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 0 毫秒
1.
The determination of enantiomeric purity of (R)- and (S)-2-hydroxy-4-phenylbutyric acid by chiral HPLC is described. Good resolution has been obtained on covalently bonded L-hydroxyproline saturated with Cu(II) ions. The method makes possible the determination of enantiomeric purity in media containing growing cells. © 1994 Wiley-Liss, Inc.  相似文献   

2.
Studies of the perturbing effect of chiral solvating agents (CSAs) 5a and mostly of 5c upon the NMR spectra of chiral Delta(2)-oxazoline 1 demonstrated the ability of these fluoroalcohols to afford diastereomeric solvates from these solutes. Thus, for all tested Delta(2)-oxazolines 1Aa-d, 1Ba, and 1e there is at least one possibility to proceed to their enantiomeric discrimination either by (1)H or (19)F NMR using these CSAs (see Fig. 1). NMR results are discussed from substrate and CSA structure standpoints and a solvation model is proposed on the basis of the inequivalence senses generally observed. Then the method was applied to extracts of incubated locust tissues obtained by solid phase extraction (SPE) after a partial unmasking of the substrate 1.  相似文献   

3.
Five optically active urea derivatives ( 1 ‐ 5 ) were used as NMR solvating agents for analysis of the optical purity of different 2‐arylpropanoic acids commonly used as nonsteroidal anti‐inflammatory drugs. These novel chiral solvating agents were more efficient at discriminating the respective enantiomers of targets than the chiral solvating agents known so far, without the need to add a base for achieving the signal splitting. The advantages and limits of the use of these novel chiral solvating agents were studied.  相似文献   

4.
Brewer BN  Zu C  Koscho ME 《Chirality》2005,17(8):456-463
The ability to use mixtures of deprotonated N-(3,5-dinitrobenzoyl)amino acids as chiral selectors for the determination of enantiomeric composition by electrospray ionization-mass spectrometry is demonstrated. For each experiment, two N-(3,5-dinitrobenzoyl)amino acids were chosen such that each would have opposite selectivity for the enantiomers of the analyte. Electrospray ionization-mass spectrometry, monitored in the negative ion mode, of solutions containing the two N-(3,5-dinitrobenzoyl)amino acids, sodium hydroxide, and the analyte, in a one-to-one mixture of methanol and water, afford peaks in the mass spectrum that correspond to the deprotonated 1:1 analyte-selector complexes. The ratio of the intensities of the complexes in the mass spectrum can be related to the enantiomeric composition of the analyte. Additionally, the sense and extent of chiral recognition is consistent with chromatographic observations, using chiral stationary phases derived from N-(3,5-dinitrobenzoyl)amino acids. Each analysis of enantiomeric composition requires less than 10 s to complete, indicating that this method has great potential for the development of fast-/high-throughput chiral analyses.  相似文献   

5.
《Chirality》2017,29(6):247-256
The enantioresolution and determination of the enantiomeric purity of 32 new xanthone derivatives, synthesized in enantiomerically pure form, were investigated on (S ,S )‐Whelk‐O1 chiral stationary phase (CSP). Enantioselectivity and resolution (α and RS) with values ranging from 1.41–6.25 and from 1.29–17.20, respectively, were achieved. The elution was in polar organic mode with acetonitrile/methanol (50:50 v/v ) as mobile phase and, generally, the (R )‐enantiomer was the first to elute. The enantiomeric excess (ee ) for all synthesized xanthone derivatives was higher than 99%. All the enantiomeric pairs were enantioseparated, even those without an aromatic moiety linked to the stereogenic center. Computational studies for molecular docking were carried out to perform a qualitative analysis of the enantioresolution and to explore the chiral recognition mechanisms. The in silico results were consistent with the chromatographic parameters and elution orders. The interactions between the CSP and the xanthone derivatives involved in the chromatographic enantioseparation were elucidated.  相似文献   

6.
Three chiral stationary phases, obtained by derivatizing γ-mercaptopropylsilanized silica gel with quinine, quinidine, and cinchonidine, have been employed in the resolution of N-acyl derivatives of β-hydroxyphenethylamines. The use of circular dichroism for detection and NMR analysis of analyte–selector mixtures provides an experimental basis for preliminary assignment of a recognition mechanism.  相似文献   

7.
Esters of 1-(1-naphthly)ethylurea derivatives of L-valine, L-leucine, L-tert-leucine, and L-proline are examined as organic-soluble chiral nuclear magnetic resonance (NMR) resolving agents. The reagents are useful for resolving the spectra of chiral sulfoxides, amines, alcohols, and carboxylic acids. Enantiomeric resolution is caused by a combination of diastereomeric effects and the different association constants of the substrates with the resolving agents. Organic-soluble lanthanide species are added to resolving agent-substrate mixtures and often enhance the enantiomeric resolution. The enhancement occurs because the substrate that exhibits weaker binding with the resolving agent is more available to bond to the lanthanide. Broadening in the spectra with lanthanides is reduced at 50°C. Enantiomeric resolution is still observed at elevated temperatures. Chirality 9:1–9, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

8.
The use of P(III) and P(V) organophosphorus derivatizing agents prepared from C(2) symmetrical (1R,2R)- and (1S,2S)-trans-N,N'-bis-[(S)-alpha-phenylethyl]-cyclohexane-1,2-diamines 1 and 2, as well as (1R,2R)- and (1S,2S)-trans-N,N'-bis-[(S)-alpha-phenylethyl]-4-cyclohexene-1,2-diamines 3 and 4 for the determination of enantiomeric composition of chiral carboxylic acids by (31)P NMR, is described.  相似文献   

9.
The determination of the enantiomeric impurity, i.e., the percentage of (+) N?0437 (= N?0924) in several batches of (??) N-0437 (= N-0923) by chiral HPLC is described. Enantiomeric impurities were calculated based on the peak areas of the two baseline separated enantiomers in the chromatogram. The enantiomeric impurities found in different batches ranged from 0.02% to 0.11%. Calibration curves of the two isomers of N-0437 (Fig. 1,) were made twice to study the reproducibility and linearity of the method. The absorbance ratio, N-0923/N-0924, was found to be 1.02 with a relative standard deviation (RSD) of 9% over the whole concentration range used for the calibration curves.  相似文献   

10.
The attachment of radiometals to monoclonal antibodies for medical applications requires extreme stability under physiological conditions, with no significant release of metal. Chelators that can hold radiometals like 111In, 67Ga, and 90Y with high stability under these conditions are essential for radiotherapy or immunoscintigraphy. 2-(p-Nitrobenzyl)-1,4,7,10-tetraazacyclododecane- N,N',N',N'-tetraacetic acid (nitrobenzyl-DOTA) is one of the most promising bifunctional chelating agents. A large-scale synthesis of nitrobenzyl-DOTA is described. The overall yield for the nine-step synthesis sequence starting from nitrophenylalanine is 5.6%. Synthesis of nitrobenzyl-DOTA according to the new procedure yields up to approximately 10 g without special apparatus. Both enantiomers of the chiral chelate nitrobenzyl-DOTA have been prepared, and their enantiomeric purity has been checked by chiral chromatography.  相似文献   

11.
Molecular dynamics simulations were performed on complexes of (S)-methyl N-(2-naphthyl)alaninate (NAP) with the enantiomers of N-(3,5-dinitrobenzoyl)leucine n-propylamide (DNB), which are used as models for chiral stationary-phase systems developed by Pirkle and co-workers. These studies were undertaken to qualitatively examine (pictorially) the role of entropic effects in these systems. The results of the dynamics calculations were used to refine the search for low-energy conformers. The structures were refined by the use of BioDesign's molecular mechanics method implemented in Biograf. The results of the structural refinements support our previous observation that the SR complex can achieve the same three primary interactions which are observed in the SS structure (i.e., two intermolecular hydrogen bonds and pi stacking) without a significant increase in energy. In addition, these primary interactions are conserved during molecular dynamics simulations with the occurrence of conformations which differ only in the rotational states of the alkyl side chains and ester group (which bears two potential hydrogen bond acceptors utilized in both the homo- and heterochiral complexes). The major difference in the two complexes is the relative position of the sec-butyl group and hydrogen atom on DNB's chiral center, both of which are outside the primary interaction region. All other local minima which have different relative pi orientations (“front–back,” “back–back,” and “back–front” as defined herein) are not sufficiently populated to make more than a negligible contribution to the statistical (time- or energy-averaged) analysis of the (SS)- and (SR)-NAP–DNB complexes. Thus the entropic effects observed in this study (e.g., alkyl side chain or ester group rotations) do not show evidence of qualitative differential effects on the maintenance of the same three primary interactions by both the homo- and heterochiral complexes. The reliability of the present study, which provides pictorial representations of the entropic effects, is not sufficient to determine whether the entropic effects observed herein are sufficient to achieve enantiomeric discrimination alone or in conjunction with other factors (e.g., conformational strain energy). Thus, all of the computational studies we have performed to date (i.e., our previous studies, which include strain energy and through-space field effects, and the present study, which includes entropic effects) show no evidence of any qualitative difference in the homo- and heterochiral complexes in terms of maintaining the same three “contact points”.  相似文献   

12.
The title compounds have been prepared from the respective 3,3,3-trifluoro-2-methoxy-2-phenylpropanoic acids (MTPA) by a three-step synthesis with MTPA chloride and MTPA amide as reaction intermediates. The requested compounds were obtained in high chemical yields without any change in optical purity during the preparation. To ascertain the usefulness of this auxiliary agent in the chiral analysis, isomeric 3,3,3-trifluoro-2-phenylpropanoyl isocyanates were subjected in NMR tubes to noncatalyzed reactions with 16 different commercially available chiral alcohols. The steric arrangement of all diastereomers prepared correlated well with their NMR spectral nonequivalence data (Deltadelta), thus demonstrating the regular sign distribution of Deltadelta of particular groups according to the devised molecular model. The usefulness of the novel derivatization is discussed.  相似文献   

13.
Carbamate diastereomers 3b-18b were prepared from easily accessible (S)-2-chloro-2-fluoroethanoyl isocyanate (1) and various secondary chiral alcohols. Compound 1 as a chiral analog of trichloroacetyl isocyanate undergoes the reaction with alcohols very fast, thus blocking the hydroxyl group for the purposes of NMR investigation. Moreover, the correlation of stereochemistry of 3b-18b with their (1)H NMR spectra revealed that the constitution as well as configuration influences regularly the values of chemical shift difference (deltadelta = delta(R) - delta(S)) except for those diastereomers bearing simple alkyl groups in the molecule. Spectral as well as crystallographic data manifest the predominant planar conformation of the central part of the molecule. Due to the good accessibility and high reactivity in particular, the acylisocyanate 1 might be considered, to some extent, an alternative for TAI giving additional information on a compound's spatial structure.  相似文献   

14.
An efficient methodology for the preparation of the α‐tetrasubstituted proline analog (S,S,S)‐2‐methyloctahydroindole‐2‐carboxylic acid, (S,S,S)‐(αMe)Oic, and its enantiomer, (R,R,R)‐(αMe)Oic, has been developed. Starting from easily available substrates and through simple transformations, a racemic precursor has been synthesized in excellent yield and further subjected to HPLC resolution using a cellulose‐derived chiral stationary phase. Specifically, a semipreparative (250 mm × 20 mm ID) Chiralpak® IC column has allowed the efficient resolution of more than 4 g of racemate using a mixture of n‐hexane/tert‐butyl methyl ether/2‐propanol as the eluent. Multigram quantities of the target amino acids have been isolated in enantiomerically pure form and suitably protected for incorporation into peptides. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

15.
Copper(II) complexes of N2-octyl-(S)-phenylalaninamide (Noc-Phe-NH2), N2-dodecyl-(S)-phenylalaninamide (Ndo-Phe-NH2), and N2-octyl-(S)-norleucinamide (Noc-NLeu-NH2), dynamically adsorbed on a reversed-phase C18 column, were able to perform the direct enantiomeric separation of unmodified amino acids, amino acid amides and esters, hydroxy acids, and dipeptides by elution with aqueous or mixed aqueous-organic solutions containing copper(II) sulphate or acetate. The role played by several parameters in the separation procedure was examined with the copper(II) complex of Noc-Phe-NH2 [concentration of the copper(II) ion in the eluent, pH and eluent polarity, amount of adsorbed selector]. The separation was shown to occur entirely on the stationary phase. The mechanism of chiral discrimination is discussed in terms of the chromatographic parameters and of the structure of the copper(II) complexes in solution and in the solid state. The chiral stationary phase maintained its separation ability for about 3 months. However, the column could be easily restored by recovering the selector with methanol and repeating the loading procedure. © 1996 Wiley-Liss, Inc.  相似文献   

16.
2(2-Bromoethyl)bromobenzene was subjected to microbial oxidation by the whole cells of Pseudomonas putida 39/D and JM109(pDTG601) yielding (3R,4S)-2-(2-bromoethyl)-bromocyclohexa-1,5-diene-3,4-diol. © 1995 Wiley-Liss, Inc.  相似文献   

17.
H Navrátilová 《Chirality》2001,13(10):731-735
S-Mosher acid 1 induced chemical-shift differences (Delta delta) in NMR spectra of chiral trans-4-fluorophenyl-3-substituted-1-methylpiperidines. The magnitude of Delta delta in (1)H and (19)F NMR spectra varied with the substituent at position 3 of the piperidine ring. The magnitudes of Delta delta observed for certain protons and for the fluorine in the 4-fluorophenyl group were sufficiently large to allow determination of enantiomeric composition.  相似文献   

18.
It has been proposed that the chiral inversion of the 2-arylpropionic acids is due to the stereospecific formation of the (-)-R-profenyl-CoA thioesters which are putative intermediates in the inversion. Accordingly, amino acid conjugation, for which the CoA thioesters are obligate intermediates, should be restricted to those optical forms which give rise to the (-)-R-profenyl-CoA, i.e., the racemates and the (-)-(R)-isomers. We have examined this problem in dogs with respect to 2-phenylpropionic acid(2-PPA). Regardless of the optical configuration of 2-phenylpropionic acid administered, the glycine conjugate was the major urinary metabolite and this was shown to be exclusively the (+)-(S)-enantiomer by chiral HPLC. Both (-)-(R)- and (+)-(S)-2-phenylpropionic acid were present in plasma after the administration of either antipode, and further evidence of the chiral inversion of both enantiomers was provided by the presence of some 25% of the opposite enantiomer in the free 2-phenylpropionic acid and its glucuronide excreted in urine after administration of (-)-(R)- and (+)-(S)-2-phenylpropionic acid. The (+)-(S)-enantiomer underwent chiral inversion to the (-)-(R)-antipode when incubated with dog hepatocytes. These data suggests that both enantiomers of 2-phenylpropionic acid are substrates for canine hepatic acyl CoA ligase(s) and thus undergo chiral inversion, but that the CoA thioester of only (+)-(S)-2-phenylpropionic acid is a substrate for the glycine N-acyl transferase. These studies are presently being extended to the structure and species specificity of the reverse inversion and amino acid conjugation of profen NSAIDs.  相似文献   

19.
Resolution of (2RS,3RS)-2-[alpha-(2-methoxymethoxyphenoxy)phenylmethyl]morpholine, 11, with (+) mandelic acid led to the formation of (+)-(2S,3S)-2-[alpha-(2-methoxymethoxyphenoxy)phenyl methyl] morpholine (11a). Compound 11 was synthesized in seven steps from (2RS,3RS)-cinnamyl alcohol-2,3-epoxide (4), with an overall yield of 17%. Cleavage of the methoxymethyl group of the Fmoc derivative 12 with catalytic amounts of p-toluenesulfonic acid in methanol afforded (+)-(2S,3S)-2-(2-morpholin-2-yl-2-phenylmethoxy)phenol 2. The synthetic utility as well as the configuration of compound 2 has been demonstrated by converting (S,S)-2-(2-morpholin-2-yl-2-phenylmethoxy)phenol 2 to (2S,3S)-2-[alpha-(2-ethoxyphenoxy)phenylmethyl]morpholine (1) and (2S,3S)-2-(2-methoxyphenoxy) benzyl)morpholine (16), two potential norepinephrine reuptake inhibitors under clinical evaluation.  相似文献   

20.
(Z)-1,1-Dichloro-2-(4-benzyloxyphenyl)-2,3-bis(4-methoxyphenyl)cyclopropane ( 5 ), a potential antitumor agent designed to treat breast cancer, was prepared in three steps. A stereospecific palladium-catalyzed cross coupling reaction which provided the intermediate (Z)-triaryl alkene 4 was a crucial step in the synthesis. Makosza phase transfer reaction on 4 gave the enantiomeric (Z)-dichlorocyclopropane derivatives 5 which were resolved by semipreparative HPLC on a chiral stationary phase consisting of amylose tris-3,5-dimethylphenyl carbamate coated on silica gel. © 1994 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