首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
3.
Riduan SN  Ying JY  Zhang Y 《Organic letters》2012,14(7):1780-1783
A simple protocol for the stereoselective copper-catalyzed hydrothiolation of alkynes under a CO(2) atmosphere has been developed. The stereoselectivity is determined by the presence/absence of a CO(2) atmosphere. The reaction system is robust and utilizes inexpensive, readily available substrates. A cyclic alkene/carboxylate copper complex intermediate is proposed as the key step in determining the stereoselectivity, and an equivalent amount of water is found to play an active role as a proton donor.  相似文献   

4.
The N-arylation of aromatic and aliphatic secondary acyclic amides, known to be poor nucleophiles, has been accomplished using a simple and cheap copper catalytic system. The corresponding tertiary acyclic amides, which can be found in numerous biologically active compounds, have been obtained in good to excellent yields.  相似文献   

5.
Jiwon Kim  Hyesu Lee  Jaesook Yun 《Tetrahedron》2019,75(32):4250-4254
Enantioselective copper-catalyzed conjugate borylation of α,β-unsaturated phosphonates with bis(pinacolato)diboron affords chiral tertiary organoboronate esters with a vicinal phosphonate group in good yields and enantiomeric excess. Linear aliphatic, aromatic, and heteroaromatic substituents at the β-position can be accommodated, and oxidation of the borylated organophosphonate products leads to β-hydroxyphosphonates.  相似文献   

6.
A simple and efficient method for the deoxygenation of amine N-oxides to corresponding amines is reported using the green and economical reagent phenylboronic acid. Deoxygenation of N,N-dialkylaniline N-oxides, trialkylamine N-oxides and pyridine N-oxides were achieved in good to excellent yields. The reduction susceptible functional groups such as ketone, amide, ester and nitro groups are well tolerated with phenylboronic acid during the deoxygenation process even at high temperature. In addition, an indirect method for identification and quantification of tert-amine N-oxide is demonstrated using UV–Vis spectrometry which may be useful for drug metabolism studies.  相似文献   

7.
8.
A methodology for the high yield and facile synthesis of isoquinolones from benzamides and alkynes via the oxidative ortho C-H activation of benzamides has been developed. Ag(2)CO(3) proved to be an optimal oxidant when MeCN was used as a solvent, and [RhCp*Cl(2)](2) was utilized as an efficient catalyst. Both N-alkyl and N-aryl secondary benzamides can be applied as effective substrates. Furthermore, primary benzamides react with two alkyne units, leading to tricyclic products via double C-H activation and oxidative coupling. The reactivity of the structurally related 1-hydroxyisoquinoline was also demonstrated, where both N- and O-containing rhodacyclic intermediates can be generated, leading to the construction of different O- or N-containing heterocycles.  相似文献   

9.
The Pd-catalyzed oxidative coupling of N-substituted sydnones and terminal alkynes offers a quick, one-step synthesis of 4-alkynylsydnones in moderate to good yields.  相似文献   

10.
The copper-catalyzed addition of bis(pinacolato)diboron to internal alkynes in the presence of methanol generates alkenylboron compounds with high levels of regio- and stereoselectivities. The catalytic efficiency is increased by using monodentate phosphine ligands, especially P(p-tolyl)(3) and a range of internal alkynes was borylated in good yields.  相似文献   

11.
This study develops a new efficient pathway for synthesis of 2,4-disubstituted 1,2,3-triazoles through regioselective direct arylation between 2-aryl-1,2,3-triazole N-oxides and Ar-B(OH)2. The reaction proceeds smoothly at room temperature and exhibits good yield and high C5 position selectivity. The possible pathway of oxidative Suzuki coupling is also discussed. This simple methodology can be used to construct 2,4-disubstituted 1,2,3-triazole moiety.  相似文献   

12.
A copper-catalyzed synthesis of N-sulfonylamidines via three-component coupling of sulfonyl azides, terminal alkynes, and trialkylamines is reported.  相似文献   

13.
Facile reduction of alkylamino-, anilino-, and pyridyl-N-oxides can be achieved via the use of diboron reagents, predominantly bis(pinacolato)- and in some cases bis(catecholato)diboron [(pinB)(2) and (catB)(2), respectively]. Reductions occur upon simply mixing the amine N-oxide and the diboron reagent in a suitable solvent, at a suitable temperature. Extremely fast reductions of alkylamino- and anilino-N-oxides occur, whereas pyridyl-N-oxides undergo slower reduction. The reaction is tolerant of a variety of functionalities such as hydroxyl, thiol, and cyano groups, as well as halogens. Notably, a sensitive nucleoside N-oxide has also been reduced efficiently. The different rates with which alkylamino- and pyridyl-N-oxides are reduced has been used to perform stepwise reduction of the N,N'-dioxide of (S)-(-)-nicotine. Because it was observed that (pinB)(2) was unaffected by the water of hydration in amine oxides, the feasibility of using water as solvent was evaluated. These reactions also proceeded exceptionally well, giving high product yields. In constrast to the reactions with (pinB)(2), triethylborane reduced alkylamino-N-oxides, but pyridine N-oxide did not undergo efficient reduction even at elevated temperature. Finally, the mechanism of the reductive process by (pinB)(2) has been probed by (1)H and (11)B NMR.  相似文献   

14.
A method for the synthesis of propargylic amines has been developed via an efficient copper(I)-catalyzed three-component coupling reaction of alkynes, benzal halides and amines through C-H and C-halogen activation. This reaction is conducted under mild conditions and provides an alternative method for the synthesis of propargylic amines.  相似文献   

15.
Irradiation (1) (253-7 nm) of Na,Nb-dimethyltryptaimne with pyridine N-oxide or benzo[c]cinnoline N-oxide in CH2Cl2 yielded 1,8-dimethyl-3a-hydroxy-1,2,3,3a,8-8a,- hexahydropyrrolo[2,3-b]indole (19), while with visible light Nb-(4-cyanobutadienyl)-Na,Nb,- dimethyltryptamine (21) was obtained. This method was applied to trimethyltryptamine and the corresponding oxindole (34) and the N-formyl derivative (20) were obtained.  相似文献   

16.
A general and efficient Cu(I)-catalyzed cross-coupling reaction of terminal alkynes and monooxalyl chloride for the synthesis of 2-oxo-3-butynoates and 2-oxo-3-butynoamides was developed. Readily available starting materials, the mild reaction conditions, wide functional group tolerance, and the obviation of stoichiometric organolithium or magnesium reagents combine to highlight this reaction.  相似文献   

17.
A silver-mediated tandem protocol for the synthesis of quinolines involving the oxidative coupling/cyclization of N-arylimines and alkynes has been developed. We demonstrated that scenario-dependent metalation could occur either at the ortho C-H bond of an N-arylimine through protonation-driven enhancement of acidity or at the terminal C-H bond of an alkyne by virtue of the carbophilic π-acidity of silver. The diverse set of mechanistic manifolds implemented with a single type of experimental protocol points toward the importance of stringent reactivity analysis of each individual potentially reactive molecular site. Importantly, the direct arene C-H bond activation provides a unique and distinct mechanistic handle for the expansion of reactivity paradigms for silver. As expected, the protocol allows for the incorporation of both internal and terminal alkynes into the products, and in addition, both electron-withdrawing and -donating groups are tolerated on N-arylimines, thus enabling the vast expansion of substituent architectures on quinoline framework. Further, an intriguing phenomenon of structural isomerization and chemical bond cleavage has been observed for aliphatic internal alkynes.  相似文献   

18.
An efficient and simple protocol of copper-catalyzed C-S bond formation between aryl halides and inexpensive and commercially available aminothiourea is reported.A variety of symmetrical diaryl sulfides can be synthesized in good to excellent yields up to 94%with the advantage of avoiding foul-smelling thiols.  相似文献   

19.
Methanol is an abundant and renewable chemical raw material, but its use as a C1 source in C–C bond coupling reactions still constitutes a big challenge, and the known methods are limited to the use of expensive and noble metal catalysts such as Ru, Rh and Ir. We herein report nickel-catalyzed direct coupling of alkynes and methanol, providing direct access to valuable allylic alcohols in good yields and excellent chemo- and regioselectivity. The approach features a broad substrate scope and high atom-, step- and redox-economy. Moreover, this method was successfully extended to the synthesis of [5,6]-bicyclic hemiacetals through a cascade cyclization reaction of alkynones and methanol.

Methanol is an abundant and renewable chemical raw material, but its use as a C1 source in C–C bond coupling reactions still constitutes a big challenge, and the known methods are limited to the use of expensive and noble metal catalysts such as Ru, Rh and Ir.

To address the sustainability issues in the production of new chemicals, the development of new catalytic processes that are free of by-products and using abundant renewable feedstocks is one of the most important challenges facing chemists today. The simplest alcohol, methanol, is very abundant, with a total annual production capacity of approximately 110 million metric tons per year,1 and is an important C1-feedstock in the chemical industry. Beller2 and Milstein3 made fundamental developments in catalytic dehydrogenation reactions of methanol.4 Krische and coworkers pioneered the study of Ir-catalyzed direct C–C coupling of methanol with reactive π-unsaturated reactants (1,3-dienes, 1,3-enynes and allenes).5 The groups of Glorius,6 Donohoe,7 Obora,8 Andersson9 and others10 demonstrated the direct methylation of ketones or amines using methanol. Despite these achievements, the catalytic C–C bond coupling reactions with methanol are still extremely rare and are limited to the use of precious and noble metal-catalysts such as Ru, Rh or Ir.11 The development and use of cheap and abundant metal catalysts for methanol activation is uphill and remains an important field that urgently needs to be developed.On the other hand, allylic alcohols are highly versatile building blocks in organic synthesis and the pharmaceutical industry, and much effort has been devoted to their synthesis. Among them, nickel-catalyzed reductive coupling of alkynes and aldehydes represents an effective and powerful method. However, this method generally requires the use of stoichiometric reducing reagents that are air-sensitive, metallic or pyrophoric (e.g. ZnR2, BEt3, and R3SiH, Scheme 1a).12 The direct cross-coupling of alcohols and alkynes to synthesize allylic alcohols without the use of any reductant or oxidant represents a significant advancement (Scheme 1b).13 However, this approach still poses many limitations that will require considerable effort to overcome. (1) Alkynes are limited to dialkyl alkynes, and poor regioselectivities were observed for unsymmetrical alkynes, which greatly limits the scope of application of the reaction. (2) Alcohols are restricted to active benzyl alcohols and higher alcohols. The direct cross-coupling of alkynes with methanol has not yet been reported.Open in a separate windowScheme 1Synthesis of allylic alcohols by Ni-catalyzed coupling reaction with alkynes.Although the alkyne–paraformaldehyde reductive coupling has been developed,14 the paraformaldehyde was itself prepared from synthesis gas (through methanol). Therefore, the development of a new strategy for the direct coupling of alkynes and methanol without the use of any reductant or oxidant is still of great value, but also extremely challenging: (1) alkynes are reactive and could rapidly dimerize to 1,3-dienes15 or cyclotrimerize to aromatic ring derivatives in the interaction with nickel.16 (2) Unsymmetric alkynes could result in a mixture of regioisomers that are difficult to separate. (3) The activation energy of methanol in the dehydrogenation process (ΔH = +84 kJ mol−1) is significantly higher than that of higher alcohols or even ethanol (ΔH = +68 kJ mol−1).17Herein we report the nickel-catalyzed direct and regioselective hydrohydroxymethylation of alkynes for the first time using methanol as a C1-feedstock, providing a broad and efficient approach for the synthesis of high added-value allylic alcohols in a high atom-, step- and redox-economic manner. In addition, a cascade cyclization reaction of alkynones and methanol has also been developed for the synthesis of [5,6]-bicyclic hemiacetals in good yields and excellent regio- and diastereoselectivity (Scheme 1c).In our initial experiments, we chose unsymmetrical internal alkyne 1a as a model substrate to optimize the reaction conditions (13a Even if the reaction temperature was increased to 100 °C, only a trace amount of allylic alcohol product 2a was observed (entry 2), which indicates that the use of methanol in the catalytic C–C coupling reactions is indeed a big challenge. Various N-heterocyclic carbene ligands (L2–L6) were investigated (entries 3–9). We found that the selectivity of allylic alcohol 2a is challenged by a number of side reactions, such as the hydrogenation (3a), dimerization (4a) and trimerization (5a) of alkyne 1a. L4 is the most effective, providing 2a with the highest yield (40%) and excellent regioselectivity (14/1), but an appreciable quantity of dimerization and trimerization by-products 4a and 5a was still obtained (entry 5). Krische14 reported that PCy3 could promote the reductive coupling of alkynes and paraformaldehyde, but we found that it is not effective for alkyne–methanol coupling (entry 8).Optimization of reaction conditionsa
EntryLigandAdditiveYieldb (2, %)Yieldb (3, %)Yieldb (4, %)Yieldb (5, %)
1c,d L1 No reaction
2c L1 66<220
3 L2 30 (14/1)e151320
4 L3 35<243
5 L4 40 (14/1)e92223
6 L5 30 (7/1)e61827
7 L6 No reaction
8Cy3PComplicated
9PPh3104<259
10 L4 A1 10<2<2<2
11 L4 A2 38 (14/1)e<2<2<2
12 L4 A3 60f (14/1)e6<2<2
13 L4 A4 No reaction
14 L4 A3 g54f (14/1)e867
Open in a separate windowaReactions conditions: 1a (0.2 mmol), Ni(COD)2 (10 mol%), ligand (20 mol%), tBuOK (12 mol%), additive (1 equiv.) in toluene (1 mL) and MeOH (3 mL) in a sealed tube at 100 °C.bDetermined by GC analysis using adamantane as the internal standard.cWithout tBuOK.dRoom temperature.eRegioselectivity (2a/2a′).fIsolated yield.g0.2 equivalent.Many examples have reported that olefins can affect the outcomes of transition metal-catalyzed cross coupling reactions through increased activity, stability, or selectivity.18 More recently, Montgomery et al.19 found that adding electron-deficient olefins to NHC–Ni(0) complexes can improve their catalytic performance. Inspired by this discovery, we examined various acrylates A1–A4 (entries 10–13). Excitingly, the addition of methyl methacrylate (A3) can indeed significantly improve the chemoselectivity of the reaction, providing the allylic alcohol 2a in 60% isolated yield and a more than 14/1 ratio of regioisomers (entry 12). The structure of acrylates has a great influence on the reaction outcome, indicating that they may act as additional ligands to coordinate with the nickel catalyst, thereby suppressing these undesired dimerization or cyclotrimerization side reactions. However, by-products formed by the reductive coupling of acrylates and alkynes have also been observed (see Section 3 in the ESI).20 It is worth mentioning that stoichiometric acrylate additives are not necessary. As shown in entry 14, even if 0.2 equivalent of A3 was used, 54% of the target product 2a can be obtained, thus showing the subtleties of our catalytic system.With the optimized reaction conditions in hand, we turned our attention to explore the substrate scope of alkynes (Scheme 2). We were pleased to find that various unsymmetrical aryl–alkyl alkynes were coupled with methanol to provide the corresponding allylic alcohols 2a–2q in moderate to good yields and high regioselectivities. Various functional groups, such as fluorine (2b), trifluoromethyl (2c), chlorine (2e), allyl (2f), bromine (2g), amine (2h–2j) and amide (2k and 2l) could all be well-tolerated. Heteroaromatic ring-substituted alkynes, such as 5-indole,21 2-dibenzothiophene and 2-dibenzofuran could also proceed smoothly to furnish allylic alcohols 2n–2p in 40–63% yield. It is worth mentioning that complex biologically active molecules such as estrone derivatives, could also be successfully incorporated into the desired product 2q in 62%, thus demonstrating the robustness and generality of this methodology for late-stage modification of complex biologically active molecules. Terminal alkynes were also found to be compatible with the reaction conditions, providing the corresponding products 2r–2s in moderate yields and excellent regioselectivity (>20/1). Symmetric diarylalkynes bearing electron-donating or electron-withdrawing groups were applicable to the reaction (2t–2y). Strikingly, both 1,2-di(furan-2-yl)ethyne and 1,2-di(thiophen-2-yl)ethyne were competent substrates and furnished the desired allylic alcohols 2x–2y in good yields.Open in a separate windowScheme 2Substrate scope of alkynes for the synthesis of allylic alcohols. Reactions were carried out with 1 (0.2 mmol), Ni(COD)2 (10 mol%), L4 (10 mol%), tBuOK (12 mol%), and methyl methacrylate (0.2 mmol) in toluene (1.0 mL) and MeOH (3.0 mL) in a sealed tube at 100 °C. Isolated yields are given. a The reaction was conducted with Ni(COD)2 (15 mol%), L4 (15 mol%), and tBuOK (18 mol%). b ((4-Bromophenyl)ethynyl)trimethylsilane was used. c The reaction was conducted with toluene (0.5 mL) and MeOH (1.5 mL).In addition, this transformation is not restricted to aryl-substituted alkynes. As shown in Scheme 2, oct-4-yne and cyclododecane were coupled with methanol to produce allylic alcohols 2z and 2aa in 87% and 54% yields, respectively. To further evaluate the influence of the electronic properties of the substituents on the regioselectivity, we tested the hydrohydroxymethylation reaction of unsymmetrical dialkyl-substituted alkynes bearing benzyloxy or dibenzylamino groups at the propargylic position. To our delight, the corresponding allylic alcohols 2ab–2ad were obtained in moderate yields, with remarkably high regioselectivity (>20/1). However, the regioselectivity of this reaction was decreased by the alkyne bearing a benzyloxy group at the homopropargylic position.To expand the potential synthetic applications of the transformation, we investigated the hydrohydroxymethylation of 1,3-enynes. The corresponding dienol 2af was obtained, which was selectively hydrohydroxymethylated on the alkyne but not on the alkene moiety. 1,6-Enyne was also compatible to give the corresponding allylic alcohol 2ag in 42% yield with >20/1 regioselectivity. This strategy can serve as a powerful supplement to the previous method reported by Krische et al.,22 in which alcohols were reacted with alkenes to obtain the corresponding homopropargylic alcohols.23Alkynone substrates were also tested, but the expected product was not detected due to their sensitivity to base. After slightly modifying the reaction conditions, we were pleased to find that various [5,6]-bicyclic hemiacetals 7 could be obtained in good yields with excellent regio- and diastereoselectivities through the cascade cyclization reaction of alkynones 6 with methanol (Scheme 3). We first explored the influence of the substituents (R1) at the terminus of the triple bond. A variety of para-substituted aromatic rings at the alkyne terminus could undergo tandem cyclization to provide the target hemiacetals 7b–7g in 54–78% yields. The structure of 7a was confirmed by an X-ray crystal diffraction study. The aryl groups with substituents at the meta and ortho position were also found to be compatible, leading to the corresponding products 7h–7j in 56–74% yields. Moreover, various (hetero)aryl rings such as naphthalene (7k), benzodioxan (7l), 3,4-dihydrobenzodioxine (7m), thiophene (7n), dibenzofuran (7o), dibenzothiophene (7p), indole (7q) and pyridine (7r) at the terminal of the triple bond could be successfully incorporated into the desired products in good yields. Strikingly, estrone was also compatible with this transformation to afford the desired product 7s in 65% yield. However, no desired product was observed when the methyl substituted alkynone substrate was used. We then investigated the influence of the substituents (R2) at the 2-position of the cyclopentane-1,3-diones. Ethyl, benzyl, and allyl were all well tolerated leading to the corresponding [5,6]-bicyclic hemiacetals 7t–7w in moderate yields.Open in a separate windowScheme 3Substrate scope of alkynones for the synthesis of [5, 6]-bicyclic hemiacetals. Reactions were carried out with 6 (0.2 mmol), Ni(COD)2 (15 mol%), IMes (15 mol%), LiF (10 mol%), and methyl methacrylate (0.2 mmol) in toluene (1.5 mL) and MeOH (0.5 mL) in a sealed tube at 40 °C. Isolated yields are given.To provide a deeper insight into the reaction mechanism, deuterium-labelling experiments were performed. 1a was reacted with CH3OD under our standard reaction conditions; however, no incorporation of deuterium was detected in product 2a (Scheme 4a), revealing that the hydroxyl of methanol is not the proton source. This result is different from the previous report by Zhou et al.,24 in which the Ni(0) catalyst underwent oxidative addition to the O–H bond of methanol to form methoxyl nickel hydride species and then migratory insertion into unsaturated bonds. Further investigation using CD3OD as solvent provided 2a-D in 41% yield, in which 99% of the deuterium was incorporated into the olefinic position, but the reaction rate is obviously slowed down (Scheme 4b). We also conducted the kinetic isotope effect (KIE) experiment. The intermolecular competition reaction between 1a and CD3OD or CH3OH under standard reaction conditions provided a KIE (kH/kD) value of 6.1 (Scheme 4c). Taken together, these results may indicate that the dehydrogenation of methanol to form the key formaldehyde intermediate is the rate-determining step of this transformation.Open in a separate windowScheme 4Deuterium-labelling experiments.On the basis of these experimental results and previous observations, a possible reaction mechanism is proposed in Scheme 5. The reaction is initiated by reducing alkyne to alkene and simultaneously oxidizing methanol to formaldehyde, as evidenced by the detection of catalytic amounts of alkene 3. Oxidative cyclization of acrylate-coordinated NHC–Ni(0) A 17 with alkyne and formaldehyde gives oxa-nickelacycle intermediate B. Subsequent protonation of nickelacycle species B with methanol affords the vinylnickel intermediate C, which can undergo β-H elimination to generate vinyl nickel hydride species D and formaldehyde.25 Reductive elimination of D will furnish allylic alcohol 2 and the catalytically active Ni(0) catalyst A. Further nucleophilic addition of the hydroxyl group to one of the ketone carbonyl groups will produce [5,6]-bicyclic hemiacetal 7. We speculate that the acrylate is used as an additional ligand, thereby inhibiting the alkyne dimerization to 1,3-dienes or cyclotrimerization to aromatic ring derivatives.Open in a separate windowScheme 5Proposed reaction mechanism.  相似文献   

20.
Na Guo  Jian-Xin Ji 《Tetrahedron letters》2012,53(36):4797-4801
The first copper-catalyzed direct three-component coupling reaction of aldehydes, alkynes, and hydroxylamines for the synthesis of propargylamines has been developed under mild conditions, which has the advantages of ready availability of catalyst as well as operation simplicity. The present protocol provides an appealing alternative for the construction of propargylamines in a simple and one-pot procedure.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号