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.
Shen K. Yang 《Chirality》1995,7(5):365-375
Oxazepam (OX), 3-O-methyloxazepam, 3-O-ethyloxazepam, temazepam (TMZ), 3-O-methyltemazepam, and 3-O-ethyltemazepam underwent acid-catalyzed nucleophilic substitution reaction (hydrolysis) in an acetonitrile–oxygen-18 water mixture to form either OX or TMZ in which the 3-hydroxyl group was either partially or fully labeled with an oxygen-18 atom. The dependence of the hydrolysis rates on solvent composition, temperature, ionic strength, and in deuterated solvent was studied by reversed-phase high-performance liquid chromatography (HPLC). The rates of racemization of enantiomeric compounds in acidic aqueous solutions were studied by both spectropolarimetry and chiral stationary phase HPLC. In acetonitrile: 2.5 M H2SO4 (4:1, v/v) at 50°C, enantiomers of OX and TMZ underwent racemization at rates ≥40-fold faster than the rates of hydrolysis. Enantiomeric 3-O-alkyl derivatives of OX and TMZ in acidic aqueous solutions did not themselves undergo racemization and it was their hydrolysis products (either OX or TMZ) that underwent racemization. © 1995 Wiley-Liss, Inc.  相似文献   

2.
Enantiomeric and racemic oxazepam (OX), 3-O-methyloxazepam (MeOX), and 3-O-ethyloxazepam (EtOX) were used to study racemization, heteronucleophilic, and homonucleophilic substitution reactions in anhydrous acidic methanol and ethanol. Kinetics of racemization and nucleophilic substitution reactions in nondeuterated and deuterated solvents were determined by circular dichroism spectropolarimetry, chiral stationary phase high-performance liquid chromatography (HPLC), reversed-phase HPLC, and mass spectrometry. Several reactions occurred when (S)-OX, for example, was dissolved in acidic methanol: (1) (S)-OX itself underwent spontaneous racemization, (2) the 3-hydroxyl group of (S)-OX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (S)-MeOX, (3) the 3-methoxy group of (S)-MeOX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (S)-MeOX, and (4) the 3-methoxy group of (R)-MeOX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (R)-MeOX. Repetitive reactions 3 and 4 eventually resulted in a racemic MeOX. Similar reactions occurred for an enantiomeric OX in acidic ethanol. © 1996 Wiley-Liss, Inc.
  • 1 This article is a US Government work and, as such, is in the public domain in the United States of America.
  •   相似文献   

    3.
    J L Hoffman 《Biochemistry》1986,25(15):4444-4449
    The chirality of biologically active S-adenosyl-L-methionine (AdoMet) is S,S, where the designations refer to the sulfur and the alpha-carbon, respectively. This paper describes a cation-exchange high-performance liquid chromatographic (HPLC) method for separating (S,S)-AdoMet from the biologically inactive (R,S)-AdoMet that results from racemization at the sulfur. This method was used to measure the rates of the degradation reactions of (S,S)-AdoMet as a function of pH. These reactions and the first-order rate constants, which were found at 37 degrees C and pH 7.5, are racemization, 1.8 X 10(-6) s-1; cleavage to homoserine lactone and 5'-(methylthio)adenosine, 4.6 X 10(-6) s-1; and hydrolysis to adenine and S-pentosylmethionine, 3 X 10(-6) s-1. Racemization showed no change in rate over the pH range from 7.5 to 1.5. The cleavage reaction persisted until the pH was lowered to 1.5, but hydrolysis ceased at pH 6. Commercial samples of nonradioactive AdoMet contained 20-30% (R,S)-AdoMet, while a sample of [methyl-3H]AdoMet had less than 1% (R,S)-AdoMet. Preparing enzyme substrates by mixing such samples will cause an underestimate of specific activity and an overestimate of the amount of product. The (R,S)-AdoMet/(S,S)-AdoMet ratio in mouse liver was 0.03, much less than the value of 0.19 calculated from the above rate constants. An enzyme extract from mouse liver did not degrade (R,S)-AdoMet, but a more thorough search may find such an activity. In any event, the cleavage and hydrolysis reactions partially balance the racemization of (S,S)-AdoMet in vivo and prevent excessive accumulation of (R,S)-AdoMet.  相似文献   

    4.
    The influence of the amount and nature of the necessary tertiary amine on racemization attending the BOP-mediated coupling of N-benzyloxycarbonylglycyl-dipeptides has been examined by determining the epimeric products by high-performance liquid chromatography. Racemization was encountered in all cases examined that included dichloromethane as solvent. The least racemization occurred when the base was diisopropylethylamine, the most when the base was N-methylmorpholine. Excess base promoted racemization. Racemization was significantly diminished but not eliminated when 1-hydroxybenzotriazole was added to the reaction mixtures. The above results pertain to the coupling of peptide segments; they do not pertain to the coupling of N-alkoxycarbonylamino acids.  相似文献   

    5.
    Because of the constantly increasing demand for optically pure drugs it is of great importance to elucidate factors affecting stereochemistry, in order to provide a stable formulation with a high chiral quality of the desired isomer. Therefore, the effects of cyclodextrins (CyDs) and their alkylated and hydroxyalkylated derivatives on racemization and hydrolysis of (?)-(S)-hyoscyamine and (?)-(S)-scopolamine were examined kinetically and spectroscopically (NMR). Direct methods, based on a chiral and achiral chromatographic phase system, were used to determine their degradation products and enantiomer composition during stability tests. All different CyDs, except α-CyD, retarded racemization and hydrolysis. The inclusion of the drug substances in CyDs inhibits the attack of hydroxyl ions and/or water molecules and thus retards the racemization and hydrolysis. The racemization of the tropic acid alkaloids is dependent on the pH and temperature. NMR studies were used to evidence the formation of a soluble 1:1 complex in aqueous solution. © 1993 Wiley-Liss, Inc.  相似文献   

    6.
    We investigated the degradation and racemization of zopiclone (ZOP) enantiomers in plasma and partially aqueous solutions (ethanol:phosphate buffer). Degradation and racemization increased with increasing pH and temperature. Degradation products were identified by means of mass spectrometry, which revealed hydrolysis of the carbamate function and opening of the pyrrolidone ring. In plasma, neither degradation nor racemization occurred after 6 months of storage at -20°C and subsequent extraction. © 1995 Wiley-Liss, Inc.  相似文献   

    7.
    Xanthan biopolymer has been produced in a single-stage continuous fermentation with Xanthomonas campestirs NRRL B-1459, using a viscostal control method instead of the conventional chemostat method. A Bendix Ultraviscoson
  • 1 The mention of firm names or trade products does not imply that they are endorsed or recommended by the U.S. Department of Agriculture over other firms or similar products not mentioned.
  • sensed the fermentor viscosity, and the recorder–controller actuated the feed medium pump in an on–off control mode. Since all continuous fermentations eventually become contaminated or suffer culture variation, this work served also to demonstrate the effectiveness of the viscostat control. Neither the presence of a mold contaminant with specific growth rates lower than that of X. campestris, nor the presence of a bacterial contaminant of specific growth rate greater than X. campestris, affected the maintenance of constant viscosity in this control system.  相似文献   

    8.
    In the enantioselective hydrolysis of the di-O-acetyl derivatives of meso-1,3-diol catalyzed by lipases, racemization of the monoacetate products occurs due to non-enzymatic general base-catalyzed acyl migration. The rate of acyl migration increases with increase of pH and buffer concentration. A mechanism of the migration has been proposed to proceed through a six-member ring transition that accounts for the experimental results. The acyl migration, however, was not observed in the enantioselective transesterification of meso-1,3-diols in neutral organic solvents.  相似文献   

    9.
    Under basic pH conditions, the heavy chain 220-light chain 214 (H220-L214) disulfide bond, found in the flexible hinge region of an IgG1, can convert to a thioether. Similar conditions also result in racemization of the H220 cysteine. Here, we report that racemization occurs on both H220 and L214 on an IgG1 with a λ light chain (IgG1λ) but almost entirely on H220 of an IgGl with a κ light chain (IgG1κ) under similar conditions. Likewise, racemization was detected at significant levels on H220 and L214 on endogenous human IgG1λ but only at the H220 position on IgG1κ. Low but measurable levels of d-cysteines were found on IgG2 cysteines in the hinge region, both with monoclonal antibodies incubated under basic pH conditions and on antibodies isolated from human serum. A simplified reaction mechanism involving reversible β-elimination on the cysteine is presented that accounts for both base-catalyzed racemization and thioether formation at the hinge disulfide.  相似文献   

    10.
    Racemization in the synthesis of tripeptide intermediates and their polymers was investigated, using L -amino acid oxidase. Stepwise investigation of peptide intermediates showed no racemization during peptide coupling steps or deprotection of benzyl esters by hydrogenolysis. Saponification of one of the methyl esters produced some racemization. Preparation of active esters from N-protected tripeptide acids containing optically active C-terminal amino acid, with one exception, produced racemization. The fractionated polymers were found to contain less racemized amino acids than the crude products or starting monomeric tripeptides, indicating that the racemized sequences gave rise to lower molecular-weight oligomers. The sequences investigated were -Pro-Pro-Ala-, -Ala-Pro-Pro-, -Val-Pro-Pro-, -Pro-Pro-Leu-, -Pro-Gly-Leu-, -Pro-Gly-Phe-, -Pro-Gly-Val-, -Gly-Val-Pro-, -Phe-Pro-Gly-, -Leu-Pro-Gly-, and Ile-Pro-Gly-.  相似文献   

    11.
    We have studied the hydrolysis of guanosine 5-phospho-2-methylimidazolide, 2-MeImpG, in aqueous buffered solutions of various pH's at 75°C and 37°C. At 75°C and pH1.0, two kinetic processes were observed spectrophotometrically: the first and more rapid one is attributed to the hydrolysis of the phosphoimidazolide P-N bond; the second and much slower one, to the cleavage of the glycosidic bond. At 37°C, pH 2.0, the spectrophotometrically determined rate constant of P–N bond hydrolysis was confirmed by using high pressure liquid chromatography, HPLC. With the latter technique it was possible to separate reactants and products and also to extend the pH-rate profile into the neutral region where rates are slower and, therefore, difficult to measure spectrophotometrically. The pH-rate profiles at both temperatures exhibit similar behavior. At pH<2 the pseudo-first-order rate constant increases with decreasing pH; in the region 27. These data are consistent with a reactivity order zwitterion>anion for P–N bond hydrolysis. It is noteworthy that P–N bond hydrolysis in phosphoimidazolides is very slow compared to other phosphoramidates. This may be one of the reasons why this compound showed extraordinary ability in forming long oligomers under template-directed conditions.  相似文献   

    12.
    Owing to slow rotation of a sterically constrained dimethylamide substituent, two slowly interconverting enantiomers of a preclinical candidate for pharmaceutical development, 1, (6-(3-Chloro-4-fluoro-benzyl)-4-hydroxy-2-methyl-3,5-dioxo-2,3,5,6,7,8-hexahydro-[2,6]naphthyridine-1-carboxylic acid dimethylamide) are observed by chiral chromatography. Isolation of pure enantiomer by preparative chiral chromatography followed by enantiopurity analysis over time allowed for a study of the kinetics of enantiomer interconversion under a variety of conditions. Relatively slow racemization was observed in alcohol solvents, with a half life on the order of 5-10 h. A dramatic influence of aqueous buffer pH on racemization was noted, with higher pH leading to rapid racemization within a few minutes, and lower pH leading to essentially no racemization for periods up to a week. A hypothesis explaining this unusual effect of pH on carboxamide bond rotation is offered, and some suggestions for potential utility of such a system are considered.  相似文献   

    13.
    The configurational lability of enantiomers can be characterized by different terms, each defining a specific process. Racemization relates to the macroscopic and statistical process of the irreversible transformation of one of the enantiomers into the racemic mixture. Enantiomerization refers to the microscopic and molecule process of the reversible conversion of one enantiomer into the other. Methods allowing the experimental determination of rate constants of racemization (krac) and enantiomerization (kenant) are discussed, and it is shown that kenant = 1/2 krac. Neglect of this fact is a source of some confusion in the literature. When two or more elements of chirality are present in a molecule and one of them is configurationally labile, epimerization occurs, a particular case of diastereomerization. These processes of interconversion between diastereomers are kinetically more complicated than racemization and enantiomerization since the rate constants of the forward and reverse reactions are always different (kdiast/A-to-B ≠ kdiast/B-to-A), however small the difference. An important aspect of the configurational lability of stereoisomeric drugs is the time scale of the phenomenon. When interconversion occurs to a significant extent during the residence time of a drug in the body, a pharmacological time scale is implied. In contrast, the pharmaceutical time scale refers to slower rates of interconversion that affect the configurational purity of a drug during its shelf-life. © 1995 Wiley-Liss, Inc.  相似文献   

    14.
    在纯乙酸溶剂中,以乙酸酐酰化消旋酪氨酸,消旋温度为90℃,L-酪氨酸和乙酸酐的摩尔比为12,每消旋1gL-酪氨酸时,溶剂乙酸的用量为5mL,效果较好。同样的方法应用于L-脯氨酸,L-苏氨酸,L-缬氨酸,L-赖氨酸,L-半胱氨酸等氨基酸,消旋率均可达到100%。  相似文献   

    15.
    Racemization in the synthesis of peptide intermediates and their polymers was investigated, using L-amino acid oxidase. The formation of N-hydroxysuccinimide esters from N-protected peptide acids yielded optically pure products in contrast to p-nitrophenyl and pentachlorophenyl active esters. The racemization in the polymerization step was found to be base sensitive. Partially racemized polymer can result from optically homogeneous monomer. Thus, the optical integrity of active monomer species carries no guarantee for that of the polymer.  相似文献   

    16.
    First enantioselective synthesis of S-(-)-1-[3-(4-tert-butylphenyl)-2-methyl]propyl-cis-3,5-dimethylmorpholine (6), biologically active enantiomer of the systematic fungicide fenpropimorph, is reported. It comprises reacting 4-tert-butylbenzylbromide with methyldiethylmalonate, decarbethoxylation of 2 into racemic 3-(4-tert-butylphenyl)-2-methylpropionic acid ethylester (3) in DMSO in the presence of alkali, then Pseudomonas sp. lipase catalyzed kinetic resolution of racemic 3 into S-(+)-acid (4), base-catalyzed racemization and recycling of the R-(-)-ester 3, acylation of cis-3,5-dimethylmorpholine, and final reduction of the intermediary amide 5 to provide enantiomerically pure S-(-)-6.  相似文献   

    17.
    Shen K. Yang 《Chirality》1994,6(3):175-184
    Enantiomers of 3-O-methyloxazepam (MeOX) and 3-O-ethyloxazepam (EtOX) were resolved by chiral stationary phase high-performance liquid chromatography (CSP-HPLC). Reaction kinetics and deuterium isotope effects of acid-catalyzed racemization of enantiomeric MeOX in ethanol and enantiomeric EtOX in methanol were studied by spectropolarimetry. The acid-catalyzed heteronucleophilic substitution reactions of racemic MeOX in ethanol and racemic EtOX in methanol were studied by reversed-phase HPLC. Thermodynamic parameters involved in the reactions were obtained by temperature-dependent reaction rates. The effects of solvent's dielectric constant on the heteronucleophilic substitution reactions were also determined. A nucleophilically solvated and transient C3 carbocation intermediate resulting from an N4-protonated enantiomer, derived from a 1,4-benzodiazcpine either in M (minus) or P (plus) conformation, is proposed to be an intermediate and responsible for the acid-catalyzed stereoselective nucleophilic substitution and the resulting racemization. © 1994 Wiley-Liss, Inc.  相似文献   

    18.
    Most amino acids contain chiral centres and exist as both D‐enantiomer and L‐enantiomer. The optically pure enantiomer is often more valuable than the racemate. Enzymatic resolution provides an effective strategy to obtain optically pure amino acids but often results in large amounts of unwanted isomer. In this study, optically pure L‐glufosinate (L‐PPT) was obtained by coupling amidase‐mediated hydrolysis of N‐phenylacetyl‐D,L‐glufosinate with racemization of N‐phenylacetyl‐D‐glufosinate (NPDG), which exclusively exhibits effective herbicidal properties compared with its D‐enantiomer. To improve the yield of L‐PPT, the racemization reaction conditions were optimized, and through single‐factor experiments, the optimal reaction temperature, reaction time, and mole ratio of phenylacetic acid to NPDG were determined to be 150°C, 30 minutes, and 1.5, respectively. The response surface methodology was applied to further optimize the racemization conditions, and the final yield of L‐PPT reached 96.13% with optimum reaction temperature of 154°C, reaction time of 23 minutes, and phenylacetic acid/NPDG mole ratio of 1.7, respectively. Moreover, adding a small amount of acetic anhydride further raised the yield of L‐PPT to 97.02%.  相似文献   

    19.
    The demand for enantiopure substances, e.g. for pharmaceutical applications or fine chemical production, continues to increase. This has led to the development of numerous stereoselective synthesis methods. Nevertheless a large number of chemical syntheses still result in racemic mixtures making a subsequent enantioseparation step necessary and thus are restricted to a maximum yield of 50%. Our work focuses on strategies to overcome this limitation by combining physicochemical separation processes with enzymatic racemization of the unwanted enantiomer in order to produce enantiopure amino acids. This paper deals with the production and characterization of a suitable amino acid racemase with broad substrate specificity (EC 5.1.1.10) from Pseudomonas putida which we cloned into Escherichia coli. Two enzyme lyophilizates of different purity were obtained from which the crude (CL) was sufficient for the racemization of methionine (Met) and the pure (PL) was used for asparagine (Asn). Racemization reactions of D-/L-Asn in H2O and D-/L-Met in 95 vol.% 100 mM KPi-buffer, 5 vol.% methanol (MeOH) at different pH values and temperatures were characterized. The studied range of reaction parameters was chosen in dependency on planned enantioseparation processes. We found increasing Vmax values when temperature was risen stepwise from 20 to 40 °C for both systems and when pH was shifted from 6 to 8 for the Met system. The presented results provide the basis for engineering enzyme-assisted physicochemical enantioseparation processes.  相似文献   

    20.
    The relative racemization rates of free and peptide-bound serine and aspartic acid have been studied as a function of pH at 100°C. These results have been used to explain the observed relative in vivo racemization rates of serine and aspartic acid in human tooth dentin and human ocular lens proteins.  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