首页 | 官方网站   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 stereoselective rearrangement of tetrahydrofuran or tetrahydropyran rings having a phenylsulfanyl group in an exo position, via the intermediate thiiranium ions, is reported. The 5‐ or 6‐exo‐tet cyclization of hydroxy sulfides gave the kinetic products while the 6‐endo‐tet or 5‐endo‐tet gave the thermodynamic products. The rearrangement of the 5‐exo product to the 6‐endo‐ one is an interesting way for the stereoselective synthesis of substituted tetrahydropyrans.  相似文献   

2.
A highly diastereoselective copper‐catalyzed multicomponent cyclization of exocyclic enol ethers/enamines with methylene malonate and aldehydes has been developed to furnish spiroheterocyclic tetrahydropyrans in high yields with greater than 95:5 d.r. This method is practical, in that 36 examples, including a range of aldehydes and exo‐vinyl heterocycles, are presented. By applying the newly developed method, the total synthesis of (+)‐broussonetine G and formal synthesis of (+)‐broussonetine H were achieved in a concise way.  相似文献   

3.
Nucleophilic substitution at the anomeric positions of tetrahydropyranyl (THP) and related carbohydrate‐derived esters that proceeded through pyridinium‐type salt intermediates have been developed. Treatment of the 6‐substituted α‐acetoxy‐tetrahydropyrans with TMSOTf (TMS=trimethylsilyl) and 2‐substitutited pyridines, such as 2‐p‐tolylpyridine and 2‐methoxypyridine, led to the efficient generation of cis‐pyridinium‐type salts. These salts reacted with various nucleophiles, such as alcohols, azides, and organozinc reagents, to form nucleophilic‐substitution products. A characteristic feature of these processes was that they took place under mild conditions, which did not affect acid‐labile protecting groups. Furthermore, the reactions that employed azides and C‐nucleophiles generated 2,6‐trans products with high degrees of stereoselectivity.  相似文献   

4.
The 1JC‐F coupling constant can be useful to probe the conformational landscape of organofluorine compounds and the intramolecular interactions governing the stereochemistry of these compounds. Neighboring oxygen electron lone pairs and a carbonyl group relative to a C─F bond affect this coupling constant in an opposite way, and therefore, analysis of the interactions involving these entities simultaneously indicates which effect dominates 1JC‐F. Spin–spin coupling constant calculations for a series of fluorinated tetrahydropyrans, cyclohexanones, and dihydropyran‐3‐ones indicated that an electrostatic/dipolar interaction between the C─F and C═O bonds is more important than the steric interaction between the C─F bond and the oxygen electron lone pairs. An intuitive consequence of such outcome is that this interaction not only drives the coupling constant but can also be taken into account when aiming at the stereochemical control of functionalized organofluorine compounds.  相似文献   

5.
The 13C NMR spectra of 62 oxanes (tetrahydropyrans) with and without methyl substituents at various ring positions, some of them bearing in addition (or instead) ethyl, vinyl, ethynyl, carbomethoxy and methylol substituents at C-2, have been recorded, and the 294 resulting chemical shifts have been correlated by multiple linear regression analysis. Axial and equatorial α-, β-, γ-, δ-, gem- and vic-parameters for shifts caused by methyl groups at all ring positions, and similar parameters for Et,—CH?CH2,—C?CH, CO2Me and CH2OH groups at C-2, are reported. Standard deviations of the parameters are, in most cases, within 0.3 ppm and the agreement of calculated and experimental shifts is excellent. This is probably the largest parameter set of this type extant. 13C NMR spectra of a number of additional substituted tetrahydropyrans, and of 3,6-dihydro-2H-pyrans and 3,4-dihydro-2H-pyrans, are tabulated and discussed.  相似文献   

6.
Novel, specially prepared, tetrahydropyran‐based γ‐amino alcohols (S)‐2‐(aminomethyl)‐3‐hydroxy‐6‐ethoxy(phenoxy)‐tetrahydropyrans ( I ) (amino = n‐Bu2N, piperidinyl, pyrrolidinyl, azetidinyl) were tested as catalysts in the asymmetric addition of Et2Zn and n‐Bu2Zn to (hetero)aromatic aldehydes. In most cases the phenoxy derivatives of I acted more enantioselectively than the ethoxy ones. The dibutylamino derivaties showed the least enantioselectivity; the pyrrolidinyl derivatives were more active as catalysts than piperidinyl and azetidinyl compounds. The highest enantioselectivity was observed in the addition of Et2Zn to benzaldehyde in the presence of (S)‐2‐(N‐pyrrolidinylmethyl)‐3‐hydroxy‐6‐phenoxytetrahydropyran. The corresponding alcohol was prepared with 72% ee (R‐configuration). The addition of dibutylzinc proceeded slowly and less selectively. The alkylation of (hetero)aromatic aldehydes with Et2Zn and n‐Bu2Zn was also studied in the presence of the known optical inductor (1S,2R)‐N,N‐dibutylnorephedrine. Some chiral aromatic secondary alcohols were synthesized in high chemical yields and up to 93% ee enantioselectivity. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

7.
The 2-(1-aziridinyl)tetrahydropyrans cannot be obtained by aminolysis of the 2-dialkylaminotetrahydropyrans. These compounds are synthesized by the Hofmann reaction between 2-chlorotetrahydropyrans and ethylenimine. The effect of acetic anhydride on the 2-(1-aziridinyl) and 2-dialkylaminotetrahydropyrans was studied. The 2-dimethylaminotetrahydropyrans were converted into 2-acetoxytetrahydropyrans and dimethylacetamide by cleavage of the carbon-nitrogen bond. Ring opening of the oxygen heterocycle gives the corresponding dialkylacetamides and polymerised products. The 2-(1-aziridinyl)tetrahydropyrans allow the synthesis of the N-(2-tetrahydropyrannyl)-N-(2-acetoxy-ethyl)acetamides. Acetic anhydride ring opens the nitrogen heterocycle.  相似文献   

8.
ABSTRACT

The anomeric effect1,2 is a well established phenomenon among carbohydrate derivatives, and more generally among 2-substituted tetrahydropyrans. The generalized anomeric effect1 refers to structural moieties R-X-C-Y (where X possesses one or two lone pairs of electrons, and Y is an atom or group of high electronegativity) in both open chain or cyclic molecules. Thus hydroxyacetonitrile prefers to exist in the gauche conformation,3,4 and several 3-cyano-perhydro-1,2-oxazine derivatives prefer conformations in which the C-3--CN bond and the lone pair of the nitrogen are coplanar.5 Investigation of conformational equilibria of all but one 2,6-anhydro-hexononitrile6,7 (pentopyranosyl cyanides) left doubt relating to the anomeric effect of the cyano group, and raised the possibility of a stabilizing interaction between 1,3-diaxial cyano and acetoxy groups.  相似文献   

9.
An unprecedented protocol has been developed for the efficient synthesis of substituted tetrahydropyrans via a bismuth‐promoted Prins cyclization of imines with homoallyl alcohols. In the presence of 40 mol% BiCl3, a wide variety of imines react smoothly with homoallyl alcohols at room temperature to give the corresponding 4‐chlorotetrahydropyran derivatives in good to excellent yields.  相似文献   

10.
The different factors that control the alkene Prins cyclization catalyzed by iron(III) salts have been explored by means of a joint experimental–computational study. The iron(III) salt/trimethylsilyl halide system has proved to be an excellent promoter in the synthesis of crossed all‐cis disubstituted tetrahydropyrans, minimizing the formation of products derived from side‐chain exchange. In this iron(III)‐catalyzed Prins cyclization reaction between homoallylic alcohols and non‐activated alkenes, two mechanistic pathways can be envisaged, namely the classical oxocarbenium route and the alternative [2+2] cycloaddition‐based pathway. It is found that the [2+2] pathway is disfavored for those alcohols having non‐activated and non‐substituted alkenes. In these cases, the classical pathway, via the key oxocarbenium ion, is preferred. In addition, the final product distribution strongly depends upon the nature of the substituent adjacent to the hydroxy group in the homoallylic alcohol, which can favor or hamper a side 2‐oxonia‐Cope rearrangement.  相似文献   

11.
2-Alkoxy-3-hydroxy tetrahydropyrans undergo ring contraction as observed with 3-hydroxy tetrahydropyrans which, however, initiates a series of fragmentation sequences (e.g., ethylene elimination).

Mit 3 Abbildungen

3. Mitt. s.  相似文献   

12.
The induced decomposition of t-butyl hex-5-enyl peroxide in good hydrogen donor solvents led to 2-substitude tetrahydropyrans and adduct peroxides. The presence of substituents on the hexenyl moinety influenced seriously the relative ratio of the heterocycle and the adduct peroxide.  相似文献   

13.
A comprehensive study of a diastereoselective Rh-catalyzed cyclization of terminal and internal allenols is reported. The methodology allows the atom economic and highly syn-selective access to synthetically important 2,4-disubstituted and 2,4,6-trisubstituted tetrahydropyrans (THP). Furthermore, its utility and versatility are demonstrated by a great functional-group compatibility and the enantioselective total synthesis of (−)-centrolobine.  相似文献   

14.
Homoallylic alcohols, carbonyl compounds and nitriles undergo a smooth tandem Prins-Ritter type cyclization in the presence of CeCl3·7H2O/AcCl at ambient temperature to produce 4-amido tetrahydropyrans in high yields with all cis-selectivity. Spirocyclic 4-amido tetrahydropyrans are obtained in the case of cyclic ketones.  相似文献   

15.
The possible biogenetic connection of the 10‐membered lactone (decanolide) and the tricyclic [5,5]‐spiroacetal‐cis‐fused‐γ‐lactone (SAFL) natural products was proposed by Hanson in 1985. Mimicking such biosynthetic hypothesis led us to develop a general synthetic strategy for the total syntheses of both decanolide‐type cephalosporolides and SAFL‐type natural products. This biomimetic strategy features two key transformations: i) oxidative ring expansion of bicyclic β‐hydroxy tetrahydropyrans to the decanolides and ii) ring contraction rearrangement of the decanolides to the tricyclic SAFLs. In particular, the phenol derivatives are exploited as the starting materials for the decanolides and then the SAFLs. The successful biomimetic total synthesis allows us to revise the structures and biosynthetic hypothesis of several natural products, along with development of an NMR analysis method for determination of the relative stereochemistry of SAFL‐type compounds. In addition, this account will summarize recent synthetic work by other research groups.  相似文献   

16.

2-Sulfinyl dienols undergo an efficient base-promoted cyclization to produce sulfinyl dihydropyrans with creation of two asymmetric centers. The configurational stability of the allylic sulfoxide in most cases is noteworthy. Simple manipulations lead to a variety of functionalized tetrahydropyrans.  相似文献   

17.
A propionate-derived polyketide building block A whose 2-methyl-1,3-diol moiety was built by a Ti(III)-mediated ring opening reaction of a trisubstituted 2,3-epoxy alcohol precursor was employed as a common starting material for the syntheses of highly substituted tetrahydropyrans 1-5, the first one being the C20-C28 fragment of cytotoxic natural products, phorboxazoles.  相似文献   

18.
O‐heterocycles have wide applications, and their efficient and green synthesis is very interesting. Herein, we report hydrogen‐bonding catalyzed ring‐closing metathesis of aliphatic ethers to O‐heterocycles over ionic liquid (IL) catalyst under metal‐ and solvent‐free conditions. The IL 1‐butylsulfonate‐3‐methylimidazolium trifluoromethanesulfonate ([SO3H‐BMIm][OTf]) is discovered to show outstanding performance, better than the reported catalysts. An interface effect plays an important role in mediating the reaction rate due to the immiscibility between the products and the IL catalyst, and the products can be spontaneously separated. NMR analysis and DFT calculation suggest that a pair of cation and anion of [SO3H‐BMIm][OTf] could form three strong H‐bonds with an ether molecule, which catalyze the ether transformation via a cyclic oxonium intermediate. A series of O‐heterocycles including tetrahydrofurans, tetrahydropyrans, morpholines and dioxane can be obtained from their corresponding ethers in excellent yields (e.g., >99 %). This work opens an efficient and metal‐free way to produce O‐heterocycles from aliphatic ethers.  相似文献   

19.
An efficient one‐pot synthesis of 3‐[(4,5‐dihydro‐1H‐pyrrol‐3‐yl)carbonyl]‐2H‐chromen‐2‐one (=3‐[(4,5‐dihydro‐1H‐pyrrol‐3yl)carbonyl]‐2H‐1‐benzopyran‐2‐one) derivatives 4 by a four‐component reaction of a salicylaldehyde 1 , 4‐hydroxy‐6‐methyl‐2H‐pyran‐2‐one, a benzylamine 2 , and a diaroylacetylene (=1,4‐diarylbut‐2‐yne‐1,4‐dione) 3 in EtOH is reported. This new protocol has the advantages of high yields (Table), and convenient operation. The structures of these coumarin (=2H‐1‐benzopyran‐2‐one) derivatives, which are important compounds in organic chemistry, were confirmed spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS) and by elemental analyses. A plausible mechanism for this reaction is proposed (Scheme 2).  相似文献   

20.
The 3-C3strategy involves (i) Pd(0)/SnX2 (X = Cl, Br) mediated generation of allyltin(IV) from allyl bromide in anhydrous DCM, (ii) formation of homoallyloxytin(IV) intermediate I from allyltin(IV) and an aldehyde, and (iii) coupling of I with an aldehyde, an aryl epoxide or an arene as the third partner to afford tetrahydropyrans, benzyl tetrahydropyrans or 4,4-diarylbut-1-enes, respectively.  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-23

京公网安备 11010802026262号