Next Article in Journal
Prominent COF, g-C3N4, and Their Heterojunction Materials for Selective Photocatalytic CO2 Reduction
Previous Article in Journal
Highly Efficient Catalytic Hydrodeoxygenation for Aliphatic Acid to Liquid Alkane: The Role of Molybdenum
Previous Article in Special Issue
Redox Chemistry of Pt(II) Complex with Non-Innocent NHC Bis(Phenolate) Pincer Ligand: Electrochemical, Spectroscopic, and Computational Aspects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CCl4-TMEDA-CuCl—A Novel Convenient Catalytic System for Dimerization of Terminal Acetylenes in Mild Conditions

by
Vasiliy M. Muzalevskiy
1,
Alexey V. Shastin
2,
Sarvinoz I. Tirkasheva
3,
Odiljon E. Ziyadullaev
3,
Askar B. Parmanov
4 and
Valentine G. Nenajdenko
2,*
1
Department of Chemistry, Lomonosov Moscow State University, 119899 Moscow, Russia
2
Institute of Problems of Chemical Physics, Russian Academy of Sciences, 142432 Chernogolovka, Russia
3
Department of Chemistry, Chirchik State Pedagogical University, A. Temur 104, Chirchik 111700, Uzbekistan
4
Deparment of General and Oil-Gas Chemistry, National University of Uzbekistan, University Street 4, Almazar District, Tashkent 100174, Uzbekistan
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(10), 1330; https://doi.org/10.3390/catal13101330
Submission received: 29 June 2023 / Revised: 8 August 2023 / Accepted: 27 September 2023 / Published: 28 September 2023
(This article belongs to the Special Issue Free Radicals in Catalysis, Organic Synthesis, and Material Science)

Abstract

:
A novel catalytic system for homocoupling terminal acetylenes was elaborated based on CuCl as a catalyst (10 mol%), TMEDA as a base and CCl4 as an oxidant. The influence of the solvent, base, amount of catalyst and CCl4 on the reaction was investigated. Methanol was found to be the solvent of choice. The broad synthetic scope of the reaction was demonstrated. Diynes with various substituents were prepared in up to 92% yields. The possible reaction mechanism is discussed.

Graphical Abstract

1. Introduction

Several years ago, a novel catalytic olefination reaction of carbonyl compounds (COR) was discovered in our laboratory [1,2]. It was found that N-unsubstituted hydrazones of carbonyl compounds transform into alkenes under treatment of polyhaloalkanes (PHA) in the presence of a base (ammonia, ethylenediamine, triethylamine) and catalytic amounts of copper salts. The reaction is accompanied by the evolution of nitrogen, and azines are formed as the only by-products (Scheme 1). The reaction has a broad synthetic scope and allows the synthesis of alkenes containing various halogens and functional groups in yields of up to 90–95% [3,4,5,6]. It is worth noting the high stereoselectivity of the reaction; in some cases, the ratio of diastereoisomers reaches 21:1 [7] (Figure 1). The use of freons as olefinating reagents allows the synthesis of fluorine-containing alkenes, convenient building blocks for the synthesis of more complex fluorinated compounds with important practical applications [8,9,10,11,12].
It was found that the PHA-nitrogen base-copper salt system also works for the transformation of a number of other types of hydrazones. In particular, the application of this system made it possible to oxidize isatin hydrazones to the corresponding diazoketones [13]. It was also found that N-substituted hydrazones can be converted into the corresponding halogenated azabutadienes [14] (Figure 1).
Conjugated diynes are under intensive investigations due to their unique properties [15,16,17,18]. They are valuable materials for various synthetic transformations [19,20]. In particular, they have found application in the preparation of natural products [21,22], pharmaceuticals [23], π-conjugated acetylene polymers [24,25], modern construction materials [26,27], heterocyclic compounds [28], electronic and optical materials [29,30]. Conjugated 1,3-diynes also possess biological activity [21], showing antifungal [31], antibacterial [32], anti-inflammatory [33], anti-HIV [34] and anticancer properties [35].
The most efficient method for the synthesis of conjugated diynes is the oxidative dimerization of terminal acetylenes, which is known as the Glaser reaction [15,36,37,38]. The original Glaser reaction used equivalent amounts of copper salts, ammonia (Glaiser [39]) or pyridine (Eglinton [40,41]) as the bases. Later, catalytic variants using various copper salts and ligands were also successfully developed. In particular, Hay proposed the O2-TMEDA-CuCl system, which allows dimerization of terminal acetylenes using catalytic amounts of copper salts [42,43]. This catalytic system has become a very popular system for the dimerization of terminal alkynes [44]. Nevertheless, the search for novel conditions for the Glaser reaction is still an ongoing process. For example, a cooperative behavior of nickel [45,46,47,48], iron [49] and palladium [50,51,52,53] based catalysts has been reported for copper-catalyzed alkyne coupling. Heterogeneous catalysis has also been used for this aim. For example, copper hydroxide on TiO2 [54], cuprous chloride-doped zeolites [55,56,57], silica-supported Cu(II)-hydrazone coordination compounds [58], Cu3(BTC)2 metal organic framework [59] and even naturally occurring copper-containing minerals, chalcocite (Cu2S) and bornite (Cu5FeS4) [60], have demonstrated high efficiency as catalysts for terminal alkyne coupling. The reaction was also performed in supercritical fluids [61,62], water [63,64], ionic liquids [65,66] and solvent-free conditions [67,68] and under photocatalysis [69]. However, no other oxidants instead of O2 were found in the literature. In our opinion, the use of other oxidants and performing the reaction in homogeneous conditions could be very attractive. It should be noted that oxidant-TMEDA-Cu systems are of great interest for the modern methodology of oxidations. Thus, these systems were employed for the synthesis of hydrazine compounds from secondary anilines via the oxidative formation of a N−N bond [70] and for the oxidation of primary alcohols into aldehydes [71,72]. This study is devoted to the investigation of the dimerization of terminal acetylenes using the CCl4-TMEDA-CuCl system.

2. Results

The study of the reaction mechanism of the catalytic olefination reaction showed that the reaction starts with the oxidation of copper (I) chloride with a polyhaloalkane in the presence of a nitrogen-containing ligand to form a complex of copper (II) chloride and a polyhalogenated alkyl radical. The regeneration of copper (I) proceeds under the action of either hydrazone or the addition of products of polyhaloalkyl radicals to hydrazone [73]. Thus, the PHA-nitrogen base-CuCl system works as an oxidizing agent. As a part of the further study of the synthetic possibilities of this system, we decided to study it for the dimerization of terminal acetylenes. TMEDA was chosen as a base and carbon tetrachloride as an oxidant. First, we investigated the influence of the nature of the solvent on the course of the reaction. The reaction was carried out at room temperature using 10 mol% of CuCl. We found that under the action of the CCl4-TMEDA-CuCl system, the model substrate (phenylacetylene) was successfully dimerized to 1,3-diyne 2a (Scheme 1). The reaction proceeds both in polar aprotic solvents and in alcohols. The best reaction yields (90–92%) were achieved in ethanol and methanol. In contrast, the yields in coordinating solvents (MeCN, THF and DMSO) did not exceed 63% due to possible poisoning of the catalytic system. Next, we studied an effect of the amount of catalyst on the reaction outcome. It turned out that the reaction proceeds even with 1 mol% of CuCl; however, in this case, the conversion time of phenylacetylene increases to about 2 days, and the yield of target 1,3-diyne 2a decreases to 51%. The use of a 10 mol% catalyst turned out to be optimal, since in this case the highest yield of 1,3-diyne 2a was observed. It should be noted that the reaction does not take place without CuCl (Scheme 1). We also tested several other nitrogen-containing bases in the reaction. In the case of ethylenediamine, triethylamine and ammonia, the yield dropped to 4% due to massive tarring of the reaction mixture. It was reported that TMEDA (which is the bidentate tertiary amine ligand) provides enhanced solubility to the reactive copper intermediate [44]. We believe that it is a reason for such a dramatic difference in efficiency between TMEDA and other investigated nitrogen-containing bases. Indeed, TMEDA works effectively as both a base and a ligand for the coordination with copper.
Next, we performed a series of experiments using various amounts of CCl4. We found that the reaction proceeds in high yield (76%) even with one equivalent of CCl4. The yield of diyne 2a increases until the amount of PHA increases to 2–2.5 equiv., after which it remains constant. As a proof of principle, we tested several other PHAs. It was found that CFBr3, CF3CCl3, CBrCl3 and CBr4 are also suitable for the dimerization of alkynes to afford diyne 2a in 78–94% yields. However, using CCl4 is preferable because of its very affordable price. Concluding this part of the investigation, we found that the optimal condition for the reaction is to perform it in methanol with TMEDA as a base, using 2.5 equivalents of carbon tetrachloride and 10 mol% of CuCl.
Having found the optimal conditions for the dimerization of phenylacetylene, we carried out a series of reactions with a number of other terminal arylacetylenes. We found that the reaction proceeds in high yield for acetylenes bearing both acceptor and donor substituents at the aromatic ring (Scheme 2). Sterically hindered 2-methoxyphenylacetylene was also successfully involved in the reaction. The corresponding 1,3-diyne 2g was obtained in a 76% yield. We found that this catalytic system also works well for the dimerization of alkylacetylenes. Thus, 1,3-diyne 2i was obtained in a good yield from dec-1-yne. Phenylpropargyl ethers were also successfully involved in the reaction to give the corresponding 1,3-diynes 2j,2k in a high yield (Scheme 2). It should be noted that the reaction is tolerated by a carbonyl function. Thus, 1,3-diyne 2k containing carbonyl groups was isolated in a 92% yield.
To further study the synthetic potential of the reaction, we investigated the dimerization of substituted propargyl alcohols under the action of the CCl4-TMEDA-CuCl system. We found that, in contrast to aryl- and alkylacetylenes, the solvent affects the reaction much more significantly. Thus, in polar aprotic solvents, the dimerization reaction of propargyl alcohol 3a proceeds in low yields, not exceeding 4–6%. On the contrary, in alcohols, the reaction proceeds in good to high yields, and methanol also resulted as the optimal solvent for dimerization, in which diyne 4a was obtained in an 89% yield (Scheme 3). More complex propargyl alcohols have also been successfully involved in the dimerization. Diynes bearing long alkyl chains (4b,4c), derivatives containing cyclic fragments (4d,4e) and aryl substituents (4f) were obtained efficiently. It should be noted that we synthesized diyne 4a in gram scale amounts (1.42 g), while the yield remained the same (85%). However, our attempts to carry out dimerization of parent propargyl alcohol were not successful due to massive tarring of the reaction.
The diacetylene diols obtained are interesting compounds from the point of view of medicinal chemistry. Compounds with a moiety of conjugated polyacetylene alcohols have been isolated from various natural sources. Thus, many polyacetylene derivatives have been isolated from Oplopanax horridus and Panax ginseng plants belonging to the Araliaceae family, which exhibit antitumor, anti-inflammatory, antibacterial, antiviral, antifungal, immunomodulatory, neuroprotective, antidressing, hypoglycemic, hepatoprotective activity, as well as activity associated with obesity control (Figure 2) [74]. Recently, three new polyacetylene alcohols extracted from the sea sponge Siphonochalina Siphonella in Egypt were found to have activity against a human cervical cancer cell line (HeLa), a human breast cancer cell line (MCF-7) and a human lung cancer cell line. (A549) [75,76].
Eventually, we focused our attention on the possible reaction mechanism. It has previously been found that single electron transfer (SET) from CuCl to CCl4 generates Cu(II) species and the CCl3 radical [14]. In order to check this possibility for the dimerization of acetylenes, we performed the reaction in the presence of one equivalent of TEMPO, which is a very effective radial scavenger. We found that TEMPO does not completely block the consumption of phenylacetylene and the formation of diyne 2a. However, noticeable tarring of the reaction was observed to give diyne 2a in a lower yield compared to the reaction without TEMPO. Next, we carried out the reaction in the presence of one equivalent of α-cyclopropylstyrene. It was found that diyne 2a is formed in a slightly lower yield (72% compared to 92%). At the same time, dehydronaphthalene 6 was found in the reaction mixture by 1H NMR. However, no formation of compound 7 was detected. This product could be formed by the addition of phenylacetylenyl radical to α-cyclopropylstyrene. Therefore, the reaction proceeds through the participation of trichloromethyl radical (Scheme 4).
Taking into account these results and the literature data about the Glaser reaction mechanism, the following scheme of the reaction can be proposed (Scheme 5) [36,44,77]. Oxidation of the Cu(I) complex I under the action of CCl4 starts the catalytic cycle to form Cu(II) II complex. Next, complex II reacts with alkyne (or the copper π-complex of alkyne) in the presence of a base (TMEDA) to give intermediate III. The subsequent reaction of III with another molecule of alkyne (or alkyne π-complex) results in the formation of copper complex IV, which is a key reaction intermediate. The reductive elimination of copper from IV provides the target bis-alkyne and Cu(I) complex I to restart the catalytic cycle.
In conclusion, we developed a new catalytic system (CCl4-TMEDA-CuCl) for the oxidative dimerization of terminal acetylenes. It was found that terminal acetylenes can be dimerized in yields up to 92% under the action of carbon tetrachloride in methanol in the presence of TMEDA and catalytic amounts of CuCl. The reaction has a wide synthetic scope, allowing the synthesis of conjugated diynes containing both aromatic substituents and aliphatic substituents. Terminal acetylenes containing functional groups were also successfully involved in the reaction to convert into conjugated diynes in up to 92% yield.

3. Materials and Methods

General remarks. 1H, and 13C NMR spectra (Supplementary Materials) were recorded on a Bruker AVANCE 400 MHz spectrometer in acetone-d6 and CDCl3 at 400.1 and 100.6 MHz, respectively. Chemical shifts (δ) in ppm are reported with the use of the residual CD3COCHD2 and chloroform signals (2.04, 7.25 for 1H and 29.80, 77.0 for 13C) as internal reference. The coupling constants (J) are given in Hertz (Hz). HRMS spectra were measured using the MicroTof Bruker Daltonics instrument. A TLC analysis was performed on “Macherey-Nagel ALUGRAM Xtra SIL G/UV254” plates. Column chromatography was performed on silica gel “Macherey-Nagel 0.063–0.2 nm (Silica 60)”. All reagents were of reagent grade and were used as such or were distilled prior to use. Terminal acetylenes were prepared by literature procedures (1b,c,d,e,h [78]; 3b,c,d,e,f [79]) or purchased from commercial suppliers (1a,f,g,i,j,k; 3a). Melting points were determined on Electrothermal 9100 apparatus. Due to the reported toxicity [80] of CCl4, all manipulations with this reagent should be carried out with care.
Screening of the optimal conditions for the dimerization of phenylacetylene (general procedures). Screening of the optimal solvent. An 8 mL vial with a screw cap was charged with phenylacetylene 1a (1 mmol), the corresponding solvent (3 mL), TMEDA (0.32 mL, 2.2 mmol), CCl4 (0.25 mL, 2.5 mmol) and CuCl (10 mg, 0.1 mmol were added at stirring by a magnetic stirrer). The reaction mixture was stirred for 1 day at room temperature and then broken by 0.1 M of HCl (30 mL). The product was extracted by CH2Cl2 (3 × 10 mL); the organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was purified by column chromatography on silica gel using gradient elution by hexane, followed by a hexane-CH2Cl2 mixture (3:1).
Screening of the optimal amount of CuCl. The procedure for screening the optimal solvent was used with the only difference that the reaction was carried out in methanol and the appropriate amount of CuCl was used instead of 0.1 mmol (10% mmol) of CuCl.
Screening of the optimal base. The procedure for screening of the optimal solvent was used with the only difference that the reaction was carried out in methanol and the corresponding base was used instead of TMEDA.
Screening of the optimal amount of CCl4. The procedure for screening of the optimal solvent was used with the only difference that the reaction was carried out in methanol and the appropriate amount of CCl4 was used instead of 0.25 mL (2.5 mmol) of CCl4.
Screening of the optimal base. The procedure for screening of the optimal solvent was used with the only difference that the reaction was carried out in methanol and the corresponding PHA was used instead of CCl4.
Dimerization of terminal acetylenes 1 (general procedure). An 8 mL vial with a screw cap was charged with corresponding acetylene 1 (1 mmol), MeOH (3 mL), TMEDA (0.32 mL, 2.2 mmol), CCl4 (0.25 mL, 2.5 mmol) and CuCl (10 mg, 0.1 mmol were added at stirring by a magnetic stirrer). The reaction mixture was stirred for 1 day and then broken by 0.1 M of HCl (30 mL). The product was extracted by CH2Cl2 (3 × 10 mL); the organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was purified by passing through a short silica gel pad of silica gel using gradient eluation by hexane, followed by a hexane-CH2Cl2 mixture (3:1) for 2a2j; and by a hexane-CH2Cl2 mixture (3:1) followed by CH2Cl2 for 2k.
1,4-Diphenylbuta-1,3-diyne (2a). Obtained from phenylacetylene 1a (102 mg, 1 mmol). Pale yellow solid, m.p. 86–88 °C, (Lit. [81] 86–88 °C), RF 0.55 (hexane-CH2Cl2, 3:1), yield 93 mg (92%). 1H NMR (CDCl3, 400.1 MHz): δ 7.32–7.43 (m, 3H), 7.55–7.58 (m, 2H). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 73.9, 81.5, 121.7, 128.4, 129.2, 132.4. NMR data are in agreement with those in the literature [82].
1,4-Bis(4-chlorophenyl)buta-1,3-diyne (2b). Obtained from 1-chloro-4-ethynylbenzene 1b (143 mg, 1.048 mmol). White powder, m.p. 259–262 °C, (Lit. [83] 258–259 °C), RF 0.63 (hexane-CH2Cl2, 3:1), yield 106 mg (75%). 1H NMR (CDCl3, 400.1 MHz): δ 7.31 (d, 2H, J = 8.4), 7.44 (d, 2H, J = 8.4). NMR data are in agreement with those in the literature [82].
1,4-Bis(4-bromophenyl)buta-1,3-diyne (2c). Obtained from 1-bromo-4-ethynylbenzene 1c (181 mg, 1 mmol). Pale beige powder, m.p. 262–264 °C, (Lit. [84] 260.1–262.3 °C), RF 0.68 (hexane-CH2Cl2, 3:1), yield 143 mg (79%). 1H NMR (CDCl3, 400.1 MHz): δ 7.37 (d, 2H, J = 8.6), 7.47 (d, 2H, J = 8.6). NMR data are in agreement with those in the literature [85].
1,4-Bis(4-(tert-butyl)phenyl)buta-1,3-diyne (2d). Obtained from 1-(tert-butyl)-4-ethynylbenzene 1d (149 mg, 0.943 mmol). Pale beige powder, m.p. 208–211 °C, (Lit. [84] 209–210 °C), RF 0.8 (hexane-CH2Cl2, 3:1), yield 104 mg (70%). 1H NMR (CDCl3, 400.1 MHz): δ 1.33 (s, 9H), 7.37 (d, 2H, J = 8.4), 7.48 (d, 2H, J = 8.4). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 31.1, 34.9, 73.5, 81.5, 118.8, 125.4, 132.2, 152.5. NMR data are in agreement with those in the literature [84].
1,4-Bis(4-methoxyphenyl)buta-1,3-diyne (2e). Obtained from 1-ethynyl-4-methoxybenzene 1e (131 mg, 0.992 mmol). Pale yellow powder, m.p. 140–142 °C, (Lit. [84] 137.5–139.2 °C), RF 0.2 (hexane-CH2Cl2, 3:1), yield 108 mg (83%). 1H NMR (CDCl3, 400.1 MHz): δ 3.80 (s, 3H), 6.85 (d, 2H, J = 8.9), 7.46 (d, 2H, J = 8.9). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 55.3, 72.9, 81.2, 113.8, 114.1, 134.0, 160.2. NMR data are in agreement with those in the literature [84].
1,4-Bis(3-methoxyphenyl)buta-1,3-diyne (2f). Obtained from 1-ethynyl-3-methoxybenzene 1f (141 mg, 1.068 mmol). White powder, m.p. 92–93 °C, (Lit. [64] 92–93 °C), RF 0.3 (hexane-CH2Cl2, 3:1), yield 105 mg (75%). 1H NMR (CDCl3, 400.1 MHz): δ 3.80 (s, 3H), 6.93 (ddd, 1H, J = 8.0, J = 2.5, J = 1.0), 7.05 (dd, 1H, J = 2.5, J = 1.4), 7.13 (dt, 1H, J = 8.0, J = 1.2), 7.46 (t, 1H, J = 8.0). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 55.2, 73.6, 81.5, 116.0, 117.0, 122.6, 125.0, 129.5, 159.2. NMR data are in agreement with those in the literature [84].
1,4-Bis(2-methoxyphenyl)buta-1,3-diyne (2g). Obtained from 1-ethynyl-2-methoxybenzene 1g (132 mg, 1 mmol). White powder, m.p. 137–139 °C, (Lit. [64] 137–138 °C), RF 0.5 (hexane-CH2Cl2, 1:1), yield 100 mg (76%). 1H NMR (CDCl3, 400.1 MHz): δ 3.88 (s, 3H), 6.86–6.92 (m, 2H), 7.13 (td, 1H, J = 7.9, J = 1.7), 7.47 (dd, 1H, J = 7.6, J = 1.7). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 55.7, 77.9, 78.6, 110.6, 111.2, 120.4, 130.5, 134.3, 161.3. NMR data are in agreement with those in the literature [84].
1,4-Bis(3,4-dimethylphenyl)buta-1,3-diyne (2h). Obtained from 4-ethynyl-1,2-dimethylbenzene 1h (130 mg, 1 mmol). White powder, m.p. 164–166 °C, RF 0.65 (hexane-CH2Cl2, 3:1), yield 96 mg (74%). 1H NMR (CDCl3, 400.1 MHz): δ 2.25 (s, 3H), 2.28 (s, 3H), 7.09 (d, 1H, J = 7.8), 7.28 (dd, 1H, J = 7.7, J = 1.4), 7.31 (s, 1H). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 19.6, 19.9, 73.2, 81.6, 119.0, 129.7, 129.9, 133.4, 136.8, 138.3. NMR data are in agreement with those in the literature [86].
Icosa-9,11-diyne (2i). Obtained from dec-1-yne 1i (137 mg, 0.992 mmol). Pale brown oil, yield 76 mg (56%). 1H NMR (CDCl3, 400.1 MHz): δ 0.87 (t, 6H, J = 6.8), 1.22–1.36 (m, 20H), 1.46–1.54 (m, 4H), 2.23 (t, 4H, J = 7.0). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 14.1, 19.2, 22.6, 28.3, 28.8, 29.05, 29.14, 31.8, 65.2, 77.6. NMR data are in agreement with those in the literature [60].
1,6-Diphenoxyhexa-2,4-diyne (2j). Obtained from (prop-2-yn-1-yloxy)benzene 1i (135 mg, 1.023 mmol). Pale beige powder, m.p. 78–80 °C, (Lit. [87] 77–79 °C), RF 0.2 (hexane-CH2Cl2, 3:1), yield 84 mg (63%). 1H NMR (CDCl3, 400.1 MHz): δ 4.75 (s, 2H), 6.96–6.99 (m, 2H), 7.02–7.05 (m, 1H), 7.31–7.36 (m, 2H). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 56.0, 70.9, 74.6, 114.7, 121.7, 129.5, 157.3. NMR data are in agreement with those in the literature [88].
4,4’-(Hexa-2,4-diyne-1,6-diylbis(oxy))dibenzaldehyde (2k). Obtained from 4-(prop-2-yn-1-yloxy)benzaldehyde 1k (159 mg, 0.994 mmol). Pale beige powder, m.p. 154–158 °C, yield 145 mg (92%). 1H NMR (CDCl3, 400.1 MHz): δ 4.83 (s, 4H), 7.03 (d, 4H, J = 8.8), 7.84 (d, 4H, J = 8.8), 9.89 (s, 2H). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 56.2, 71.4, 73.9, 115.1, 130.7, 131.9, 162.0, 190.7. NMR data are in agreement with those in the literature [87].
Dimerization of terminal acetylenes 3 (general procedure). An 8 mL vial with a screw cap was charged with corresponding acetylene 3 (1 mmol), MeOH (3 mL), TMEDA (0.32 mL, 2.2 mmol), CCl4 (0.25 mL, 2.5 mmol) and CuCl (10 mg, 0.1 mmol were added at stirring by a magnetic stirrer). The reaction mixture was stirred for 1 day and then broken by 0.1 M of HCl (30 mL). The product was extracted by CH2Cl2 (3 × 10 mL); the organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was purified by passing through a short silica gel pad of silica gel using gradient eluation by CH2Cl2, followed by CH2Cl2-MeOH mixture (30:1).
2,7-Dimethylocta-3,5-diyne-2,7-diol (4a). Obtained from 2-methylbut-3-yn-2-ol 3a (84 mg, 1 mmol). White powder, m.p. 124–126 °C, yield 74 mg (89%). 1H NMR (acetone-d6, 400.1 MHz): δ 1.43 (s, 12H, 4CH3), 4.59 (s, 2H, 2OH). 13C{1H} NMR (acetone-d6, 100.6 MHz): δ 31.4, 65.0, 66.0, 85.5. NMR data are in agreement with those in the literature [45].
5,10-Diethyltetradeca-6,8-diyne-5,10-diol (4b). Obtained from 3-ethylhept-1-yn-3-ol 3b (143 mg, 1.021 mmol). Yellow-brown oil, yield 77 mg (54%). 1H NMR (CDCl3, 400.1 MHz): 0.91 (t, 6H, 2CH3, J = 7.3), 1.02 (t, 6H, 2CH3, J = 7.3), 1.29–1.38 (m, 4H, 2CH2), 1.41–1.49 (m, 4H, 2CH2), 1.57–1.73 (m, 8H, 4CH2), 2.01 (s, 2H, 2OH). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 8.5, 14.0, 22.8, 26.3, 34.7, 41.1, 68.4, 72.2, 82.0.
8,13-Dimethylicosa-9,11-diyne-8,13-diol (4c). Obtained from 3-methyldec-1-yn-3-ol 3c (169 mg, 1.004 mmol). Pale brown oil, yield 138 mg (82%). 1H NMR (CDCl3, 400.1 MHz): δ 0.87 (t, 6H, 2CH3, J = 6.9), 1.21–1.34 (m, 16H, 8CH2), 1.40–1.52 (m, 10H, 2CH2, 2CH3), 1.59–1.71 (m, 4H, 2CH2), 2.02 (s, 2H, 2OH). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 14.1, 22.6, 24.6, 29.2, 29.5, 29.6, 31.8, 43.5, 67.4, 68.7, 83.2. HRMS (ESI-TOF): m/z [M-OH]+ Calcd for C22H37O+: 317.2839; found: 317.2848.
1,1’-(Buta-1,3-diyne-1,4-diyl)bis(cyclohexan-1-ol) (4d). Obtained from 1-ethynylcyclohexan-1-ol 3d (124 mg, 1 mmol). Colorless oil, yield 88 mg (72%). 1H NMR (CDCl3, 400.1 MHz): δ 1.15–1.30 (m, 2H, CH2), 1.44–1.75 (m, 14H, 7CH2), 1.83–1.95 (m, 4H, 2CH2), 2.23 (s, 2H, 2OH). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 23.1, 25.0, 39.6, 68.7, 69.1, 83.0. NMR data are in agreement with those in the literature [89].
2,2’-(Buta-1,3-diyne-1,4-diyl)bis(adamantan-2-ol) (4e). Obtained from 2-ethynyladamantan-2-ol 3e (175 mg, 0.994 mmol). White solid, m.p. 240–242 °C, yield 124 mg (71%). 1H NMR (acetone-d6, 400.1 MHz): δ 1.48–1.53 (m, 4H), 1.70–1.80 (m, 12H), 1.87–1.92 (m, 4H), 2.03–2.08 (m, 4H, 3CH), 2.16–2.23 (m, 4H), 2.97 (s, 2H, 2OH). 13C{1H} NMR (acetone-d6, 100.6 MHz): δ 27.57, 27.64, 32.0, 35.8, 38.1, 39.3, 69.0, 72.6, 85.2. HRMS (ESI-TOF): m/z [M-OH]+ Calcd for C24H29O+: 333.2213; found: 333.2223.
2,7-Bis(4-chlorophenyl)octa-3,5-diyne-2,7-diol (4f). Obtained from 2-(4-chlorophenyl)but-3-yn-2-ol 3f (181 mg, 1 mmol). Yellow-brown oil, yield 115 mg (64%). 1H NMR (CDCl3, 400.1 MHz): δ 1.77 (s, 6H, 2CH3), 2.62 (s, 2H, 2OH), 7.32 (d, 4H, 4CH, J = 8.6), 7.53 (d, 4CH, J = 8.6). 13C{1H} NMR (CDCl3, 100.6 MHz): δ 32.9, 69.0, 70.0, 82.9, 126.3, 128.6, 133.9, 142.9. NMR data are in agreement with those in the literature [90].
Dimerization of terminal acetylene 3a in gram scale. A 50 mL round-bottomed flask was charged with 2-methylbut-3-yn-2-ol 3a (1684 mg, 20.05 mmol), MeOH (60 mL), TMEDA (6.4 mL, 44 mmol), CCl4 (5 mL, 50 mmol) and CuCl (20 mg, 2 mmol were added at stirring by a magnetic stirrer). The reaction mixture was stirred for 1 day; volatiles were evaporated in vacuo. The residue was dispersed between 0.1 M of HCl (30 mL) and CH2Cl2 (20 mL). The organic layer was separated; the water phase was extracted by CH2Cl2 (20 mL). The combined organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was by passing through a short silica gel pad of silica gel using gradient eluation by CH2Cl2, followed by CH2Cl2-MeOH mixture (30:1). White powder, m.p. 124–126 °C, yield 1420 mg (85%). 1420 mg (85%). For NMR spectral data see above.
Dimerization of phenylacetylene 1a in the presence of TEMPO. An 8 mL vial with a screw cap was charged with phenylacetylene 1a (57 mg, 0.56 mmol), MeOH (1.5 mL), TMEDA (0.16 mL, 1.1 mmol), CCl4 (79 mg, 0.51 mmol), TEMPO (77 mg, 0.49 mmol) and CuCl (5.2 mg, 0.052 mmol), which were added at stirring by a magnetic stirrer. The reaction mixture was stirred for 1 day and then broken by 0.1 M of HCl (30 mL). The product was extracted by CH2Cl2 (3 × 10 mL); the organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was purified by passing through a short silica gel pad of silica gel using gradient eluation by hexane, followed by hexane-CH2Cl2 mixture (3:1). White powder, yield 16 mg (28%). For NMR spectral data see above.
Dimerization of phenylacetylene 1a in the presence of α-cyclopropylstyrene. An 8 mL vial with a screw cap was charged with phenylacetylene 1a (107 mg, 1.05 mmol), MeOH (3 mL), TMEDA (0.32 mL, 2.2 mmol), CCl4 (174 mg, 1.13 mmol), α-cyclopropylstyrene (155 mg, 1.07 mmol) and CuCl (9 mg, 0.09 mmol), which were added at stirring by a magnetic stirrer. The reaction mixture was stirred for 1 day and then broken by 0.1 M of HCl (30 mL). The product was extracted by CH2Cl2 (3 × 10 mL); the organic phase was washed with water (10 mL) and dried over Na2SO4. Volatiles were evaporated in vacuo; the residue formed was purified by passing through a short silica gel pad of silica gel using gradient eluation by hexane, followed by hexane-CH2Cl2 mixture (3:1). White powder, yield 76 mg (72%). For NMR spectral data see above. Compound 6 was observed in the NMR spectra of crude product 2a. The yield of 6 was calculated by comparison with the amount of 2a in the 1H NMR of crude 2a. The NMR data of 6 are in agreement with those in the literature [14].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13101330/s1, Copies of all NMR spectra.

Author Contributions

Conceptualization, V.M.M. and A.V.S.; methodology, V.M.M. and A.V.S.; validation, V.M.M., investigation, V.M.M., A.V.S. and S.I.T.; writing—original draft preparation, V.M.M.; writing—review and editing, V.M.M., S.I.T., O.E.Z., A.B.P. and V.G.N.; visualization, V.M.M.; supervision, V.M.M.; project administration, V.M.M. and V.G.N.; funding acquisition, V.G.N. All authors have read and agreed to the published version of the manuscript.

Funding

The research was conducted in terms of the state contract of the chair of organic chemistry of Moscow State University, entitled “Synthesis and study of physical, chemical and biological properties of organic and organometallic compounds”—CITIC No—AAAA-A21-121012290046-4.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Acknowledgments

The authors acknowledge partial support from M. V. Lomonosov Moscow State University Program of Development.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shastin, A.V.; Korotchenko, V.N.; Nenaidenko, V.G.; Balenkova, E.S. A Novel Reaction of Double Carbon-Carbon Bond Formation: Synthesis of 2,2-Dichlorostyrenes. Russ. Chem. Bull. 1999, 48, 2184–2185. [Google Scholar] [CrossRef]
  2. Shastin, A.V.; Korotchenko, V.N.; Nenajdenko, V.G.; Balenkova, E.S. A Novel Synthetic Approach to Dichlorostyrenes. Tetrahedron 2000, 56, 6557–6563. [Google Scholar] [CrossRef]
  3. Nenajdenko, V.G.; Korotchenko, V.N.; Shastin, A.V.; Balenkova, E.S. Catalytic Olefination of Carbonyl Compounds. A New Versatile Method for the Synthesis of Alkenes. Russ. Chem. Bull. 2004, 53, 1034–1064. [Google Scholar] [CrossRef]
  4. Shixaliyev, N.Q.; Gurbanov, A.V.; Maharramov, A.M.; Mahmudov, K.T.; Kopylovich, M.N.; Martins, L.M.D.R.S.; Muzalevskiy, V.M.; Nenajdenko, V.G.; Pombeiro, A.J.L. Halogen-Bonded Tris(2,4-Bis(Trichloromethyl)-1,3,5-Triazapentadienato)-M(iii) [M = Mn, Fe, Co] Complexes and Their Catalytic Activity in the Peroxidative Oxidation of 1-Phenylethanol to Acetophenone. New J. Chem. 2014, 38, 4807–4815. [Google Scholar] [CrossRef]
  5. Shastin, A.V.; Korotchenko, V.N.; Nenajdenko, V.G.; Balenkova, E.S. A Novel Synthesis of β,β-Dibromostyrenes. Synthesis 2001, 2001, 2081–2084. [Google Scholar] [CrossRef]
  6. Motornov, V.A.; Muzalevskiy, V.M.; Tabolin, A.A.; Novikov, R.A.; Nelyubina, Y.V.; Nenajdenko, V.G.; Ioffe, S.L. Radical Nitration-Debromination of α-Bromo-α-Fluoroalkenes as a Stereoselective Route to Aromatic α-Fluoronitroalkenes—Functionalized Fluorinated Building Blocks for Organic Synthesis. J. Org. Chem. 2017, 82, 5274–5284. [Google Scholar] [CrossRef] [PubMed]
  7. Shastin, A.V.; Muzalevsky, V.M.; Balenkova, E.S.; Nenajdenko, V.G. Stereoselective Synthesis of 1-Bromo-1-Fluorostyrenes. Mendeleev Commun. 2006, 16, 179–180. [Google Scholar] [CrossRef]
  8. Goldberg, A.A.; Muzalevskiy, V.M.; Shastin, A.V.; Balenkova, E.S.; Nenajdenko, V.G. Novel Efficient Synthesis of β-Fluoro-β-(Trifluoromethyl)Styrenes. J. Fluor. Chem. 2010, 131, 384–388. [Google Scholar] [CrossRef]
  9. Nenajdenko, V.G.; Muzalevskiy, V.M.; Shastin, A.V.; Balenkova, E.S.; Kondrashov, E.V.; Ushakov, I.A.; Rulev, A.Y. Fragmentation of Trifluoromethylated Alkenes and Acetylenes by N,N-Binucleophiles. Synthesis of Imidazolines or Imidazolidines (Oxazolidines) Controlled by Substituent. J. Org. Chem. 2010, 75, 5679–5688. [Google Scholar] [CrossRef]
  10. Nenajdenko, V.G.; Shastin, A.V.; Korotchenko, V.N.; Varseev, G.N.; Balenkova, E.S. A Novel Approach to 2-Chloro-2-Fluorostyrenes. Eur. J. Org. Chem. 2003, 2003, 302–308. [Google Scholar] [CrossRef]
  11. Muzalevskiy, V.M. Synthesis of Heterocyclic Compounds Using the Nenajdenko-Shastin Reaction. Chem. Heterocycl. Compd. 2012, 48, 117–125. [Google Scholar] [CrossRef]
  12. Nenajdenko, V.G.; Muzalevskiy, V.M.; Shastin, A.V. Polyfluorinated Ethanes as Versatile Fluorinated C2-Building Blocks for Organic Synthesis. Chem. Rev. 2015, 115, 973–1050. [Google Scholar] [CrossRef]
  13. Muzalevskiy, V.M.; Balenkova, E.S.; Shastin, A.V.; Magerramov, A.M.; Shikhaliev, N.G.; Nenajdenko, V.G. New Method for the Preparation of 3-Diazo-1,3-Dihydroindol-2-Ones. Russ. Chem. Bull. 2011, 60, 2343–2346. [Google Scholar] [CrossRef]
  14. Nenajdenko, V.G.; Shastin, A.V.; Gorbachev, V.M.; Shorunov, S.V.; Muzalevskiy, V.M.; Lukianova, A.I.; Dorovatovskii, P.V.; Khrustalev, V.N. Copper-Catalyzed Transformation of Hydrazones into Halogenated Azabutadienes, Versatile Building Blocks for Organic Synthesis. ACS Catal. 2017, 7, 205–209. [Google Scholar] [CrossRef]
  15. Akhtar, R.; Zahoor, A.F. Transition Metal Catalyzed Glaser and Glaser-Hay Coupling Reactions: Scope, Classical/Green Methodologies and Synthetic Applications. Synth. Commun. 2020, 50, 3337–3368. [Google Scholar] [CrossRef]
  16. Zhang, G.; Wu, K.; Wen, C.; Li, Q. Nickel-Catalyzed Cross-Coupling of Organoaluminum Reagents with Alkynylhalides for the Synthesis of Symmetrical and Unsymmetrical Conjugated 1,3-Diynes Derivatives. J. Organomet. Chem. 2020, 906, 121040. [Google Scholar] [CrossRef]
  17. Matsuda, Y.; Naoe, S.; Oishi, S.; Fujii, N.; Ohno, H. Formal [4+2] Reaction between 1,3-Diynes and Pyrroles: Gold(I)-Catalyzed Indole Synthesis by Double Hydroarylation. Chem.—Eur. J. 2015, 21, 1463–1467. [Google Scholar] [CrossRef] [PubMed]
  18. Zhu, Y.; Shi, Y. A Facile Copper(i)-Catalyzed Homocoupling of Terminal Alkynes to 1,3-Diynes with Diaziridinone under Mild Conditions. Org. Biomol. Chem. 2013, 11, 7451. [Google Scholar] [CrossRef] [PubMed]
  19. Singha, R.; Nandi, S.; Ray, J.K. Bromine-Mediated Cyclization of 1,4-Diaryl Buta-1,3-Diyne to 1,2,3-Tribromo-4-Aryl Naphthalene. Tetrahedron Lett. 2012, 53, 6531–6534. [Google Scholar] [CrossRef]
  20. Tang, J.; Zhao, X. Synthesis of 2,5-Disubstituted Thiophenes via Metal-Free Sulfur Heterocyclization of 1,3-Diynes with Sodium Hydrosulfide. RSC Adv. 2012, 2, 5488. [Google Scholar] [CrossRef]
  21. Shi Shun, A.L.K.; Tykwinski, R.R. Synthesis of Naturally Occurring Polyynes. Angew. Chem. Int. Ed. 2006, 45, 1034–1057. [Google Scholar] [CrossRef] [PubMed]
  22. Yun, H.; Chou, T.-C.; Dong, H.; Tian, Y.; Li, Y.; Danishefsky, S.J. Total Synthesis as a Resource in Drug Discovery: The First In Vivo Evaluation of Panaxytriol and Its Derivatives. J. Org. Chem. 2005, 70, 10375–10380. [Google Scholar] [CrossRef] [PubMed]
  23. Nicolaou, K.C.; Bulger, P.G.; Sarlah, D. Metathesis Reactions in Total Synthesis. Angew. Chem. Int. Ed. 2005, 44, 4490–4527. [Google Scholar] [CrossRef] [PubMed]
  24. Martin, R.E.; Diederich, F. Linear Monodisperse π-Conjugated Oligomers: Model Compounds for Polymers and More. Angew. Chem. Int. Ed. 1999, 38, 1350–1377. [Google Scholar] [CrossRef]
  25. Tour, J.M. Conjugated Macromolecules of Precise Length and Constitution. Organic Synthesis for the Construction of Nanoarchitectures. Chem. Rev. 1996, 96, 537–554. [Google Scholar] [CrossRef] [PubMed]
  26. Diederich, F.; Stang, P.J.; Tykwinski, R.R. (Eds.) Acetylene Chemistry: Chemistry, Biology and Material Science; Wiley-VCH GmbH KGaA: Weinheim, Germany, 2005. [Google Scholar]
  27. Gholami, M.; Tykwinski, R.R. Oligomeric and Polymeric Systems with a Cross-Conjugated π-Framework. Chem. Rev. 2006, 106, 4997–5027. [Google Scholar] [CrossRef] [PubMed]
  28. Nenajdenko, V.G.; Sumerin, V.V.; Chernichenko, K.Y.; Balenkova, E.S. A New Route to Annulated Oligothiophenes. Org. Lett. 2004, 6, 3437–3439. [Google Scholar] [CrossRef]
  29. Marsden, J.A.; Haley, M.M. Carbon Networks Based on Dehydrobenzoannulenes. 5. Extension of Two-Dimensional Conjugation in Graphdiyne Nanoarchitectures. J. Org. Chem. 2005, 70, 10213–10226. [Google Scholar] [CrossRef]
  30. Eisler, S.; Slepkov, A.D.; Elliott, E.; Luu, T.; McDonald, R.; Hegmann, F.A.; Tykwinski, R.R. Polyynes as a Model for Carbyne: Synthesis, Physical Properties, and Nonlinear Optical Response. J. Am. Chem. Soc. 2005, 127, 2666–2676. [Google Scholar] [CrossRef]
  31. Stütz, A. Allylamine Derivatives—A New Class of Active Substances in Antifungal Chemotherapy. Angew. Chem. Int. Ed. Engl. 1987, 26, 320–328. [Google Scholar] [CrossRef]
  32. Lechner, D.; Stavri, M.; Oluwatuyi, M.; Pereda-Miranda, R.; Gibbons, S. The Anti-Staphylococcal Activity of Angelica Dahurica (Bai Zhi). Phytochemistry 2004, 65, 331–335. [Google Scholar] [CrossRef]
  33. Schmidt, R.; Thorwirth, R.; Szuppa, T.; Stolle, A.; Ondruschka, B.; Hopf, H. Fast, Ligand- and Solvent-Free Synthesis of 1,4-Substituted Buta-1,3-Diynes by Cu-Catalyzed Homocoupling of Terminal Alkynes in a Ball Mill. Chem.—Eur. J. 2011, 17, 8129–8138. [Google Scholar] [CrossRef]
  34. Lerch, M.L.; Harper, M.K.; Faulkner, D.J. Brominated Polyacetylenes from the Philippines Sponge Diplastrella sp. J. Nat. Prod. 2003, 66, 667–670. [Google Scholar] [CrossRef] [PubMed]
  35. Mayer, S.F.; Steinreiber, A.; Orru, R.V.A.; Faber, K. Chemoenzymatic Asymmetric Total Syntheses of Antitumor Agents (3R,9R,10R)- and (3S,9R,10R)-Panaxytriol and (R)- and (S)-Falcarinol from Panax ginseng Using an Enantioconvergent Enzyme-Triggered Cascade Reaction. J. Org. Chem. 2002, 67, 9115–9121. [Google Scholar] [CrossRef] [PubMed]
  36. Li, J.J. Glaser Coupling. In Name Reactions; Springer International Publishing: Cham, Switzerland, 2014; pp. 282–286. ISBN 978-3-319-03978-7. [Google Scholar]
  37. Sindhu, K.S.; Anilkumar, G. Recent Advances and Applications of Glaser Coupling Employing Greener Protocols. RSC Adv. 2014, 4, 27867–27887. [Google Scholar] [CrossRef]
  38. Jover, J.; Spuhler, P.; Zhao, L.; McArdle, C.; Maseras, F. Toward a Mechanistic Understanding of Oxidative Homocoupling: The Glaser–Hay Reaction. Catal. Sci. Technol. 2014, 4, 4200–4209. [Google Scholar] [CrossRef]
  39. Glaser, C. Beiträge Zur Kenntniss Des Acetenylbenzols. Berichte Dtsch. Chem. Ges. 1869, 2, 422–424. [Google Scholar] [CrossRef]
  40. Eglinton, G.; Galbraith, A.R. Cyclic dyines. Chem. Ind. 1956, 28, 736–737. [Google Scholar]
  41. Eglinton, G.; Galbraith, A.R. Macrocyclic Acetylenic Compounds. Part I. Cyclotetradeca-1:3-Diyne and Related Compounds. J. Chem. Soc. Resumed 1959, 889–896. [Google Scholar] [CrossRef]
  42. Hay, A. Communications- Oxidative Coupling of Acetylenes. J. Org. Chem. 1960, 25, 1275–1276. [Google Scholar] [CrossRef]
  43. Hay, A.S. Oxidative Coupling of Acetylenes. II. J. Org. Chem. 1962, 27, 3320–3321. [Google Scholar] [CrossRef]
  44. Allen, S.E.; Walvoord, R.R.; Padilla-Salinas, R.; Kozlowski, M.C. Aerobic Copper-Catalyzed Organic Reactions. Chem. Rev. 2013, 113, 6234–6458. [Google Scholar] [CrossRef]
  45. Yin, W.; He, C.; Chen, M.; Zhang, H.; Lei, A. Nickel-Catalyzed Oxidative Coupling Reactions of Two Different Terminal Alkynes Using O2 as the Oxidant at Room Temperature: Facile Syntheses of Unsymmetric 1,3-Diynes. Org. Lett. 2009, 11, 709–712. [Google Scholar] [CrossRef]
  46. Bedard, A.-C.; Collins, S.K. Phase Separation As a Strategy Toward Controlling Dilution Effects in Macrocyclic Glaser-Hay Couplings. J. Am. Chem. Soc. 2011, 133, 19976–19981. [Google Scholar] [CrossRef]
  47. Crowley, J.D.; Goldup, S.M.; Gowans, N.D.; Leigh, D.A.; Ronaldson, V.E.; Slawin, A.M.Z. An Unusual Nickel−Copper-Mediated Alkyne Homocoupling Reaction for the Active-Template Synthesis of [2]Rotaxanes. J. Am. Chem. Soc. 2010, 132, 6243–6248. [Google Scholar] [CrossRef] [PubMed]
  48. Muesmann, T.W.T.; Wickleder, M.S.; Christoffers, J. Preparation of linear aromatic disulfonic acids: New linker molecules for metal-organic frameworks. Synthesis 2011, 17, 2775–2780. [Google Scholar] [CrossRef]
  49. Meng, X.; Li, C.; Han, B.; Wang, T.; Chen, B. Iron/copper promoted oxidative homo-coupling reaction of terminal alkynes using air as the oxidant. Tetrahedron 2010, 66, 4029–4031. [Google Scholar] [CrossRef]
  50. Li, J.-H.; Liang, Y.; Xie, Y.-X. Efficient Palladium-Catalyzed Homocoupling Reaction and Sonogashira Cross-Coupling Reaction of Terminal Alkynes under Aerobic Conditions. J. Org. Chem. 2005, 70, 4393–4396. [Google Scholar] [CrossRef] [PubMed]
  51. Batsanov, A.S.; Collings, J.C.; Fairlamb, I.J.S.; Holland, J.P.; Howard, J.A.K.; Lin, Z.; Marder, T.B.; Parsons, A.C.; Ward, R.M.; Zhu, J. Requirement for an Oxidant in Pd/Cu Co-Catalyzed Terminal Alkyne Homocoupling To Give Symmetrical 1,4-Disubstituted 1,3-Diynes. J. Org. Chem. 2005, 70, 703–706. [Google Scholar] [CrossRef] [PubMed]
  52. Merkul, E.; Urselmann, D.; Müller, T.J.J. Consecutive One-Pot Sonogashira–Glaser Coupling Sequence—Direct Preparation of Symmetrical Diynes by Sequential Pd/Cu Catalysis. Eur. J. Org. Chem. 2011, 2011, 238–242. [Google Scholar] [CrossRef]
  53. Hoshi, M.; Okimoto, M.; Nakamura, S.; Takahashi, S. Homo- and Heterocoupling of Terminal Conjugated Enynes: One-Pot Synthesis of Alka-1,7-diene-3,5-diynes and Alk-1-ene-3,5-diynes via Two Types of Coupling Reaction. Synthesis 2011, 23, 3839–3847. [Google Scholar] [CrossRef]
  54. Oishi, T.; Katayama, T.; Yamaguchi, K.; Mizuno, N. Heterogeneously Catalyzed Efficient Alkyne–Alkyne Homocoupling by Supported Copper Hydroxide on Titanium Oxide. Chem. Eur. J. 2009, 15, 7539–7542. [Google Scholar] [CrossRef] [PubMed]
  55. Kuhn, P.; Alix, A.; Kumarraja, M.; Louis, B.; Pale, P.; Sommer, J. Copper–Zeolites as Catalysts for the Coupling of Terminal Alkynes: An Efficient Synthesis of Diynes. Eur. J. Org. Chem. 2009, 2009, 423–429. [Google Scholar] [CrossRef]
  56. Kuhn, P.; Pale, P.; Sommer, J.; Louis, B. Probing Cu-USY Zeolite Reactivity: Design of a Green Catalyst for the Synthesis of Diynes. J. Phys. Chem. C 2009, 113, 2903–2910. [Google Scholar] [CrossRef]
  57. Chassaing, S.; Alix, A.; Boningari, T.; Sido, K.S.S.; Keller, M.; Kuhn, P.; Louis, B.; Sommer, J.; Pale, P. Copper(I)-Zeolites as New Heterogeneous and Green Catalysts for Organic Synthesis. Synthesis 2010, 2010, 1557–1567. [Google Scholar] [CrossRef]
  58. Heydari, N.; Bikas, R.; Siczek, M.; Lis, T. Green carbon–carbon homocoupling of terminal alkynes by a silica supported Cu(ii)-hydrazone coordination compound. Dalton Trans. 2023, 52, 421–433. [Google Scholar] [CrossRef] [PubMed]
  59. Devarajan, N.; Karthik, M.; Suresh, P. Copper catalyzed oxidative homocoupling of terminal alkynes to 1,3-diynes: A Cu3(BTC)2 MOF as an efficient and ligand free catalyst for Glaser–Hay coupling. Org. Biomol. Chem. 2017, 15, 9191–9199. [Google Scholar] [CrossRef]
  60. 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]
  61. Li, J.H.; Jiang, H.F. Glaser coupling reaction in supercritical carbon dioxide. Chem. Commun. 1999, 2369–2370. [Google Scholar] [CrossRef]
  62. Jiang, H.F.; Tang, J.Y.; Wang, A.Z.; Deng, G.H.; Yang, S.R. Cu(II)-Promoted Oxidative Homocoupling Reaction of Terminal Alkynes in Supercritical Carbon Dioxide. Synthesis 2006, 2006, 1155–1161. [Google Scholar] [CrossRef]
  63. Zhou, L.; Zhan, H.Y.; Liu, H.L.; Jiang, H.F. An Efficient and Practical Process for Pd/Cu CocatalyzedHomocoupling Reaction of Terminal Alkynes Using Sodium Percarbonate as a Dual Reagent in Aqueous Media. Chin. J. Chem. 2007, 25, 1413–1416. [Google Scholar] [CrossRef]
  64. 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]
  65. Yadav, J.S.; Reddy, B.V.S.; Reddy, B.K.; Uma, K.; Prasad, G.A.R. Glaser oxidative coupling in ionic liquids: An improved synthesis of conjugated 1,3-diynes. Tetrahedron Lett. 2003, 44, 6493–6496. [Google Scholar] [CrossRef]
  66. Ranu, B.C.; Banerjee, S. Homocoupling of terminal alkynes to 1,4-disubstituted 1,3-diynes promoted by copper(I) iodide and a task specific ionic liquid, [bmim]OH -A green procedure. Lett. Org. Chem. 2006, 3, 607–609. [Google Scholar] [CrossRef]
  67. Kabalka, G.W.; Wang, L.; Pagni, R.M. Microwave Enhanced Glaser Coupling Under Solvent Free Conditions. Synlett 2001, 2001, 108–110. [Google Scholar] [CrossRef]
  68. Stolle, A.; Ondruschka, B. Solvent-free reactions of alkynes in ball mills: It is definitely more than mixing. Pure Appl. Chem. 2011, 83, 1343–1349. [Google Scholar] [CrossRef]
  69. Bag, S.S.; Sinha, S.; Singh, S.; Golder, A.K. Greener photocatalytic route to the hetero-selective Glaser coupling reaction: Role of hole/oxygen in air. Catal. Sci. Technol. 2023, 13, 1281–1287. [Google Scholar] [CrossRef]
  70. Yan, X.-M.; Chen, Z.-M.; Yang, F.; Huang, Z.-Z. A Dehydrogenative Homocoupling Reaction for the Direct Synthesis of Hydrazines from N-Alkylanilines in Air. Synlett 2011, 2011, 569–572. [Google Scholar] [CrossRef]
  71. Cicco, L.; Roggio, M.; López-Aguilar, M.; Ramos-Martín, M.; Perna, F.M.; García-Álvarez, J.; Vitale, P.; Capriati, V. Selective Aerobic Oxidation of Alcohols in Low Melting Mixtures and Water and Use for Telescoped One-Pot Hybrid Reactions. ChemistryOpen 2022, 11, e202200160. [Google Scholar] [CrossRef]
  72. Silva, E.D.; Alves, O.A.L.; Ribeiro, R.T.; Chagas, R.C.R.; Villar, J.A.F.P.; Princival, J.L. Homogeneous CuCl2/TMEDA/TEMPO-Catalyzed chemoselective base- and halogen- free aerobic oxidation of primary alcohols in mild conditions. Appl. Cat. A-Gen. 2021, 623, 118289. [Google Scholar] [CrossRef]
  73. Shastin, A.V.; Muzalevskii, V.M.; Korotchenko, V.N.; Balenkova, E.S.; Nenaidenko, V.G. Effect of the Catalyst Nature and Quantity on Catalytic Olefination. Russ. J. Org. Chem. 2006, 42, 183–189. [Google Scholar] [CrossRef]
  74. Chen, Y.; Peng, S.; Luo, Q.; Zhang, J.; Guo, Q.; Zhang, Y.; Chai, X. Chemical and Pharmacological Progress on Polyacetylenes Isolated from the Family Apiaceae. Chem. Biodivers. 2015, 12, 474–502. [Google Scholar] [CrossRef]
  75. Ki, D.-W.; El-Desoky, A.H.; Wong, C.P.; Abdel-Ghani, M.; El-Beih, A.A.; Mizuguchi, M.; Morita, H. New Cytotoxic Polyacetylene Alcohols from the Egyptian Marine Sponge Siphonochalina Siphonella. J. Nat. Med. 2020, 74, 409–414. [Google Scholar] [CrossRef] [PubMed]
  76. Xie, Q.; Wang, C. Polyacetylenes in Herbal Medicine: A Comprehensive Review of Its Occurrence, Pharmacology, Toxicology, and Pharmacokinetics (2014–2021). Phytochemistry 2022, 201, 113288. [Google Scholar] [CrossRef] [PubMed]
  77. Seavill, P.W.; Holt, K.B.; Wilden, J.D. Investigations into the mechanism of copper-mediated Glaser–Hay couplings using electrochemical techniques. Faraday Discuss. 2019, 220, 269–281. [Google Scholar] [CrossRef] [PubMed]
  78. Muzalevskiy, V.; Sizova, Z.; Shastin, A.; Nenajdenko, V.G.; Diusenov, A.I. Efficient multi gram approach to acetylenes and CF3-ynones starting from dichloroalkenes prepared by catalytic olefination reaction (COR). Eur. J. Org. Chem. 2020, 2020, 4161–4166. [Google Scholar] [CrossRef]
  79. Hosseini, A.; Seidel, D.; Miska, A.; Schreiner, P.R. Fluoride-Assisted Activation of Calcium Carbide: A Simple Method for the Ethynylation of Aldehydes and Ketones. Org. Lett. 2015, 17, 2808–2811. [Google Scholar] [CrossRef] [PubMed]
  80. Carbon Tetrachloride. Available online: https://www.epa.gov/sites/default/files/2016-09/documents/carbon-tetrachloride.pdf (accessed on 20 June 2023 ).
  81. Rao, M.L.N.; Dasgupta, P.; Ramakrishna, B.S.; Murty, V.N. Domino Synthesis of 1,3-Diynes from 1,1-Dibromoalkenes: A Pd-Catalyzed Copper-Free Coupling Method. Tetrahedron Lett. 2014, 55, 3529–3533. [Google Scholar] [CrossRef]
  82. Feng, L.; Hu, T.; Zhang, S.; Xiong, H.-Y.; Zhang, G. Copper-Mediated Deacylative Coupling of Ynones via C–C Bond Activation under Mild Conditions. Org. Lett. 2019, 21, 9487–9492. [Google Scholar] [CrossRef]
  83. Shi, X.-L.; Hu, Q.; Wang, F.; Zhang, W.; Duan, P. Application of the Polyacrylonitrile Fiber as a Novel Support for Polymer-Supported Copper Catalysts in Terminal Alkyne Homocoupling Reactions. J. Catal. 2016, 337, 233–239. [Google Scholar] [CrossRef]
  84. Liu, D.-X.; Li, F.-L.; Li, H.-X.; Gao, J.; Lang, J.-P. Synthesis of 1,4-Diarylsubstituted 1,3-Diynes through Ligand-Free Copper-Catalyzed Oxidative Decarboxylative Homocoupling of Aryl Propiolic Acids. Tetrahedron 2014, 70, 2416–2421. [Google Scholar] [CrossRef]
  85. Stein, P.M.; Pascher, J.; Stracke, J.; Levacher, V.S.; Wagner, J.A.; Rominger, F.; Oeser, T.; Rudolph, M.; Hashmi, A.S.K. Gold Catalysis: 2,3- or 1,4-Addition to Butadiynes. Adv. Synth. Catal. 2022, 364, 3817–3839. [Google Scholar] [CrossRef]
  86. 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] [CrossRef]
  87. Sontakke, G.S.; Ghosh, C.; Pal, K.; Volla, C.M.R. Regioselective Dichotomy in Ru(II)-Catalyzed C–H Annulation of Aryl Pyrazolidinones with 1,3-Diynes. J. Org. Chem. 2022, 87, 14103–14114. [Google Scholar] [CrossRef]
  88. Jasiobedzki, W.; Moszczynski-Petkowski, R.; Wozniak-Kornacka, J. Diacetylene ω-glycols. Etherates—Adducts of 1,6-di(p-halophenyl)-1,6-diphenylhexa-2,4-diyne-1,6-diols with diethyl ether and dioxane. Bull. Pol. Acad. Sci. Chem. 2001, 49, 225–234. [Google Scholar]
  89. Ali, M.; Latif, A.; Bibi, S.; Ali, S.; Ali, A.; Ahmad, M.; Ahmad, R.; Khan, A.A.; Khan, A.; Ribeiro, A.I.; et al. Facile Synthesis of the Shape-Persistent 4-Hydroxybenzaldehyde Based Macrocycles and Exploration of their Key Electronic Properties: An Experimental and DFT Approach. ChemistrySelect 2022, 7, e202102715. [Google Scholar] [CrossRef]
  90. Ye, X.; Zhao, P.; Zhang, S.; Zhang, Y.; Wang, Q.; Shan, C.; Wojtas, L.; Guo, H.; Chen, H.; Shi, X. Facilitating Gold Redox Catalysis with Electrochemistry: An Efficient Chemical-Oxidant-Free Approach. Angew. Chem. Int. Ed. 2019, 58, 17226–17230. [Google Scholar] [CrossRef]
Figure 1. The use of the PHA-nitrogen base-CuCl system in the organic synthesis.
Figure 1. The use of the PHA-nitrogen base-CuCl system in the organic synthesis.
Catalysts 13 01330 g001
Scheme 1. Screening of the conditions for the dimerization of phenylacetylene.
Scheme 1. Screening of the conditions for the dimerization of phenylacetylene.
Catalysts 13 01330 sch001
Scheme 2. The reaction scope.
Scheme 2. The reaction scope.
Catalysts 13 01330 sch002
Scheme 3. The dimerization of propargyl alcohol derivatives.
Scheme 3. The dimerization of propargyl alcohol derivatives.
Catalysts 13 01330 sch003
Figure 2. Selected examples of physiologically active diacetylene diols.
Figure 2. Selected examples of physiologically active diacetylene diols.
Catalysts 13 01330 g002
Scheme 4. Control experiments.
Scheme 4. Control experiments.
Catalysts 13 01330 sch004
Scheme 5. Possible reaction mechanism.
Scheme 5. Possible reaction mechanism.
Catalysts 13 01330 sch005
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

Muzalevskiy, V.M.; Shastin, A.V.; Tirkasheva, S.I.; Ziyadullaev, O.E.; Parmanov, A.B.; Nenajdenko, V.G. CCl4-TMEDA-CuCl—A Novel Convenient Catalytic System for Dimerization of Terminal Acetylenes in Mild Conditions. Catalysts 2023, 13, 1330. https://doi.org/10.3390/catal13101330

AMA Style

Muzalevskiy VM, Shastin AV, Tirkasheva SI, Ziyadullaev OE, Parmanov AB, Nenajdenko VG. CCl4-TMEDA-CuCl—A Novel Convenient Catalytic System for Dimerization of Terminal Acetylenes in Mild Conditions. Catalysts. 2023; 13(10):1330. https://doi.org/10.3390/catal13101330

Chicago/Turabian Style

Muzalevskiy, Vasiliy M., Alexey V. Shastin, Sarvinoz I. Tirkasheva, Odiljon E. Ziyadullaev, Askar B. Parmanov, and Valentine G. Nenajdenko. 2023. "CCl4-TMEDA-CuCl—A Novel Convenient Catalytic System for Dimerization of Terminal Acetylenes in Mild Conditions" Catalysts 13, no. 10: 1330. https://doi.org/10.3390/catal13101330

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