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 289 毫秒
1.
The cycloaddition of organic azides with some conjugated enamines of the 2‐amino‐1,3‐diene, 1‐amino‐1,3‐diene, and 2‐aminobut‐1‐en‐3‐yne type is investigated. The 2‐morpholinobuta‐1,3‐diene 1 undergoes regioselective [3+2] cycloaddition with several electrophilic azides RN3 2 ( a , R=4‐nitrophenyl; b , R=ethoxycarbonyl; c , R=tosyl; d , R=phenyl) to form 5‐alkenyl‐4,5‐dihydro‐5‐morpholino‐1H‐1,2,3‐triazoles 3 which are transformed into 1,5‐disubstituted 1H‐triazoles 4a , d or α,β‐unsaturated carboximidamide 5 (Scheme 1). The cycloaddition reaction of 4‐[(1E,3Z)‐3‐morpholino‐4‐phenylbuta‐1,3‐dienyl]morpholine ( 7 ) with azide 2a occurs at the less‐substituted enamine function and yields the 4‐(1‐morpholino‐2‐phenylethenyl)‐1H‐1,2,3‐triazole 8 (Scheme 2). The 1,3‐dipolar cycloaddition reaction of azides 2a – d with 4‐(1‐methylene‐3‐phenylprop‐2‐ynyl)morpholine ( 9 ) is accelerated at high pressure (ca. 7–10 kbar) and gives 1,5‐disubstituted dihydro‐1H‐triazoles 10a , b and 1‐phenyl‐5‐(phenylethynyl)‐1H‐1,2,3‐triazole ( 11d ) in significantly improved yields (Schemes 3 and 4). The formation of 11d is also facilitated in the presence of an equimolar quantity of tBuOH. The three‐component reaction between enamine 9 , phenyl azide, and phenol affords the 5‐(2‐phenoxy‐2‐phenylethenyl)‐1H‐1,2,3‐triazole 14d .  相似文献   

2.
The 3‐heteroaryl‐1‐methylpyridazino[3,4‐b]quinoxalin‐4(1H)‐ones 6a‐e were synthesized by the oxidative‐hydrolytic ring transformation of the 3‐heteroaryl‐1,2‐diazepino[3,4‐b]]quinoxaline‐5‐carbonitriles 9a‐c , which were obtained by the 1,3‐dipolar cycloaddition reaction of the 2‐(2‐heteroarylmethylene‐1‐methylhydrazino)quinoxaline 4‐oxides with 2‐chloroacrylonitrile. The assignment of the thiophene and furan ring protons was carried out through the data of the NOE, decoupling, and coupling constants.  相似文献   

3.
An effective and practical method has been developed for the diversity‐oriented synthesis of 1‐alkyl‐3‐aroylindolizines via the 1,3‐dipolar cycloaddition of pyridinium ylides and aliphatic aldehydes in the presence of molecular iodine and a catalytic amount of MnO2. The synthesis proceeds by tandem reactions involving [3+2] cycloaddition, dehydration of the cycloadduct, and dehydroaromatization. Molecular iodine served both as a catalyst and a dehydroaromatization reagent in the reaction.  相似文献   

4.
An efficient method for the synthesis of N‐alkylated 2‐(4‐substituted‐1H‐1,2,3‐triazol‐1‐yl)‐1H‐indole‐3‐carbaldehyde has been developed starting from oxindole and indole using Huisgen's 1,3‐dipolar cycloaddition reaction of organic azides to alkynes. The effect of catalysts and solvent on these reactions has been investigated. Among all these conditions, while using CuSO4·5H2O, DMF was found to be the best system for this reaction. It could also be prepared in a one‐pot three‐component manner by treating equimolar quantities of halides, azides, and alkynes. The Huisgen's 1,3‐dipolar cycloaddition reaction was performed using CuSO4·5H2O in DMF with easy work‐up procedure.  相似文献   

5.
The [3,3′(4H,4′H)‐bi‐2H‐1,3‐oxazine]‐4,4′‐diones 3a – 3i were obtained by [2+4] cycloaddition reactions of furan‐2,3‐diones 1a – 1c with aromatic aldazines 2a – 2d (Scheme 1). So, new derivatives of bi‐2H‐1,3‐oxazines and their hydrolysis products, 3,5‐diaryl‐1H‐pyrazoles 4a – 4c (Scheme 3), which are potential biologically active compounds, were synthesized for the first time.  相似文献   

6.
Regioselective Pd0‐catalyzed cross‐coupling of substrates, which bear bispropargylic leaving groups with silyl‐protected alkynes, has provided access to a variety of 1,3‐diethynylallenes, a new family of modules for three‐dimensional acetylenic scaffolding. In enantiomerically pure form, these C‐rich building blocks could provide access – by oxidative oligomerization – to a fascinating new class of helical oligomers and polymers with all‐carbon backbones (Fig. 2). In the first of two routes, a bispropargylic epoxide underwent ring opening during Sn 2′‐type cross‐coupling, and the resulting alkoxide was silyl‐protected, providing 1,3‐diethynylallenes (±)‐ 8 , (±)‐ 12 (Scheme 3), and (±)‐ 15 (Scheme 5). A more general approach involved bispropargylic carbonates or esters as substrates (Scheme 68), and this route was applied to the preparation of a series of 1,3‐diethynylallenes to investigate how their overall stability against undesirable [2+2] cycloaddition is affected by the nature of the substituents at the allene moiety. The investigation showed that the 1,3‐diethynylallene chromophore is stable against [2+2] cycloaddition only when protected by steric bulk and when additional π‐electron delocalization is avoided. The regioselectivity of the cross‐coupling to the bispropargylic substrates is entirely controlled by steric factors: attack occurs at the alkyne moiety bearing the smaller substituent (Schemes 9 and 10). Oxidative Hay coupling of the terminally mono‐deprotected 1,3‐diethynylallene (±)‐ 49 afforded the first dimer 50 , probably as a mixture of two diastereoisomers (Scheme 12). Attempts to prepare a silyl‐protected tetraethynylallene by the new methodology failed (Scheme 13). Control experiments (Schemes 1416) showed that the Pd0‐catalyzed cross‐coupling to butadiyne moieties in the synthesis of this still‐elusive chromophore requires forcing conditions under which rapid [2+2] cycloaddition of the initial product cannot be avoided.  相似文献   

7.
Lijuan Lu  Chaoguo Yan 《中国化学》2015,33(10):1178-1188
The base mediated cycloaddition reactions of 4‐dimethylamino‐1‐phenacylpyridinium bromides with two molecular 3‐phenacylideneoxindoles in methylene dichloride afforded functionalized dispirocyclopentyl‐3,3′‐bisoxindoles in good yields and with high diastereoselectivity. The similar cycloaddition reactions of 1‐(N,N‐dialkylcarbamoylmethyl) and 1‐cyanomethyl 4‐dimethylamino‐pyridinium bromide in refluxing ethanol in the presence of triethylamine also resulted in dispirocyclopentyl‐3,3′‐bisoxindoles with high diastereoselectivity. The stereochemistry of dispirocyclopentyl‐3,3′‐bisoxindoles was elucidated on the basis of 1H NMR data and single crystal structures.  相似文献   

8.
Thioether 4‐[(1′E,3′E)‐4′‐phenylsulfanyl‐1,3′‐butadienyl]pyridine 8 and sulfone 4‐(4′‐phenylsulfonyl‐1′,3′‐butadienyl)pyridine 14 were prepared by reaction of the carbanions derived from allylic thioether or allylic sulfone with isonicotinaldehyde. The reaction with the sulfonyl carbanion occurred at the α position and on heating the alcolate gave the dienic sulfone 14 . The corresponding pyridinium iodide 10 and 15 were prepared by reaction with methyl iodide, respectively, on pyridine derivates 8 and 14 . The dienic pyridinium thioether 10 showed a long wavelength absorption band centered at 420 nm. The reaction of dienic pyridinium sulfone 15 with thiophenol gave the dienic pyridinium thioether 10 by a nucleophilic vinylic substitution. The reaction of sulfone 15 with glutathione was of second order and the rate constant was 8.5 M?1s?1 at 30°C and pH 7, about 500 times smaller than the rate constant observed with (E)‐1‐methyl‐4‐(2‐methylsulfonyl‐1‐ethenyl)pyridinium iodide 1 . The dienic pyridinium thioether 10 was a negative solvatochrome.  相似文献   

9.
The formal 1,3‐cycloaddition of 2‐diazocyclohexane‐1,3‐diones 1a –1 d to acyclic and cyclic enol ethers in the presence of RhII‐catalysts to afford dihydrofurans has been investigated. Reaction with a cis/trans mixture of 1‐ethoxyprop‐1‐ene ( 13a ) yielded the dihydrofuran 14a with a cis/trans ratio of 85 : 15, while that with (Z)‐1‐ethoxy‐3,3,3‐trifluoroprop‐1‐ene ( 13b ) gave the cis‐product 14b exclusively. The stereochemical outcome of the reaction is consistent with a concerted rather than stepwise mechanism for cycloaddition. The asymmetric cycloaddition of 2‐diazocyclohexane‐1,3‐dione ( 1a ) or 2‐diazodimedone (=2‐diazo‐5,5‐dimethylcyclohexane‐1,3‐dione; 1b ) to furan and dihydrofuran was investigated with a representative selection of chiral, nonracemic RhII catalysts, but no significant enantioselectivity was observed, and the reported enantioselective cycloadditions of these diazo compounds could not be reproduced. The absence of enantioselectivity in the cycloadditions of 2‐diazocyclohexane‐1,3‐diones is tentatively explained in terms of the Hammond postulate. The transition state for the cycloaddition occurs early on the reaction coordinate owing to the high reactivity of the intermediate metallocarbene. An early transition state is associated with low selectivity. In contrast, the transition state for transfer of stabilized metallocarbenes occurs later, and the reactions exhibit higher selectivity.  相似文献   

10.
Cephalosporin sulfoxides 1 and 2 containing an enone‐ or dienone‐type moiety at position 2 were treated with 2,3‐dimethylbuta‐1,3‐diene or diethyl azodicarboxylate to synthesize, in Diels? Alder reactions, the new cephalosporin derivatives 4 and 5 with a cyclic substituent (Scheme 1). Under the same conditions, ethyl diazoacetate and diazomethane reacted differently: while reactions of 1 and 3 with the former lead to compounds 7 – 10 corresponding to the 1,3‐dipolar cycloaddition route (Scheme 2), diazomethane produced only enol ethers 12 and 13 , respectively (Scheme 3). This difference could be rationalized by assuming two different reaction pathways: an orbital‐symmetry‐controlled concerted cycloaddition and an ionic one.  相似文献   

11.
Aryl azides 1 were treated with allenylmagnesium bromide ( 2 ) to generate 1,5‐disubstituted butynyl‐1H‐1,2,3‐triazoles 3 in a domino fashion, which upon CuI‐catalyzed 1,3‐dipolar cycloaddition with aryl azides 4 afforded novel bis‐1H‐1,2,3‐triazoles 5 in quantitative yields (Scheme 1 and Table).  相似文献   

12.
Nitrile oxide 1,3 dipolar cycloaddition is a simple and powerful coupling methodology. However, the self‐dimerization of nitrile oxides has prevented the widespread use of this strategy for macromolecular coupling. By combining an in situ nitrile oxide generation with a highly reactive activated dipolarophile, we have overcome these obstacles and present a metal‐free macromolecular coupling strategy for the modular synthesis of several ABA triblock copolymers. Nitrile oxides were generated in situ from chloroxime terminated poly(dimethylsiloxane) B‐blocks and coupled with several distinct hydrophilic (poly(2‐methyloxazoline) and poly(ethylene glycol)), and poly(N‐isopropylacrylamide) or hydrophobic (poly(l ‐lactide) A‐blocks terminated in activated dipolarophiles in a rapid fashion with high yield. This methodology overcomes many drawbacks of previously reported metal‐free methods due to its rapid kinetics, versatility, scalability, and ease of introduction of necessary functionality. Nitrile oxide cycloaddition should find use as an attractive macromolecular coupling strategy for the synthesis of biocompatible polymeric nanostructures. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 3134–3141  相似文献   

13.
The azomethine N‐oxides ( 1 ) on reacting with N‐benzylmaleimide ( 2 ) provide a mixture of stereoisomers 2,3‐diphenyl‐5‐benzyl‐4H‐2,3,3a,5,6,6a‐hexahydropyrrolo[3,4‐d]isoxazole‐4,6–dione derivatives ( 3 ) in good yields. These isomers have been assigned cis and trans configurations ( 3‐A and 3‐B ) with respect to proton C3‐H on the azomethinic carbon on the basis of their PMR and H‐NMR COSY data. The ratio between cis and trans isomers has been found to be dependent on substituents present at ortho position of C‐phenyl aldehydic moiety. The salient feature of these 1,3‐dipolar cycloaddition reactions lies in that the benzylic protons on N‐benzyl moiety suffer gem coupling, indicating magnetic nonequivalence. J. Heterocyclic Chem.,, (2012).  相似文献   

14.
The 1,3‐dipolar cycloaddition reaction of (E ,E )‐1,3‐bis(arylidene)indan‐2‐one 1a , 1b , 1c , 1d , 1e with diarylnitrilimines, generated in situ via dehydrohalogenation of the corresponding hydrazonoyl chlorides 2a , 2b , 2c , affords predominantly monospiropyrazolines 3 and 4 as a mixture of diastereoisomers. Also dispiropyrazolines 5 are formed in moderate yields. The structure and stereochemistry of cycloadducts 3 , 4 , 5 were confirmed by 1H and 13C‐NMR spectroscopy, elemental analyses data, and single‐crystal X‐ray diffraction studies of 3ba and 5ca .  相似文献   

15.
The reactivity of 3‐nitro‐4‐pyridyl isocyanate ( 7 ) and 5‐nitropyridin‐2‐yl isocyanate ( 9 ) in 1,3‐dipolar cycloaddition reactions with azides and pyridine N‐oxides has been investigated. 1,3‐Dipolar cycloaddition to trimethylsilylazide (TMSA) afforded the respective tetrazolinones, 1‐(3‐nitropyridin‐4‐yl)‐1H‐tetrazol‐5(4H)one ( 8 , 50 %) and 1‐(5‐nitropyridin‐2‐yl)‐1H‐tetrazol‐5(4H)one ( 11 , 64 %). Respectively, 1,3‐dipolar cycloaddition of nitropyridyl isocyanates 7 and 9 to 3,5‐dimethylpyridine N‐oxide ( 14 ), 3‐methylpyridine N‐oxide ( 21 ) and pyridine N‐oxide ( 22 ) gave the substituted amines, 3,5‐dimethyl‐N‐(3‐nitropyridin‐4‐yl)pyridin‐2‐amine ( 17 ), 3,5‐dimethyl‐N‐(5‐nitropyridin‐2‐yl)pyridin‐2‐amine ( 20 ), N‐(5‐nitropyridin‐2‐yl)pyridin‐2‐amine ( 24 ), 5‐methyl‐N‐(5‐nitropyridin‐2‐yl)pyridin‐2‐amine ( 23 ) and 3‐methyl‐N‐(5‐nitropyridin‐2‐yl)pyridin‐2‐amine ( 25 ) in 65 ‐ 80 % yield, obtained by cycloaddition, rearrangement and decarboxylation. The results demonstrate that the nitropyridyl isocyanates ( 7,9 ) readily undergo 1,3‐dipolar cyloaddition reactions similar to phenyl isocyanates.  相似文献   

16.
A new total synthesis of the natural carbazole murrayanine ( 1 ) was developed by using the 4,5‐dimethyleneoxazolidin‐2‐one 12 as starting material. The latter underwent a highly regioselective Diels–Alder cycloaddition with acrylaldehyde (=prop‐2‐enal; 13 ) to give adduct 14 (Scheme 3). Conversion of this adduct into diarylamine derivative 9 was carried out via hydrolysis and methylation (Scheme 4). Differing from our previous synthesis, in which such a diarylamine derivative was transformed into 1 by a PdII‐stoichiometric cyclization, this new approach comprised an improved cyclization through a more efficient Pd0‐catalyzed intramolecular diaryl coupling which was applied to 9 , thus obtaining the natural carbazole 1 in a higher overall yield.  相似文献   

17.
We describe here, for the first time, that 1‐cyano‐3‐benzoyl‐1,2,3,10b‐tetrahydropyrrolo[2,1‐a]phthalazine (8), prepared by 1,3‐dipolar cycloaddition of 2‐phenacyl phthalazinium bromide ( 6a ) with acrylonitrile ( 7a ), can be aromatized by tetrakis‐pyridine cobalt (II) dichromate (TPCD) to give 1‐cyano‐3‐benzoylpyrrolo[2,1‐α]phthalazine ( 9a ) in good yield. Furthermore, a general and convenient one‐pot procedure for preparation of pyrrolo[2,1‐α]phthalazines ( 9a‐p ) was developed by 1,3‐dipolar cycloaddition of phthala zinium N‐ylides ( 6a‐c ) with alkenes ( 7a‐g ) in the presence of TPCD.  相似文献   

18.
The application of microwave technique has been extended successfully for the first time to the synthesis of a representative class of azaphospholes, viz. 1,3‐bis(alkoxycarbonyl)‐1,3‐azaphospholo[5,1‐a]isoquinolines ( 2 ), which occurs rapidly giving higher yields. Stereoselectivity is observed in the reaction with 2,3‐dimethyl‐1,3‐butadiene, and isoprene reacts regioselectively as well. 1‐Methyl‐3‐ethoxycarbonyl‐1,3‐azaphospholo[1,5‐a]pyridine ( 4 ) remains inert toward [2+4] cycloaddition. The nonoccurrence of the Diels–Alder reaction in the latter case has been supported by semiempirical PM3 calculations. © 2003 Wiley Periodicals, Inc. Heteroatom Chem 14:560–563, 2003; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.10193  相似文献   

19.
The chemoselectivity in the reaction of 2‐diazo‐3‐oxo‐3‐phenylpropanal ( 1 ) with aldehydes and ketones in the presence of Et3N was investigated. The results indicate that 1 reacts with aromatic aldehydes with weak electron‐donating substituents and cyclic ketones under formation of 6‐phenyl‐4H‐1,3‐dioxin‐4‐one derivatives. However, it reacts with aromatic aldehydes with electron‐withdrawing substituents to yield 1,3‐diaryl‐3‐hydroxypropan‐1‐ones, accompanied by chalcone derivatives in some cases. It did not react with linear ketones, aliphatic aldehydes, and aromatic aldehydes with strong electron‐donating substituents. A mechanism for the formation of 1,3‐diaryl‐3‐hydroxypropan‐1‐ones and chalcone derivatives is proposed. We also tried to react 1 with other unsaturated compounds, including various olefins and nitriles, and cumulated unsaturated compounds, such as N,N′‐dialkylcarbodiimines, phenyl isocyanate, isothiocyanate, and CS2. Only with N,N′‐dialkylcarbodiimines, the expected cycloaddition took place.  相似文献   

20.
An efficient enantioselective synthesis of 3‐acetoxy transβ‐lactams 7a and 7b via [2+2] cycloaddition reactions of imines 4a and 4b , derived from a polycyclic aromatic amine and bicyclic chiral acid obtained from (+)‐car‐3‐ene, is described. The cycloaddition was found to be highly enantioselective, producing only trans‐(3R,4R)‐N‐azetidin‐2‐one in very good yields. This is the first report of the synthesis of enantiomerically pure transβ‐lactams 7a and 7b with a polycyclic aromatic substituent at N(1) of the azetidin ring.  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