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 31 毫秒
1.
通过酒石酸乙酯和不同胺的酰胺化, 相应地合成到(+)N,N'-二苯基酒石酸二酰胺(2), (+)N,N'-二苄基酒石酸二酰胺(3),(+)N,N'-二环己基酒石酸二酰胺(4)和(+)N,N'-二正辛基酒石酸二酰胺(5).氯化亚锡与2当量的酒石酸二酰胺反应, 再加入烯丙基溴可制得手性锡络合物6, 锡络合物与苯甲醛反应,得到光学活性的高烯丙基化合物, 产率42~64% , e.e.值24~62%.  相似文献   

2.
The preparation of an improved liquid-membrane periodate-selective electrode is described. The electrode exhibits rapid and near-Nernstian response in the range 2 × 10−6-10−2 M. The electrode was used to monitor the course of the tartaric acid-periodate reaction. A potentiometric reaction rate method for the rapid and accurate determination of tartaric acid has been developed. A total of 0.4–120 μmol of tartaric acid has been determined with relative error of about 2.0%. The method has been applied to the determination of tartaric acid in pharmaceutical preparations and to the determination of small amounts of tartaric acid in impure citric acid.  相似文献   

3.
Protocatechuic acid esters (= 3,4‐dihydroxybenzoates) scavenge ca. 5 equiv. of radical in alcoholic solvents, whereas they consume only 2 equiv. of radical in nonalcoholic solvents. While the high radical‐scavenging activity of protocatechuic acid esters in alcoholic solvents as compared to that in nonalcoholic solvents is due to a nucleophilic addition of an alcohol molecule at C(2) of an intermediate o‐quinone structure, thus regenerating a catechol (= benzene‐1,2‐diol) structure, it is still unclear why protocatechuic acid esters scavenge more than 4 equiv. of radical (C(2) refers to the protocatechuic acid numbering). Therefore, to elucidate the oxidation mechanism beyond the formation of the C(2) alcohol adduct, 3,4‐dihydroxy‐2‐methoxybenzoic acid methyl ester ( 4 ), the C(2) MeOH adduct, which is an oxidation product of methyl protocatechuate ( 1 ) in MeOH, was oxidized by the DPPH radical (= 2,2‐diphenyl‐1‐picrylhydrazyl) or o‐chloranil (= 3,4,5,6‐tetrachlorocyclohexa‐3,5‐diene‐1,2‐dione) in CD3OD/(D6)acetone 3 : 1). The oxidation mixtures were directly analyzed by NMR. Oxidation with both the DPPH radical and o‐chloranil produced a C(2),C(6) bis‐methanol adduct ( 7 ), which could scavenge additional 2 equiv. of radical. Calculations of LUMO electron densities of o‐quinones corroborated the regioselective nucleophilic addition of alcohol molecules with o‐quinones. Our results strongly suggest that the regeneration of a catechol structure via a nucleophilic addition of an alcohol molecule with a o‐quinone is a key reaction for the high radical‐scavenging activity of protocatechuic acid esters in alcoholic solvents.  相似文献   

4.
Treatment of acetylated phenyl thioglycoside of N-acetylneuraminic acid with m-chloroperbenzoic acid (MCPBA) in CH2Cl2 affords quantitatively mixtures of the respective sulfone and glycal free from the sulfoxide. The outcome of the reaction does not depend on the anomeric configuration of the starting thioglycoside. The sulfone can be selectively prepared (yield 100%) by oxidation with an excess of MCPBA and NaHCO3. In the presence of pyridine (2 equiv.) and MCPBA (2 equiv.), the major product is glycal (yields 81—88%). This version of the reaction can be regarded as a new method for the preparation of sialic acid glycals.  相似文献   

5.
Reaction of 2-(3-hydroxy-3-methyl-1-butynyl)adamantan-2-ol with acetonitrile under Ritter reaction conditions is accompanied by isomerization and partial hydration where the water addition to the triple bond occurs nonselectively. As a result of reaction carried out in the presence of 8 equiv of sulfuric acid a mixture was obtained of N 2-[4-(1-acetylamino-2-adamantyl)-2-methyl-3-butyn-2-yl]acetamide, N 3-[1-(1-acetylamino-2-adamantyl)-3-methyl-2-oxo-3-butyl]-acetamide, and N 3-[1-(1-acetylamino-2-adamantyl)-3-methyl-1-oxo-3-butyl]acetamide in ~10:3:2 ratio. In the presence of 2 equiv of the acid the mixture obtained consisted of N 2-[4-(1-acetylamino-2-adamantyl)-2-methyl-3-butyn-2-yl]acetamide, N 3-[1-(1-acetylamino-2-adamantyl)-3-methyl-2-oxo-3-butyl]acetamide, and 1-(1-acetylamino-2-adamantyl)-3-methyl-2-buten-1-one in the same ratio. In Rupe reaction conditions we obtained instead of the expected ,-unsaturated ketones a mixture of 1-(1-hydroxy-2-adamantyl)-3-hydroxy-3-methylbutan-1-one and 1-(1-hydroxy-2-adamantyl)-3-hydroxy-3-methylbutan-2-one in a 5:3 ratio.  相似文献   

6.
Protocatechuic acid (= 3,4‐dihydroxybenzoic acid; 1 ) exhibits a significantly slow DPPH (= 2,2‐diphenyl‐1‐picrylhydrazyl) radical‐scavenging reaction compared to its esters in alcoholic solvents. The present study is aimed at the elucidation of the difference between the radical‐scavenging mechanisms of protocatechuic acid and its esters in alcohol. Both protocatechuic acid ( 1 ) and its methyl ester 2 rapidly scavenged 2 equiv. of radical and were converted to the corresponding o‐quinone structures 1a and 2a , respectively (Scheme). Then, a regeneration of catechol (= benzene‐1,2‐diol) structures occurred via a nucleophilic addition of a MeOH molecule to the o‐quinones to yield alcohol adducts 1f and 2c , respectively, which can scavenge additional 2 equiv. of radical. However, the reaction of protocatechuic acid ( 1 ) beyond the formation of the o‐quinone was much slower than that of its methyl ester 2 . The results suggest that the slower radical‐scavenging reaction of 1 compared to its esters is due to a dissociation of the electron‐withdrawing carboxylic acid function to the electron‐donating carboxylate ion, which decreases the electrophilicity of the o‐quinone, leading to a lower susceptibility towards a nucleophilic attack by an alcohol molecule.  相似文献   

7.
(+)‐Tartaric acid‐catalyzed aminobromination of α,β‐unsaturated ketones, α,β‐unsaturated esters and simple olefins utilizing TsNH2/NBS as the nitrogen/halogen sources at room temperature without protection of inert gases achieved good yields (up to 92% yield) of vicinal haloamino products with excellent regio‐ and stereoselectivity, even just 10% of (+)‐tartaric acid was used as catalyst. The regio‐ and stereochemistry was unambiguously confirmed by X‐ray structural analysis of products 2b and 12c . The electron‐rich and deficient olefins show significant differences in activity to the aminobromination reaction and give the opposite regioselectivities. The 21 cases have been investigated which indicated that our protocol has the advantage of a large scope of olefins. Additionally, tartaric acid as catalyst has the advantage of avoiding any hazardous metals retained in products.  相似文献   

8.
The reaction of dihydroquinopimaric acid methyl ester with bromine was found to be chemo- and stereoselective. Regardless of the solvent (acetic acid, methanol, dioxane), bromination of the title compound with an equimolar amount of bromine occurs as electrophilic addition at the double C19=C20 bond with formation of 14α-hydroxy- or 14α-methoxy-19R-bromo derivatives. The reaction with excess bromine (3 equiv) leads to the formation of 16S-bromo derivatives. The bromination process is accompanied by formation of epoxy bridge between the C14 and C20 atoms. X-Ray analysis revealed two polymorphic modifications of (16S,19R)-16,19-dibromo-14β,20-epoxydihydroquinopimaric acid methyl ester.  相似文献   

9.
A preparation of glycosylphosphonates ( 27 , 28 , 36 , 38 , and 39 ) from 2-azido-2-deoxy-glycoses ( 26 , 35 , and 37 ) and the synthesis of the non-isosteric phosphonate analogue 3a of lipid X( 2 ) are described. The 2-azido group was introduced by azidonitration. Treatment of the 1-O-acetyl-2-azido-2-deoxy-β-D-galactopyranose 22 with 1.5-3 equiv. of P(OMe)3 and 1.2-2.5 equiv. of TfOSiMe3 gave mainly recovered starting material. In P(OMe)3 as the solvent, the dimethyl phosphoramidate 24 was obtained by way of a Staudinger reaction, even in the presence of TfOSiMe3. Treatment of the benzylated α-D-galacto-trichloroacetimidate 26 , however, with P(OMe)3 and TfOSiMe3 gave a 1:1 mixture of the α- and β-D-galacto-phosphonates 27 and 28 , while the acetylated α-D-gluco- imidate 35 led to the α-D-gluco-configurated phosphonate 36 . The stereoselectivity of the phosphonate formation is related to the relative ease of formation of oxonium-ion intermediates from 26 and 35 . Starting from the phosphonate 36 , deacetylation, benzylidenation, reduction of the azido group, acylation with (R)-3-(benzyloxy)tetradecanoic acid and deprotection yielded the desired compound 3a which was crystallized in the presence of 2 equiv. of (aminomethylidyne)trimethanol (Tris.). The structure of the phosphonates was deduced from their 1H-, 13C-, and 31P-NMR spectra.  相似文献   

10.
In this study, enantioselective Biginelli reaction of aldehyde, β‐ketoester, and urea catalyzed by natural (2R, 3R)‐tartaric acid has been investigated using density functional theory calculations. The results indicate that the most favorable pathway involves a protonated imine from aldehyde and urea in the first step. Tartaric acid forms H‐bonds network with substrates enhancing the electrophilicity of protonated imine and the nucleophilicity of β‐ketoester. (R)‐3,4‐Dihydropyrimidin‐2‐(1H)‐ones is preferable for the reaction. The solvent effect is discussed in the prediction of enantiomeric excess (ee) values in ethanol and water. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

11.
Pasteur carried out pioneering work on cinchona alkaloids and their derivatives and his studies led to important discoveries. He examined crystals of cinchona alkaloids for his correlation of crystal hemihedrism with molecular chirality, studies that led Pasteur to the discovery of physicochemical differences between diastereoisomeric salts of tartaric acids with optically active cinchona bases, an important insight into fundamentals of molecular chirality. These physicochemical differences also led to Pasteur’s invention of the vital method of racemate resolution through diastereoisomeric derivatives. Pasteur clarified the confusion around the cinchona alkaloids by elucidating their identities and relations. He discovered the conversion of the major cinchona alkaloids to quinicine and cinchonicine, a finding subsequently of considerable importance in studies of the structure and synthesis of the major cinchona alkaloids. The reaction producing quinicine and cinchonicine led Pasteur to the discovery of the racemization of tartaric acid and the finding of meso‐tartaric acid, fundamental breakthroughs in the development of stereochemistry.  相似文献   

12.
We report a novel 1:1 cocrystal of β‐alanine with dl ‐tartaric acid, C3H7NO2·C4H6O6, (II), and three new molecular salts of dl ‐tartaric acid with β‐alanine {3‐azaniumylpropanoic acid–3‐azaniumylpropanoate dl ‐tartaric acid–dl ‐tartrate, [H(C3H7NO2)2]+·[H(C4H5O6)2], (III)}, γ‐aminobutyric acid [3‐carboxypropanaminium dl ‐tartrate, C4H10NO2+·C4H5O6, (IV)] and dl ‐α‐aminobutyric acid {dl ‐2‐azaniumylbutanoic acid–dl ‐2‐azaniumylbutanoate dl ‐tartaric acid–dl ‐tartrate, [H(C4H9NO2)2]+·[H(C4H5O6)2], (V)}. The crystal structures of binary crystals of dl ‐tartaric acid with glycine, (I), β‐alanine, (II) and (III), GABA, (IV), and dl ‐AABA, (V), have similar molecular packing and crystallographic motifs. The shortest amino acid (i.e. glycine) forms a cocrystal, (I), with dl ‐tartaric acid, whereas the larger amino acids form molecular salts, viz. (IV) and (V). β‐Alanine is the only amino acid capable of forming both a cocrystal [i.e. (II)] and a molecular salt [i.e. (III)] with dl ‐tartaric acid. The cocrystals of glycine and β‐alanine with dl ‐tartaric acid, i.e. (I) and (II), respectively, contain chains of amino acid zwitterions, similar to the structure of pure glycine. In the structures of the molecular salts of amino acids, the amino acid cations form isolated dimers [of β‐alanine in (III), GABA in (IV) and dl ‐AABA in (V)], which are linked by strong O—H…O hydrogen bonds. Moreover, the three crystal structures comprise different types of dimeric cations, i.e. (AA)+ in (III) and (V), and A+A+ in (IV). Molecular salts (IV) and (V) are the first examples of molecular salts of GABA and dl ‐AABA that contain dimers of amino acid cations. The geometry of each investigated amino acid (except dl ‐AABA) correlates with the melting point of its mixed crystal.  相似文献   

13.
Acetophenones containing a methoxycarbonylamino group in position 2, 3, or 4 of the aromatic ring reacted with phenylglycine in the presence of 2 equiv of iodine and 0.5 equiv of sulfanilic acid in DMSO at 100°C for 6 h to give methyl [2(3,4)-(2-phenyl-1,3-oxazol-5-yl)phenyl]carbamates. The reaction was presumed to involve intermediate formation of methyl [(iodoacetyl)phenyl]carbamate. This was confirmed by the isolation of methyl [2-(iodoacetyl)phenyl]carbamate in the reaction of methyl (2-acetylphenyl)carbamate with iodine in glacial acetic acid and its subsequent transformation to methyl [2-(2-phenyl-1,3-oxazol-5-yl)-phenyl]carbamate.  相似文献   

14.
《Tetrahedron: Asymmetry》2001,12(5):719-724
1-(4′-Hydroxymethyl-phenyl)-2-methyl-3-(piperidine-1-yl)-propane-1-one M2, a metabolite of tolperisone, was synthesised in a solvent-free Mannich reaction. The optical resolution was carried out by diastereoisomeric salt formation and separation, for which three resolving agents ((2R,3R)-O,O′-dibenzoyl tartaric acid, (2R,3R)-O,O′-di-p-toluoyl tartaric acid and (R)-2-hydroxy-4-(2-methoxyphenyl)-5,5-dimethyl-1,3,2-dioxaphosphorinane-2-oxide (anicyphos)) were found. The absolute configuration of M2 was determined by the single-crystal X-ray diffraction method.  相似文献   

15.
The reaction of ethynylmagnesium bromide with chloroisopropylgermanes (i-Pr4 - n GeCl/sub> n , n = 1-3) was used to prepare previously unknown ethynylisopropylgermanes i-Pr4 - n Ge(CCH) n (n = 1-3). The reaction of Me3SiCCMgBr with i-PrGeCl3 afforded i-Pr(Me3SiCC)3 - n GeCl n (n = 1, 2). The reaction of the monochloride with BrMdCCH gave i-Pr(HCC)2GeCCSiMe3, while with the dichloride, i-Pr(HCC)·Ge(CSiMe3)2 formed. The latter compounds were obtained by independent synthesis from i-PrGe(CCH)3, EtMgBr, and ClSiMe3. The reaction of (bromomagnesioethynyl)triisopropylgermane with Me3SiCl gave i-Pr3GeCSiMe3.  相似文献   

16.

Abstract  

Chiral α-ethylphenylamine tartaric acid salts were synthesized from α-ethylphenylamine by direct reaction with chiral tartaric acid. The crystal structure of S-(−)-α-ethylphenylamine-(2R,3R)-(−)-dihydroxybutanedioic acid was determined. The crystal is monoclinic, of space group P21/n , with a = 6.331(5) ?, b = 14.209(11) ?, c = 7.495(6) ?, α = 90.00o, β = 107.000(13)o, γ = 90.00o, λ = 0.7103 Ǻ, V = 644.7(9), Z = 2, D c = 1.397 g/cm3, M r  = 271.27 and F(000) = 288, R = 0.0477, and ωR = 0.0838 for 1388 observed reflections with I > 2σ(I). We then used the chiral α-ethylphenylamine tartaric acid salts as catalysts in the cyanosilylation of prochiral ketones, and moderate conversions were obtained.  相似文献   

17.
Total synthesis of the polyhydroxy caprolactam amide natural product, bengamide E, is accomplished starting from tartaric acid. Key reactions in the synthesis include desymmetrization of the bis(dimethylamide) unit of tartaric acid, Zn(BH4)2‐mediated anti‐selective reduction, and a Horner–Wadsworth–Emmons olefination.  相似文献   

18.
NMR.-spectroscopic Methods as Criteria of the Purity of Isomeric Tartaric Acids and Their Esters meso-Tartaric acid ( 2 ) can be distinguished either from the natural (+)-(2R, 3R)-tartaric acid ( 1 ) or the ‘unnatural’ (?)-(2S, 3S)-tartaric acid ( 1 ′) or their racemic mixture, by 1H-NMR.-spectral resolution using europium chloride in aqueous solution. Diastereomeric esters have been prepared from different esters of tartaric acid 3 and the Mosher reagent 4 and the purities of the enantiomers of 3 have been checked by 1H-NMR. spectroscopy.  相似文献   

19.
The stereoselective total synthesis of a naturally occurring α‐pyrone (=2H‐pyran‐2‐one) derivative, synargentolide A ( 1 ), and of its epimer 2 (with the originally proposed structure of synargentolide A) was efficiently accomplished involving D ‐tartaric acid as the starting material and an olefin cross‐metathesis reaction as the key step.  相似文献   

20.
2‐Amino‐1,2,3,4‐tetrahydronaphthalene‐6,7‐diol ( 2 ; 6,7‐ADTN) was synthesized starting from naphthalene‐2,3‐diol in seven steps and with an overall yield of 44%. Methylation of naphthalene‐2,3‐diol with dimethyl sulfate, followed by Friedel? Crafts acylation with AcCl, gave 2‐acetyl‐6,7‐dimethoxynaphthalene. 2‐Acetyl‐6,7‐dimethoxynaphthalene was converted to 6,7‐dimethoxynaphthalene‐2‐carboxylic acid by a haloform reaction. Birch reduction of the carboxylic acid with 4 mol‐equiv. of Na in liquid ammonia afforded 1,2,3,4‐tetrahydro‐6,7‐dimethoxynaphthalene‐2‐carboxylic acid, from which 2 was obtained by a Curtius reaction, followed by hydrogenolysis and demethylation.  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