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 475 毫秒
1.
The antipsoriatic 10-(3-chlorophenyl)-6,8,9,10-tetrahydrobenzo[b][1,8] naphthyridin-5(7H)-one, Sch 40120 , is chiral only because it lacks planarity and possesses a stereogenic axis. It comprises short-lived, interconverting atropisomeric enantiomers distinguished by the chlorine substitutent. The atropisomers form diastereomeric complexes with the shift reagent (R)-(−)-2,2,2-trifluoro-1-(9-anthryl)ethanol, which were detected by 1H NMR spectroscopy. Liquid chromatography on an ovomucoid chiral column isolated each enantiomer from the racemic mixture. Re-injections of the separated enantiomers onto the same column held constant at 10°C established that each enantiomer formed the other. Under identical chromatographic conditions, both stereoisomers independently recreated the racemic mixture. The calculated enantiomer half-life lasted 1.6 min at the physiological temperature of 37°C. Simulations of dynamic liquid chromatograms acquired with a chiral stationary phase indirectly yielded values of the half-lives. The chromatograms were modeled with the computer program SIMUL. Also determined were the rate constants for enantiomerization and the corresponding Gibbs free energies of activation, all at varying temperatures. At 37°C, the rate constant and activation energy respectively equaled 0.213 min−1 and 21.6 kcal mole−1. An Arrhenius plot was linear. The intractably brief life spans necessitated development of the racemic drug, rather than advancement of one enantiomer only. The pharmacological, biological, and chemical consequences of molecular asymmetry inherent to the drug were therefore nil. © 1996 Wiley-Liss, Inc.  相似文献   

2.
The addition reaction to N-methyl-(S)-alanine or N-methyl-(S)-phenylalanine N-car-boxyanhydride (NCA) of 3-methyl-5-substituted hydantoin (HDT) catalyzed by a tertiary amine was investigated as a model reaction for the propagation reaction of NCA according to the activated-NCA mechanism. Several activated HDTs having the (S)-configuration of the asymmetric carbon atom were found to react more rapidly than their activated enantiomers. This experimental result indicates that the enantiomer selection by terminal-unit control takes place in the propagation reaction according to the activated-NCA mechanism in which an activated NCA is added to a terminal acylated NCA ring of the growing chain. The enantiomer excess of the HDT recovered from the reaction mixture of N-methyl-(S)-phenylalanine NCA and racemic HDTs activated by a tertiary amine was determined. The extent of the enantiomer selection in the polymerization was found to be 3–10 times as large as that in the model reaction. From these results, it was concluded that the chirality of the penultimate unit, as well as that of the terminal NCA ring, plays an important role in determining the enantiomer selection in the NCA polymerization.  相似文献   

3.
《Chirality》2017,29(3-4):130-133
Differential scanning calorimetry and nuclear magnetic resonance were used to investigate thermal effects in the guest–host systems where homochiral metal–organic sorbent [Zn2(bdc)(S ‐lac)(dmf)] is considered as a host while 1‐phenylethanol enantiomers and their racemic mixture serve as guest molecules. A maximum energy gain from the guest–host interaction was observed in the system with the racemic mixture. The effect of host–guest recognition was revealed for the case of the host and guest having a similar type of chirality in the presence of antipode guest molecules.  相似文献   

4.
Stereoselective metabolism of propranolol side‐chain glucuronidation was studied for two recombinant human uridine diphosphate glucuronosyltransferases (UGTs), UGT1A9 and UGT2B7. The S‐ and R‐propranolol side‐chain glucuronides produced in the incubation mixtures were assayed simultaneously by RP‐HPLC with fluorescent detector. The excitation and emission wavelengths were set at 310 nm and 339 nm, respectively. UGT1A9 prefers catalyzing S‐enantiomer to R‐enantiomer and the intrinsic clearance (CLint) ratios of S‐enantiomer to R‐enantiomer are 3.8 times and 6.5times for racemic propranolol and individual enantiomers, respectively. UGT2B7, however, catalyzes slightly less S‐enantiomer than R‐enantiomer and the CLint ratio of S‐enantiomer to R‐enantiomer is 0.8 times. The high concentration of racemic propranolol (>0.57 mmol/l) and individual enantiomers (>0.69 mmol/l) exhibited substrate inhibition of glucuronidation for UGT2B7, but only the S‐enantiomer (>0.44 mmol/l) in racemic propranolol exhibited substrate inhibition for UGT1A9. The substrate inhibition constants (Ksi) were all similar (P > 0.05). Drug–drug interactions were also found between S‐ and R‐enantiomer glucuronidation metabolisms by UGT1A9 and UGT2B7. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

5.
《Chirality》2017,29(1):14-18
The enantiomer ratios of chiral volatile organic compounds in fruit distillates were determined by multidimensional gas chromatography using solid‐phase microextraction (SPME) as a sample treatment procedure. Linalool and its oxides, limonene, α‐terpineol, and nerolidol, were present at the highest concentration levels, while significantly lower amounts of β‐citronellol and lactones were found in the studied samples. However, almost all terpenoids mainly occur as a racemic or near‐racemic mixture; enantiomer distribution of some chiral organic compounds in fruit distillates correlated to a botanical origin. In particular, a significant enantiomeric excess of (R)‐linalool and (S)‐α‐terpineol was found only for pear brandy, and likewise the dominance (R)‐limonene and the second eluted enantiomer of nerolidol for Sorbus domestica and strawberry, respectively. The distribution of γ‐lactones stereoisomers was more nonspecific, with a general excess of the R‐enantiomer.  相似文献   

6.
N,N′-Dialkylaminoalkylcarbonyl (DAAC) and aminoalkylcarbonyl (AAC) prodrugs of phenolic drugs acetaminophen (APAP) and naltrexone (NTX) are reported. The effects of incorporation of a basic amine group into the promoiety of an acyl prodrug of a phenolic drug on its skin permeation properties are also presented. DAAC-APAP prodrugs were synthesized via a three-step procedure starting with haloalkylcarbonyl esters which were reacted with five different amines: dimethylamine, diethylamine, dipropylamine, morpholine, and piperidine. The spacing between the amino group and the carbonyl group of the acyl group was 1-3 CH2. After the hydrolysis of the ester, the carboxylic acid product was subsequently coupled with the parent drug via a dicyclohexyl carbodiimide (DCC) mediated coupling to yield the DAAC-APAP-HCl prodrugs in excellent yields. The AAC prodrugs were synthesized using commercially available Boc-protected amino acids using DCC or EDCI as coupling agents. The yields of the prodrugs synthesized using these two different methods have been compared. Half-lives (t1/2) of a few members of the DAAC and AAC series were measured in buffer (pH 6.0, 20 mM). The members evaluated in hydrolysis experiments exhibit a t1/2 range of 15-113 min. Among AAC-APAP prodrugs, the isopropyl group in valinate-APAP-HCl exerted a steric effect that increased the t1/2 value for this prodrug compared to alaninate-APAP-HCl or prolinate-APAP-HCl. The 2-morpholinylacetate-APAP prodrug was able to achieve twice the flux of APAP in in vitro diffusion cell experiments through hairless mouse skin.  相似文献   

7.
Using the Fmoc methodology, we report the chemical synthesis of surfactin and of four of its analogues, by stepwise solid-phase peptide synthesis (SPPS) on Sasrin resin. Formation of depsipeptide bond was performed with EDC (1-ethyl-3-(3-dimethylaminopropyl) carbodiimide. In developing our strategy, surfactin was used as a model and we synthesized both its racemic mixture and its R isoform. (R) 3-hydroxy fatty acid was obtained using Candida antarctica lipase from the racemic fatty acid, allowing a further identification of both R and S isoforms in the racemic mixture. Analogues were synthesized as racemic linear lipopeptides. Then, both enantiomers were separated and purified by adsorption chromatography on silicic acid, following cleavage from the resin. Linear R lipoheptapeptides were identified by TLC. They exhibit, in all cases, higher Rf values than those of the corresponding S isoforms. Cyclization was then performed independently for each enantiomer, using a HATU/DIEA coupling in solution. The yields were highly dependent on the position and on the nature of the modified amino acids.  相似文献   

8.
The 50‐residue snake venom protein L ‐omwaprin and its enantiomer D ‐omwaprin were prepared by total chemical synthesis. Radial diffusion assays were performed against Bacillus megaterium and Bacillus anthracis; both L ‐ and D ‐omwaprin showed antibacterial activity against B. megaterium. The native protein enantiomer, made of L ‐amino acids, failed to crystallize readily. However, when a racemic mixture containing equal amounts of L ‐ and D ‐omwaprin was used, diffraction quality crystals were obtained. The racemic protein sample crystallized in the centrosymmetric space group P21/c and its structure was determined at atomic resolution (1.33 Å) by a combination of Patterson and direct methods based on the strong scattering from the sulfur atoms in the eight cysteine residues per protein. Racemic crystallography once again proved to be a valuable method for obtaining crystals of recalcitrant proteins and for determining high‐resolution X‐ray structures by direct methods.  相似文献   

9.
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   

10.
Competition between the (+)- and (?) enantiomers of 2,2,2-trifluoro-1-(9-anthryl) ethanol as mobile phase additives was indicated by the chromatographic behavior of their system peaks. Two types of chiral stationary phases were used, one based on dinitrobenzoylphenylglycine and the other on dinitrobenzylphenylethylamine plus tartaric acid. The racemic mixture was used as the mobile phase additive and k′ of their system peaks was studied as a function of the mixture concentration in the mobile phase in both cases. A shift in k′ of the two system peaks was observed and considered as an indication that competition occurred. The areas of the two system peaks were also studied as a function of the concentration of the enantiomers in the samples, using two different compositions of the mobile phase. The dependency of system peaks' area on the sample composition indicated whether competition between the enantiomers occurred. One mobile phase contained 0.1 mM of the racemic mixture, where the area of the two retained system peaks behaved independently, i.e., only the peak corresponding to the enantiomer was affected by its presence in the sample. The other mobile phase contained 0.75 mM of the racemic mixture, and both peaks were affected by the injection of any one of the enantiomers. The interdependency of the system peaks' area on both the enantiomers indicated that their distribution in the chiral system was interrelated due to mutual interactions. A quantitative treatment of the interdependency and competition was excluded, due to the irreversible adsorption of the two enantiomers on the chiral stationary phase after using overloading concentrations. This irreversible adsorption was visualized by the appearance of two retained system peaks of the two residual enantiomers. These system peaks were detected only when the sample contained pure enantiomers due to competition between the enantiomer in the sample with the residual enantiomers in the stationary phase. © 1994 Wiley-Liss, Inc.  相似文献   

11.
Amlodipine, 3-ethyl 5-methyl-2-[(2-aminoethoxymethyl]-4-(2-chlorophenyl)-1,4-dihydro-6-methyl-3,5-pyridinedicarboxylate, is a chiral calcium antagonist, currently on the market and in therapeutic use as a racemate. The pharmacokinetic behaviour of R-(+)- and S-(−)-amlodipine after single enantiomer administration to healthy male human volunteers together with comparative administration of the racemic mixture of both enantiomers were studied. Plasma levels were studied as a function of time and assayed using an enantioselective chromatographic method (coupled chiral and achiral HPLC) with on-line solid-phase extraction and UV absorbance detection. The method was validated separately for the R-(+)- and S-(−)-enantiomer, respectively. Results of the study indicate that the pharmacokinetic behaviour of R-(+)- and S-(−)-amlodipine after single enantiomer administration is comparable to that of each enantiomer after administration of the racemate. No racemization occurs in vivo in human plasma after single enantiomer administration.  相似文献   

12.
Summary It was previously shown that nuclei of-sheets surrounded by unordered segments are formed in polypeptide chains built up with alternating hydrophobic and hydrophilic residues and containing both L- and D-enantiomers. It was also established that segments of residues having the same configuration tend to segregate in these nuclei when the starting composition of stereomonomers departs from the racemic mixture.Soft acidic hydrolysis of these polymers has been studied. Kinetic measurements show two pseudo first order rate constants, in agreement with the existence of two conformational species. The unordered part of the chains is hydrolyzed more rapidly, allowing the isolation of a-fraction enriched in one enantiomer. Thus, a plausible process of enrichment in enantiomer during prebiotic evolution has been described, which however does not explain the preference of one enantiomer over the other one.  相似文献   

13.
Zhang Y  Shi K  Wen J  Fan G  Chai Y  Hong Z 《Chirality》2012,24(3):239-244
Tetrahydroberberine (THB), a racemic mixture of (+)‐ and (?)‐enantiomer, is a biologically active ingredient isolated from a traditional Chinese herb Rhizoma corydalis (yanhusuo). A chiral high performance liquid chromatography method has been developed for the determination of THB enantiomers in rat plasma. The enantioseparation was carried out on a Chiral®‐AD column using methanol:ethanol (80:20, v/v) as the mobile phase at the flow rate 0.4 ml/min. The ultraviolet detection was set at 230 nm. The calibration curves were linear over the range of 0.01–2.5 μg/ml for (+)‐THB and 0.01‐5.0 μg/ml for (?)‐THB, respectively. The lower limit of quantification was 0.01 μg/ml for both (+)‐THB and (?)‐THB. The stereoselective pharmacokinetics of THB enantiomers in rats was studied after oral and intravenous administration at a dose of 50 and 10 mg/kg racemic THB (rac‐THB). The mean plasma levels of (?)‐THB were higher at almost all time points than those of (+)‐THB. (?)‐THB also exhibited greater Cmax, and AUC0–∞, smaller CL and Vd, than its antipode. The (?)/(+)‐enantiomer ratio of AUC0–∞ after oral and intravenous administration were 2.17 and 1.43, respectively. These results indicated substantial stereoselectivity in the pharmacokinetics of THB enantiomers in rats. Chirality, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

14.
A method has been developed for the determination of total celiprolol (sum of enantiomers) or the enantiomers (R)-celiprolol and (S)-celiprolol in plasma by high-performance liquid chromatography with UV and fluorescence detection. After extraction from alkalinized plasma with methyl-tert-butyl ether and back-extraction into 0.01 M HCl (for total celiprolol determination) or after evaporation of the organic phase and derivatisation with R(−)-1-(1-naphthyl)ethyl isocyanate (enantiomer determination), total celiprolol or its diastereomeric derivatives were chromatographed on a reversed-phase HPLC column with a mixture of acetonitrile and phosphate buffer pH 3.5 (+0.05% triethylamine). Acebutolol was used as internal standard. Linearity was obtained in the range of 5 to 2000 ng/ml for total and 2.5 to 500 ng/ml for enantiomer determination. Intra-day and inter-day variation was lower than 10%. The method can be applied for analysis of plasma samples obtained from patients treated with oral racemic celiprolol doses.  相似文献   

15.
Fluoxetine is used clinically as a racemic mixture of (+)‐(S) and (–)‐(R) enantiomers for the treatment of depression. CYP2D6 catalyzes the metabolism of both fluoxetine enantiomers. We aimed to evaluate whether exposure to gasoline results in CYP2D inhibition. Male Wistar rats exposed to filtered air (n = 36; control group) or to 600 ppm of gasoline (n = 36) in a nose‐only inhalation exposure chamber for 6 weeks (6 h/day, 5 days/week) received a single oral 10‐mg/kg dose of racemic fluoxetine. Fluoxetine enantiomers in plasma samples were analyzed by a validated analytical method using LC‐MS/MS. The separation of fluoxetine enantiomers was performed in a Chirobiotic V column using as the mobile phase a mixture of ethanol:ammonium acetate 15 mM. Higher plasma concentrations of the (+)‐(S)‐fluoxetine enantiomer were found in the control group (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.68). In animals exposed to gasoline, we observed an increase in AUC0‐∞ for both enantiomers, with a sharper increase seen for the (–)‐(R)‐fluoxetine enantiomer (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.07), resulting in a loss of enantioselectivity. Exposure to gasoline was found to result in the loss of enantioselectivity of fluoxetine, with the predominant reduction occurring in the clearance of the (–)‐(R)‐fluoxetine enantiomer (55% vs. 30%). Chirality 25:206–210, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

16.
An enantioselective HPLC bioassay has been developed relying on extraction of (R)- and (S)-atenolol from alkalinized plasma or serum (pH > 12) into dichloromethane containing 5% (v/v) 1-butanol followed by an achiral derivatization of the drug with phosgene leading to (R)- and (S)-oxazolidine-2-one derivatives. Under these conditions there was quantitative conversion of the acetamido group to the corresponding nitrile. These stable derivatives were separated on a (R,R)-diaminocylohexane-dinitrobenzoyl chiral stationary phase [(R,R)-DACH-DNB] using dichloromethane/methanol 98/2 as mobile phase. Determination limits of 0.5 ng for (R)- and 0.6 ng for (S)-atenolol could be achieved using fluorimetric detection. The assay was applied to a human pharmacokinetic study which was performed in a randomized cross-over, double-blind fashion in 12 healthy volunteers, administering single oral doses of 100 mg (R,S)-, 50 mg (R)-, and 50 mg (S)-atenolol AUC0–24 and Cmax values of (R)-atenolol were slightly but significant higher than those of (S)-atenolol. The R/S ratios were 1.09 for AUC(R)/AUC(S) and 1.03 for Cmax (R)/Cmax(S) (P < 0.01) respectively after administration of the racemic drug. However, there were no differences between AUC, Cmax, and t½ values of each enantiomer, whether they were administered as single enantiometers or in the form of its racemic mixture. © 1993 Wiley-Liss, Inc.  相似文献   

17.
Although it has been assumed that the effects of nonsteroidal antiinflammatory drugs (NSAIDs) are mainly the result of their action on local synthesis of prostaglandins, there is growing evidence to suggest that they may also exert a central analgesic action. Some authors have suggested that inhibition of prostaglandin synthesis in the brain could contribute to the analgesic action. The effect of dexketoprofen trometamol (tromethamine salt of the enantiomer (+)-S-ketoprofen) on prostaglandin synthesis was investigated in rat brain fragments and in cyclooxygenase preparations from rat brain microsomes. Effects of the (-)-R-enantiomer and the racemic mixture were also evaluated. Significant levels of prostaglandin F (PGF) were synthesized in rat brain fragments after 10 min of incubation at 37°C. Dexketoprofen was found to be a potent inhibitor of this PGF production in rat brain (IC50 = 6.2 nM), and it completely suppressed PGF production at 1 μM concentration. In addition, inhibition of PGF synthesis by dexketoprofen was highly stereoselective since the enantiomer (-)-R-ketoprofen was significantly less potent (IC50 = 294 nM); with this enantiomer, even at high concentrations such as 1 μM, less than 60% inhibition was achieved. These results correlated with those obtained in the study of racemic ketoprofen and its enantiomers on cyclooxygenase activity of rat brain microsomes, where dexketoprofen also inhibited enzymatic activity stereoselectively. IC50 values obtained for dexketoprofen, (-)-R-ketoprofen, and rac-ketoprofen were 3.5 μM, 45.3 μM, and 5.8 μM, respectively. The above results could be related to the potent analgesic effect of dexketoprofen observed in vivo, which was also stereoselective. Taken together, these findings suggest that prostaglandin synthesis inhibition in rat brain by dexketoprofen could be associated, at least in part, with the analgesic effect of this NSAID. Chirality 9:281–285, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

18.
Elevation of endogenous GABA by the racemic mixture of gamma vinyl-GABA (GVG, Vigabatrin) decreases extracellular nucleus accumbens (NAc) dopamine (DA) levels and diminishes the response to many drugs of abuse known to elevate DA in the mesocorticolimbic system. We investigated the effects of the individual enantiomers (S(+)-GVG, R(-)-GVG) on cocaine-induced NAc DA in rodents as well as the effects of nicotine-induced increases in primates. In a series of microdialysis experiments in freely moving animals, S(+)-GVG (150 mg/kg), R(-)-GVG (150 mg/kg) or racemic (R, S) GVG (300 mg/kg) was administered 2.5 hours prior to cocaine (20 mg/kg) administration. When compared with cocaine alone, the R(-) enantiomer did not significantly inhibit cocaine induced NAc DA release. S(+)-GVG, at half the dose of the racemic mixture (150 mg/kg), inhibited cocaine-induced DA elevation by 40%, while the racemic mixture (300 mg/kg) inhibited cocaine-induced DA release by 31%. In addition, our PET studies in primates demonstrated that S(+)-GVG completely inhibits nicotine-induced increases in the corpus striatum, again at half the dose of the racemic mixture. The R(-) enantiomer was ineffective. Although the S(+) enantiomer has been well established as the active compound in the treatment of epilepsy, the efficacy of this enantiomer with regard to mesolimbic DA inhibition generates a complex series of clinical and neurochemical issues. Further investigations will determine the locus of action and physiologic properties of each enantiomer.  相似文献   

19.
The occurrence of pharmaceuticals in the environment represents a challenge of emerging concern. Many pharmaceuticals are chiral compounds; however, few studies have examined the relative toxicity of pharmaceutical enantiomers to wildlife. Further, our understanding of stereospecific pharmacokinetics remains largely informed by research on humans and a few well‐studied laboratory test animals, and not by studies conducted with environmentally relevant species, including fish. The objective of this study was to investigate whether rainbow trout display stereospecific in vitro metabolism of three common chiral pharmaceuticals. Metabolism by trout liver S9 fractions was evaluated using a substrate depletion approach, which provides an estimate of intrinsic hepatic clearance (CLIN VITRO,INT). No biotransformation was observed for rac‐, R‐, or S‐fluoxetine. Ibuprofen, including both enantiomers and the racemic mixture, appeared to undergo slow metabolism, but the resulting substrate depletion curves did not differ significantly from those of inactive controls. Contrary to relative clearance rates in humans, S(?)‐propranolol was more rapidly cleared than the R(+)‐ enantiomer. This work demonstrates that relative clearance rates and the effects of racemic mixtures in trout could not have been predicted based on human data. Additional research describing species differences and exploring tools for species extrapolation in biomedical and environmental studies is needed. Chirality 25:763–767, 2013, © 2013 Wiley Periodicals, Inc.  相似文献   

20.
The five‐steps synthesis of a hemicryptophane cage combining a benzene‐1,3,5‐tricarboxamide unit and a cyclotriveratrylene (CTV) moiety is described. Chiral high‐performance liquid chromatography (HPLC) was used to resolve the racemic mixture. The absolute configuration of the isolated enantiomers was assigned by comparison of the experimental electronic circular dichroism (ECD) spectra with the calculated ones. X‐ray molecular structures reveal that the capped benzene‐1,3,5‐tricarboxamide unit adopts a structurally chiral conformation in solid state: the chirality of CTV moiety controls the Λ or Δ orientation of the three amides.  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