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 921 毫秒
1.
Vinyl substituted (1R,2S)-amino alcohols 5 were obtained by addition of vinyl magnesium bromide to the corresponding cyanohydrin O-trimethylsilyl ethers (R)-2. The O- and N-protected vinyl amino alcohols 6 were ozonized at −78°C in methanol yielding (1R,2S)-2-amino-1,3-diols7 in high enantiomeric and diastereomeric excesses. For purification, compounds 7 in some cases were acetylated to give the derivatives (1R,2S)-8. Racemic 6a was converted by oxidative ozonolysis at −78°C in methanolic NaOH solution to the corresponding methyl N-acetyl-β-hydroxy propanoate 9a. The configuration of (1R,2S)-8a was confirmed by x-ray crystallographic analysis.  相似文献   

2.
A practical synthetic method for a xanthine dehydrogenase inhibitor, (S)-(−)-BOF-4272, was established utilizing an asymmetric oxidation of diaryl sulfide BOF-4269. The oxidation of the sulfide with 1-chlorobenzotriazole carried out in the presence of 4-cyanopyridine and chiral 2-phenylcyclohexanol gave a high enantiomeric excess (73%ee). The sulfoxides in each enantiomerically pure form could be obtained by treating with alkaline hydrolysis or thermolysis of one the diastereomeric intermediate sulfonium salts (>99%de). Thus the transformation into the sulfoxides occur with virtually perfect inversion (alkaline hydrolysis) or retention (thermolysis). It is therefore possible to obtain the target sulfoxide, (S)-(−)-BOF-4272, from both the two diastereomeric sulfonium intermediates.  相似文献   

3.
The reaction of racemic chloro-dibenzo[c.e][1,2]oxaphosphorine with (S)-(−)--methyl-benzylamine and (1R,2S,5R)-(−)-menthol led to a mixture of optically active diastereomers of the corresponding phosphonous derivatives. The isomers were separated in a diastereomeric excess of 71–93% at the phosphonic oxide and/or at the phosphonous borane stage. The P(III) boranes are suitable precursors of the P-ligands. The absolute P-configuration in one of the menthyl phosphonates isolated was determined by single crystal X-ray analysis.  相似文献   

4.
A number of ,β-unsaturated sulfinyl chlorides 1 has been separately prepared and treated with (−)-cholesterol under various conditions some of which incorporated chiral amines quinine or quinidine. Some (RS) vinylic sulfinates could be isolated in enantiopure form following one or two recrystallizations of the resulting diastereomeric mixtures of (−)-cholesteryl 1-alkenesulfinates 2. Access to diastereomerically enriched (SS) vinylic sulfinates (66–75% de) was achieved in three instances. Absolute stereochemical assignments were made with the assistance of the chiral solvating agent (R)-2,2,2-trifluoro-1-(9-anthryl)ethanol.  相似文献   

5.
Optically active o-substituted diphenyl N-substituted sulphilimines are readily synthesised by the reaction of the corresponding sulphides and t-butyl hypochlorite in the presence of l-menthol and amide anions. (−)-N-p-Tolylsulphonylsulphilimines (1, 2) obtained were converted to the corresponding (−)-N-unsubstituted sulphilimines (8, 9) by treating them with concentrated sulphuric acid. When (−)-S-o-anisyl S-phenyl N-(unsubstituted) sulphilimine (8) was treated with acylating agents or acrylonitrile, the corresponding optically active (−)-N-substituted sulphilimines were prepared with complete retention at sulphur. The absolute configuration of (−)-S-o-anisyl S-phenyl N-p-tolylsulphonylsulphilimine (1) was determined by converting it to (+)-S-o-anisyl sulphoxide (17). CD curves of (−)-o-substituted diarylsulphilimines exhibited a negative Cotton effect at around 270–285 nm, which was assigned to (S)-configuration at sulphur by comparing with the analogous sulphoxides. The substituent on the imino group of the sulphilimine gave no appreciable effect on the CD behavior and the lack of substituent effect was considered to be due to the semi-polar character of the S(IV)-N bond. Unusual effect of o-methoxy group on the CD curves was discussed in connection with solvent effect. Mechanism of this asymmetric synthesis has been investigated, and it has become apparent that the diastereomeric menthoxysulphonium chloride was an excess of (RR)-configuration was formed initially and the amide anion attacks the S atom of the salt with net inversion.  相似文献   

6.
4-tert-Butyl-2-cyclohexen-1-one dimerizes in THF solution via its kinetic enolate, leading to di-tert-butylditwistane 8 in up to 36% yield (−78 °C → room temp., protonolysis, flash chromatography). X-ray crystallography shows that 8 incorporates one R and one S enantiomer of the starting ketone; none of the diastereomeric ditwistanes epi-8, epi’-8 or iso-8 was isolated. This means that the formation of 8 proceeds with mutual kinetic resolution and 100% induced diastereoselectivity.  相似文献   

7.
The antipodes of 1-aryl-, 1-alkyl- and 1-alkoxy-3-methyl-3-phospholene 1-oxides 1a–h and 1-phenyl-3-methyl-3-phospholene 1-sulfide 1i were separated in good yields and high enantiomeric excesses (up to >99% ee) by resolution via formation of diastereomeric complexes with either (−)-(4R,5R)-4,5-bis(diphenylhydroxymethyl)-2,2-dimethyldioxolane 2 (TADDOL) or (−)-(2R,3R)-,,′,′-tetraphenyl-1,4-dioxaspiro[4.5]decan-2,3-dimethanol 3. The stereostructure of the supramolecular formations and the absolute configurations of the 3-phospholene oxides 1a, 1e and 1f were elucidated by single crystal X-ray crystallography. CD spectroscopy was also useful in determining the absolute configurations of some phospholene oxides 1b, 1c, 1g and 1h.  相似文献   

8.
Daniela Fattori  Pierre Vogel   《Tetrahedron》1992,48(48):10587-10602
(1S,4S)-7-Oxabicyclo[2.2.1]hept-5-en-2-one ((−)-5, a “naked sugar”) has been converted to (−)-(1R,4S,6S)-6-endo-benzyloxy-2-bromo-7-oxabicyclo[2.2.1]hept-2-ene ((−)-12) in a highly stereoselective fashion. Double hydroxylation of the C=C double bond of (−)-12, followed by acetylation and Baeyer-Villiger oxidation of the resulting -acetoxyketone (−)-14 afforded (−)-5-O-acetyl-2-O-benzyl-3-deoxy-β-D-arabino-hexofuranurono-6,1-lactone ((−)-15). This compound was converted readily into (+)-methyl 3-deoxy--D-arabino-hexofuranoside ((+)-6 and (+)-methyl 3-deoxy-β-L-xylo-hexofuranoside ((+)-7) and partially protected derivatives. (−)-15 was also converted into 4-deoxy-D-lyxo-hexopyranose (34) and several partially protected derivatives such as (+)-methyl 4-deoxy-2,3-O-isopropylidene--D-lyxo-hexopyranoside ((+)-8).  相似文献   

9.
Reaction of CpCoI2(P(OMe)3) 8 with the chiral aminophosphine (S)-(−)-diphenyl-phenylethylaminophosphine affords the diastereomeric phosphonate complexes (R,S)Co,SC-CpCoI(P(0)(OMe)2)(PPh2NHCH(Me)Ph) (10a,10b) via Arbuzov dealkylation. 10a,10b are separable and configurationally stable in solution for extended periods. The structure and absolute configuration of the lower Rf diastereomer (−)-436-10b were determined via single-crystal X-ray diffraction. It crystallizes as a toluene solvate in space group P21 with a 13.194(6), b 9.062(4), c 17.023(5) Å, β 108.78(3)°, Z = 2, and was refined to R = 0.067 for 6318 reflections. Spectroscopic and structural evidence demonstrate a strong 1,6 intramolecular NH O=P hydrogen bond between the aminophosphine NH and the basic phosphoryl oxygen, which establishes a quasi-boat conformation. Proton nuclear Overhauser difference spectra show that the conformation in solution is the same as that observed in the solid state.  相似文献   

10.
Asymmetric cyclomercuration of chiral ferrocenylimines (+)-(R)-2 and (−)-(S)-2 with mercuric acetate and subsequent treatment by lithium chloride gave two different pairs of diastereomeric cyclomercurated ferrocenylimines. The four enantiomerically pure compounds with both planar and central chirality were isolated from the two pairs by thin layer chromatography. The structures and absolute configurations of these complexes were determined by single-crystal X-ray analysis and characterized by IR, 1H NMR, specific rotation and elemental analysis.  相似文献   

11.
(±)-Piperidine-2-carboxylic acid ((±)-pipecolic acid) has been resolved by the fractional crystallisation of diastereomeric palladium(II) complexes containing orthometallated (S)-(−)-1-[1-dimethylamino)ethyl]naphthalene. The enantiomers of the acid were liberated from the individual configurationally homogeneous diastereomers of the complex in high yield with []D ± 26.0 (c 1.00, H2O). The crystal and molecular structures of both diastereomers of the complex have been determined.  相似文献   

12.
Asymmetric hydroboration of [E]- and [Z]-2-methoxy-2-butene, using (−)-diisopinocampheylborane at −25°C in THF solvent, followed by oxidation using H2O2/NaOH, gave (−)-[2R,3R]- and (+)-[2R,3S]-3-methoxy-2-butanols in >97 and 90% ee, respectively. (−)-[2R,3R]-3-Methoxy-2-butanol was converted to (−)-[2R,3R]-butane-2,3-diol (>97% ee, in an overall yield of 65%).  相似文献   

13.
Antigen I/II can be found on streptococcal cell surfaces and is involved in their interaction with salivary proteins. In this paper, we determine the adsorption enthalpies of salivary proteins to Streptococcus mutans LT11 and S. mutans IB03987 with and without antigen I/II, respectively, using isothermal titration calorimetry. In addition, protein adsorption to the cell surfaces was determined spectrophotometrically. S. mutans LT11 with antigen I/II, yielded a much higher, exothermic adsorption enthalpy at pH 6.8 (ranging from −2073 × 10−9 to −31707 × 10−9 μJ per bacterium) when mixed with saliva than did S. mutans IB03987 (−165 × 10−9 to −1107 × 10−9 μJ per bacterium) at all bacterial concentrations studied (5 × 109, 5 × 108, and 5 × 107 ml−1), largest effects per bacterium being observed for the lowest concentration. However, the enthalpy of salivary protein adsorption to S. mutans LT11 became smaller at pH 5.8. Adsorption isotherms for the S. mutans LT11 showed considerable protein adsorption at pH 6.8 (1.2–2.1 mg/m2), that decreased only slightly at pH 5.8 (1.1–1.6 mg/m2), with the largest amount adsorbed at the lowest bacterial concentration. This suggests that the protein(s) in the saliva with the strongest affinity for antigen I/II is (are) readily depleted from saliva. In conclusion, antigen I/II surface proteins on S. mutans play a determinant role in adsorption of salivary proteins through the creation of enthalpically favorable adsorption sites.  相似文献   

14.
Norfloxacin, 1-ethyl-6-fluoro-1,4-dihydro-4-oxo-7-(1-piperazinyl)-3-quinoline carboxylic acid (NORH), reacts with aluminium(III) ion forming the strongly fluorescent complex [Al(HNOR)]3+, in slightly acidic medium. The complex shows maximum emission at 440 nm with excitation at 320 nm. The fluorescence intensity is enhanced upon addition of 0.5% sodium dodecylsulphate. Fluorescence properties of the Al-NOR complex were used for the direct determination of trace amounts of NOR in serum. The linear dependence of fluorescence intensity on NOR concentration, at a NOR to Al concentration ratio of 1:10, was found in the concentration range 0.001–2 μg/ml NOR with a detection limit of 0.1 ng/ml. The ability of aluminium (III) ion to form complexes with NOR was investigated by titrations in 0.1 M LiCl medium, using a glass electrode, at 298 K, in the concentration range: 2 × 10−4 ≤ [Al] ≤ 8 × 10−4; 5 × 10−4 ≤ [NOR] ≤ 9 × 10−4 mol/dm3; 2.8 ≤ pH ≤ 8.3. The experimental data were explained by the following complexes and their respective stability constants, log(β ± σ): [Al(HNOR)], (14.60 ± 0.05); [Al(NOR)], (8.83 ± 0.08); [A1(OH)3(NOR)], (−14.9 ± 0.1), as well as several pure hydrolytic complexes of A13+. The structure of the [Al(HNOR)] complex is discussed, with respect to its fluorescence properties.  相似文献   

15.
A chemo-enzymatic synthesis of (−)-epipentenomycin I is reported using a lipase-catalysed kinetic resolution of the racemic pentacyclic alcohol 8. Flash vacuum pyroloysis of (−)-8 so obtained gave (−)-(4R)-4-hydroxy-5-methylene-2-cyclopentenone. Epoxidation of this compound with dimethyldioxirane followed by hydrolytic ring-opening of the resulting epoxide gave (−)-epipentenomycin I.  相似文献   

16.
The structures of an isomeric pair: 1-[2-(2,6-dichlorophenoxy)-ethyl]-3-methylthio-5-amino-1H-1,2,4-triazole (IV) and 2-[2-(2,6-dichlorophenoxy)-ethyl]-3-methylthio-5-amino-2H-1,2,4-triazole (V) have been established by X-ray crystallography from diffractometer data. Crystals of IV are monoclinic, space group P21/c with a = 11.220(1), b = 10.005(3), c = 12.784(3) Å, β = 91.99(1)°, Z = 4, Dc = 1.478 g cm−3. Crystals of V are orthorhombic, space group Pbca with a = 7.618(2), b = 14.289(4), c = 26.494(7) Å, Z = 8, Dc = 1.470 g cm−3. The structures determined by direct methods were refined to R = 0.060 for 1931 reflections of IV and R = 0.061 for 1315 reflections of V. The X-ray analysis corroborated the structures assigned to the isomeric pair on the basis of proton coupled 13C-NMR spectra. The features (i.e. the sequence of the endocyclic bond angle magnitudes) of the planar s-triazole rings are governed by the same rules reported recently by Kálmán and Argay [J. Mol. Struct., 102 (1983) 391] discussing the structures of the related I, II and III compounds. In contrast to I, II, III and IV the S(6)---C(7) bond of V is almost perpendicular to the best plane of the s-triazole ring C(7)---S(6)---C(3)---N(4) = −79.0(7)°. This unusual orientation of the S-methyl bond is created by the steric hindrance of the very same vicinal 2-ethyl (C(9)H2---C(10)H2) moiety once within the molecule V and the second time as the part of a symmetry equivalent molecule at a distance of −a (7.618 Å). These molecules are fixed together by two infinite helices of hydrogen bonds formed around the screw axis 2, (X, 1/4, 1/2) via an intermediate V molecule at (x − 1/2, 1/2 − y, 1 − z).  相似文献   

17.
Raman and infrared spectra of propylgermane, CH3CH2CH2GeH3, and its Ge-deuterated analog, CH3CH2CH2GeD3, were investigated in their gaseous, liquid and solid states. The normal coordinate treatment was carried out by density functional theory (DFT) calculation, using B3LYP/6-31G* and 6-311++G** basis sets, and the corresponding fundamental vibrations were assigned. The trans (T) and gauche (G) forms around the central C–C bond coexisted in the gaseous and liquid states and only the T form existed in the solid state. From the temperature dependent measurements of the Raman spectra in the liquid state, the enthalpy difference was found to be ΔH(TG)=−0.36±0.02 kcalmol−1 with the T form being more stable. The energy differences between the isomers obtained by DFT calculations were ΔE(TG)=−0.46 kcalmol−1 and ΔE(TG)=−0.87 kcalmol−1 by the 6-31G* basis set and 6-311++G** basis set, respectively.  相似文献   

18.
The Schiff base compound, N-N′-bis(4-methoxybenzylidene)ethylenediamine (C18H20N2O2) has been synthesized and its crystal structure has been investigated by X-ray analysis and PM3 method. The compound crystallizes in monoclinic space group P21/n with a=10.190(1), b=7.954(1), c=10.636(1) Å, β=111.68(1)°, V=801.1(1) Å3, Z=2 and Dcal=1.229 Mgm−3. The title structure was solved by direct methods and refined to R=0.056 for 2414 reflections [I>3.0σ(I)] by full-matrix anisotropic least-squares methods. The energy profile of the compound was calculated by PM3 method as a function of θ[N1′–C9′–C9–N1]. The most stable molecular structure of the title compound is the anti conformation, which is different in energy by 5.0 and 1.0 kcal mol−1 from the eclipsed conformation I and gauche conformations, (III and V), respectively.  相似文献   

19.
Two novel Cd(II)-citrate complexes were obtained with different metal/ligand ratios through hydrothermal method. Their structures were determined by single-crystal X-ray diffraction analysis. Although their topological structures are both 2-D layer network assemblies, both central Cd(II) ions and Hcit3− ligands display completely different coordination modes. In polymeric complex 1, Hcit3− serves as a μ10-bridged and central Cd(II) ions adopt 6- and 8-coordinated configurations. In contrast, a μ9-bridged and 6- and 7-coordinated environments between Cd(II) and Hcit3− are established in the polymeric complex 2. Two Complexes remain stable up to approximately 300 °C. The complex 1 exhibits strong fluorescent emission band at 450 nm (λ=346 nm) as well as complex 2 exhibits strong fluorescent emission band at 430 (λ=346 nm).  相似文献   

20.
Racemic 1,1′-methylene[(1RS,1′RS,3RS,3′RS,5RS,5′RS)-8-oxabicyclo[3.2.1]oct-6-en-3-ol] ((±)-6) derived from 2,2′-methylenedifuran has been resolved kinetically with Candida cyclindracea lipase-catalysed transesterification giving 1,1′-methylenedi[(1R,1′R,3R,3′R,5R,5′R)-8-oxabicyclo[3.2.1]oct-6-en-3-ol] (−)-6 (30% yield, 98% ee) and 1,1′-methylenedi[(1S,1′S,3S,3′S,5S,5′S)-8-oxabicyclo[3.2.1]oct-6-en-3-yl] diacetate (+)-8, (40% yield, 98% ee). These compounds have been converted into 1,1′-methylenedi[(4S,4′S,6S,6′S)- and (4R,4′R,6R,6′R)-cyclohept-1-en-4,6-diyl] derivatives.  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