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 46 毫秒
1.
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   

2.
The design of 6‐azido‐6‐deoxy‐l ‐idose for use as a hetero‐bifunctional spacer is reported. The hemiacetal at one terminus is an equivalent of an aldehyde and can react with nucleophiles, such as amino groups and electron‐rich aromatics. The azido group at the other terminus bio‐orthogonally undergoes a Hüisgen [3+2] cycloaddition with an acetylene. The idose derivative exhibited a higher level of reactivity towards oxime formation than a corresponding glucose derivative. The 13C NMR spectrum of the uniformly 13C‐labeled 6‐azido‐idose indicated that the acyclic forms of the sugar totaled 0.3 % of all the isomers, whereas those of glucose totaled 0.01 %. The larger population of the acyclic forms of the idose derivative would result in higher reactivity towards electrophilic addition in comparison with glucose derivatives. Finally, we prepared a C‐idosyl epigallocatechin gallate (EGCG) that bears an azido group through C‐glycosylation of EGCG with 6‐azido‐idose. This glycosyl form of the C‐idosyl EGCG exhibited a cytotoxicity against U266 cells that was comparable to that of EGCG. These results suggested that the EGCG derivative could be used as an effective chemical probe for the elucidation of EGCG biological functions.  相似文献   

3.
The 1,3‐dipolar cycloaddition of arylnitriloxides on 1,2‐dihydroisoquinoline derivatives led to new 3‐aryl, 3a‐8,9,9a‐tetrahydro[5,4‐c]‐isoxazoloisoquinoline adducts. The regioselectivity of the cycloaddition reactions is discussed on the basis of 1H and 13C NMR data.  相似文献   

4.
A synthetic strategy for catalytic asymmetric conjugate addition‐protonation and diastereoselective switch between 5H‐oxazol‐4‐ones and 5‐methylene 1,3‐oxazolidine‐2,4‐diones was established. An array of chiral conjugate addition‐protonation products bearing 1,3‐O‐heterotertiary‐O‐heteroquarternary nonadjacent stereocenters were obtained in excellent yields, moderate to good diastereoselectivities, and excellent enantioselectivities (up to 97% yield, 11: 1 dr, and 98% ee). Induction by 2,2’‐biphenol could effectively promote the production of the corresponding diastereoisomers via cycloaddition intermedia.  相似文献   

5.
Enantioselective catalytic intermolecular 1,3‐dipolar cycloadditions are powerful methods for the synthesis of heterocycles. In contrast, intramolecular enantioselective 1,3‐dipolar cycloadditions are virtually unexplored. A highly enantioselective synthesis of natural‐product‐inspired pyrrolidino‐piperidines by means of an intramolecular 1,3‐dipolar cycloaddition with azomethine ylides is now reported. The method has a wide scope and yields the desired cycloadducts with four tertiary stereogenic centers with up to 99 % ee. Combining the enantioselective catalytic intramolecular 1,3‐dipolar cycloaddition with a subsequent diastereoselective intermolecular 1,3‐dipolar cycloaddition yielded complex piperidino‐pyrrolizidines with very high stereoselectivity in a one‐pot tandem reaction.  相似文献   

6.
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   

7.
Synthesis of some new class of regioselective spiro isoxazolidine derivatives have been described using N‐benzyl‐C‐fluoro substituted‐phenyl nitrones with new dipolarophiles via 1,3‐dipolar cycloaddition reaction in ionic liquid. The novel spiro cycloadducts found to exhibit good synthetic potentiality as they could be converted into synthetically more important spiro 1,3‐amino alcohols. Simple reaction methodology, noninvolvent of catalysts, good to excellent yields, and greener approaches are the important features noticed in this syntheses.  相似文献   

8.
The reactions of 5‐benzylidene‐3‐phenylrhodanine ( 2 ; rhodanine=2‐thioxo‐1,3‐thiazolidin‐4‐one) with diazomethane ( 7a ) and phenyldiazomethane ( 7b ) occurred chemoselectively at the exocyclic C?C bond to give the spirocyclopropane derivatives 9 and, in the case of 7a , also the C‐methylated products 8 (Scheme 1). In contrast, diphenyldiazomethane ( 7c ) reacted exclusively with the C?S group leading to the 2‐(diphenylmethylidene)‐1,3‐thiazolidine 11 via [2+3] cycloaddition and a ‘two‐fold extrusion reaction’. Treatment of 8 or 9b with an excess of 7a in refluxing CH2Cl2 and in THF at room temperature in the presence of [Rh2(OAc)4], respectively, led to the 1,3‐thiazolidine‐2,4‐diones 15 and 20 , respectively, i.e., the products of the hydrolysis of the intermediate thiocarbonyl ylide. On the other hand, the reactions with 7b and 7c in boiling toluene yielded the corresponding 2‐methylidene derivatives 16, 21a , and 21b . Finally, the reaction of 11 with 7a occurred exclusively at the electron‐poor C?C bond, which is conjugated with the C?O group. In addition to the spirocyclopropane 23 , the C‐methylated 22 was formed as a minor product. The structures of the products (Z)‐ 8, 9a, 9b, 11 , and 23 were established by X‐ray crystallography.  相似文献   

9.
A novel series of (9Z)‐9‐arylmethylidene‐3‐(2,6‐dichlorophenyl)‐5,6‐dihydro[1,3]thiazolo[2′,3′:2,3]imidazo [1,2‐d][1,2,4]oxadiazol‐8(9H)‐one derivatives were prepared in moderate yields by the 1,3‐dipolar cycloaddition reaction of a nitrile oxide with (2Z)‐2‐arylmethylidene‐5,6‐dihydroimidazo [2,1‐b][1,3]thiazol‐3(2H)‐ones. The reaction site of the dipolarphile is the C═N of imidazo[2,1‐b][1,3]thiazole rather than the expected C═C of the arylmethylidene. The product structures were characterized thoroughly by IR, MS, NMR spectroscopy, and elemental analysis. The results indicate that this reaction proceeds with chemoselectivity and regioselectivity.  相似文献   

10.
Addition reactions of acid chlorides with various 2‐substituted 4,5‐dihydro‐4,4‐dimethyl‐5‐(methylsulfanyl)‐1,3‐thiazoles under basic conditions were studied. Two kinds of products were obtained from these additions, β‐lactams and non‐β‐lactam adducts. When the reaction was carried out with 4,5‐dihydro‐1,3‐thiazoles with a Ph substituent at C(2), the reaction proceeded via formal [2+2] cycloaddition and led to the correspoding β‐lactam. On the other hand, acid chlorides and 4,5‐dihydro‐1,3‐thiazoles bearing an α‐H‐atom at the C(2)‐substituent underwent C(α)‐ and/or N‐addition reactions and furnished non‐β‐lactam adducts, i.e., C(α)‐ and/or N‐acylated 1,3‐thiazolidines. The attempted transformations of sulfonyl esters of exo‐6‐hydroxy penams to endo‐6‐azido penams failed, although they were successful with mono‐β‐lactams under the same conditions.  相似文献   

11.
A series of 2‐acyl‐2H‐1,2,3‐diazaphospholes 3 underwent ready 1,3‐dipolar cycloaddition reactions with 9‐diazofluorenes as the 1,3‐dipole, yielding the respective bicyclic phosphiranes 5 or trimers 7 depending on the reaction conditions employed. The reaction is believed to proceed via the formation of the [3+2]‐cycloaddition adducts followed by elimination of nitrogen from the cyclic azo moiety. In the case of 3c , the phosphatetraazabicyclooctadiene compound 6 has been isolated with no loss of nitrogen. Likewise, the dipolar cycloaddition reaction of diphenyldiazomethane with the >C?P‐ moiety as the 1,3‐dipolarophile gave phosphadiazabicyclohexenes 8 in 32–68% yields.  相似文献   

12.
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   

13.
π‐Extended TCBD‐porphyrins that contained a 1,1,4,4‐tetracyanobuta‐1,3‐diene unit were prepared by a highly efficient [2+2] cycloaddition of tetracyanoethene (TCNE) or 7,7,8,8‐tetracyano‐p‐quinodimethane (TCNQ) with meso‐substituted trans‐A2B2‐porphyrins that contained two phenylethynyl groups, followed by a retro‐electrocyclization reaction. Depending on the electronic properties of the arylethynyl groups, the cycloaddition reaction took place exclusively on either one or two ethynyl moieties with high yield. The addition of TCNQ proceeded with complete regioselectivity. The resulting π‐expanded TCBD‐porphyrins had a hypsochromically shifted Soret band and showed unique, broad absorption in the visible region.  相似文献   

14.
Enantiomerically pure α‐oxo diazo compounds derived from (S)‐proline were used for 1,3‐dipolar cycloaddition with aryl and hetaryl thioketones, as well as with cycloalkanethiones. Whereas the reactions with hetaryl thioketones in boiling THF yield α,β‐unsaturated ketones via a cascade of cycloaddition, 1,3‐dipolar electrocyclization, and desulfurization, the analogous reactions with thiobenzophenone and cycloalkanethiones result in the formation of 1,3‐oxathiole derivatives. In the latter case, the 1,5‐dipolar electrocyclization of the intermediate thiocarbonyl ylide is the key step of the reaction sequence. In all cases, the isolated products are optically active, i.e., the multistep processes occur with retention of the stereogenic center incorporated via the use of (S)‐proline as the precursor of the diazo compounds.  相似文献   

15.
We report a synthetic strategy for a chemoselective switch and a diastereo‐divergent approach for the asymmetric reaction of 5H‐oxazol‐4‐ones and N‐itaconimides catalyzed by l ‐tert‐leucine‐derived tertiary amine–urea compounds. The reaction was modulated to harness either tandem conjugate addition–protonation or [4+2] cycloaddition as major product with excellent enantio‐ and diastereoselectivities. Subjecting the enantio‐enriched cycloaddition products to a basic silica gel reagent yields the diastereomer vis‐à‐vis the product directly obtained under conditions for addition–protonation, thus opening a diastereo‐divergent route for creating 1,3‐tertiary‐hetero‐quaternary stereocenters. Quantum chemical studies further provide stereochemical analysis for the [4+2] process and a plausible mechanism for this chemoselective switch is proposed.  相似文献   

16.
The 1,3‐dipolar cycloaddition reactions of nitrilimine with thiazolo[3,2‐a]pyrimidine derivatives was investigated. Bis‐cycloadducts were obtained through a domino 1,3‐dipolar cycloaddition/ring‐opening/ring‐opening/1,3‐dipolar cycloaddition processes. The structures of the products were characterized thoroughly by NMR, IR, MS, elemental analysis together with X‐ray crystallographic analysis.  相似文献   

17.
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   

18.
A concise and efficient approach to the spiro‐tetrahydroisoquinoline derivatives has been developed by 1,4‐dipolar cycloaddition of zwitterions resulting from isoquinoline and acetylene esters and (1,3‐dihydro‐1,3‐dioxo‐2H‐inden‐2‐ylidene)malononitrile in MeCN at room temperature. The significance of this method lies in good yields and ease of product purification, and no inert atmosphere is required. The structures of the products were confirmed spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS) and by elemental analyses. A plausible mechanism for this reaction is proposed (Scheme).  相似文献   

19.
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   

20.
It has been shown previously that the reaction of diazomethane with 5‐benzylidene‐3‐phenylrhodanine ( 1 ) in THF at ?20° occurs at the exocyclic C?C bond via cyclopropanation to give 3a and methylation to yield 4 , respectively, whereas the corresponding reaction with phenyldiazomethane in toluene at 0° leads to the cyclopropane derivative 3b exclusively. Surprisingly, under similar conditions, no reaction was observed between 1 and diphenyldiazomethane, but the 2‐diphenylmethylidene derivative 5 was formed in boiling toluene. In the present study, these results have been rationalized by calculations at the DFT B3LYP/6‐31G(d) level using PCM solvent model. In the case of diazomethane, the formation of 3a occurs via initial Michael addition, whereas 4 is formed via [3+2] cycloaddition followed by N2 elimination and H‐migration. The preferred pathway of the reaction of 1 with phenyldiazomethane is a [3+2] cycloaddition, subsequent N2 elimination and ring closure of an intermediate zwitterion to give 3b . Finally, the calculations show that the energetically most favorable reaction of 1 with diphenyldiazomethane is the initial formation of diphenylcarbene, which adds to the S‐atom to give a thiocarbonyl ylide, followed by 1,3‐dipolar electrocyclization and S‐elimination.  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