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.
The use of ortho‐phthalaldehyde (OPA) for the derivatization of amino acids (AA) is well known. It enables the separation of the derivatives on common reversed phase columns and improves the sensitivity with fluorescence detection. With the use of a chiral thiol an indirect enantioseparation of chiral amines and AAs is feasible. The major drawback of the OPA‐derivatization is the poor stability of the products. Here, a method with an in‐needle derivatization procedure is optimized to facilitate a quantitative conversion of the AA with OPA and the chiral thiols N‐acetyl‐L‐cysteine or N‐isobutyryl‐L‐cysteine, followed by a subsequent analysis, eluding the stability issue. Both enantiomers of a single AA were separated as OPA‐derivatives with a pentafluorophenyl column and a gradient program consisting of 50 mM sodium acetate buffer pH = 5.0 and acetonitrile. Fluorescence detection is commonly used to achieve sufficient sensitivity. In this study, the enantiomeric impurity of an AA can be detected indirectly with common UV spectrophotometric detection with a limit of quantitation of 0.04%. Seventeen different L‐AAs were tested and the amount of D‐AA for each individual AA was calculated by means of area normalization, which ranged from not detectable up to 4.29%. The recovery of the minor enantiomer of L‐ and D‐AA was demonstrated for three AAs at a 0.04% level and ranged between 92.3 and 113.3%, with the relative standard deviation between 1.7 and 8.2%.  相似文献   

2.
For the first time, a method for enantiomer resolution of the anticonvulsant Galodif (1‐((3‐chlorophenyl)(phenyl)methyl) urea) by chiral HPLC was developed, whereas the enantiomeric composition of 1‐((3‐chlorophenyl)(phenyl)methyl) amine—precursor in Galodif synthesis—cannot be resolved by this method. However, starting 1‐((3‐chlorophenyl)(phenyl)methyl) amine quantitatively forms diastereomeric N‐((3‐chlorophenyl)(phenyl)methyl)‐1‐camphorsulfonamides in reaction with chiral (1R)‐(+)‐ or (1S)‐(?)‐camphor‐10‐sulfonyl chlorides. The diastereomeric ratio of obtained camphorsulfonamides can be easily determined by NMR 1H and 13C spectroscopy. The DFT calculations of specific rotation of Galodif enantiomers showed good agreement with experimental data. The absolute configuration of enantiomers was proposed for the first time.  相似文献   

3.
The fluorogenic reaction involving three species, namely, a primary amine, o-phthaldialdehyde (OPA), and a thiol compound was studied with the fluorescence stopped-flow technique. The results are consistent with the reaction of the amine with a 1:1 adduct of OPA and the thiol compound. The equilibrium constant for the formation of the adduct, OPAME, from OPA and mercaptoethanol (ME) was determined to be 164 m?1. A survey of the rates of reaction of OPAME with various amino acids demonstrated that with OPA: ME:amine equal to 1:2.4:1 (total OPA concentration 0.5 to 3.0 × 10?3m), the reaction followed second-order kinetics, with k = 150 to 450 m?1s?1 at pH 9.O. The differences in rates are discussed in relation to structural differences between the amines. The reaction, when conducted under conditions of excess OPAME yielded pseudo-first-order kinetics, with rates consistent with the second-order rate constants. The rate of reaction of OPAME with alanine was maximal at pH 10.5–11, and a great excess of ME resulted in a slower rate. Slower rates were also observed if ME was replaced by dithiothreitol or 1-propanethiol.  相似文献   

4.
A sensitive and convenient method for the simultaneous determination of d- and l-aspartic acid in amino acid mixtures is described. The method involves derivatization of the mixture with a chiral fluorogen, followed by high-performance liquid chromatography on a reverse-phase column. The fluorogen used is an adduct of o-phthaldialdehyde with an optically active thiol, N-acetyl-l-cysteine. The sensitivity and accuracy of this method is similar to that using adducts of o-pthaldialdehyde with the achiral thiol, 2-mercaptoethanol. Five picomoles of d-aspartate can be accurately detected in the presence of a 100-fold excess of l-aspartate with a total analysis time (including derivatization) of 10 min.  相似文献   

5.
Derivatization reactions aimed at creating the basis for the chromatographic resolution of biologically and pharmaceutically important enantiomers are reviewed, with emphasis on the literature published in the last 10 years. Three main aspects of chiral derivatization are discussed. (a) Enantiomers containing suitable functional groups (amino, carboxyl, hydroxyl, epoxy, etc.) are transformed into covalently bonded diastereomeric derivatives using homochiral derivatizing agents. The diastereomers formed (esters, amides, urethanes, urea and thiourea, etc., derivatives) can be separated on achiral stationary phases. The derivatization reactions often afford further advantages, such as the improvement of chromatographic properties and the detectability of the solutes using UV and fluorimetric detectors. (b) Covalent but achiral derivatization is often necessary even with the use of chiral stationary phases enabling in principle direct enantioseparations (Pirkle-type columns, cyclodextrin-bonded phases, glycoprotein column and functionalized cellulose columns). The main goals of these derivatization reactions (which are analogous to those discussed above), are to introduce functional groups into the molecule of the enantiomers that improve the possibilities for chiral interactions or block functional groups to avoid non-specific interactions. (c) In the broader sense, the dynamic formation of diastereomers using chiral mobile phase additives (cyclodextrins, various reagents to form diastereomeric ion pairs, adducts, mixed metal complexes) can also be considered to be chiral derivatization reactions and is therefore briefly discussed also.  相似文献   

6.
Chiral α‐methylbenzyl amine is a well known and often used chiral auxiliary, e.g., in the resolution of racemates or asymmetric catalysis. In this work, α‐methylbenzyl amine and its derivatives N,α‐dimethylbenzyl amine, N,N,α‐trimethylbenzyl amine, and bis[α‐methylbenzyl] amine were investigated by vibrational circular dichroism (VCD) spectroscopy and density functional theory (DFT). For all compounds, stable low energy conformers were obtained by the DFT calculations and based on those, the theoretical vibrational absorption (VA) and VCD spectra were calculated and compared with experimental spectra. Hence, the absolute configurations and conformational preferences were determined. A qualitative comparison of all the experimental VCD spectra of the investigated chiral molecules supported by the calculated ones is given which clearly shows similarities between the spectra of the different chiral amines. These can be assigned to vibrations of the unchanged chiral center. Chirality 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

7.
The compound (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid was evaluated as a chiral nuclear magnetic resonance (NMR) solvating agent for a series of diamines and bicyclic β‐amino acids. The amine must be protonated for strong association with the crown ether. An advantage of (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid over many other crown ethers is that it undergoes a neutralization reaction with neutral amines to form the protonated species needed for binding. Twelve primary diamines in neutral and protonated forms were evaluated. Diamines with aryl and aliphatic groups were examined. Some are atropisomers with equivalent amine groups. Others have two nonequivalent amine groups. Association equilibria for these systems are complex, given the potential formation of 2:1, 1:1, and 1:2 crown‐amine complexes and given the various charged species in solution for mixtures of the crown ether with the neutral amine. The crown ether produced enantiomeric differentiation in the 1H NMR spectrum of one or more resonances for every diamine substrate. Also, a series of five bicyclic β‐amino acids were examined and (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid caused enantiomeric differentiation in the 1H NMR spectrum of three or more resonances of each compound. Chirality 27:708–715, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

8.
A detailed study of diastereomeric complexes of chiral ureido‐1,1′‐binaphthalene derivatives with chiral 1‐phenylethanol showed that a derivative bearing only one urea unit makes five times more stable complex with (S)‐enantiomer than with (R)‐enantiomer of the alcohol. This phenomenon could be used in chiral discrimination processes. The influence of individual parts of the structure on the complexation properties is shown. The probable structure of diastereomeric complexes based on experimental results and computational methods is proposed.  相似文献   

9.
An improved method, which is highly reproducible, was developed for the enantioseparation of racemic O‐ethyl phenylphosphonothioic acid ( 1a ) with brucine by introducing seeding to a supersaturated solution of the diastereomeric salt mixture. The present method gave both diastereomeric salts in high yields with a diastereomeric ratio of >99.5:0.5 upon choosing the crystallization solvent (MeOH for the ( (R)-1a salt and MeOH/H2O for the ( (S)-1a salt). The enantiopure acid (R)-1a , (S)-1a showed a good chirality recognition ability for not only strong bases, such as amines and amino alcohols, but also weakly basic alcohols and was applicable as a solvating agent to the 1H NMR determination of the enantiomeric excess of chiral amines, amino alcohols, and alcohols, including aliphatic substrates. Chirality 26:614–619, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
The thermodynamics underlying enantioselective complexation and partitioning behaviour are poorly understood. This paper presents a model that decouples the effects of enantioselective complexation and subsequent diastereomer partitioning. Regular solution theory is applied in a semi‐empirical manner to describe the diastereomer partitioning process, which is reported to be governed by hydrophobic interactions. The model was shown to give a good fit to experimental partitioning for the enantioselective extraction of phenylalanine isomers by two chiral extractants; a modified amino acid [copper (II) N‐decyl‐(L)‐hydroxyproline] and a chiral crown ether [(S)‐bis(phenylnaphtho)‐20‐crown‐6]. A variety of aliphatic and aromatic solvents were tested. The predicted and observed experimental enantioselectivities were found to vary exponentially with the difference in the solubility parameters of the aqueous and organic phases and with those of the two diastereomeric complexes formed. This model provides the basis for a better understanding of enantioselective partitioning effects. Chirality 11:241–248, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

11.
The synthesis and antibacterial activity of a series of nocathiacin I derivatives (4-20) containing polar water solubilizing groups is described. Thiol-Michael adducts containing acidic polar groups have reduced antibacterial activity whereas those with basic polar groups have retained very good antibacterial activity.  相似文献   

12.
Although the organocatalytic direct asymmetric Michael reactions of carbonyl compounds to nitroalkenes have been investigated intensely, the Michael reaction of the thioether‐based donors remains a rather undeveloped field. This work concerns the asymmetric Michael addition of aryl sulfanyl‐propan‐2‐one to nitroalkenes with benzoic acid as an additive, and chiral amine‐thiourea as a bifunctional organocatalyst. The reactions provided the highly functionalized chiral adducts with excellent enantioselectivities (up to 96% ee) and good yields. Moreover, the further transformed products exhibited excellent diastereoselectivity. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
Full NMR spectral assignments of the phospholene chalcogenides 1-12 are presented and their stereochemistry proven. The enantiomeric ratio of any of these compounds can be monitored easily by adding one mole equivalent of the chiral auxiliary Rh(2)[(R)-MTPA](4) (MTPA-H identical with Mosher's acid) and subsequent NMR inspection. Some surprisingly large diastereomeric signal dispersion is observed in the (1)H NMR spectra of the adducts, leading to the conclusion that intramolecular anisotropy interaction between groups inside the ligand molecules exists. The dependence of dispersion effects on the nature of the chalcogenide atom is investigated.  相似文献   

14.
This article reports, for the first time, on the absolute configuration of (+)-9-benzyloxy-α-dihydrotetrabenazine ( 8 ), as determined from the perspective of X-ray crystallography. Compound 8 was prepared by a six-step reaction using 3-benzyloxy-4-methoxybenzaldehyde ( 1 ) as a starting material. The X-ray crystal diffraction structure of two compounds, racemic 9-benzyloxy-tetrabenazine ( 5 ) and the diastereomeric salt of compound 8 , is also described for the first time in this article. The X-ray results and the chiral HPLC helped elucidate that compound 8 has an absolute configuration as 2R,3R,11bR. The crystal structure of racemic compound 5 contains two symmetry- independent molecules in the unit cell. Interestingly, while they are structural isomers, they are enantiomers, too, i.e., in solution, because they are not mirror images of each other in the crystal lattice. In order to elucidate the intermolecular interaction mechanism of the diastereomeric salt of compound 8 , its crystal packing was investigated with regard to the weak interactions, such as salt bridge, OH…O and CH…O hydrogen bonds, and intermolecular CH…π interaction. The results showed that the carbonyl-assisted salt bridges and the OH…O hydrogen bonds formed polar columns in the crystal structure of the diastereomeric salt of compound 8 , resembling butterflies with open wings as viewed along the c-axis. These polar columns were extended to three-dimensional network by intermolecular CH…O hydrogen bonds and intermolecular CH…π interactions. Chirality, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
Chiral derivatization reagents for resolution of biologically important compounds, such as chiral drugs by high-performance liquid chromatography (HPLC), based upon pre-column derivatization and diastereomer formation, are reviewed. The derivatization reagents for various functional groups, i.e., amine, carboxyl, carbonyl, hydroxyl and thiol, are evaluated in terms of reactivity, stability, wavelength, handling, versatility, sensitivity, and selectivity. The applicability of the reagents to the analyses of drugs and bioactive compounds are included in the text.  相似文献   

16.
This paper presents a systematic comparison of liquid chromatography (LC) and supercritical fluid chromatography (SFC) for Chiralcel OD and Chiralpak AD chiral stationary phases (CSPs), performed using various chiral compounds having a known or potential pharmaceutical activity. The chiral recognition mechanisms involved in LC and SFC for the enantiomeric separation of β-blockers have been studied more particularly. As a general rule, it appears that the presence of polar functions, like primary or secondary hydroxyl or amine functions, may result in marked discrepancies in selectivity between LC and SFC. This result is peculiar to cellulose- and amylose-derived CSPs, for which the interactions involved in chiral recognition mechanism are not always well balanced, contrary to what happens for independent CSPs. In the case of chiral resolution of polar solutes or polymer-type CSPs, the analyst should try both the LC and SFC techniques to be able to choose the more stereoselective one. © 1995 Wiley-Liss, Inc.  相似文献   

17.
Separation of racemic amino drugs (α-methylbenzeneethanamine, 6-amino-2-methyl-2-heptanol and 1-aminoethyl-benzenemethanol) and thiol drugs [N-(2-mercapto-1-oxopropyl) glycine, 2-mercaptopropanoic acid, and N-acetyl-3-mercaptovaline] has been evaluated after derivatization. ortho-Phthalaldehyde (OPA) and naphthalene-2,3-dicarboxaldehyde (NDA) were used with either homochiral thiols (N-acetyl-L-cysteine and N-acetyl-D-penicillamine) or amines [(-)-(1R,2S)-norephedrine, L-phenylalanine, L-tyrosine, and 3-hydroxy-L-tyrosine] as chiral selectors according to the analyte reactive group. The resulting 36 diastereoisomeric derivatives were studied using reversed-phase high-performance liquid chromatography (RP-HPLC) and capillary electrophoresis (CE). Of the CE modes, micellar electrokinetic chromatography (MEKC) using sodium dodecyl sulfate (SDS) as surfactant, β-cyclodextrin (β-CD)-modified capillary zone electrophoresis (β-CD-CZE), and β-CD-MEKC were applied. Results highlight respective performance of the reagents and separative techniques. All OPA derivatives of racemic amino drugs were resolved either by MEKC or β-CD-MEKC. In the case of racemic thiol drugs, 10 of the 12 OPA derivatives were resolved in β-CD-CZE. © 1995 Wiley-Liss, Inc.  相似文献   

18.
The synthesis of two fluorescent coumarin‐type chiral derivatization agents ( 4 and 11 ) is reported. A chiral side chain was introduced at position 7 of the coumarin via Mitsunobu reaction. The two coumarins bear in this side chain either a free amino group or a carboxyl group, making them useful for further transformations. Conjugates of chiral prototype drugs with 4 or 11 were prepared by amide coupling of the analyte's carboxyl group to the reagent's amine group, or vice versa. The separation of seven diastereomeric conjugates through achiral high‐performance liquid chromatography (HPLC) on a common C18 column is demonstrated. Chirality 25:957–964, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

19.
Methanethiosulfonate reagents may be used to introduce virtually unlimited structural modifications in enzymes via reaction with the thiol group of cysteine. The covalent coupling of enantiomerically pure (R) and (S) chiral auxiliary methanethiosulfonate ligands to cysteine mutants of subtilisin Bacillus lentus induces spectacular changes in catalytic activity between diastereomeric enzymes. Amidase and esterase kinetic assays using a low substrate approximation were used to establish kcat/KM values for the chemically modified mutants, and up to 3-fold differences in activity were found between diastereomeric enzymes. Changing the length of the carbon chain linking the phenyl or benzyl oxazolidinone ligand to the mutant N62C by a methylene unit reverses which diastereomeric enzyme is more active. Similarly, changing from a phenyl to benzyl oxazolidinone ligand at S166C reverses which diastereomeric enzyme is more active. Chiral modifications at S166C and L217C give CMMs having both high esterase kcat/KM's and high esterase to amidase ratios with large differences between diastereomeric enzymes.  相似文献   

20.
The application of porous graphitic carbon as adsorbing phase for direct separation of enantiomeric acids and amines using chiral ion-pair chromatography is described. The enantiomeric amines were separated as diastereomeric ion pairs with N-benzyloxycarbonylglycyl-L -proline, N-benzyloxycarbonylglycylglycyl-L -proline, or captopril as the chiral counterion. High enantioselectivities were obtained for amines having a hydrogen bonding function in the vicinity of the asymmetrical carbon atom. Quinine was the chiral counterion used to separate the enantiomeric acids. The strongly UV-absorbing quinine improved detection of solutes having low UV-absorbing properties, e.g., (R,S)-2-chloropropionic acid, by “indirect detection.” Retention and stereoselectivity of enanticmeric acids were regulated by the quinine concentration and by the addition of carboxylic acids as well as polar modifiers, e.g., methanol and 2-propanol, to the mobile phase. © 1992 Wiley-Liss, Inc.  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