Next Article in Journal
S-Ethyl-Isothiocitrullin-Based Dipeptides and 1,2,4-Oxadiazole Pseudo-Dipeptides: Solid Phase Synthesis and Evaluation as NO Synthase Inhibitors
Next Article in Special Issue
Systematic Assessment of the Catalytic Reactivity of Frustrated Lewis Pairs in C-H Bond Activation
Previous Article in Journal
Zinc Oxide Nanoparticles Blunt Potassium-Bromate-Induced Renal Toxicity by Reinforcing the Redox System
Previous Article in Special Issue
Effect of Controlling Thiophene Rings on D-A Polymer Photocatalysts Accessed via Direct Arylation for Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Simple and Practical Bis-N-Heterocyclic Carbene as an Efficient Ligand in Cu-Catalyzed Glaser Reaction

Hunan Provincial Key Laboratory of Xiangnan Rare-Precious Metals Compounds Research and Application, School of Chemistry and Environmental Science, Xiangnan University, Chenzhou 423000, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(13), 5083; https://doi.org/10.3390/molecules28135083
Submission received: 29 May 2023 / Revised: 17 June 2023 / Accepted: 27 June 2023 / Published: 29 June 2023
(This article belongs to the Special Issue π-Conjugated Functional Molecules & Polymers)

Abstract

:
Conjugated diyne derivatives are important scaffolds in modern organic synthetic chemistry. Using the Glaser reaction involves the coupling of terminal alkynes which can efficiently produce conjugated diyne derivatives, while the use of a stoichiometric amount of copper salts, strong inorganic base, and excess oxidants is generally needed. Developing an environmentally friendly and effective method for the construction of symmetrical 1,3-diynes compounds by Glaser coupling is still highly desirable. In this study, we present an economical method for the production of symmetric diynes starting from various terminal acetylenes in a Glaser reaction. A simple and practical bis-N-heterocyclic carbene ligand has been introduced as efficient ligands for the Cu-catalyzed Glaser reaction. High product yields were obtained at 100 °C for a variety of substrates including aliphatic and aromatic terminal alkynes and differently substituted terminal alkynes including the highly sterically hindered substrate 2-methoxy ethynylbenzene or 2-trifluoromethyl ethynylbenzene and a series of functional groups, such as trifluoromethyl group, ester group, carboxyl group, and nitrile group. The established protocol is carried out in air under base-free condition and is operationally simple. These research work suggest that bis-N-heterocyclic carbene could also an appealing ligand for Glaser reaction and provide a reference for the preparation of symmetric 1,3-diynes in industrial filed.

1. Introduction

Alkynes are important structural units existing in a wide range of natural products, pharmaceuticals, and material functional molecules [1,2,3]. Alkynes also as building blocks play generally an important role in organic synthesis due to the diverse reactivity [4,5,6,7]. Particularly, conjugate 1,3-dialkynes are valuable building blocks for the synthesis of linearly π-conjugated acetylenic oligomers and polymers [8,9], supramolecular materials [10,11], and substituted heterocyclic compounds [12,13,14]. In this context, the efficient synthesis method toward this goal is always the key point of chemists’ attention [15,16,17,18,19,20]. Conventionally, the widely known method of producing synthetically important 1,3-diyne derivatives by the homocoupling of various alkynyl organometallics [21,22,23]. This reliable route, however, is not atom efficient, generating a stoichiometric amount of salt byproduct. In contrast, the homocoupling of terminal alkynes, such as Glaser-Hay coupling [24,25,26] and Cadiot-Chodkiewicz coupling [27,28,29], is widely believed to be an economic and practical method to construct 1,3-diynes.
Since the pioneering work by Glaser in 1869 [30], Glaser coupling and its modifications have become one of the most widely utilized and powerful strategies for the synthesis of diverse 1,3-dialkynes [31,32,33,34]. Numerous metal catalytic systems such as palladium [35,36], cobalt [37,38], ruthenium [39], and gold, etc., [40,41,42] have been studied on this reaction. However, most of their wide applications are limited because of their higher price or instability in air. Therefore, copper-catalyzed aerobic homocoupling of terminal alkynes remains the most popular method due to copper being an abundant and inexpensive metal. Very recently, Tiwari and co-workers [25] reported a benzotriazole ligand for the cross-coupling of terminal alkynes catalyzed by CuI in the presence of K2CO3. Lankalapalli and co-workers, using a catalytic combination of a 2-azidopyridine analogue, 4-azido-5H-pyrrolo [3,2-d]pyrimidine, and CuI, afforded homocoupled products of terminal alkynes [43]. However, it usually needs a stoichiometric amount of copper salts, excess oxidants, rigid inorganic bases, and high temperature to afford low to moderate yields for aliphatic diynes. On the other hand, Glaser coupling of terminal alkynes catalyzed using copper-containing heterogeneous catalyst has also reported in very recently [44,45]. Despite the great progress, to the best of our knowledge, few catalyst systems have been reported for the homocoupling of terminal alkynes under a neutral condition with a mild condition [46]. From both academic and industrial standpoints, it is still highly desirable to develop a more simple, effective, and environmentally friendly method for the construction of symmetrical 1,3-diyne compounds by Glaser coupling.
Using N-heterocyclic carbine (NHC) as ligands, various cross coupling reactions have been achieved under mild and environmentally friendly conditions [47,48,49,50]. So far, N-heterocyclic carbene have emerged as one of the most powerful green ligands in transition-metal catalyzed coupling reactions [51,52,53,54]. Recently, we have also demonstrated a bis-NHC as a ligand that combines a nonprecious metal that would efficiently catalyze C-C and C-N cross-coupling reactions under mild conditions [55,56]. Considering NHC ligands’ easy accessibility, environmental benignancy, and stronger σ-donors and that their steric bulkiness is tunable through variation of the N-substituents, we therefore hypothesized that we can utilize NHC as ligand for a Cu-catalyzed Glaser reaction under external-base-free conditions to produce important conjugate 1,3-diyne derivatives with broad functional group tolerance. In this context, herein, we report an efficient catalytic system with a simple and practical bis-NHC and CuCl for the Glaser coupling reaction under an air atmosphere which avoids the utilization of any base (Scheme 1). This protocol has the advantages of broad substrate scope, green reaction conditions, easy operation, and good yields.

2. Results and Discussion

We initially examined the reaction of phenyl acetylene 1a under a range of reaction conditions in the presence of a CuCl catalyst (Table 1). A series of ligands were screened to evaluate their suit abilities (Scheme 2). All results are summarized in Table 1. We were pleased to find that the desired product 2a was formed at an 86% yield when ligand L2 was used. Other ligands, such as L1, L3, L4, L5, and L6 were also examined and afford product 2a with yields of 78%, 34%, 42%, 49%, and 47%, respectively (Table 1, entries 1–6). These results indicated that L2 was most effective among the ligands L1L6. We next studied the solvent effect because the solvent is also an important factor in this reaction. Only traces of the product were seen on the TLC plate when 1,4-dioxane was used as the solvent (Table 1, entry 7). To our delight, DMF gives the best result with a yield of 95% (Table 1, entry 8). Other solvents, such as DMAc, toluene, and xylene, also show similar reaction activities (Table 1, entry 9–11). There was no significant reduction of the yield when the temperature was 100 °C, and the reaction time could decrease to 4 h (Table 1, entry 12–16). However, further reduction of the temperature resulted in decreased product yield (Table 1, entry 17–19). Finally, control experiments indicated that both the CuCl and the ligand L2 are essential for the reaction (Table 1, entry 20–25). Thus, it could be concluded that the optimal conditions were 5 mol% CuCl, 10 mol% L2, DMF as a solvent, and air as an oxidant at 100 °C for 4 h.
With the optimized reaction conditions in hand, the substrate scope was next explored. We first study the effect of substituents on the homocoupling of terminal alkynes. As shown in Scheme 3, various aromatic terminal alkynes with either electron-donating or electron-withdrawing substituents were well tolerated under the identified conditions, affording the desired products 2b2j with 89–96% yields. A series of functional groups, such as the trifluoromethyl group, ester group, carboxyl group, and nitrile group, were compatible with the present catalytic system. For example, 4-ethynyl-α,α,α-trifluorotoluene, methyl 4-ethynylbenzoate, 4-ethynylbenzoic acid, and 4-ethynylbenzonitrile were suitable for the couplings, and good to excellent yields of the products were obtained (86–90%). Delightedly, the sterically hindered ortho-methyl ethynylbenzene or their derivatives could provide the corresponding homocoupling product 2o and 2p in 96% and 92% yield, respectively. Even the highly sterically hindered substrate 2-methoxy ethynylbenzene or 2-trifluoromethyl ethynylbenzene could also react smoothly to afford the desired products 2q and 2r in a good yield of 93% and 89%.
Encouraged by the above results, we explore further the scope of aliphatic terminal alkynes and heterocyclic acetylenes. It is well known that aliphatic terminal alkynes are less reactive compared to aromatic terminal alkynes. To our delight, a satisfying yield was obtained when the reaction was carried out in 120 °C with 12 h. As shown in Scheme 4, heterocyclic acetylenes, such as 3-ethynylpyridine, 2-ethynylpyridine, and 3-ethynylthiophene, can be successfully reacted to afford the corresponding compounds 4a4c in 81–90% yields. It is noteworthy that, when we employed aliphatic terminal alkynes, they performed well and provided the corresponding conjugated diyne products 4d4h in moderate yields. These results show that heterocyclic or aliphatic acetylenes are also tolerated in this present protocol.
With regard to our results, the reaction mechanism is not fully understood at present. After referring to a lot of the relevant literature [14,24,25,40,43,46], we proposed that this Cu-catalyzed Glaser homocoupling might take place as shown in Scheme 5, which possibly involves the typical Glaser reaction steps. According to our assumptions, halogen bromide anions perhaps act as Lewis bases which have a weak ability to deproton in this catalysis system. The ligand L was coordinated with CuCl to form complex A. Then, the intermediate A reacts with terminal alkyne to form alkynyl copper intermediate B, and then it involved the oxidative reaction with O2 to give the conjugated 1,3-diynes at the last step via intermediates C and transition state TS.

3. Materials and Methods

3.1. General Information

Unless otherwise stated, all chemicals and reagents were commercially available in analytical grade without further purification. All terminal alkynes were purchased from Aldrich Chemical Co. Ltd. (St. Louis, MI, USA). All solvents were purchased from Shanghai Macklin Chemical Co. Ltd. (Shanghai, China). L1L6 were prepared according to our previous reported procedures [49,56]. All reactions were performed under an atmosphere of air unless otherwise stated. Analytical thin layer chromatography (TLC) was performed on silica gel GF254 (layer thickness 0.20–0.25 mm). Column chromatography was carried out on silica gel (300−400 mesh) using petroleum ether as eluent. 1H, 13C NMR spectra were performed at room temperature on a Bruker Avance 400 MHz spectrometer using the residual solvent signal as internal standard (CDCl3: 7.26 ppm (1H), 77.16 ppm (13C); DMSO-d6: 2.50 ppm (1H), 39.52 ppm (13C)) (Supplementary Materials).

3.2. Typical Experimental Procedure for the Synthesis of 1,3-Diyne

Unless otherwise noted, the Glaser reaction was carried out under aerobic conditions. All solvents were used as received, and no further purification was needed. A parallel reactor containing a stir bar was charged with alkynes (1.0 mmol), CuCl (5% mol), ligands (10 mol %), and 1 mL of solvent. The reaction mixture was carried out at 100 °C for 4 h. After completion of the reaction, the reaction mixture was cooled to ambient temperature, and 10 mL of water was added. The mixture was diluted with dichloromethane (5 mL), followed by extraction three times (3 × 5 mL) with dichloromethane. The organic layer was dried with anhydrous magnesium sulfate, filtered, and evaporated under reduced pressure. The crude cross-coupling products were purified by silica-gel column chromatography using petroleum ether as eluent, and the isolated yield was then calculated. The isolated cross-coupling products were characterized by 1H NMR and 13C NMR.

3.3. Characterization Data of the Products

  • L1 [56]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 9.60 (s, 2H), 8.12 (s, 2H), 7.83 (s, 2H), 6.81 (s, 2H), 3.91 (s, 6H). 13C NMR (100 MHz, DMSO-d6) δ 138.03, 124.27, 121.88, 57.79, 36.28.
  • L2 [57]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 10.52 (s, 2H), 8.50 (d, J = 17.3 Hz, 4H), 7.86 (d, J = 7.8 Hz, 4H), 7.70 (t, J = 7.6 Hz, 4H), 7.63 (t, J = 7.3 Hz, 2H), 7.01 (s, 2H). 13C NMR (100 MHz, DMSO-d6) δ 137.33, 134.46, 130.28, 130.17, 123.04, 121.92, 121.50, 58.24.
  • L3 [57]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 10.22 (s, 2H), 8.56 (s, 2H), 8.25 (s, 2H), 7.65 (d, J = 7.7 Hz, 2H), 7.56 (q, J = 7.4 Hz, 4H), 7.49 (t, J = 7.2 Hz, 2H), 7.05 (s, 2H), 2.31 (s, 6H). 13C NMR (100 MHz, DMSO-d6) δ 138.79, 133.93, 133.20, 131.76, 130.79, 127.38, 126.31, 124.12, 122.46, 58.08, 17.30.
  • L4 [57]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 10.28 (s, 2H), 8.52 (s, 2H), 8.24 (s, 2H), 7.72 (dd, J = 7.9, 1.3 Hz, 2H), 7.66–7.57 (m, 2H), 7.41 (d, J = 8.3 Hz, 2H), 7.21 (t, J = 7.7 Hz, 2H), 7.08 (s, 2H), 3.93 (s, 6H). 13C NMR (100 MHz, DMSO-d6) δ 151.78, 138.77, 131.88, 125.78, 123.96, 123.03, 122.00, 121.13, 113.42, 57.97, 56.53.
  • L5 [49]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 10.62 (s, 2H), 8.81 (s, 2H), 8.60 (s, 1H), 8.25 (s, 2H), 8.07 (s, 2H), 4.05 (s, 6H). 13C NMR (100 MHz, DMSO-d6) δ 145.23, 144.89, 136.32, 125.00, 119.10, 114.04, 36.51.
  • L6 [49]. White solid. 1H NMR (400 MHz, DMSO-d6) δ 10.70 (s, 2H), 8.93 (s, 2H), 8.61 (t, J = 8.1 Hz, 1H), 8.42–8.11 (m, 4H), 4.90 (dt, J = 13.3, 6.6 Hz, 2H), 1.62 (d, J = 6.7 Hz, 12H). 13C NMR (100 MHz, DMSO-d6) δ 145.31, 144.62, 134.73, 121.78, 119.66, 114.30, 53.42, 22.20.
  • 1,4-diphenylbuta-1,3-diyne (2a) [24]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.54 (dd, J = 7.8, 1.4 Hz, 4H), 7.44–7.30 (m, 6H). 13C NMR (100 MHz, CDCl3) δ 132.64, 129.35, 128.59, 121.95, 81.70, 74.06.
  • 1,4-di-p-tolylbuta-1,3-diyne (2b) [24]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.42 (d, J = 7.9 Hz, 4H), 7.14 (d, J = 7.8 Hz, 4H), 2.37 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 139.63, 132.53, 129.35, 118.95, 81.69, 73.60, 21.76.
  • 1,4-bis(4-isopropylphenyl)buta-1,3-diyne (2c) [5]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.45 (d, J = 8.2 Hz, 4H), 7.19 (d, J = 8.1 Hz, 4H), 2.91 (dt, J = 13.8, 6.9 Hz, 2H), 1.25 (d, J = 6.9 Hz, 12H). 13C NMR (100 MHz, CDCl3) δ 150.46, 132.66, 126.75, 119.30, 81.70, 73.56, 34.32, 23.86.
  • 1,4-bis(4-(tert-butyl)phenyl)buta-1,3-diyne (2d) [24]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.45 (t, J = 10.7 Hz, 4H), 7.39–7.31 (m, 4H), 1.32 (s, 18H). 13C NMR (100 MHz, CDCl3) δ 152.71, 132.40, 125.61, 118.98, 81.65, 73.63, 35.06, 31.25.
  • 1,4-bis(4-ethylphenyl)buta-1,3-diyne (2e) [45]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.45 (d, J = 8.0 Hz, 4H), 7.17 (d, J = 8.0 Hz, 4H), 2.67 (q, J = 7.6 Hz, 4H), 1.24 (t, J = 7.6 Hz, 6H). 13C NMR (100 MHz, CDCl3) δ 145.88, 132.63, 128.16, 119.18, 81.71, 73.61, 29.06, 15.37.
  • 1,4-bis(4-propylphenyl)buta-1,3-diyne (2f) [43]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.44 (d, J = 8.1 Hz, 4H), 7.14 (d, J = 8.1 Hz, 4H), 2.82–2.41 (m, 4H), 1.83–1.52 (m, 4H), 0.94 (t, J = 7.3 Hz, 6H). 13C NMR (100 MHz, CDCl3) δ 144.37, 132.54, 128.76, 119.20, 81.73, 73.64, 38.19, 24.40, 13.89.
  • 1,4-bis(4-butylphenyl)buta-1,3-diyne (2g) [25]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.44 (d, J = 8.1 Hz, 4H), 7.15 (d, J = 8.2 Hz, 4H), 2.78–2.47 (m, 4H), 1.60 (dt, J = 12.9, 7.5 Hz, 4H), 1.35 (dq, J = 14.6, 7.3 Hz, 4H), 0.93 (t, J = 7.3 Hz, 6H). 13C NMR (100 MHz, CDCl3) δ 144.60, 132.55, 128.70, 119.15, 81.73, 73.63, 35.83, 33.45, 22.45, 14.05.
  • 1,4-bis(4-pentylphenyl)buta-1,3-diyne (2h) [24]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.44 (d, J = 8.0 Hz, 4H), 7.15 (d, J = 8.0 Hz, 4H), 2.98–2.39 (m, 4H), 1.78–1.52 (m, 4H), 1.32 (dt, J = 10.6, 3.3 Hz, 8H), 0.90 (t, J = 6.8 Hz, 6H). 13C NMR (100 MHz, CDCl3) δ 144.63, 132.55, 128.69, 119.14, 81.73, 73.64, 36.11, 31.58, 30.99, 22.65, 14.14.
  • 1,4-di([1,1′-biphenyl]-4-yl)buta-1,3-diyne (2i) [5]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.64–7.53 (m, 12H), 7.45 (t, J = 7.5 Hz, 4H), 7.37 (t, J = 7.3 Hz, 2H). 13C NMR (100 MHz, CDCl3) δ 141.74, 140.41, 132.70, 129.02, 127.87, 127.20, 127.15, 121.13, 83.70, 77.88.
  • 1,4-bis(4-methoxyphenyl)buta-1,3-diyne (2j) [24]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.46 (d, J = 8.9 Hz, 4H), 6.85 (d, J = 8.8 Hz, 4H), 3.82 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 160.39, 134.18, 114.28, 114.11, 81.38, 73.10, 55.48.
  • 1,4-bis(4-(trifluoromethyl)phenyl)buta-1,3-diyne (2k) [5]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.63 (q, J = 8.5 Hz, 8H). 13C NMR (100 MHz, CDCl3) δ 132.96, 131.27 (d, J = 33.0 Hz), 125.61 (d, J = 3.7 Hz), 125.31 (d, J = 22.8 Hz), 122.49, 81.12, 75.79.
  • dimethyl 4,4′-(buta-1,3-diyne-1,4-diyl)dibenzoate (2l) [24]. White solid. Rf = 0.7 (10% ethyl acetate/n-hexane) 1H NMR (400 MHz, CDCl3) δ 8.02 (d, J = 8.4 Hz, 4H), 7.59 (d, J = 8.4 Hz, 4H), 3.93 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 166.41, 132.62, 130.72, 129.73, 126.28, 82.00, 76.42, 52.50.
  • 4,4′-(buta-1,3-diyne-1,4-diyl)dibenzoic acid (2m) [58]. White solid. Rf = 0.6 (ethyl acetate). 1H NMR (400 MHz, CDCl3) δ 8.06 (d, J = 8.2 Hz, 4H), 7.58 (d, J = 8.2 Hz, 4H), 3.26 (s, 2H). 13C NMR (100 MHz, CDCl3) δ 170.03, 132.33, 130.19, 80.63.
  • 4,4′-(buta-1,3-diyne-1,4-diyl)dibenzonitrile (2n) [43]. White solid. Rf = 0.4 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.74–7.52 (m, 8H). 13C NMR (100 MHz, CDCl3) δ 132.82, 132.17, 127.16, 118.39, 112.50, 82.01, 81.67.
  • 1,4-di-o-tolylbuta-1,3-diyne (2o) [36]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.51 (d, J = 7.6 Hz, 2H), 7.25 (dt, J = 17.1, 7.3 Hz, 4H), 7.16 (t, J = 7.4 Hz, 2H), 2.51 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 141.78, 133.06, 129.71, 129.24, 125.80, 121.88, 81.29, 77.67, 20.88.
  • 1,4-dimesitylbuta-1,3-diyne (2p) [59]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 6.87 (s, 4H), 2.41 (s, 12H), 2.28 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 140.96, 138.25, 127.71, 119.09, 84.64, 81.52, 21.45, 21.02.
  • 1,4-bis(2-methoxyphenyl)buta-1,3-diyne (2q) [36]. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.48 (d, J = 7.6 Hz, 2H), 7.32 (t, J = 7.9 Hz, 2H), 6.90 (dd, J = 15.5, 8.0 Hz, 4H), 3.89 (s, 6H). 13C NMR (100 MHz, Chloroform-d) δ 161.45, 134.49, 130.65, 120.61, 111.40, 110.80, 78.78, 78.10, 55.93.
  • 1,4-bis(2-(trifluoromethyl)phenyl)buta-1,3-diyne (2r) [24]. White solid. White solid. Rf = 0.6 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.70 (t, J = 8.3 Hz, 4H), 7.50 (dt, J = 22.3, 7.5 Hz, 4H). 13C NMR (100 MHz, CDCl3) δ 135.29, 131.63, 129.27, 126.20 (q, J = 4.9 Hz), 124.75, 122.04, 119.90 (d, J = 2.8 Hz), 78.85, 78.75.
  • 1,4-di(pyridin-3-yl)buta-1,3-diyne (4a) [24]. White solid. White solid. Rf = 0.4 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 8.77 (s, 2H), 8.60 (d, J = 3.9 Hz, 2H), 7.93–7.70 (m, 2H), 7.30 (dd, J = 7.6, 5.0 Hz, 2H). 13C NMR (100 MHz, CDCl3) δ 153.12, 149.48, 139.69, 123.31, 119.07, 79.27.
  • 1,4-di(pyridin-2-yl)buta-1,3-diyne (4b) [20]. White solid. White solid. Rf = 0.4 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 8.62 (d, J = 4.6 Hz, 2H), 7.69 (td, J = 7.7, 1.4 Hz, 2H), 7.54 (d, J = 7.8 Hz, 2H), 7.29 (dd, J = 7.0, 5.5 Hz, 2H). 13C NMR (100 MHz, CDCl3) δ 150.53, 142.06, 136.35, 128.56, 123.93, 81.04, 73.35.
  • 1,4-di(thiophen-3-yl)buta-1,3-diyne (4c) [24]. White solid. White solid. Rf = 0.4 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 7.59 (s, 2H), 7.37–7.24 (m, 2H), 7.21–7.03 (m, 2H). 13C NMR (100 MHz, CDCl3) δ 131.37, 130.31, 125.74, 121.04, 76.70, 73.66.
  • 1,4-di(cyclohex-1-en-1-yl)buta-1,3-diyne (4d) [25]. White solid. White solid. Rf = 0.5 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 6.53–6.06 (m, 2H), 2.35–1.95 (m, 8H), 1.66–1.53 (m, 8H). 13C NMR (100 MHz, CDCl3) δ 138.23, 120.10, 82.84, 60.55, 28.83, 26.01, 22.27, 21.45.
  • 1,1′-(buta-1,3-diyne-1,4-diyl)bis(cyclohexan-1-ol) (4e) [24]. White solid. Rf = 0.4 (ethyl acetate) 1H NMR (400 MHz, CDCl3) δ 2.14 (s, 2H), 1.88 (dd, J = 9.9, 5.8 Hz, 4H), 1.67 (dq, J = 11.4, 5.4, 4.9 Hz, 4H), 1.61–1.45 (m, 10H), 1.32–1.11 (m, 2H). 13C NMR (100 MHz, CDCl3) δ 87.89, 72.18, 68.61, 39.86, 25.20, 23.21.
  • 1,4-dicyclopropylbuta-1,3-diyne (4f) [20]. Colorless oil. White solid. Rf = 0.5 (n-hexane). 1H NMR (400 MHz, CDCl3) δ 1.24 (ddp, J = 11.4, 5.4, 3.0, 2.5 Hz, 2H), 0.90–0.64 (m, 8H). 13C NMR (100 MHz, CDCl3) δ 87.87, 63.51, 8.26, −0.67.
  • hexa-2,4-diyne-1,6-diyl diacetate (4g) [20]. Colorless oil. Rf = 0.6 (ethyl acetate). 1H NMR (400 MHz, CDCl3) δ 4.63 (s, 4H), 2.06 (s, 6H). 13C NMR (100 MHz, CDCl3) δ 170.14, 77.74, 74.88, 51.97, 20.68.
  • 2,7-dimethylocta-3,5-diyne-2,7-diol (4h) [20]. Colorless oil. Rf = 0.4 (ethyl acetate) 1H NMR (400 MHz, CDCl3) δ 2.13 (s, 2H), 1.49 (s, 12H). 13C NMR (100 MHz, CDCl3) δ 88.91, 70.23, 65.05, 31.35.

4. Conclusions

In summary, using a simple and practical bis-N-heterocyclic carbene as ligand, an efficient and green method for the Cu-catalyzed Glaser reaction to produce 1,3-conjugated dialkynes was developed. This method is carried out in simple procedure and without base additives in air conditions. Various terminal alkynes containing functional groups, such as the ester group, carboxyl group, and nitrile group, etc., could be effectively coupled to produce the corresponding conjugated 1,3-diynes with good yields. Aliphatic alkynes were also compatible in our protocol. This easily accessible system implies that N-heterocyclic carbene can serve as an alternative ligand in the Glaser reaction and offers a reference for the preparation of symmetric 1,3-diynes. This work also demonstrates the great potential of bis-N-heterocyclic carbene in the green ligand for the transition metal-catalyzed coupling reactions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28135083/s1, 1H and 13C NMR spectra of the compounds.

Author Contributions

Conceptualization, X.L. and J.C.; methodology, X.L. and J.C.; validation, J.L. (Jie Liu) and Y.Z.; formal analysis, J.L. (Jun Luo) and Z.Z.; investigation, J.L. (Jie Liu) and Y.Z.; data curation, L.Z. and X.Z.; writing—original draft preparation, J.L. (Jie Liu) and Y.Z.; writing—review and editing, D.L. and X.L.; supervision, J.C.; project administration, X.L.; funding acquisition, X.L. and J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Hunan Provincial Natural Science Foundation of China (No. 2021JJ40519), the Outstanding Youth Project of Hunan Education Department (No. 21B0750), and the Hunan Students’ program for innovation and entrepreneurship training (No. S202110545008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the people from the Hunan Provincial Key Laboratory of Xiangnan Rare-Precious Metals Compounds Research and Application, School of Chemistry and Environmental Science, Xiangnan University.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Li, X.; Liu, L.; Huang, T.; Tang, Z.; Li, C.; Li, W.; Zhang, T.; Li, Z.; Chen, T. Palladium-Catalyzed Decarbonylative Sonogashira Coupling of Terminal Alkynes with Carboxylic Acids. Org. Lett. 2021, 23, 3304–3309. [Google Scholar] [CrossRef]
  2. Jin, L.; Hao, W.; Xu, J.; Sun, N.; Hu, B.; Shen, Z.; Mo, W.; Hu, X. N-Heterocyclic carbene copper-catalyzed direct alkylation of terminal alkynes with non-activated alkyl triflates. Chem. Commun. 2017, 53, 4124–4127. [Google Scholar] [CrossRef]
  3. Shi Shun, A.L.K.; Tykwinski, R.R. Synthesis of Naturally Occurring Polyynes. Angew. Chem. Int. Ed. 2006, 45, 1034–1057. [Google Scholar] [CrossRef]
  4. Yang, B.; Lu, S.; Wang, Y.; Zhu, S. Diverse synthesis of C2-linked functionalized molecules via molecular glue strategy with acetylene. Nat. Commun. 2022, 13, 1858. [Google Scholar] [CrossRef]
  5. Morri, A.K.; Thummala, Y.; Ghosh, S.; Doddi, V.R. Urea-Promoted Metal-Free Homolytic Alkynyl Substitution (HAS): Metal-Free C−C Coupling of Alkynyl Bromides Formed In Situ from 1,1-Dibromoalkenes. ChemistrySelect 2021, 6, 2387–2393. [Google Scholar] [CrossRef]
  6. Sagadevan, A.; Pampana, V.K.K.; Hwang, K.C. Copper Photoredox Catalyzed A3′ Coupling of Arylamines, Terminal Alkynes, and Alcohols through a Hydrogen Atom Transfer Process. Angew. Chem. Int. Ed. 2019, 58, 3838–3842. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, T.; Zhao, C.-Q.; Han, L.-B. Hydrophosphorylation of Alkynes Catalyzed by Palladium: Generality and Mechanism. J. Am. Chem. Soc. 2018, 140, 3139–3155. [Google Scholar] [CrossRef]
  8. Wu, X.; Wei, B.; Hu, R.; Tang, B.Z. Polycouplings of Alkynyl Bromides and Sulfonamides toward Poly(ynesulfonamide)s with Stable Csp–N Bonds. Macromolecules 2017, 50, 5670–5678. [Google Scholar] [CrossRef]
  9. Wang, Y.; Wang, W.; Wang, X.; Cheng, X.; Qin, A.; Sun, J.Z.; Tang, B.Z. Polymerization of 1-chloro-2-benzaldehyde-acetylene using an NHC-Pd/AgOTf catalyst and post-polymerization modification. Polym. Chem. 2017, 8, 5546–5553. [Google Scholar] [CrossRef]
  10. Haque, A.; Al-Balushi, R.A.; Al-Busaidi, I.J.; Khan, M.S.; Raithby, P.R. Rise of Conjugated Poly-ynes and Poly(Metalla-ynes): From Design Through Synthesis to Structure–Property Relationships and Applications. Chem. Rev. 2018, 118, 8474–8597. [Google Scholar] [CrossRef]
  11. Wang, X.; Sun, J.Z.; Tang, B.Z. Poly(disubstituted acetylene)s: Advances in polymer preparation and materials application. Prog. Polym. Sci. 2018, 79, 98–120. [Google Scholar]
  12. Wei, J.; Liu, M.; Ye, X.; Zhang, S.; Sun, E.; Shan, C.; Wojtas, L.; Shi, X. Facile synthesis of diverse hetero polyaromatic hydrocarbons (PAHs) via the styryl Diels–Alder reaction of conjugated diynes. Org. Chem. Front. 2022, 9, 4301–4308. [Google Scholar]
  13. Day, D.P.; Chan, P.W.H. Gold-Catalyzed Cycloisomerizations of 1,n-Diyne Carbonates and Esters. Adv. Synth. Catal. 2016, 358, 1368–1384. [Google Scholar] [CrossRef]
  14. Shi, W.; Lei, A. 1,3-Diyne chemistry: Synthesis and derivations. Tetrahedron Lett. 2014, 55, 2763–2772. [Google Scholar] [CrossRef] [Green Version]
  15. Ghosh, S.; Kumar Chattopadhyay, S. Transition-Metal-Free Synthesis of Symmetrical 1,4-Diarylsubstituted 1,3-Diynes by Iodine-Mediated Decarboxylative Homocoupling of Arylpropiolic Acids. Tetrahedron Lett. 2022, 102, 153908. [Google Scholar]
  16. Bakhoda, A.; Okoromoba, O.E.; Greene, C.; Boroujeni, M.R.; Bertke, J.A.; Warren, T.H. Three-Coordinate Copper(II) Alkynyl Complex in C–C Bond Formation: The Sesquicentennial of the Glaser Coupling. J. Am. Chem. Soc. 2020, 142, 18483–18490. [Google Scholar] [CrossRef]
  17. Albrecht, F.; Rey, D.; Fatayer, S.; Schulz, F.; Pérez, D.; Peña, D.; Gross, L. Intramolecular Coupling of Terminal Alkynes by Atom Manipulation. Angew. Chem. Int. Ed. 2020, 59, 22989–22993. [Google Scholar] [CrossRef]
  18. Tang, S.; Li, L.; Ren, X.; Li, J.; Yang, G.; Li, H.; Yuan, B. Metallomicelle catalyzed aerobic tandem desilylation/Glaser reaction in water. Green Chem. 2019, 21, 2899–2904. [Google Scholar]
  19. Xu, D.; Sun, Q.; Quan, Z.; Wang, X.; Sun, W. Cobalt-Catalyzed Dimerization and Homocoupling of Terminal Alkynes. Asian J. Org. Chem. 2018, 7, 155–159. [Google Scholar]
  20. Zhang, S.; Liu, X.; Wang, T. An Efficient Copper-Catalyzed Homocoupling of Terminal Alkynes to Give Symmetrical 1,4-Disubstituted 1,3-Diynes. Adv. Synth. Catal. 2011, 353, 1463–1466. [Google Scholar]
  21. Singh, F.V.; Amaral, M.F.Z.J.; Stefani, H.A. Synthesis of symmetrical 1,3-diynes via homocoupling reaction of n-butyl alkynyltellurides. Tetrahedron Lett. 2009, 50, 2636–2639. [Google Scholar] [CrossRef]
  22. Maji, M.S.; Pfeifer, T.; Studer, A. Oxidative Homocoupling of Aryl, Alkenyl, and Alkynyl Grignard Reagents with TEMPO and Dioxygen. Angew. Chem. Int. Ed. 2008, 47, 9547–9550. [Google Scholar] [CrossRef]
  23. Nishihara, Y.; Ikegashira, K.; Hirabayashi, K.; Ando, J.-I.; Mori, A.; Hiyama, T. Coupling Reactions of Alkynylsilanes Mediated by a Cu(I) Salt:  Novel Syntheses of Conjugate Diynes and Disubstituted Ethynes. J. Org. Chem. 2000, 65, 1780–1787. [Google Scholar] [CrossRef]
  24. Mishra, N.; Singh, S.K.; Singh, A.S.; Agrahari, A.K.; Tiwari, V.K. Glycosyl Triazole Ligand for Temperature-Dependent Competitive Reactions of Cu-Catalyzed Sonogashira Coupling and Glaser Coupling. J. Org. Chem. 2021, 86, 17884–17895. [Google Scholar] [CrossRef] [PubMed]
  25. Singh, M.; Singh, A.S.; Mishra, N.; Agrahari, A.K.; Tiwari, V.K. Benzotriazole as an Efficient Ligand in Cu-Catalyzed Glaser Reaction. ACS Omega 2019, 4, 2418–2424. [Google Scholar] [CrossRef] [Green Version]
  26. Zhang, L.-J.; Wang, Y.-H.; Liu, J.; Xu, M.-C.; Zhang, X.-M. Efficient and environmentally friendly Glaser coupling of terminal alkynes catalyzed by multinuclear copper complexes under base-free conditions. RSC Adv. 2016, 6, 28653–28657. [Google Scholar] [CrossRef]
  27. Radhika, S.; Harry, N.A.; Neetha, M.; Anilkumar, G. Recent trends and applications of the Cadiot–Chodkiewicz reaction. Org. Biomol. Chem. 2019, 17, 9081–9094. [Google Scholar] [CrossRef] [PubMed]
  28. Knutson, P.C.; Fredericks, H.E.; Ferreira, E.M. Synthesis of 1,3-Diynes via Cadiot–Chodkiewicz Coupling of Volatile, in Situ Generated Bromoalkynes. Org. Lett. 2018, 20, 6845–6849. [Google Scholar] [CrossRef] [PubMed]
  29. Li, X.; Xie, X.; Sun, N.; Liu, Y. Gold-Catalyzed Cadiot–Chodkiewicz-type Cross-Coupling of Terminal Alkynes with Alkynyl Hypervalent Iodine Reagents: Highly Selective Synthesis of Unsymmetrical 1,3-Diynes. Angew. Chem. Int. Ed. 2017, 56, 6994–6998. [Google Scholar] [CrossRef]
  30. Glaser, C. Beiträge zur Kenntniss des Acetenylbenzols. Ber. Der Dtsch. Chem. Ges. 1869, 2, 422–424. [Google Scholar] [CrossRef] [Green Version]
  31. Kaldhi, D.; Vodnala, N.; Gujjarappa, R.; Kabi, A.K.; Nayak, S.; Malakar, C.C. Transition-metal-free variant of Glaser- and Cadiot-Chodkiewicz-type Coupling: Benign access to diverse 1,3-diynes and related molecules. Tetrahedron Lett. 2020, 61, 151775. [Google Scholar] [CrossRef]
  32. Lampkowski, J.S.; Uthappa, D.M.; Halonski, J.F.; Maza, J.C.; Young, D.D. Application of the Solid-Supported Glaser–Hay Reaction to Natural Product Synthesis. J. Org. Chem. 2016, 81, 12520–12524. [Google Scholar] [CrossRef]
  33. Sindhu, K.S.; Anilkumar, G. Recent advances and applications of Glaser coupling employing greener protocols. RSC Adv. 2014, 4, 27867–27887. [Google Scholar] [CrossRef]
  34. Jia, X.; Yin, K.; Li, C.; Li, J.; Bian, H. Copper-catalyzed oxidative alkyne homocoupling without palladium, ligands and bases. Green Chem. 2011, 13, 2175–2178. [Google Scholar] [CrossRef]
  35. Hsiao, T.-H.; Wu, T.-L.; Chatterjee, S.; Chiu, C.-Y.; Lee, H.M.; Bettucci, L.; Bianchini, C.; Oberhauser, W. Palladium acetate complexes bearing chelating N-heterocyclic carbene (NHC) ligands: Synthesis and catalytic oxidative homocoupling of terminal alkynes. J. Organomet. Chem. 2009, 694, 4014–4024. [Google Scholar] [CrossRef]
  36. Chen, S.-N.; Wu, W.-Y.; Tsai, F.-Y. Homocoupling reaction of terminal alkynes catalyzed by a reusable cationic 2,2′-bipyridyl palladium(II)/CuI system in water. Green Chem. 2009, 11, 269–274. [Google Scholar] [CrossRef]
  37. Han, J.-F.; Guo, P.; Chen, L.; Ye, K.-Y. Cobalt-Catalyzed Glaser-type Homocoupling Reaction. Synthesis 2022, 54, 1989–1995. [Google Scholar] [CrossRef]
  38. Krafft, M.E.; Hirosawa, C.; Dalal, N.; Ramsey, C.; Stiegman, A. Cobalt-catalyzed homocoupling of terminal alkynes: Synthesis of 1,3-diynes. Tetrahedron Lett. 2001, 42, 7733–7736. [Google Scholar] [CrossRef]
  39. Das, U.K.; Jena, R.K.; Bhattacharjee, M. Synthesis, structure and catalytic properties of [Ru(dppp)2(CH3CN)Cl][BPh4] and isolation of catalytically active [Ru(dppp)2Cl][BPh4]: Ruthenium catalysed alkyne homocoupling and tandem alkyne–azide cycloaddition. RSC Adv. 2014, 4, 21964–21970. [Google Scholar] [CrossRef]
  40. Mo, G.; Tian, Z.; Li, J.; Wen, G.; Yang, X. Silver-catalyzed Glaser coupling of alkynes. Appl. Organomet. Chem. 2015, 29, 231–233. [Google Scholar] [CrossRef]
  41. Boronat, M.; Laursen, S.; Leyva-Pérez, A.; Oliver-Meseguer, J.; Combita, D.; Corma, A. Partially oxidized gold nanoparticles: A catalytic base-free system for the aerobic homocoupling of alkynes. J. Catal. 2014, 315, 6–14. [Google Scholar] [CrossRef]
  42. Alonso, F.; Yus, M. Heterogeneous Catalytic Homocoupling of Terminal Alkynes. ACS Catal. 2012, 2, 1441–1451. [Google Scholar] [CrossRef] [Green Version]
  43. Thangarasu, A.K.; Yadhukrishnan, V.O.; Krishnakumar, K.A.; Varma, S.S.; Lankalapalli, R.S. Cu(I)-azidopyrrolo [3,2-d]pyrimidine Catalyzed Glaser–Hay Reaction under Mild Conditions. ACS Org. Inorg. Au 2022, 2, 3–7. [Google Scholar] [CrossRef]
  44. Györke, G.; Dancsó, A.; Volk, B.; Hunyadi, D.; Szalóki, I.; Milen, M. Copper-Containing Mineral Mediated Glaser Coupling of Terminal Alkynes. ChemistrySelect 2022, 7, e202200480. [Google Scholar] [CrossRef]
  45. Wang, B.; Gao, L.; Zheng, G. Leaf-like CuO nanosheets on rGO as an efficient heterogeneous catalyst for Csp-Csp homocoupling of terminal alkynes. Catal. Commun. 2021, 150, 106260. [Google Scholar] [CrossRef]
  46. Chen, X.; Zhang, H.; Chen, J.; Gong, H. A Mild CuCl-catalyzed Glaser-type Homocoupling Reaction: Air Oxidation and Room Temperature. Chem. Lett. 2014, 44, 129–131. [Google Scholar] [CrossRef]
  47. Peris, E. Smart N-Heterocyclic Carbene Ligands in Catalysis. Chem. Rev. 2018, 118, 9988–10031. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, F.-Y.; Lan, X.-B.; Xu, C.; Yao, H.-G.; Li, T.; Liu, F.-S. Rigid hindered N-heterocyclic carbene palladium precatalysts: Synthesis, characterization and catalytic amination. Org. Chem. Front. 2019, 6, 3292–3299. [Google Scholar] [CrossRef]
  49. Lan, X.-B.; Ye, Z.; Liu, J.; Huang, M.; Shao, Y.; Cai, X.; Liu, Y.; Ke, Z. Sustainable and Selective Alkylation of Deactivated Secondary Alcohols to Ketones by Non-bifunctional Pincer N-heterocyclic Carbene Manganese. ChemSusChem 2020, 13, 2557–2563. [Google Scholar] [CrossRef] [PubMed]
  50. Li, D.-H.; Lan, X.-B.; Song, A.X.; Rahman, M.M.; Xu, C.; Huang, F.-D.; Szostak, R.; Szostak, M.; Liu, F.-S. Buchwald-Hartwig Amination of Coordinating Heterocycles Enabled by Large-but-Flexible Pd-BIAN-NHC Catalysts**. Chem. Eur. J. 2022, 28, e202103341. [Google Scholar] [CrossRef]
  51. Chen, C.; Liu, F.-S.; Szostak, M. BIAN-NHC Ligands in Transition-Metal-Catalysis: A Perfect Union of Sterically Encumbered, Electronically Tunable N-Heterocyclic Carbenes? Chem. Eur. J. 2021, 27, 4478–4499. [Google Scholar] [CrossRef]
  52. Ma, B.-B.; Lan, X.-B.; Shen, D.-S.; Liu, F.-S.; Xu, C. Direct C-H bond (Hetero)arylation of thiazole derivatives at 5-position catalyzed by N-heterocyclic carbene palladium complexes at low catalyst loadings under aerobic conditions. J. Organomet. Chem. 2019, 897, 13–22. [Google Scholar] [CrossRef]
  53. Lan, X.-B.; Li, Y.; Li, Y.-F.; Shen, D.-S.; Ke, Z.; Liu, F.-S. Flexible Steric Bulky Bis(Imino)acenaphthene (BIAN)-Supported N-Heterocyclic Carbene Palladium Precatalysts: Catalytic Application in Buchwald–Hartwig Amination in Air. J. Org. Chem. 2017, 82, 2914–2925. [Google Scholar] [CrossRef] [PubMed]
  54. Lan, X.-B.; Chen, F.-M.; Ma, B.-B.; Shen, D.-S.; Liu, F.-S. Pd-PEPPSI Complexes Bearing Bulky [(1,2-Di-(tert-butyl)acenaphthyl] (DtBu-An) on N-Heterocarbene Backbones: Highly Efficient for Suzuki–Miyaura Cross-Coupling under Aerobic Conditions. Organometallics 2016, 35, 3852–3860. [Google Scholar] [CrossRef]
  55. Li, W.; Huang, M.; Liu, J.; Huang, Y.-L.; Lan, X.-B.; Ye, Z.; Zhao, C.; Liu, Y.; Ke, Z. Enhanced Hydride Donation Achieved Molybdenum Catalyzed Direct N-Alkylation of Anilines or Nitroarenes with Alcohols: From Computational Design to Experiment. ACS Catal. 2021, 11, 10377–10382. [Google Scholar] [CrossRef]
  56. Lan, X.-B.; Ye, Z.; Huang, M.; Liu, J.; Liu, Y.; Ke, Z. Nonbifunctional Outer-Sphere Strategy Achieved Highly Active α-Alkylation of Ketones with Alcohols by N-Heterocyclic Carbene Manganese (NHC-Mn). Org. Lett. 2019, 21, 8065–8070. [Google Scholar] [CrossRef]
  57. Huang, M.; Li, Y.; Li, Y.; Liu, J.; Shu, S.; Liu, Y.; Ke, Z. Room temperature N-heterocyclic carbene manganese catalyzed selective N-alkylation of anilines with alcohols. Chem. Commun. 2019, 55, 6213–6216. [Google Scholar] [CrossRef]
  58. Colazzo, L.; Sedona, F.; Moretto, A.; Casarin, M.; Sambi, M. Metal-Free on-Surface Photochemical Homocoupling of Terminal Alkynes. J. Am. Chem. Soc. 2016, 138, 10151–10156. [Google Scholar] [CrossRef] [PubMed]
  59. Ren, P.; Li, Q.; Song, T.; Yang, Y. Facile Fabrication of the Cu-N-C Catalyst with Atomically Dispersed Unsaturated Cu-N2 Active Sites for Highly Efficient and Selective Glaser–Hay Coupling. ACS Appl. Mater. Interfaces 2020, 12, 27210–27218. [Google Scholar] [CrossRef]
Scheme 1. Outline of conjugate 1,3-dialkynes and they synthesis.
Scheme 1. Outline of conjugate 1,3-dialkynes and they synthesis.
Molecules 28 05083 sch001
Scheme 2. Structure of ligands.
Scheme 2. Structure of ligands.
Molecules 28 05083 sch002
Scheme 3. Substrate scope of aromatic terminal alkynes. Reaction conditions: 1 (1.0 mmol), Cu (5 mol%), L2 (10 mol%), DMF (1 mL), 100 °C, under air atmosphere for 4 h. Isolated yields.
Scheme 3. Substrate scope of aromatic terminal alkynes. Reaction conditions: 1 (1.0 mmol), Cu (5 mol%), L2 (10 mol%), DMF (1 mL), 100 °C, under air atmosphere for 4 h. Isolated yields.
Molecules 28 05083 sch003
Scheme 4. Substrate scope of aliphatic and heterocyclic terminal alkynes. Reaction conditions: 3 (1.0 mmol), Cu (5 mol%), L2 (10 mol%), DMF (1 mL), 120 °C, under air atmosphere for 12 h. Isolated yields.
Scheme 4. Substrate scope of aliphatic and heterocyclic terminal alkynes. Reaction conditions: 3 (1.0 mmol), Cu (5 mol%), L2 (10 mol%), DMF (1 mL), 120 °C, under air atmosphere for 12 h. Isolated yields.
Molecules 28 05083 sch004
Scheme 5. Proposed mechanism for the copper-mediated Glaser coupling.
Scheme 5. Proposed mechanism for the copper-mediated Glaser coupling.
Molecules 28 05083 sch005
Table 1. Optimization of Reaction Conditions [a].
Table 1. Optimization of Reaction Conditions [a].
Molecules 28 05083 i001
Entry[Cu]LigandSolventTemperature (°C)Time (h)Yield(%) [b]
1CuClL1DME1201278
2CuClL2DME1201286
3CuClL3DME1201234
4CuClL4DME1201242
5CuClL5DME1201249
6CuClL6DME1201247
7CuClL21,4-dioxane12012trace
8CuClL2DMF1201295
9CuClL2DMAc1201293
10CuClL2toluene1201292
11CuClL2xylene1201288
12CuClL2DMF1001294
13CuClL2DMF801289
14CuClL2DMF100895
15CuClL2DMF100693
16CuClL2DMF100494
17CuClL2DMF80486
18CuClL2DMF60466
19CuClL2DMF40437
20CuCl-----DMF100461
21-----L2DMF10040
22----------DMF10040
23CuClL2DMF60467
24CuCl-----DMF60428
25-----L2DMF6040
[a] Reaction conditions: 1a (1.0 mmol), Cu (5 mol %), ligand (10 mol %), Solvent (1 mL), 100 °C, under air atmosphere for 4 h. [b] Isolated yields.
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

Liu, J.; Zhu, Y.; Luo, J.; Zhu, Z.; Zhao, L.; Zeng, X.; Li, D.; Chen, J.; Lan, X. A Simple and Practical Bis-N-Heterocyclic Carbene as an Efficient Ligand in Cu-Catalyzed Glaser Reaction. Molecules 2023, 28, 5083. https://doi.org/10.3390/molecules28135083

AMA Style

Liu J, Zhu Y, Luo J, Zhu Z, Zhao L, Zeng X, Li D, Chen J, Lan X. A Simple and Practical Bis-N-Heterocyclic Carbene as an Efficient Ligand in Cu-Catalyzed Glaser Reaction. Molecules. 2023; 28(13):5083. https://doi.org/10.3390/molecules28135083

Chicago/Turabian Style

Liu, Jie, Yao Zhu, Jun Luo, Ziyi Zhu, Lin Zhao, Xiaoyan Zeng, Dongdong Li, Jun Chen, and Xiaobing Lan. 2023. "A Simple and Practical Bis-N-Heterocyclic Carbene as an Efficient Ligand in Cu-Catalyzed Glaser Reaction" Molecules 28, no. 13: 5083. https://doi.org/10.3390/molecules28135083

Article Metrics

Back to TopTop