and Badr Jismy. 2025. "Pd/Ligand-Free Synthesis of 2-Alkynylated Pyrano[4,3-d]imidazol-4-ones via One-Pot Cu-Mediated Tandem Sonogashira Coupling/Regioselective 6-endo-dig Oxacyclization Reaction" Molecules 30, no. 14: 3045. https://doi.org/10.3390/molecules30143045

APA Style

Ayachi, A., Tikad, A., Lazeran, V., Allouchi, H., Bletry, M., Besbes, R., Abarbri, M., & Jismy, B. (2025). Pd/Ligand-Free Synthesis of 2-Alkynylated Pyrano[4,3-d]imidazol-4-ones via One-Pot Cu-Mediated Tandem Sonogashira Coupling/Regioselective 6-endo-dig Oxacyclization Reaction. Molecules, 30(14), 3045. https://doi.org/10.3390/molecules30143045

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop