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.
Clobazam, a 1,5‐benzodiazepin‐2,4‐dione, is a chiral molecule because its ground state conformation features a nonplanar seven‐membered ring lacking reflection symmetry elements. The two conformational enantiomers of clobazam interconvert at room temperature by a simple ring‐flipping process. Variable temperature HPLC on the Pirkle type (R)‐N‐(3,5‐dinitronenzoyl)phenylglycine and (R,R)‐Whelk‐O1 chiral stationary phases (CSPs) allowed us to separate for the first time the conformational enantiomers of clobazam and to observe peak coalescence‐decoalescence phenomena due to concomitant separation and interconversion processes occurring on the same time scale. Clobazam showed temperature dependent dynamic high‐performance liquid chromatography (HPLC) profiles with interconversion plateaus on the two CSPs indicative of on‐column enantiomer interconversion. (enantiomerization) in the column temperature range between Tcol = 10°C and Tcol = 30°C, whereas on‐column interconversion was absent at temperature close to or lower than Tcol = 5°C. Computer simulation of exchange‐deformed HPLC profiles using a program based on the stochastic model yielded the apparent rate constants for the on‐column enantiomerization and the corresponding free energy activation barriers. At Tcol = 20°C the averaged enantiomerization barriers, ΔG?, for clobazam were found in the range 21.08–21.53 kcal mol?1 on the two CSPs. The experimental dynamic chromatograms and the corresponding interconversion barriers reported in this article are consistent with the literature data measured by DNMR at higher temperatures and in different solvents. Chirality 28:17–21, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

2.
The enantiomerization and enantioselective bioaccumulation of metalaxyl by a single dose of exposure to Tenebrio molitor larvae under laboratory condition were studied by high‐performance liquid chromatography tandem mass spectroscopy (HPLC‐MS/MS) based on a ChiralcelOD‐3R [cellulosetris‐tris‐(3, 5‐dichlorophenyl‐carbamate)] column. Exposure of enantiopure R‐metalaxyl and S‐metalaxyl in Tenebrio molitor larvae exhibited significant enantiomerization, with formation of the R enantiomers from the S enantiomers, and vice versa, which might be attributed to the chiral pesticide catalyzed by a certain enzyme in Tenebrio molitor larvae. Enantiomerization was not observed in wheat bran during the period of 21 d. In addition, bioaccumulation of rac‐metalaxyl in Tenebrio molitor larvae was enantioselective with a preferential accumulation of S‐metalaxyl. These results showed that enantioselectivity was caused not only by actual degradation and metabolism but also by enantiomerization, which was an important process in the environmental fate and behavior of metalaxyl enantiomers. Chirality 26:88–94, 2014. © 2013 Wiley Periodicals, Inc.  相似文献   

3.
pKa1 values of 3-methoxy-N-desmethyldiazepam in acetonitrile and methanol containing various acid concentrations were determined by spectrophotometry to be 3.5 and 1.3, respectively. Temperature-dependent racemization of enantiomeric 3-methoxy-N-desmethyldiazepam in methanol containing 0.5 M H2SO4 was studied by circular dichroism spectropolorimetry and the racemization reactions were found to follow apparent first-order kinetics. Thermodynamic parameters of the racemization reaction were found to be: Eact = 18.8 kcal/mol, and at 25°C: ΔH? = 18.3 kcal/mol, ΔS? = ?14.8 entropy unit, and ΔG? = 22.7 kcal/mol, respectively. The racemization had an isotope effect (kH/kD) of 1.6 at 42°C. Based on the results of this report and those of earlier reports by other investigators, a nucleophilically solvated C3 carbocation intermediate resulting from either a P (plus) or an M (minus) conformation is proposed to be an intermediate and responsible for the stereoselective nucleophilic substitution and the subsequent racemization of 3-methoxy-N-desmethyldiazepam enantiomers. © 1993 Wiley-Liss, Inc.  相似文献   

4.
A series of chiral tris-(1,10)-phenanthroline iron(II) complexes have been resolved by HPLC on chiral stationary phases based on either cellulose tris-(3,5-dimethylphenylcarbamate) or teicoplanin. At sub ambient temperatures, baseline separation of the enantiomers was observed for five different iron(II) complexes featuring substituted phenanthroline ligands. Dynamic HPLC profiles were observed near or above room temperature, indicating on-column Delta/Lambda enantiomerization. Rate constants for the Delta/Lambda interconversion in free solution and during chromatography were obtained by thermal racemization experiments and by computer simulation of the HPLC dynamic plots, respectively.  相似文献   

5.
An on-column HPLC procedure using a chiral stationary phase (CSP) was developed for the determination of rate constants and free energy barriers of enantiomerization of (+/-)IDRA21. Subsequently, the HPLC method was applied for investigation of two structurally related chiral compounds. The individual enantiomers of the studied compounds were isolated in parallel by preparative HPLC and rate constants and free energy barriers of enantiomerization were determined in different solvents. The on-column enantiomerization data revealed that CSP induces different rate constants for the two enantiomers. The results generated off-line were used to determine the influence of solvents on the racemization of (+) and (-) IDRA21 and to gain further insight into the enantiomerization mechanism of chiral 3,4-dihydro-1,2,4-benzothiadiazine 1,1-dioxide type compounds.  相似文献   

6.
The reversible enantiomerization of axially chiral 2'-dodecyloxy-6-nitrobiphenyl-2-carboxylic acid was studied in the presence of a brush type chiral stationary phase based on O-(tert-butylcarbamoyl) quinine as chiral selector unit by stopped-flow high-performance liquid chromatography (sfHPLC) and capillary electrochromatography (sfCEC). After initial separation of the enantiomers in the first section of the column, the flow was stopped and the resolved species allowed to enantiomerize on-column. From this conversion, which could be determined from the enantiomeric ratios at different enantiomerization times, kinetic rate constants were calculated. By sfHPLC at a constant temperature of 15 degrees C, kinetic rate constants in the presence of the CSP were found to be 4.1 x 10(-5) s(-1) and 2.2 x 10(-5) s(-1) for the (-) and (+)-enantiomers, respectively, corresponding to half-lives of 279 and 530 min. Thus, apparent activation energies of enantiomerization were calculated to be 93.0 and 94.6 kJ mol(-1) for the (-) and (+)-enantiomers. On the macroscopic level, the apparent difference of rotational energy barriers and kinetic rate constants for both enantiomers is reflected as deracemization. For example, starting from a racemic mixture, an enantiomeric excess (ee) of 14% was seen in the stopped-flow HPLC experiment described after an enantiomerization time of 220 min at 15 degrees C, and a maximal ee of 17% can be approximated after infinite enantiomerization time. There is good agreement between HPLC and CEC results as well as their experimental errors, confirming that the new sfCEC technique may be a valuable and convenient tool to study interconversion processes.  相似文献   

7.
The enantiomers of rac-2,2′-diiodobiphenyl were separated by liquid chromatography on microcrystalline triacetylcellulose. The conformational lability, a large separation factor α, and a suitable capacity factor k′(+) of this biphenyl allowed us to convert the racemate into 90% of enantiomerically pure (-)-2,2′-diiodobiphenyl and 10% of pure (+)-2,2′-diiodobiphenyl, respectively, by a series of in situ racemization-elution cycles. The much better retained (+)-enantiomer was racemized on the chromatographic column at 50°C after the less retained (-)-enantiomer has already been eluted at 8°C. © 1995 Wiley-Liss, Inc.  相似文献   

8.
Herein we present design, synthesis, chiral HPLC resolution, and kinetics of racemization of axially chiral Ni(II) complexes of glycine and di‐(benzyl)glycine Schiff bases. We found that while the ortho‐fluoro derivatives are configurationally unstable, the pure enantiomers of corresponding axially chiral ortho‐chloro‐containing complexes can be isolated by preparative HPLC and show exceptional configurational stability (t1/2 from 4 to 216 centuries) at ambient conditions. Synthetic implications of this discovery for the development of new generation of axially chiral auxiliaries, useful for general asymmetric synthesis of α‐amino acids, are discussed.  相似文献   

9.
The configurational stability of (+)- and (−)-diethylpropion [(+)- and (−)-2-(diethyl)-1-phenyl-1-propanone or (+)- and (−)-DEP ] was investigated systematically from chemical, pharmaceutical, and pharmacological aspects. The enantiomeric ratio was monitored directly with a recently developed stability-indicating enantioselective HPLC method. In aqueous solutions, the rate of racemization increased non-linearly with increasing pH and with increasing phosphate buffer concentration. The racemization rate showed a positive slope with increasing temperature and decreasing ionic strength. The racemization rates of (+)- and (−)-DEP in the presence of cyclodextrins (CDs) did not differ significantly. CDs that were added to (+)- and (−)-DEP in a molar ratio 5:1 showed the following effects after dissolution in 10 mM phosphate buffer (final pH 6.7): sulfobutyl ether-β-CD (SBE-β-CD) and methylated-β-CD (Me-β-CD) retarded racemization; whereas hydroxypropyl-β-CD (HP-β-CD), acetyl-γ-CD (Ac-γ-CD), acetyl-β-CD (Ac-β-CD), γ-CD, and β-CD showed a weak destabilising effect. In contrast to the described CDs, α-CD distinctly accelerated the rate of racemization. The configurational stability of (+)- and (−)-DEP was also studied under physiological conditions. The half-life of racemization in heparinised human plasma was for both enantiomers determined to be approximately 23–25 min. In phosphate buffer (10 mM, pH 7.4), rac-DEP showed a high, but unselective affinity towards human α1-acid glycoprotein (orosomucoid) immobilised on silica (Chiral AGP). The rate of racemization of the free base of (−)-DEP dissolved in organic solutions generally increases with the polarity of the solvating agent. Chirality 10:307–315, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
Enantiomers of chiral molecules commonly exhibit differing pharmacokinetics and toxicities, which can introduce significant uncertainty when evaluating biological and environmental fates and potential risks to humans and the environment. However, racemization (the irreversible transformation of one enantiomer into the racemic mixture) and enantiomerization (the reversible conversion of one enantiomer into the other) are poorly understood. To better understand these processes, we investigated the chiral fungicide, triadimefon, which undergoes racemization in soils, water, and organic solvents. Nuclear magnetic resonance (NMR) and gas chromatography / mass spectrometry (GC/MS) techniques were used to measure the rates of enantiomerization and racemization, deuterium isotope effects, and activation energies for triadimefon in H2O and D2O. From these results we were able to determine that: 1) the alpha‐carbonyl carbon of triadimefon is the reaction site; 2) cleavage of the C‐H (C‐D) bond is the rate‐determining step; 3) the reaction is base‐catalyzed; and 4) the reaction likely involves a symmetrical intermediate. The B3LYP/6–311 + G** level of theory was used to compute optimized geometries, harmonic vibrational frequencies, nature population analysis, and intrinsic reaction coordinates for triadimefon in water and three racemization pathways were hypothesized. This work provides an initial step in developing predictive, structure‐based models that are needed to identify compounds of concern that may undergo racemization. Chirality 28:633–641, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

11.
It has been found that the interconversion of enantiomers on a chromatographic column during the separation process can be studied by the first-order kinetic equations derived both for reversible and irreversible reactions in a stationary system if the extent of interconversion is not too high. The equation derived for irreversible reactions gives, however, results also for higher degrees of enantiomerization while that derived for reversible interconversion failed. The irreversible equation was used to determine the enantiomerization barrier of N-(p-methoxybenzyl)-l,3,2-benzodithiazol-l-oxide enantiomers by supercritical fluid chromatography. The racemate of N-(p-methoxybenzyl)-l,3,2-benzodithiazol-l-oxide was separated by supercritical fluid chromatography on the (R,R)-Whelk-Ol column with supercritical carbon dioxide containing 20% methanol as a mobile phase. Peak areas of enantiomers prior to and after the separation used for the calculation of the enantiomerization barrier were determined by computer-assisted peak deconvolution of peak clusters registered on chromatograms using commercial software.  相似文献   

12.
From achiral imidazolinium salts, chiral transition metal complexes containing an N-heterocyclic carbene (NHC) ligand were prepared (metal = palladium, copper, silver, gold, rhodium). Axial chirality in these complexes results from the formation of the metal-carbene bond leading to the restriction of rotation of dissymmetric N-aryl substituents about the C–N bond. When these complexes exhibited a sufficient configurational stability, a resolution by chiral high-performance liquid chromatography (HPLC) on preparative scale enabled isolation of enantiomers with excellent enantiopurities (>99% ee) and good yields. A study of the enantiomerization barriers revealed the effect of the backbone nature as well as the type of transition metal on its values. Nevertheless, the evaluation of palladium-based complexes in asymmetric intramolecular α-arylation of amides demonstrated that the ability to induce an enantioselectivity cannot be correlated to the configurational stability of the precatalysts.  相似文献   

13.
The low configurational stability of two model compounds, the anorectic drug amfepramone and one of its basic metabolites, rac-cathinone, was examined for its mechanism and kinetics. Assuming a common intermediate for chiral inversion and deuteration, the rates of racemization were determined by the indirect method of proton–deuterium substitution monitored by 1H-NMR. The rate of racemization increased linearily with increasing pD. At a pD of 7.4, the rate of racemization of both aminoketones was markedly dependent on buffer concentration, indicating a mechanism of general-base catalysis. The reactions were some five to six times faster in phosphate buffers than in hydroxylamine buffers of identical molarity. Amfepramone racemized about five times faster than the primary amine cathinone. One implication of these findings is that amfepramone and cathinone may be subject in vivo to chiral inversion catalyzed by numerous endogenous bases. As a result, it will be misleading to extrapolate rates of inversion expected in plasma from those measured in buffer solutions. © 1995 Wiley-Liss, Inc.  相似文献   

14.
Chiral 5-arylbenzothiadiazine derivatives have recently attracted particular attention because they exhibit an interesting pharmacological activity as AMPA receptor (AMPAr) positive modulators. However, investigations on their configurational stability suggest a rapid enantiomerization in physiological conditions. In order to enhance configurational stability, preserving AMPAr activity, we have designed the novel compound (R,S)-7-chloro-9-(furan-3-yl)-2,3,3a,4-tetrahydro-1H-benzo[e]pyrrolo[2,1-c][1,2,4]thiadiazine 5,5-dioxide bearing a pyrrolo moiety coupled with the 5-(furan-3-yl) substituent on benzothiadiazine core. A stereoselective synthesis was projected to obtain single enantiomer of the latter compound. Absolute configuration was assigned by X-ray crystal structure. Patch clamp experiments evaluating the activity of single enantiomers as AMPAr positive allosteric modulator showed that R stereoisomer is the active component. Molecular modeling studies were performed to explain biological results. An on-column stopped-flow bidimensional recycling HPLC procedure was applied to obtain on a large scale the active enantiomer with enantiomeric enrichment starting from the racemic mixture of the compound.  相似文献   

15.
Benalaxyl (BX), methyl‐N‐phenylacetyl‐N‐2,6‐xylyl alaninate, is a potent acylanilide fungicide and consist of a pair of enantiomers. The stereoselective metabolism of BX was investigated in rat and rabbit microsomes in vitro. The degradation kinetics and the enantiomer fraction (EF) were determined using normal high‐performance liquid chromatography with diode array detection and a cellulose‐tris‐(3,5‐dimethylphenylcarbamate)‐based chiral stationary phase (CDMPC‐CSP). The t1/2 of (?)‐R‐BX and (+)‐S‐BX in rat liver microsomes were 22.35 and 10.66 min of rac‐BX and 5.42 and 4.03 of BX enantiomers. However, the t1/2 of (?)‐R‐BX and (+)‐S‐BX in rabbit liver microsomes were 11.75 and 15.26 min of rac‐BX and 5.66 and 9.63 of BX enantiomers. The consequence was consistent with the stereoselective toxicokinetics of BX in vitro. There was no chiral inversion from the (?)‐R‐BX to (+)‐S‐BX or inversion from (+)‐S‐BX to (?)‐R‐BX in both rabbit and rat microsomes. These results suggested metabolism of BX enantiomers was stereoselective in rat and rabbit liver microsomes. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

16.
Diaziridines are very interesting representatives of organic compounds containing stereogenic nitrogen atoms. In particular, 1,2-dialkyldiaziridines show extraordinarily high stereointegrity. The lone electron pairs of the nitrogen atoms are in trans configuration, avoiding a four-electron repulsive interaction. Furthermore, the trans configuration of the substituents at the nitrogen atoms is energetically favored because of reduced steric interactions. Therefore only two stereoisomers (enantiomers) are observed. At elevated temperatures the enantiomers are interconverting because of the limited stereointegrity of the chirotopic nitrogen atoms. The enantiomerization rate constants and the activation parameters of interconversion are of great interest. Here, we investigated the stereodynamics of a set of small 1,2-dialkyldiaziridines bearing short substituents (Me, Et, iPr, tBu), using enantioselective dynamic gas chromatography (DGC). Separation of enantiomers of all compounds, including the highly volatile 1,2-dimethyldiaziridine, was achieved using heptakis(2,3-di-O-ethyl-6-O-tert-butyldimethylsilyl)-β-cyclodextrin in 50% PS086 (w/w) as chiral stationary phase in fused silica capillaries with a length of up to 50 m. Measurements at variable temperatures were performed and reaction rate constants were determined using the unified equation of chromatography implemented in the software DCXplorer. The activation barriers at room temperature for 1-(tert-butyl)-2-ethyldiaziridine, ΔG298K = 123.8 kJ mol–1H = 115.5 ± 2.9 kJ mol–1, ΔS = –28 ± 1 J mol–1 K–1), and 1-ethyl-2-isopropyldiaziridine, ΔG298K = 124.2 kJ mol–1H = 113.1 ± 2.4 kJ mol–1, ΔS = –37 ± 2 J mol–1 K–1), were determined, representing some of the highest values observed for nitrogen inversion in diaziridines. Chirality 00:000–000, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

17.
The residual enantiomers of three tris‐(3‐indolyl)‐phosphane oxides bearing different alkyl groups (methyl, ethyl or i‐propyl) in position 2 of the indole rings constituting the blades were separated on the immobilized type Chiralpak IC column in polar organic and reversed‐phase modes. The good enantioselectivity and versatility of the IC CSP allowed easy isolation of the enantiomerically highly enriched samples suitable for configurational stability studies. The enantiomerization barriers of residual phosphane oxides were evaluated both by off‐column techniques (CD signal and enantiomeric purity decay kinetics) and by dynamic enantioselective high‐performance liquid chromatography (HPLC). Chirality 27:888–899, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
Direct enantiomer separation of hypericin, pseudohypericin, and protohypericin was accomplished by high‐performance liquid chromatography (HPLC) using immobilized polysaccharide‐type chiral stationary phases (CSPs). Enantioselectivities up to 1.30 were obtained in the polar‐organic elution mode whereby for hypericin and pseudohypericin Chiralpak IC [chiral selector being cellulose tris(3,5‐dichlorophenylcarbamate)] and for protohypericin Chiralpak IA (chiral selector being the 3,5‐dimethylphenylcarbamate of amylose) gave favorable results. Enantiomers were distinguished by on‐line electronic circular dichroism detection. Optimized enantioselective chromatographic conditions were the basis for determining stereodynamic parameters of the enantiomer interconversion process of hypericin and pseudohypericin. Rate constants delivered by computational simulation of dynamic HPLC elution profiles (stochastic model, consideration of peak tailing) were used to calculate averaged enantiomerization barriers (ΔG) of 97.6–99.6 kJ/mol for both compounds (investigated temperature range 25–45°C). Complementary variable temperature off‐column (i.e., in solution) racemization experiments delivered ΔG = 97.1–98.0 kJ/mol (27–45°C) for hypericin and ΔG = 98.9–101.4 kJ/mol (25–55°C) for pseudohypericin. An activation enthalpy of ΔH# = 86.0 kJ/mol and an activation entropy of ΔS# = ?37.7 J/(K mol) were calculated from hypericin racemization kinetics in solution, whereas for pseudohypericin these figures amounted to 74.1 kJ/mol and ?82.6 J/(K mol), respectively. Although the natural phenanthroperylene quinone pigments hypericin and pseudohypericin as well as their biological precursor protohypericin are chiral and can be separated by enantioselective HPLC low enantiomerization barriers seem to prevent the occurrence of an excess of one enantiomer under typical physiological conditions—at least as long as stereoselective intermolecular interactions with other chiral entities are absent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
Two different enantioselective chiral chromatographic methods were developed and validated to investigate the disposition of the β1-receptor antagonist atenolol in blood and in brain extracellular fluid of rats (tissue dialysates). System A for the plasma samples was a one-column chromatographic system with a Chiral CBH column with an aqueous buffer as mobile phase into which cellobiose was added for selective regulation of the retention of the internal standard, (S)-metoprolol. The plasma samples were analysed after a simple extraction procedure. The limit of quantitation was 0.2 μg/ml for the atenolol enantiomers. The repeatability of the medium concentration quality control plasma sample (6.0 μg rac-atenolol/ml) was 11–18% for the enantiomers. The dynamic linear range of the plasma samples was 0.5–20 μg/ml. For system B, since atenolol is an extremely hydrophilic drug, the tissue dialysate sample required a much more sensitive system as compared to the plasma samples. A coupled column system was used for peak compression of the enantiomers in the eluate after the separation on the Chiral CBH column, hence increasing the detection sensitivity. The limit of quantification was 0.045 μg/ml for the atenolol enantiomers in artificial CSF. The repeatability of the medium concentration quality control samples (0.1 and 4.0 μg rac-atenolol/ml in artificial CSF and Hepes Ringer, respectively) was 2.8–9.3% for the two enantiomers. The dynamic linear range of the brain samples was 0.05–1.0 and 0.5–20 μg/ml in artificial CSF and Hepes Ringer, respectively. Chirality 9:329–334, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

20.
Two racemic tetrathiafulvalene‐[2.2]paracyclophane electron donors EDT‐TTF‐[2.2]paracyclophane 1 and (COOMe)2‐TTF‐[2.2]paracyclophane 2 have been synthesized via the phosphite mediated cross coupling strategy. Chiral HPLC allowed the optical resolution of the (RP) and (SP) enantiomers for both compounds. Solid‐state structures of (RP)‐ 1 and (rac)‐ 2 have been determined by single crystal X‐ray analysis. Intermolecular π‐π and S???S interactions are disclosed in the packing. Single crystal X‐ray analysis of (RP)‐ 1 combined with experimental and theoretical circular dichroism spectra allowed the assignment of the absolute configuration of the enantiomers of 1 and 2 .  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