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 780 毫秒
1.
Asymmetric synthesis of L -γ-carboxyglutamic acid derivatives A modified Strecker synthesis according to Patel & Worsley was used to prepare γ,γ′-di-t-butyl L (?)-N-phthaloyl-γ-carboxyglutamate with almost 100% optical purity and an overall yield of about 10% relative to di-t-butyl malonate, or 20% relative to (?)-α-methyl-benzylamine (s. Scheme).  相似文献   

2.
The chemical synthesis of γ,γ′-tert.-butyl-γ-carboxyglutamic acid is accompanied by extensive racemization, and very careful resolution is needed to obtain and -γ,γ′-di-tert.-butyl-γ-carboxyglutamic acids in high chiral purity. A novel method was devised for the separation of enantiomers of γ,γ′-di-tert.-butyl-γ-carboxyglutamic acid and γ-carboxyglutamic acid, applying precolumn derivatization with 1-fluoro-2,4-dinitrophenyl-5- -alanine amide and 2,3,4,6-tetra-O-acetyl-β- -glucopyranosyl isothiocyanate as chiral reagents, with subsequent reversed-phase high-performance liquid chromatographic separation of diastereomeric compounds. The effects of organic modifiers, of the mobile-phase composition and of the pH on the separation of the diastereomers were investigated.  相似文献   

3.
The chemical resolution of γ,γ′-di-t-butyl DL -N-benzyloxycarbonyl-γ-carboxy-glutamate is described in detail (preliminary account see [1]). The D (?)-derivative was obtained as a crystalline quinine salt, and the L (+)-derivative as a crystalline salt with (?)-ephedrine in yields of 44 and 70%, respectively. Physical data are indicated for the enantiomers of γ,γ′-di-t-butyl N-benzyloxycarbonyl-γ-carboxyglutamate, γ,γ′-di-t-butyl γ-carboxyglutamate, and γ-carboxyglutamic acid. The absolute configurations and optical purities of the γ,γ′-di-t-butyl (+)- and (?)-N-benzyloxycarbonyl-γ-carboxyglutamates were determined by removal of the protecting groups and decarboxylation to optically active glutamic acid.  相似文献   

4.
Synthesis of Diastereo- and Enantioselectively Deuterated β,ε-, β,β-, β,γ- and γ,γ-Carotenes We describe the synthesis of (1′R, 6′S)-[16′, 16′, 16′-2H3]-β, εcarotene, (1R, 1′R)-[16, 16, 16, 16′, 16′, 16′-2H6]-β, β-carotene, (1′R, 6′S)-[16′, 16′, 16′-2H3]-γ, γ-carotene and (1R, 1′R, 6S, 6′S)-[16, 16, 16, 16′, 16′, 16′-2H6]-γ, γ-carotene by a multistep degradation of (4R, 5S, 10S)-[18, 18, 18-2H3]-didehydroabietane to optically active deuterated β-, ε- and γ-C11-endgroups and subsequent building up according to schemes \documentclass{article}\pagestyle{empty}\begin{document}${\rm C}_{11} \to {\rm C}_{14}^{C_{\mathop {26}\limits_ \to }} \to {\rm C}_{40} $\end{document} and C11 → C14; C14+C12+C14→C40. NMR.- and chiroptical data allow the identification of the geminal methyl groups in all these compounds. The optical activity of all-(E)-[2H6]-β,β-carotene, which is solely due to the isotopically different substituent not directly attached to the chiral centres, is demonstrated by a significant CD.-effect at low temperature. Therefore, if an enzymatic cyclization of [17, 17, 17, 17′, 17′, 17′-2H6]lycopine can be achieved, the steric course of the cyclization step would be derivable from NMR.- and CD.-spectra with very small samples of the isolated cyclic carotenes. A general scheme for the possible course of the cyclization steps is presented.  相似文献   

5.
Good yields of some crystalline γ-alkyl esters of L -glutamic acid were obtained by carrying out the esterfication with a small (20–50 mole-%) excess of alcohol in aqueous hydrochloric acid or 60–80% sulfuric acid followed by neutralization with an alkaline solution. This new method made it possible to synthesize various γ-alkyl L -glutamates, including those higher than ethyl, and consequently, various poly(γ-alkyl L -glutamates) such as methyl, ethyl, n-propyl, n-butyl, isobutyl, and isoamyl. The conformation of these poly-L -glutamates in the solid state was determined by the infrared absorption method. The molecular motions of the polymers of γ-methyl, -ethyl, -n-propyl, -n-butyl, and-isoamyl L -glutamates and poly(γ-methyl-D -glutamate) in the solid state were studied by NMR, and dielectric and mechanical measurements. At temperatures up to 400°K., the NMR spectra of poly(γ-methyl D -glutamate) can be explained only by rotational motion of the side chain. Also, from NMR results, rotational motion of C?O groups in the side chain of poly(γ-methyl D -glutamate) is expected near room temperature, and such a motion was examined by dielectric measurements. Rotation of C?O groups in the side chains of polymers of γ-methyl, γ-ethyl, γ-n-propyl, γ-n-butyl, and γ-isoamyl L -glutamate was also observed near room temperature by dielectric measurements in the frequency range from 102 to 106 cps. Activation energies obtained by dielectric and mechanical measurements were similar to those for the side chain motions of the corresponding esters of poly(methacrylic acid). Although it has been noted that the molecular motion of poly(γ-benzyl L -glutamate) in the solid state at room temperature may be related to the motion of its back bone, the molecular motion in these poly-L -glutamates at these temperatures can be explained only in terms of side-chain rotation.  相似文献   

6.
The dissociation equilibria of aqueous solutions of β‐glutamic acid were studied by potentiometric titration and the three pK values determined under standard conditions. The hydrogen β‐glutamate anion β‐GluH was found to be the dominating species in the physiologically relevant pH range 4.0–9.4. Neutralization of β‐glutamic acid by magnesium oxide affords magnesium bis(hydrogen β‐glutamate) Mg (β‐GluH)2, which crystallizes as the hexahydrate from dilute aqueous solution. A single‐crystal X‐ray study showed that the β‐GluH anions are not part of the coordination sphere of the magnesium ion. Instead hexahydrated dications [Mg(H2O)6]2+ are intimately associated with free β‐GluH anions through a three‐dimensional network of H‐bonds. This study provides the first structural and conformational reference data for the β‐GluH anion.  相似文献   

7.
Macroinitiator‐amino terminated poly(ethylene glycol) (PEG) (NH2‐PEO‐NH2) was prepared by converting both terminal hydroxyl groups of PEG to more reactive primary amino groups. The synthetic route involved reactions of chloridize, phthalimide and finally hydrazinolysis. Furthermore, poly(γ‐benzyl‐L ‐glutamate)‐poly(ethylene oxide)‐poly(γ‐benzyl‐L ‐glutamate) (PBLG‐PEO‐PBLG) triblock copolymer was synthesized by polymerization of γ‐benzyl‐L ‐glutamate N‐carboxyanhydride (Bz‐L‐GluNCA) using NH2‐PEO‐NH2 as macroinitiator. The resultant NH2‐PEO‐NH2 and triblock copolymer were characterized by FT‐IR, 1H‐NMR and gel permeation chromatography (GPC) techniques. The results demonstrated that the degree of amination of the NH2‐PEO‐NH2 could be up to 1.95. The molecular weight of the PBLG‐PEO‐PBLG triblock copolymer could be adjusted easily by controlling the molar ratio of Bz‐L ‐Glu NCA to the macroinitiator NH2‐PEO‐NH2. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

8.
In this article, a novel “off–on” fluorescent probe 2-(pyren-1-ylmethylene)-1H-indene-1,3(2H)-dione ( PID ) for Hg2+ was designed and synthesized. The selectivity, concentration titration, pH titration, time dependence, limit of detection, and recognition mechanism of PID for Hg2+ in CH3CH2OH/H2O solution were also investigated. The results indicated that PID exhibited high selectivity, sensitivity, and fast response to Hg2+, and the limit of detection was as low as 20.7 nmol/L. In addition, PID could work in a wide pH range, and the determination of Hg2+ in water samples showed that it could be used as a potential detection tool in practical application.  相似文献   

9.
By heating with iron powder at 120–150° some γ-bromo-α, β-unsaturated carboxylic methyl esters, and, less smothly, the corresponding acids, were lactonized to Δ7alpha;-butenolides with elimination of methyl bromide. The following conversions have thus been made: methyl γ-bromocrotonate ( 1c ) and the corresponding acid ( 1d ) to Δα-butenolide ( 8a ), methyl γ-bromotiglate ( 3c ) and the corresponding acid ( 3d ) to α-methyl-Δα-butenolide ( 8b ), a mixture of methyl trans- and cis-γ-bromosenecioate ( 7c and 7e ) and a mixture of the corresponding acids ( 7d and 7f ) to β-methyl-Δα-butenolide ( 8c ). The procedure did not work with methyl trans-γ-bromo-Δα-pentenoate ( 5c ) nor with its acid ( 5d ). Most of the γ-bromo-α, β-unsaturated carboxylic esters ( 1c, 7c, 7e and 5c ) are available by direct N-bromosuccinimide bromination of the α, β-unsaturated esters 1a, 7a and 5a ; methyl γ-bromotiglate ( 3c ) is obtained from both methyl tiglate ( 3a ) and methyl angelate ( 4a ), but has to be separated from a structural isomer. The γ-bromo-α, β-unsaturated esters are shown by NMR. to have the indicated configurations which are independent of the configuration of the α, β-unsaturated esters used; the bromination always leads to the more stable configuration, usually the one with the bromine-carrying carbon anti to the carboxylic ester group; an exception is methyl γ-bromo-senecioate, for which the two isomers (cis, 7e , and trans, 7d ) have about the same stability. The N-bromosuccinimide bromination of the α,β-unsaturated carboxylic acids 1b , 3b , 4b , 5b and 7b is shown to give results entirely analogous to those with the corresponding esters. In this way γ-bromocrotonic acid ( 1 d ), γ-bromotiglic acid ( 3 d ), trans- and cis-γ-bromosenecioic acid ( 7d and 7f ) as well as trans-γ-bromo-Δα-pentenoic acid ( 5d ) have been prepared. Iron powder seems to catalyze the lactonization by facilitating both the elimination of methyl bromide (or, less smoothly, hydrogen bromide) and the rotation about the double bond. α-Methyl-Δα-butenolide ( 8b ) was converted to 1-benzyl-( 9a ), 1-cyclohexyl-( 9b ), and 1-(4′-picoly1)-3-methyl-Δα-pyrrolin-2-one ( 9 c ) by heating at 180° with benzylamine, cyclohexylamine, and 4-picolylamine. The butenolide 8b showed cytostatic and even cytocidal activity; in preliminary tests, no carcinogenicity was observed. Both 8b and 9c exhibited little toxicity.  相似文献   

10.
γ4‐Tripeptides and γ4‐hexapeptides, 1 – 4 , with OH groups in the 2‐ or 3‐position on each residue have been prepared. The corresponding 2‐hydroxy amino acids were obtained by Si‐nitronate (3+2) cycloadditions to the acryloyl derivative of Oppolzer's sultam and Raney‐Ni reduction of the resulting 1,2‐oxazolidines (Scheme 1). The 3‐hydroxy amino acid derivatives were prepared by chain elongation via Claisen condensation of Boc‐Ala‐OH, Boc‐Val‐OH, and Boc‐Leu‐OH, and NaBH4 reduction of the methyl 4‐amino 3‐oxo carboxylates formed (Scheme 2). The N‐Boc hydroxy amino acids were coupled in solution to give the γ‐peptides. CD Spectra of the new types of γ‐peptides were recorded and compared with those of simple γ2‐, γ3‐, γ4‐, and γ2,3,4‐peptides (Figs. 3, 4, and 5). An intense Cotton effect at ca. 200 nm ([Θ]=−2⋅105 deg⋅cm2⋅dmol−1) indicates that the hexapeptide built of (3R,4S)‐4‐amino‐3‐hydroxy acids (with the side chains of Val, Ala, Leu) folds to a secondary structure so far unknown. The stability of peptides from β‐ and γ‐amino acids, which carry heteroatoms on their backbones is discussed (Fig. 1). Positions on the γ‐peptidic 2.614 helix are identified at which non‐H‐atoms are `allowed' (Fig. 2).  相似文献   

11.
Complexation in the Fe2+–Fe3+N-(carboxymethyl)aspartic acid (H3L) system in aqueous solutions was studied by pH- and redox-potentiometric titration at 25°C and at an ionic strength of 0.1 (KCl). Depending on the H3L concentration and pH, neutral, protonated, and hydroxo complexes of iron(III) can be formed in the solutions. The stability constants for all the detected complexes were calculated, and the distribution plots for the fractions of complexes vs. the solution pH were constructed.  相似文献   

12.
The dinuclear ZnII complex [Zn2L(DNBA)2]BPh4 · EtOH ( 1 ) (DNBA = 3,5‐dinitrobenzonic acid) with an asymmetric dinuclear ligand, N‐4‐methyl‐homopiperazine‐N′‐[N‐(2‐pyridylmethyl)‐N‐2‐(2‐pyridylethyl)amine]‐1,3‐diamino‐propan‐2‐ol (HL), was synthesized and characterized. Single crystal X‐ray crystallographic analysis shows that the coordination around the two ZnII ions in 1 is significantly asymmetric, and the distance between both atoms is 3.426 Å, which is close to the Zn···Zn distance in related natural dinuclear metalloenzymes. Phosphodiesterase activity of Zn2L in situ formed from a 2:1 mixture of Zn2+ ion and HL was investigated using bis(4‐nitrophenyl) phosphate (BNPP) as substrate. The pH dependence of the BNPP cleavage in aqueous buffer media reveals a bell‐shaped pH‐kobs profile with an optimum at about pH 7.9, which is parallel to the formation of the dinuclear species Zn2L‐OH obtained from the potentiometric titration. The catalytic rate constant (kcat) is 6.30 × 10–4 s–1 at pH 7.9 and 25 °C, which is approx. 108‐fold higher than that of the uncatalyzed reaction. The homopiperazine bound deprotonated Zn‐OH group is responsible for the hydrolysis reaction. The possible mechanism for the BNPP cleavage promoted by Zn2L is proposed on the basis of kinetic and spectral analysis.  相似文献   

13.
On the Biosynthesis of γ-Dodecanolactone in Ripening Fruits: Flavor Constituents from Strawberries (Fragaria ananassa) and Peaches (Prunus persica) Administration of deuterium-labelled 9,10-expoxy[8,8-2H2]heptadecanoic acid 8a / b and 9,10-dihydroxy-[8,8-2H2]methylheptadecanoate 9 as lower analogues of oleic acid 1 to ripening fruits of strawberries (Fragaria ananassa) and peaches (Prunus persica) results in the emission of labelled γ-undecanolactone ( 5 ) as the lower analog of γ-dodecanolactone ( 2 ). The transformation proceeds with loss of a single D-atom from C(8) of the precursors. Early precursors, like the C17-epoxy-acids 8a / b yield (4R)-γ-undecanolactone ( 5 ) of high enantiomeric purity, while later intermediates results in (4R)-γ-undecanolactone ( 5 ) of low purity. The data support a biosynthetic sequence involving the consecutive action of an epoxide hydrolase and β-oxidation to generate the correct chain length of the lactone percursor. The final steps proceed via cyclization of the 3,4-dihydroxyundecanoic acid 13 to the 3-hydroxy-γ-undecanolactone 14 . Elimination of H2O and reduction of the intermediate γ-undec-2-enolactone 15 terminate the biosynthesis of 5 . The sequence is representative for the biosynthesis of naturally occurring γ-dodecanolactone ( 2 ).  相似文献   

14.
To determine some of the structural features of geiparvarin that account for its cytostatic activity in vitro, certain geiparvarin analogues modified in the furan-3(2H)-one moiety and the alkenyloxy substituent were synthesized and tested against the growth of 60 human cancer cell lines derived from nine cancer-cell types. These compounds demonstrated a strong growth-inhibitory activity against leukemia cell lines but were relatively inactive against non-small-cell lung cancers and CNS cancers. Comparison of the mean log GI50 values of γ-[(E)-1-methylprop-1-enyl]-α-methylidene-γ-butyrolactones 7 – 9 revealed that 7-[(E)-3-(2,3,4,5-tetrahydro-4-methylidene-5-oxofuran-2-yl)but-2-enyloxy]-2H- 1-benzopyran-2-one ( 8 ; −5.47) was more active than its 6-substituted counterpart 7 (−5.21) and its 3-chloro-4-methyl derivative 9 (−5.31) and had a potency similar to that of geiparvarin (log GI50=−5.41). These results indicated that the furan-3(2H)-one moiety of geiparvarin could be replaced by an α-methylidene-γ-butyrolactone unit without losing the anticancer potency, and that the best substitution site at the coumarin moiety was C(7). The alkenyloxy substituent of 8 was also replaced by a methoxy substituent. Among these α-methylidene-γ-butyrolactones, 7-[(2,3,4,5-tetrahydro-4-methylidene-5-oxo-2-phenylfuran-2-yl)methoxy]-2H-1-benzopyran-2-one ( 11 ) was the most potent with a mean log GI50 value of −5.83 and a range value of 132 (102.12).  相似文献   

15.
A number of multi-N?-poly(γ-benzyl-L -glutamyl)copoly(L -lysine γ-methyl-L -glutamate)s with branches having various degrees of polymerization and with various intervals of the grafting sites in the core molecule were prepared in N,N-dimethylformamide containing dimethyl sulfoxide by the reaction of N-carboxy anhydride of γ-benzyl L -glutamate with random copoly(L -lysine γ-methyl-L -glutamate)s of different composition with various anhydride-initiator ratios. The relationship between the intrinsic viscosity measured in a coil solvent, dichloroacetic acid (DCA), and the number-average molecular weight determined by osmometry was found to be expressed by the Mark–Houwink–Sakurada equation for the multichain copoly(α-amino acid)s which were made from the same polymeric initiator. The observed α values of the multichain copoly(α-amino acid)s in the equation were lower than that of linear poly(γ-benzyl-L -glutamate). The solvent induced helix–coil transition of the multichain copolymer was investigated in the chloroform?DCA system by the ORD technique. Two kinds of transition regions were clearly distinguished: The α-helices of the core molecules underwent the transition at lower DCA concentration and those of the branch chains at higher DCA concentration. The reduced viscosity of the multichain copoly-(α-amino acid) increased slightly between the two transition regions, in contrast to the large decrease in the reduced viscosity of linear poly(γ-benzyl-L -glutamate) during the helix–coil transition.  相似文献   

16.
Asymmetric telechelic α‐hydroxyl‐ω‐(carboxylic acid)‐poly(ε‐caprolactone) (HA‐PCL), α‐hydroxyl‐ω‐(benzylic ester)‐poly(ε‐caprolactone) (HBz‐PCL), and an asymmetric telechelic copolymer α‐hydroxyl‐ω‐(carboxylic acid)‐poly(ε‐caprolactone‐co‐γ‐butyrolactone) (HA‐PCB) were synthesized by ring‐opening polymerization of ε‐caprolactone (CL). CL and CL/γ‐butyrolactone mixture were used to obtain homopolymers and copolymer respectively at 150°C and 2 hr using ammonium decamolybdate (NH4) [Mo10O34] (Dec) as a catalyst. Water (HA‐PCL and HA‐PCB) or benzyl alcohol (HBz‐PCL) were used as initiators. The three polylactones reached initial molecular weights between 2000 and 3000 Da measured by proton nuclear magnetic resonance (1H‐NMR). Compression‐molded polylactone caplets were allowed to degrade in 0.5 M aqueous p‐toluenesulfonic acid at 37°C and monitored up to 60 days for weight loss behavior. Data showed that the copolymer degraded faster than the PCL homopolymers, and that there was no difference in the weight loss behavior between HA‐PCL and HBz‐PCL. Caplets of the three polylactones containing 1% (w/w) hydrocortisone were placed in two different buffer systems, pH 5.0 with citrate buffer and pH 7.4 with phosphate buffer at 37°C, and monitored up to 50 days for their release behavior. The release profiles of hydrocortisone presented two stages. The introduction of a second monomer in the polymer chain significantly increased the release rate, the degradation rate for HA‐PCB being faster than those for HBz‐PCL and HA‐PCL. At the pH studied, only slight differences on the liberation profiles were observed. SEM micrographs indicate that hydrolytic degradation occurred mainly by a surface erosion mechanism. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
Photochemistry of γ,δ-Methano-α-enones Direct excitation (λ = 254 or ≥ 347 nm) converts the γ,δ-methano-α-enone (E)- 10 into the isomeric ether 23 and the isomeric diene-ketone 24 . Furthermore, on 1π,π*-excitation (λ = 254 nm) (E)- 10 undergoes an 1,3-homosigmatropic rearrangement yielding the enone (E)- 25 . In addition (E → Z)-isomerization of (E)- 10 and conversion of 10 to the isomeric furan 28 is observed. The isomerization (E)- 10 → 23 , 24 and (E)- 25 proceeds by photocleavage of the C(γ), C(δ)-bond, whereas the formation of 28 occurs by photocleavage the C(γ), C(δ)-bond together with that of the C(γ), C(δ′)-bond of 10 . On direct excitation the bicyclic diene-ether 23 yields the methano-enone 10 , the dieneketone 24 and the tricyclic ether 29 . Evidence is given, that the conversion 23 → 10 is a singulet process. On the other hand, the isomerization 23 → 24 and the intramolecular [2 + 2]-photocycloaddition 23 → 29 are shown to be triplet reactions. Irradiation (λ = 254 nm) of the homoconjugated ketone 24 yields the isomeric ketone 27 by an 1,3-acyl shift. The excitation of the (E)-enone 25 induces (E → Z)-isomerization and photoenolization to give the homoconjugated ketone 26 .  相似文献   

18.
γ-Nitro-γ-butyrolactone By oxidation of 3-(1-nitro-2-oxocyclohexyl)propanal ( 1 ) with KMnO4, besides 3-(1′-nitro-2′-oxocyclohexyl)pripionic acid ( 2 ), the complete hydrolysis product 4-oxononanedioic acid ( 4 ) and the oxidized semi-hydrolysis product 5-(2-nitro-5-oxotetrahydro-2-furyl)pentanoic acid ( 3 ) were formed. The crystalline 3 decomposes at r.t. forming 4 and nitrous gases; its structure was established by X-ray determination.  相似文献   

19.
Spin-lattice relaxation times (T1), spin-spin relaxation times (T2), and nuclear Overhauser enhancements (NOE), at 75.5 MHz are reported for α-13C nuclei of poly (γ-benzyl-L -glutamate) in deuterated dimethylformamide at 60°C and of poly(γ-hexyl-L -glutamate) in cyclohexanone at 48 and 79°C. It is shown that for molecular weights above 105, the polypeptides cannot be considered as essentially rigid helices with internal librational motions; additional backbone flexing motions contribute to the relaxation behavior.  相似文献   

20.
The stability constants of some 1:1 Me2+-complexes of the following five-membered heterocyclic carboxylic acids have been measured in 50 perc. aqueous dioxane (I = 0,1; t = 25°): thiophene-2- (I), 3-phenylisothiazole-5- (II), tetrahydrothiophene-2- (III), furan-2- (IV), pyrrole-2- (V), and tetrahydrofuran-2-carboxylic acid (VI) (table 1 and 2). A comparison of the stability constants of the Cu2+-complexes of acetic acid (VII), benzoic acid (VIII), m-chlorobenzoic acid (IX), p-nitrobenzoic acid (X), and chloroacetic acid (VI) shows that the heterocyclic S and O atoms coordinate with Cu2+, i.e. Cu2+ chelates (structure XII) are formed (Figure 1). NMR. spectra (Fig. 2) give evidence for the coordination of the «aromatic» S atom in the Cu2+ complexes of thiophene-2-carboxylic acid (I), i.e. at least a part of the complexes are chelates. The NMR. spectra of furan-2-carboxylic acid (IV) gave no unequivocal results; in the case of pyrrole-2-carboxylic acid (V) the interaction between Cu2+ and the NH-group is very small (Fig. 4), i.e. a simple carboxylic acid complex is formed.  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