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 100 毫秒
1.
The synthesis of methyl (2S,4R)‐4‐(benzyloxy)‐N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)prolinate ( 10 ), a novel 2H‐azirin‐3‐amine (`3‐amino‐2H‐azirine'), is described (Scheme 1). The reaction of methyl (2S,4R)‐N‐(2‐methylpropanoyl)‐4‐(benzyloxy)prolinate ( 7 ) with Lawesson reagent gave methyl (2S,4R)‐4‐(benzyloxy)‐N‐[2‐(methylthio)propanoyl]prolinate ( 8 ) and consecutive treatment with COCl2, 1,4‐diazabicyclo[2.2.2]octane (DABCO), and NaN3 led to 10 . The use of 10 as a building block of the dipeptide Aib‐Hyp (Aib=2‐aminoisobutyric acid, Hyp=(2S,4R)‐4‐hydroxyproline) is demonstrated by the syntheses of several model peptides (Scheme 2 and Table). The benzyl protecting group of the 4‐OH function in Hyp in the model peptides has been removed in good yields.  相似文献   

2.
A series of new C2‐symmetric (1S,2S)‐cyclohexane‐1,2‐dicarboxamides was synthesized from (1S,2S)‐cyclohexane‐1,2‐dicarbonyl dichloride and N‐benzyl‐substituted aromatic amines, which were prepared from 2‐aminopyridine, 2‐chloroaniline, and 2‐aminophenol via imine formation with benzaldehyde and subsequent reduction with NaBH4. (1S,2S)‐N,N′‐Dibenzyl‐N,N′‐bis[2‐(benzyloxy)phenyl]cyclohexane‐1,2‐dicarboxamide was converted to (1S,2S)‐N,N′‐dibenzyl‐N,N′‐bis(2‐hydroxyphenyl)cyclohexane‐1,2‐dicarboxamide via hydrogenolysis in the presence of Pd(OH)2 on active carbon powder.  相似文献   

3.
徐晖  叶飞  代会苓 《中国化学》2008,26(8):1465-1468
本文报道了一种在离子液[Bmim]BF4中,由无水碳酸钾参与的串联甲磺酸取代苯酯脱保护和芳香亲核取代反应来合成不对称二芳醚的方法。离子液可以循环使用多次,而且此实验操作具有便捷、实用与环境友好的特点。  相似文献   

4.
Twelve novel compounds of 2‐acetyl‐1H‐benzimidazole oxime‐ethers and 2‐acetyl‐6‐chloro‐1H‐benzimidazole oxime‐ethers were synthesized with o‐phenylenediamine (or 4‐chloro‐o‐phenylenediamine), 2‐hydroxypropyl acid, alkoxy (or benzyloxy) amines hydrochloride as starting materials. The structures of the target compounds were characterized by IR, 1H NMR spectra and elemental analyses. The in vitro fungicidal activities against Botrytis cinerea Pers and Alternaria alternata were also evaluated by mycelium growth rate method. The results indicate that the compounds 3b , 3c , 3f , 3g and 3h exhibit good activities against Botrytis cinerea Pers, while 3b and 3f possess excellent activities against Alternaria alternate, and their fungicidal activities are all higher than that of carbendazim.  相似文献   

5.
The multicomponent domino Knoevenagel hetero‐Diels? Alder hydrogenation process of N‐[(benzyloxy)carbonyl(Cbz)‐protected amino aldehydes with N,N‐dimethylbarbituric acid and the trimethylsilyl enol ethers 1 – 3 leads to the formation of the substituted pyrrolidines 12 – 15 . Under the same conditions, reaction of the trimethylsilyl enol ether 4 , obtained from acetophenone, gave the primary amines 18a , b probably due to a hydrogenolytic cleavage of the intermediately formed pyrrolidines. The zwitterionic products were obtained in high purity simply by precipitation with Et2O.  相似文献   

6.
Only one out of the four possible trans isomers of the important perfumery alcohol Norlimbanol® ( 1 ) possesses a very strong amber‐woody smell, the isomer 1A with (1′R,3S,6′S) absolute configuration. Its enantiomer 1B is almost odorless and devoid of amber‐woody character, whereas the diastereoisomers 1C and 1D are considerably weaker and perceptible only by the most‐sensitive persons. The same is true for a whole series of perceptual analogs of 1 , including β‐alkoxy alcohols. These ethers belong to two structural classes: [(2,2,6‐trimethylcyclohexyl)oxy]‐ (see 3, 4 , and 16 ) or {[2‐(tert‐butyl)cyclohexyl]oxy}alkan‐2‐ol derivatives (see 19 and 20 ; Table). A superimposition model allowing for good overlap of the respective hydroxylated side chains offers a tentative explanation for the shared perceptual characteristics of the two classes (Fig. 5). The lipophilic cyclohexane moieties present only a minimal overlap in this model, suggesting that quite larger molecules might possess the same smell. (S)‐Configured β‐alkoxy alcohols can conveniently be obtained on a larger scale by enantioselective reduction of the corresponding ketones (Scheme 9).  相似文献   

7.
Treatment of ethyl (E)‐5,5‐bis[(benzyloxy)methyl]‐8‐(N,N‐diethylcarbamoyl)‐2‐octen‐7‐ynoate with an iron reagent generated from FeCl2 and tBuMgCl in a ratio of 1:4 (abbreviated as FeCl2/4 tBuMgCl) afforded ethyl [4,4‐bis[(benzyloxy)methyl]‐2‐[(E)‐(N,N‐diethylcarbamoyl)methylene]cyclopent‐1‐yl]acetate in good yield. Deuteriolysis of an identical reaction mixture afforded the bis‐deuterated product ethyl [4,4‐bis[(benzyloxy)methyl]‐2‐[(E)‐(N,N‐diethylcarbamoyl)deuteriomethylene]cyclopent‐1‐yl]deuterioacetate, thus confirming the existence of the corresponding dimetalated intermediate. The latter intermediate can react with halogens or aldehydes to facilitate further synthetic transformations. The amount of FeCl2 was reduced to catalytic levels (10 mol % relative to enyne), and catalytic cyclizations of this sort proceeded with yields comparable to those of the aforementioned stoichiometric reactions. The cyclization of diethyl (E,E)‐2,7‐nonadienedioate with a stoichiometric amount of FeCl2/4 tBuMgCl, followed by the addition of sBuOH as a proton source, afforded a mixture of 2‐(ethoxycarbonyl)‐3‐bicyclo[3.3.0]octanone and its enol form in good yield. The use of aldehyde or ketone in place of sBuOH afforded 2‐(ethoxycarbonyl)‐3‐bicyclo[3.3.0]octanone, which has an additional hydroxyalkyl side chain. Additionally, the metalation of a carbon–carbon unsaturated bond in N,N‐diethyl‐5,5‐bis[(benzyloxy)methyl]‐7,8‐epoxy‐2‐octynamide or (E)‐3,3‐dimethyl‐6‐(N,N‐diethylcarbamoyl)‐5‐hexenyl p‐toluenesulfonate with FeCl2/4 tBuMgCl or FeCl2/4 PhMgBr was followed by an intramolecular alkylation with an epoxide or alkyl p‐toluenesulfonate to afford 5,5‐bis[(benzyloxy)methyl]‐3‐[(E)‐(N,N‐diethylcarbamoyl)methylene]‐1‐cyclohexanol or N,N‐diethyl(3,3‐dimethylcyclopentyl)acetamide after hydrolysis. In both cases, the remaining metalated portion α to the amide group was confirmed by deuteriolysis and could be utilized for an alkylation with methyl iodide.  相似文献   

8.
An efficient and versatile method for the stereospecific construction of α,ω-cis- and α,ω-trans-disubstituted eight- and nine-membered cyclic ethers was developed. Cyclization of the hydroxy epoxides promoted by Eu(fod)3 proceeded via an SN2 process and exo mode to provide the corresponding cyclic ethers in excellent yields.  相似文献   

9.
The formation of (1R)‐1‐methylheptyl phenyl ether from (2S)‐octan‐2‐ol via its isourea derivative (S)‐ 1 follows a borderline mechanism. The intermediacy of a carbocation (see (S)‐ 2 ) can be demonstrated (Scheme 1). However, the extremely high inversion of configuration and the olefinic by‐products are also indicative of an SN2 mechanism.  相似文献   

10.
CHIANG  Liwu  PAN  Sider  LO  Jemmau  YU  Chungshan 《中国化学》2009,27(11):2296-2299
The protected ceramide: N‐((2S,3S,4R)‐3,4‐bis(benzyloxy)‐1‐hydroxyoctadecan‐2‐yl)tetracosanamide, was attempted to introduce a triflate as a leaving group followed by a nucleophilic substitution with azido group in one‐pot manner. Unexpectedly, the oxazole ring formed via a thermodynamically favored intramolecular cyclization was opened to generate the original ceramide by triflic acid. In addition, the residual acid promoted a formylation of the primary hydroxy group in DMF.  相似文献   

11.
Four derivatives of diethylenetriaminepentaacetic acid (=3,6,9‐tris(carboxymethyl)‐3,6,9‐triazaundecanedioic acid (H5dtpa)), potential contrast agents for magnetic resonance imaging (MRI), carrying benzyl groups at various positions of the parent structure were synthesized and characterized by a thorough multinuclear NMR study, i.e., the (S)‐ and (R)‐stereoisomers 1a and 1b of 4‐benzyl‐3,6,9‐tris(carboxymethyl)‐3,6,9‐triazaundecanedioic acid (H5[(S)‐(4‐Bz)dtpa] and H5[(R)‐(4‐Bz)dtpa], the diamide derivative N,N″‐bis[(benzylcarbamoyl)methyl]diethylenetriamine‐N,N′,N″‐triacetic acid (=3,9‐bis[2‐(benzylamino)‐2‐oxoethyl]‐6‐(carboxymethyl)‐3,6,9‐triazaundecanedioic acid; H3[dtpa(BzA)2]; 2 ), and the diester derivative N,N″‐bis{[(benzyloxy)carbonyl]methyl}diethylenetriamine‐N,N′,N″‐triacetic acid (=3,9‐bis[2‐(benzyloxy)‐2‐oxoethyl]‐6‐(carboxymethyl)‐3,6,9‐triazaundecanedioic acid; H3[dtpa(BzE)2]; 3 ). From the 17O‐NMR chemical shift of H2O induced by their dysprosium complexes with ligands 1 – 3 , it was concluded that only one H2O molecule is contained in the first coordination sphere of these lanthanide complexes. The rotational correlation times (τR) of the complexes were estimated from the 2H‐NMR longitudinal relaxation rate of the deuterated diamagnetic lanthanum complexes. The exchange time of the coordinated H2O molecule (τM) was studied through the temperature dependence of the 17O‐NMR transverse relaxation rate. As compared to [Gd(dtpa)]2−, the H2O‐exchange rate is faster for [Gd{(S)‐(4‐Bz)dtpa}]2− and [Gd{(R)‐(4‐Bz)dtpa}]2−‐, slower for [Gd{dtpa(BzA)2}], and almost identical for [Gd{dtpa(BzE)2}]. The analysis of the 1H‐relaxivity of the gadolinium complexes recorded from 0.02 to 300 MHz established that i) the relaxivity of [Gd{dtpa(BzE)2}] is similar to that of [Gd(dtpa)]2−, ii) the slightly slower molecular rotation of [Gd{dtpa(BzA)2}] induces a mild enhancement of its relaxivity, and iii) the marked increase of relaxivity of [Gd{(S)‐(4‐Bz)dtpa}]2− and [Gd{(R)‐(4‐Bz)dtpa}]2− mainly results from an apparently shorter distance between the gadolinium ion and the H2O protons of the coordinated H2O molecule.  相似文献   

12.
The ability of urea anions to react as nucleophiles with alkoxy derivatives of 1,3,7‐triazapyrenes has been investigated. It was found that against all expectations, the products of the substitution of an alkoxy groups (SNipso ) by amino group were isolated in good yields. The reactions proceed in anhydrous dimethyl sulfoxide solution at room temperature. But when anions of the mono‐substituted ureas containing bulky substituents were used, the first products of the earlier unknown SNAr reactions of alkyl carbamoyl amination were obtained.  相似文献   

13.
Ferrocenylamines are excellent catalysts for the enantioselective substitution of unsymmetrical allyl chlorides with dialkylzinc compounds in the presence of a copper(I ) salt. For example, in the reaction shown in Equation (1), the product is formed with 87 % ee and a SN2′:SN2 ratio of 97:3.  相似文献   

14.
The first LA‐catalyzed [3 + 2]IMCC of GDA‐epoxides with carbon‐carbon double bonds has been developed. This provides an efficient and general strategy for construction of bridged oxa‐[n.2.1] skeletons. A novel SN‐like mechanism through a carbon‐carbon bond cleavage of epoxide ring has been proposed.  相似文献   

15.
This article presents the first detailed account of the discovery that substituted epoxides can initiate the carbocationic polymerization of isobutylene. α‐Methylstyrene epoxide (MSE), 2,4,4‐trimethyl‐pentyl‐epoxide‐1,2 (TMPO‐1), 2,4,4‐trimethyl‐pentyl‐epoxide‐2,3 (TMPO‐2), and hexaepoxi squalene (HES) initiated isobutylene polymerization in conjunction with TiCl4. MSE, TMPO‐2, and HES initiated living polymerizations. A competitive reaction mechanism is proposed for the initiation and propagation. According to the proposed mechanism, initiator efficiency is defined by the competition between the SN1 and SN2 reaction paths. A controlled initiation with external epoxides such as MSE should yield a primary hydroxyl head group and a tert‐chloride end‐group. The presence of tert‐chloride end‐groups was verified by NMR spectroscopy, whereas the presence of primary hydroxyl groups was implied by model experiments. Multiple initiation by HES was verified by diphenyl ethylene end‐capping and NMR analysis; the resulting star polymer had an average of 5.2 arms per molecule. A detailed investigation of the reaction mechanism and the characterization of the polymers are in progress. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 444–452, 2000  相似文献   

16.
For the unsymmetrical title compound, 1‐bromo‐1′‐[(2S)‐N‐(1‐hydroxy‐3‐methylbutane‐2‐yl)]‐ferroceneamide, two independent molecules were found in the asymmetric unit. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

17.
Rather than lead to the usual deoxygenation pathway, metal carbenes derived from α‐diazo‐β‐ketoesters undergo three‐atom insertions into epoxides using a combination of 1,10‐phenanthroline and [CpRu(CH3CN)3][BArF] as the catalyst. Original 1,4‐dioxene motifs are obtained as single regio‐ and stereoisomers. A perfect syn stereochemistry (retention, e.r. up to 97:3) is observed for the ring opening, which behaves as an SN1‐like transformation.  相似文献   

18.
Alkyl aryl ethers are an important class of compounds in medicinal and agricultural chemistry. Catalytic C(sp3)?O cross‐coupling of alkyl electrophiles with phenols is an unexplored disconnection strategy to the synthesis of alkyl aryl ethers, with the potential to overcome some of the major limitations of existing methods such as C(sp2)?O cross‐coupling and SN2 reactions. Reported here is a tandem photoredox and copper catalysis to achieve decarboxylative C(sp3)?O coupling of alkyl N‐hydroxyphthalimide (NHPI) esters with phenols under mild reaction conditions. This method was used to synthesize a diverse set of alkyl aryl ethers using readily available alkyl carboxylic acids, including many natural products and drug molecules. Complementarity in scope and functional‐group tolerance to existing methods was demonstrated.  相似文献   

19.
Ahmad Farouk Eweas 《合成通讯》2013,43(10):1541-1552
The ring‐closure metathesis of the diene (2S,3R,4S)‐1‐(tert‐butyldiphenylsilyloxy)‐2,4‐dimethylhex‐5‐en‐3‐yl acrylate produced the dihydropyrone with the correct stereochemistry for Soraphen A synthesis. The C2,C3 stereocenters were introduced by the addition of the (Z)‐crotyl‐n‐butylstannane to the β‐alkoxyaldehyde(S)‐3‐(benzyloxy)‐2‐methylpropanal in presence of TiCl4 as a chelating catalyst to give the desired anti,syn homoallyic alcohol (2S,3R,4S)‐1‐(tert‐butyldiphenylsilyloxy)‐2,4‐dimethylhex‐5‐en‐3‐ol.  相似文献   

20.
Propagation mechanism in the cationic polymerization of alkenyl ethers was investigated through the effect of the bulkiness of alkoxy groups on the steric structure of a polymer. In polymerization with BF3O(C2H5)2 in toluene at ?78°C, trans-propenyl ethers having less bulky alkoxy groups–methyl, ethyl, and benzyl propenyl ethers–produced a stereoregular polymer having a threo-meso structure, and the cis isomer a nonstereoregular one having threo-meso and racemic structures. On the other hand, in the polymerization of propenyl ethers having bulky alkoxy groups–isopropyl and 1-methylpropyl propenyl ethers–the trans isomer yielded a nonstereoregular polymer with threo-meso and racemic structures, and the cis isomer a stereoregular one with a erythro-meso structure. This result suggests that a bulky alkoxy group plays an important role in determining the steric structure of the polymer by repulsion between the alkoxy groups of a growing chain end and of a monomer. The effect of solvent polarity on the steric structure of a polymer was also studied.  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