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.
Glycosidation by the diazirine 1 , the trichloroacetimidate 4 , and the bromide 5 of the altro-diol 2 , possessing an intramolecular H-bond (HO? C(3) to O? C(1)) in solution, but not in the solid state, proceeds with high and complementary regioselectivity. From 2 and 1 , one obtains mostly the 1,2-linked disaccharides 10 and 11 (β-D > α-D ), together with the 1,3-linked isomers 12 and 13 (α-D > β-D ; 1,2-/1,3-linked products ca. 9:1), the demethylated 1,3-linked disaccharides 24–27 , the trisaccharides 19–22 , the lactone azines 23 , and the hydroxyglucal 18 , while 2 reacted with 4 or 5 to yield mostly the 1,3-linked disaccharides (1,2-/1,3-linked products ca. 1:9). The disaccharides were additionally characterized as acetates (→ 14–17, 28–31 ). Yields and stereoselectivity depended upon the donor, stoichiometry, solvent, temperature, and concentration. Glycosidation of the 1,3-linked disaccharides with 1 yielded the trisaccharides 19–22 . Reaction of the β-D -altro-diol 3 with 1 gave the 1,2- and 1,3-linked disaccharides 32/33 and 34/35 in a 1:1 ratio, characterized as the acetates 36–39 , while glycosidation with 5 according to Lemieux proceeded regioselectively (1,2-/1,3-linked products 91:9). The monotosylates 6 and 7 reacted with 1 to yield the anomeric pairs 40/41 , and 42/43 of the tosylated disaccharides; the oxiranes 44 and 45 were not observed.  相似文献   

2.
The α-D -allo-diol 9 possesses an intramolecular H-bond (HO? C(3) to O? C(1)) in solution and in the solid state (Fig. 2). In solution, it exists as a mixture of the tautomers 9a and 9b (Fig. 3), which possess a bifurcated H-bond, connecting HO? C(2) with both O? C(1) and O? C(3). In addition, 9a possesses the same intramolecular H-bond as in the solid state, while 9b is characterized by an intramolecular H-bond between HO? C(3) and O? C(4). In solution, the β-D -anomer 12 is also a mixture of tautomers, 12a and presumably a dimer. The H-bonding in 9 and 12 is evidenced by their IR and 1H-NMR spectra and by a comparison with those of 3–8, 10 , and 11 . The expected regioselectivity of glycosidation of 9 and 12 by the diazirine 1 or the trichloroacetimidate 2 is discussed on the basis of the relative degree of acidity/nucleophilicity of individual OH groups, as governed by H-bonding. Additional factors determining the regioselectivity of glycosidation by 1 are the direction of carbene approach/proton transfer by H-bonded OH groups, and the stereoelectronic control of both the proton transfer to the alkoxy-alkyl carbene (in the σ-plane) and the combination of the thereby formed ions (π-plane of the oxycarbenium ion). Glycosidation of 9 by the diazirine 1 or the trichloroacetimidate 2 proceeded in good yields (75–94%) and with high regioselectivity. Glycosidation of 9 and 12 by 1 or 2 gave mixtures of the disaccharides 14–17 and 18–21 , respectively (Scheme 2). As expected, glycosidation of 12 by 1 or by 2 gave a nearly 1:1 mixture of regioisomers and a slight preference for the β-D -anomers (Table 4). Glycosidation of the α-D -anomer 9 gave mostly the 1,3-linked disaccharides 16 and 17 (α-D β-D ) along with the 1,2-linked disaccharides 14 and 15 (α-D < β-D , 1,2-/1,3-linked glycosides ca. 1:4), except in THF and at low temperature, where the β-D -configurated 1,2-linked disaccharide 15 is predominantly formed. Similarly, glycosidation of 9 with 2 yielded mainly the 1,3-linked disaccharides (1,2-/1,3-linked products ca. 1:3 and α-D /β-D ca. 1:4). Yields and selectivity depend upon the solvent and the temperature. The regioselectivity and the unexpected stereoselectivity of the glycosidation of 9 by 1 evidences the combined effect of the above mentioned factors, which also explain the lack of regio-complementarity in the glycosidation of 9 by 1 and by 2 (Scheme 3). THF solvates the intermediate oxycarbenium ion, as evidenced by the strong influence of this solvent on the regio- and stereoselectivity, particularly at low temperatures, where kinetic control leads to a stereoelectronically preferred axial attack of THF on the oxycarbenium ion.  相似文献   

3.
The regio- and stereoselectivity of the glycosidation of the partially protected mono-alcohols 3 and 7 , the diols 2 and 8 , and the triol 4 by the diazirine 1 have been investigated. Glycosidation of the α-D -diol 2 (Scheme 2) gave regioselectively the 1,3-linked disaccharides 11 and 12 (80%, α-D /β-D 9:1), whereas the analogous reaction with the βD -anomer 8 led to a mixture of the anomeric 1,3- and 1,4-linked disaccharides 13 (12.5%), 14 (16%), 15 (13%), and 16 (20.5%; Table 2). Protonation of the carbene by OH–C(4) of 2 is evidenced by the observation that the α-D -mono-alcohol 3 did not react with 1 under otherwise identical conditions, and that the β-D -alcohol 7 yielded predominantly the β-D -glucoside 18 (52%) besides 14% of 17 . Similarly as for the glycosidation of the diol 2 , the influence of the H-bond of HO? C(4) on the direction of approach of the carbene, the role of HO? C(4) in protonating the carbene, and the stereoelectronic control in the interception of the ensuring oxycarbenium cation are evidenced by the reaction of the triol 4 with 1 (Scheme 3), leading mostly to the α-D -configurated 1,3-linked disaccharide 19 (41%), besides its anomer 20 (16%), and some 4-substituted β-D -glucoside 21 (9%). No 1,6-linked disaccharides could be detected. In agreement with the observed reactivity, the 1H-NMR and IR spectra reveal a strong H-bond between HO? C(3) and the phthalimido group in the α-D -, but not in the β-D -allosides. The different H-bonds in the anomeric phthalimides are in keeping with the results of molecular-mechanics calculations.  相似文献   

4.
Hydrogen bonding of the triol 4 in chlorinated solvents was studied by IR (CH2Cl2 and CCl4) and 1H-NMR spectroscopy (CDCl3), and the regioselectivity of the glycosidation of the triol 4 by the diazirine 1 is predicted on the basis of two assumptions: preferred protonation of the intermediate glycosylidene carbene by the OH group involved in the weakest intramolecular H-bond, and attack in the π-plane of the thereby generated oxycarbenium cation either by the reoriented oxy anion, or by a properly oriented vicinal OH group. Glycosidation led to the disaccharides 5–10 (Scheme) which were separated and characterized as their acetates 11–16 , to the lactone azines 17 and to the 2-(benzyloxy)glucal 18 . In agreement with the predictions, glycosidation in non-coordinating solvents gave the 1,2-, 1,3-, and 1,4-linked disaccharides in decreasing relative amounts. Glycosidation in THF proceeded with a lower degree of regioselectivity and led preferentially to the β -D -anomers, except for the minor, 1,4-linked disaccharides, where THF had only a weak influence on stereoselectivity at room temperature and led to a slight increase of the α -D -anomer at ?80°.  相似文献   

5.
Aryl-halo-diazirines react under basic conditions with 1,3-cis-, 1,2-cisand 1,2-trans-diols to give acetals. Yields are high. Diastereoselectivities depend upon the diol and upon the reaction conditions. Thus, reaction of the 1,3-cis-diol 1 (Scheme 1) with 2 gave 3 as a single diastereoisomer. The 1,2-cis-diols 4 and 7 led to the endo- and exo-acetals 5 / 6 (93:7) and 8 / 9 (ca.10:1), respectively, The 1,2-trans-diols 10 , 16 , and 19 reacted with 2 to afford 11 / 12 (90:10), 17 / 18 (1:1), and 20 / 21 (6:1), respectively. Reaction of the (4-nitrophenyl)diazirine 13 with 10 at higher temperatures yielded 14 / 15 (6:4). The uracil moiety, the acetamido group, and the enol-ether moiety are compatible with the reaction conditions. The diastereoselectivity is rationalized on the basis of a reaction sequence involving alkoxy-halogen exchange, which is regioselective or not, thermolysis of the ensuing alkoxydiazirine(s), protonation of the alkoxycarbene to form an (E)-configurated oxycarbenium ion, and attack of the neighboring oxy or hydroxy group, which is only possible for a limited range of conformers.  相似文献   

6.
The H-bonds of the enantiomeric ribosides 4 and 5 and their glycosidation by the diazirine 1 are described. HO–C(2) and HO? C(4) of 4 and 5 form a ‘flip-flop’ H-bonding system, with HO? C(3) acting as a H-bond donor to O? C(2) or O? C(4). HO? C(2) and HO? C(4) of monomeric 4 and 5 are thus the most strongly acidic OH groups. Glycosidation of 4 and 5 by 1 depends on the solvent, the temperature, and the concentration. It yields up to 91% of a mixture of anomeric pairs of the 1,2-, 1,3-, and 1,4-linked disaccharides 8–13 and 20–25 , respectively, which were characterized as their diacetates 14–19 and 26–31 (Scheme). Glycosidation in CH2Cl2 and in dioxane yielded mostly the 1,3-linked disaccharides 10/11 and 22/23 (α/β ca. 4:1), while glycosidation in THF leads mostly to the 1,2- and 1,4-linked regioisomers (β>α). There are small, but significant differences in the glycosidation of 4 and 5 . These, the regio-, and the stereoselectivities are rationalized as the consequences of the stereoelectronic control of both the H-transfer from HO? C(2) or HO? C(4) to the intermediate carbene and of the formation of the glycosidic C? O bond, and of the coordination of the intermediate oxycarbenium ion with THF.  相似文献   

7.
Insertion of the glycosylidene carbenes, derived from the gluco- and the manno-diazirines 1 and 2 , into the Sn? H bond of R3SnH (R = Bu or Ph) leads to the fully substituted stannoglycosides 3 – 8 (53–77%). The 1,2-cis-configurated products are formed preferentially (α-D /β -D ranges from 2.5:1 to 5.1:1 with 1 , and 1:1.3 to 1:4.2 with 2 ). Relative to CH2Cl2, THF favors formation of the equatorial Sn-glycosides. The stannylated (benzyloxy)glucals 9 and 10 were isolated as side products. The reaction of 1 with (Bu3Sn)2 yielded 9 (17% in CH2Cl2; 36% in CCl4) together with the azines 11 and the benzyloxyglucal 12 . NMR Data of the Sn-glycosides 3 – 8 show evidence for an anomeric effect, 1J(C(1),H) being larger in the axial and 1J(Sn,C(1)) larger in the equatorial anomers.  相似文献   

8.
Glycosidation of the myo-inositol derivatives 2 and 3 by the diazirine 1 yields 90% of a diastereoisomer pair of β-D -glycosides in a 1:1 ratio, i.e. 5/6 and 7/8 , respectively (Scheme 1). The crystal structure of 3 shows a strong intramolecular H-bond, which persists in solution, as indicated by FT-IR and 1H-NMR spectra. Yields and diastereoselectivity are lower for the glycosidation of 24 by 1 (Scheme 3). The resulting 1,2- and 1,4-linked disaccharides 25–28 were isolated as their acetates 29–32 . The previously determined crystal structure of 24 shows no intramolecular H-bonds. The yield of the glycosidation of 24 , but not of 3 , depends upon the concentration, indicating that activation of 24 by intermolecular H-bonds is required. Glycosidation of 2 and 3 with the trichloroacetimidate 14 gave mixtures of four ( 5,6,15 , and 16 ), and six ( 7,8 , and 17–20 ) disaccharides, respectively (Scheme 2).  相似文献   

9.
A synthesis of 1,2-cis-configurated, non-isosteric phosphonate analogues of aldose-1-phosphates is described. Treatment of 1-O-acyl-glycoses 1 , 7 , 13 , and 19 with trialkyl phosphite in the presence of trimethylsilyl trifluoromethanesulfonate gave the 1,2-cis-configurated glycosylphosphonates 2 , 4 , 8 , 10 , 14 , 16 , 20 , and 22 as the major anomers and the 1,2-trans-configurated glycosylphosphonates 3 , 5 , 9 , 11 , 15 , 17 , 21 , and 23 as the minor anomers. The 1,2-cis-configurated phosphonates 4 , 10 , 16 , and 22 were deprotected to give the (β-D -glucopyranosyl)phosphonate 6 , the (β-D -mannopyranosyl)phosphonate 12 , the (β-D -ribofuranosyl)phosphonate 18 , and the (β-D -arabinofuranosyl)phosphonate 24 , respectively, in high yields. The preferred formation of 1,2-cis-configurated phosphonates is explained by postulating an equilibrium between the anomeric phosphonium-salt intermediates (such as 25 and 26 ) and a stabilization of the cis-configurated salts through formation of a pentacoordinated species (such as 28 ).  相似文献   

10.
Asymmetric aerobic oxidation of a range of meso- and prochiral diols with chiral bifunctional Ir catalysts is described. A high level of chiral discrimination ability of CpIr complexes derived from (S,S)-1,2-diphenylethylenediamine was successfully demonstrated by desymmetrization of secondary benzylic diols such as cis-indan-1,3-diol and cis-1,4-diphenylbutane-1,4-diol, providing the corresponding (R)-hydroxyl ketones with excellent chemo- and enantioselectivities. Enantiotopic group discrimination in oxidation of symmetrical primary 1,4- and 1,5-diols gave rise to chiral lactones with moderate ees under similar aerobic conditions.  相似文献   

11.
The first syntheses of four stereoisomeric diaminocyclooctane diols, as well as a chlorocyclooctane aminodiol, are reported. In the first part, photooxygenation of cis,cis-1,3-cyclooctadiene gave a bicyclic endoperoxide, which was reduced with zinc followed by mesylation of the hydroxyl groups. Treatment with sodium azide afforded 1,4- and 1,2-cyclooctene diazides. Oxidation of the double bonds in the isomeric diazides with OsO4, followed by hydrogenation of the azide groups, led to 3,8-diaminocyclooctane-1,2-diol and 3,4-diaminocyclooctane-1,2-diols. In the second part, cis-3,8-diazidocyclooctene was converted into the corresponding epoxide. Stereospecific hydrolysis of the epoxide ring with HCl(g) in methanol, and hydrogenation of the azide groups gave 3,8-diamino-2-chloro-cyclooctan-1-ol. Bromination of the double bond in cyclooctene diacetate, followed by acetate deprotection, azidolysis of the bromides, and hydrogenation of the azide groups resulted in the formation of 2,3-diaminocyclooctane-1,4-diol.  相似文献   

12.
两种邻、间手性二醇的区域选择性取代反应研究   总被引:1,自引:0,他引:1  
对两种邻、间手性二醇的区域选择性取代反应进行了研究. 以(R)-(-)-1,3-丁二醇(1)为起始原料, 在各种有机碱存在下, 先合成C1-位单磺酸酯, 进行C1-位的单醚化取代, 均有非区域选择性的单(双)取代物生成. 若用SOCl2处理1, 使其首先生成环状亚硫酸酯中间体5, 苯硫酚在2% Na2CO3存在下可和其发生高区域选择性取代反应, 生成 (R)-(-)-3-羟基-丁基苯硫醚(3). 以同样的方法使(S,R)-(+)-1-(6-氟-2-色满基)-1,2-乙二醇(6)和SOCl2反应生成环状亚硫酸酯(9), 再和苄胺反应, 也可发生高区域选择性取代, 生成 (S,R)-(+)-2-苄氨基-1-(6-氟-2-色满基)-乙醇(8). 该反应方法具有高区域选择性, 产物有高旋光纯度、高产率.  相似文献   

13.
Reaction of cyclic sulfates of vic-diols with magnesium iodide in acetonitrile produced the corresponding olefins in excellent isolated yields at room temperature. Cyclic sulfates of trans-diols gave trans-alkenes exclusively. Cyclic sulfates of cis-diols gave a mixture of cis- and trans- alkenes. However, the cyclic sulfate of cyclic cis-diol afforded the corresponding cis-alkene only.  相似文献   

14.
In 1,3-disubstitued cyclohexanes, in general, the diaxial conformation of the cis isomer is, energetically, the least favored conformation. An interspacial electronic interaction in the ground state of a cis-1,3-disubstituted cyclohexane would be expected to increase the proportion of this conformer in the equilibrium mixture. Such an interaction would provide an energetically favorable pathway for cyclopolymerization. From nuclear magnetic resonance studies on cis-and trans-1,3-diisocyanatocyclohexane the conformational equilibrium in the cis isomer was determined. It is shown that cis-1,3-diisocyanatocyclohexane exists in the diequatorial conformation; this is taken as evidence that a ground-state interaction between isocyanato groups in this monomer, which readily cyclopolymerizes, is not a significant factor in the cyclopolymerization mechanism. The value of the free energy barrier, ΔG?, for trans-1,3-diisocyanatocyclohexane was calculated as 11.1 kcal/mole.  相似文献   

15.
The chiral center in ricinoleic acid methyl ester (ee ~100%) strongly affects the regioselectivity of its hydroboration-oxidation, so that the resulting 1,3-diol dominates by 74% over the 1,4-isomer. Furthermore, new asymmetric centers are formed preferentially with (S)-configuration, up to 87% for 1,3-diols and up to 100% for 1,4-diols.  相似文献   

16.
Reduction of l-hydroxy-3-imidazoline-3-oxides with NaBH4 gives 1,3-dihydroxyimidazolidines. The latter on subsequent treatment with NH2OH · HCl in EtOH or in hydrochloric acid afford 1,2-bishydroxylamines. An X-ray study of 1,2-bishydroxylaminocycloalkanes has shown them to becis-configurated.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 896–901, May, 1993.  相似文献   

17.
Glycosylidene carbenes derived from the GlcNAc and AllNAc diazirines 1 and 3 were generated by the thermolysis or photolysis of the diazirines. The reaction of 1 with i-PrOH gave exclusively the isopropyl α-D -glycoside of 5 besides some dihydrooxazole 9 (Scheme 2). A similar reaction with (CF3)2CHOH yielded predominantly the α-D -anomer of 6 , while glycosidation of 4-nitrophenol (→ 7 ) proceeded with markedly lower diastereoselectivity. Similarly, the Allo-diazirine 3 gave the corresponding glycosides 12–14 , but with a lower preference for the α-D -anomers (Scheme 3). The reactions of the carbene derived from 1 with Ph3COH (→ 8 ) and diisopropylideneglucose 10 (→ 11 ) gave selectively the α-D -anomers (Scheme 2). The αD -selectivity increases with increasing basicity (decreasing acidity) of the alcohols. It is rationalized by an intermolecular H-bond between the acetamido group and the glycosyl acceptor. This H-bond increases the probability for the formation of a 1,2-cis-glycosidic C–O bond. The gluco-intermediates are more prone to forming a N–H…?(H)OR bond than the allo-isomers, since the acetamido group in the N-acetylallosamine derivatives forms an intramolecular H-bond to the cis-oriented benzyloxy group at C(3), as evidenced by δ/T and δ/c experiments.  相似文献   

18.
In the presence of InI, Ni(acac)2 and PPh3, several 1,3-dienes were reacted with two molecules of aldehyde to give the corresponding 1,4- and 1,6-diols. The regioselectivity of the 1,4-/1,6-diol was efficiently regulated by the addition of water; the 1,6-diol was obtained selectively in dry THF, whereas the 1,4-diol was obtained predominantly in DMI containing a small amount of water.  相似文献   

19.
The microstructure of diene units was investigated in radical homopolymers of the cis and trans isomers of 1,3-pentadiene and copolymers with acrylonitrile, synthetized in bulk and emulsion. Experiments were carried out by infrared spectroscopy, 100 MHz 1H-NMR, and 25 MHz 13C-NMR studies. No difference between the bulk and emulsion samples was noted. The microstructure of poly(1,3-pentadiene) is practically independent of the cis or trans configuration of the diene monomer and is as follows: 56–59% trans-1,4, 15–17% cis-1,4, 16–20% trans-1,2 7–10% cis-1,2 and 0% 3,4. On the other hand, up to about 30% of incorporated acrylonitrile (10% in the feed), the microstructure of the pentadiene fraction in the copolymers is not affected. This finding suggests that the penultimate unit has very little influence on the polymerization process involving the terminal pentadienly unit. Beyond 10% of acrylonitrile in the feed, the proportions of the structural units were linearly dependent upon the acrylonitrile content: trans-1,4 content increased whereas the amounts of cis-1,4 trans-1,2 and cis-1,2 decreased (except the cis-1,2 fraction, constant in the copolymers from the cis-diene). These results are discussed on the assumption that the microstructure of pentadiene residues is strongly associated with the acrylonitrile comonomer in the feed.  相似文献   

20.
Summary. The transesterification of chiral (−)-pinanediol methylboronic ester was studied with various structurally modified diols by 1H NMR to understand the factors influencing the unusual stability of this boronic ester as well as to find ways of recovering pinanediol from its methylboronic ester. In all the cases, reactions were allowed to proceed to equilibrium. The preliminary experiments indeed have shown some encouraging results (displacement of pinanediol up to 40–53%). Amongst cyclopentane-based cis-1,2-diols, endo-2-phenyl-exo,exo-2,3-norbornane-diol appeared to be the most effective diol in displacing pinanediol (38%). In the cases of pinane-based diols, the best result was obtained with 2-ethyl-6,6-dimethylbicyclo[3.1.1]heptane-cis-2,3-diol (53%). It was interesting to observe that the transesterification with 2-phenyl-6,6-dimethylbicyclo[3.1.1]heptane-cis-2,3-diol resulted in a 50% conversion after 4 days only, whereas the former diol took 24 days to reach equilibrium.  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