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 15 毫秒
1.
1. Cell-free extracts of the marine bacterium Beneckea natriegens, derived by sonication, were separated into particulate and supernatant fractions by centrifugation at 150 000 × g.2. NADH, succinate, d(?)- and l(+)-lactate oxidase and dehydrogenase activities were located in the particles, with 2- to 3-fold increases in specific activity over the cell free extract. The d(?)- and l(+)-lactate dehydrogenases were NAD+ and NADP+ independent. Ascorbate-N,N,N′,N′-tetramethylphenylenediamine (TMPD) oxidase was also present in the particulate fraction; it was 7–12 times more active than the physiological substrate oxidases.3. Ascorbate-TMPD oxidase was completely inhibited by 10 μM cyanide. Succinate, NADH, d(?)-lactate and l(+)-lactate oxidases were inhibited in a biphasic manner, with 10 μM cyanide causing only 10–50 % inhibition; further inhibition required more than 0.5 mM cyanide, and 10 mM cyanide caused over 90 % inhibition. Low sulphide (5 μM) and azide (2 mM) concentrations also totally inhibited ascorbate-TMPD oxidase, but only partially inhibited the other oxidases. High concentrations of sulphide but not azide caused a second phase inhibition of NADH, succinate, d(?)-lactate and l(+)-lactate oxidases.4. Low oxidase activities of the physiological substrates, obtained by using non-saturating substrate concentrations, were more inhibited by 10 μM cyanide and 2 mM azide than high oxidase rates, yet ascorbate-TMPD oxidase was completely inhibited by 10 μM cyanide over a wide range of rates of oxidation.5. These results indicate terminal branching of the respiratory system. Ascorbate-TMPD is oxidised by one pathway only, whilst NADH, succinate, d(?)-lactate and l(+)-lactate are oxidised via both pathways. Respiration of the latter substrates occurs preferentially by the pathway associated with ascorbate-TMPD oxidase and which is sensitive to low concentrations of cyanide, azide and sulphide.6. The apparent Km for O2 for each of the two pathways was detected using ascorbate-TMPD and NADH or succinate plus 10 μM cyanide respectively. The former pathway had an apparent Km of 8–17 (average 10.6) μM and the latter 2.2–4.0 (average 3.0) μM O2.  相似文献   

2.
Five-month-old Datura innoxia plants were fed via the roots with either d(+)-hygrine-[2′-14C] or l(?)-hygrine-[2′-14C]. After 7 days the root alkaloids 3α,6β-ditigloyloxytropane, 3α,6β-ditigloyloxytropan-7β-ol, hyoscine, hyoscyamine and cuscohygrine were isolated from both groups of plants. d(+) but not l(?)-hygrine acts as a precursor for the tropane alkaloids whereas both enantiomers appeared to serve equally well in the biosynthesis of cuscohygrine.  相似文献   

3.
Huang L  Lu D  Zhang P  Diao J  Zhou Z 《Chirality》2012,24(8):610-614
Enantioselectivity in ecotoxicity of chiral pesticides in the aquatic environment has been a subject of growing interest. In this study, the toxicological impacts of hexaconazole enantiomers were investigated with freshwater algae Scenedesmus obliquus. After 96 h of exposure, the EC50 values for rac‐hexaconazole, (+)‐hexaconazole, and (?)‐hexaconazole were 0.178, 0.355, and 0.065 mg l?1, respectively. Therefore, the acute toxicities of hexaconazole enantiomers were enantioselective. In addition, the different toxic effects were evaluated when S. obliquus were exposed to 0.2, 0.5, and 1.0 mg l?1 of rac‐hexaconazole, (+)‐hexaconazole, and (?)‐hexaconazole during 96 h, respectively. The chlorophyll a and chlorophyll b contents of S. obliquus treated by (?)‐hexaconazole were lower than those exposed to (+)‐hexaconazole, whereas the malondialdehyde contents of S. obliquus treated by (?)‐form were higher than those exposed to (+)‐form at higher concentrations. In general, catalase activities were significantly upregulated by exposure to (?)‐enantiomer than (+)‐enantiomer at all three concentrations. However, superoxide dismutase activities exposed to (?)‐hexaconazole were lower than that exposed to (+)‐hexaconazole at 0.2 mg l?1 and 0.5 mg l?1. On the basis of these data, the acute toxicity and toxic effects of hexaconazole against S. obliquus were enantioselective, and such enantiomeric differences must be taken into consideration in pesticide risk assessment. Chirality 24:610–614, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

4.
The purposes of this work were (1) to develop a high performance liquid chromatographic (HPLC) assay for the enantiomers of thalidomide in blood, (2) to study their inversion and degradation in human blood, and (3) to study the pharmacokinetics of (+)-(R)- and (?)-(S)-thalidomide after oral administration of the separate enantiomers or of the racemate to healthy male volunteers. The enantiomers of thalidomide were determined by direct resolution on a tribenzoyl cellulose column. Mean rate constants of chiral inversion of (+)-(R)-thalidomide and (?)-(S)-thalidomide in blood at 37°C were 0.30 and 0.31 h?1, respectively. Rate constants of degradation were 0.17 and 0.18 h?1. There was rapid interconversion in vivo in humans, the (+)-(R)-enantiomer predominating at equilibrium. The pharmacokinetics of (+)-(R)- and (?)-(S)-thalidomide could be characterized by means of two one-compartment models connected by rate constants for chiral inversion. Mean rate constants for in vivo inversion were 0.17 h?1 (R to S) and 0.12 h?1 (S to R) and for elimination 0.079 h?1 (R) and 0.24 h?1 (S), i.e., a considerably faster rate of elimination of the (?)-(S)-enantiomer. Putative differences in therapeutic or adverse effects between (+)-(R)- and (?)-(S)-thalidomide would to a large extent be abolished by rapid interconversion in vivo. © 1995 Wiley-Liss, Inc.  相似文献   

5.
Tetrahydropalmatine (THP) is one of the active alkaloid ingredients of Rhizoma Corydalis. THP has a chiral center, and the stereoselective pharmacokinetics and tissue distribution have been reported. The aim of the present article is to study the stereoselective protein binding of THP using equilibrium dialysis followed by HPLC‐UV analysis. The results showed that THP stereoselectively binds to human serum albumin (HSA), α1‐acid glycoprotein (AGP), and proteins in human plasma. The fraction binding of (+)‐THP was significantly higher than that of (?)‐THP, whereas such stereoselectivity was not found in rat plasma. The affinity of HSA and AGP to (+)‐THP, expressed as nKA, were 9.0 × 103 M?1 and 2.34 × 105 M?1, respectively, which were notablely higher than to (?)‐THP, with the nKA of 3.4 × 103 M?1 and 1.44 × 105 M?1, respectively. The binding site of HSA for (?)‐THP was Site I, whereas for (+)‐THP was both Site I and Site II. The F1/S variants of AGP were proved to be the key variants (?)‐ and (+)‐THP binding to both. Finally, the AGP binding drugs, such as mifepristone, were demonstrated to reduce the fraction binding of (?)‐ and (+)‐THP with pure AGP (1 mg/ml) but did not affect the fraction binding of both (?)‐ and (+)‐THP with proteins in human plasma. It can be concluded that protein binding of THP is species dependent and stereoselective, both HSA and AGP contribute to the stereoselective binding to THP enatiomers, and AGP binding drugs may not cause the drug–drug interaction on THP in healthy human plasma. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
(R,S)-trans-8-Hydroxy-2-[N-n-propyl-N-(3′-iodo-2′-propenyl)amino]tetralin 7 , a new radioiodinated ligand based on 8-OH-DPAT, was reported as a potential ligand for 5-HT1A receptors. The optically active (+)-(R)- and (?)-(S)- 7 were prepared to investigate the stereoselectivity of (R,S)- 7 . Racemic intermediate 8-methoxy-2-N-n-propyltetralin was reacted with the acyl chloride of (?)-(R)-O-methylmandelic acid to form a mixture of (S,R)- and (R,R)-diastereoisomers, which were separated by flash column chromatography. After removing the N-acyl group from the diastereoisomers, the desired (+)-(R)-or (?)-(S)- 7 was obtained by adding an N-iodopropenyl group. In vitro homogenate binding studies showed the stereoselectivity of this new compound for 5-HT1A receptors. (+)-(R)- 7 isomer displayed 100-fold higher affinity than the (?)-(S)- 7 isomer. Biochemical study indicated that (+)-(R)- 7 potently inhibited forskolin-stimulated adenylyl cyclase activity in hippocampal membranes (Emax and EC50 were 24.5% and 5.4 nM, respectively), while (?)-(S)- 7 showed no effect at 1 μM. The radioiodinated (+)-(R)- and (?)-(S)-[125I] 7 were confirmed by coelution with the resolved unlabeled compound on HPLC (reverse phase column PRP-1, acetonitrile/pH 7.0 buffer, 80/20). The active isomer, (+)-(R)-[125I] 7 , displayed high binding affinity to 5-HT1A receptors (Kd = 0.09 ± 0.02 nM). In contrast, the (?)-(S)- 7 isomer displayed a significantly lower affinity to the 5-HT1A receptor (Kd > 10 nM). Thus, (+)-(R)-[125I]trans-8-OH-PIPAT, (+)-(R)- 7 , an iodinated stereoselective 5-HT1A receptor agonist, is potentially useful for study of in vivo and in vitro function and pharmacology of 5-HT1A receptors in the central nervous system. © 1995 Wiley-Liss, Inc.  相似文献   

7.
The effects of (±)-, (+)-, and (?)-atenolol, sotalol, and amosulalol alone on the rat left atria and portal vein and on the respective β1- and β2-adrenoceptor-mediated responses to isoprenaline have been determined. (±)-Atenolol at 10?6 M had no effect whereas high concentrations of (+)- and (?)-sotalol, 10?5–10?4 M, and (±)-, (+)-, and (?)-amosulalol depressed the response of the rat left atria to cardiac stimulation which indicates membrane stabilizing activity. None of the drugs tested had any effect alone on the rat portal vein. The order of potency as antagonists was (±)-amosulalol > (±)-atenolol > (±)-sotalol at β1-adrenoceptors and (±)-amosulalol > (±)-sotalol > (±)-atenolol at β2-adrenoceptors. (±)-Atenolol and (±)-amosulalol are β1-selective whereas (±)-sotalol is β2-selective. For each of the racemic β-blockers, the β1- and β2-adrenoceptor blocking activity was predominantly due to the (?)-enantiomer. © 1993 Wiley-Liss, Inc.  相似文献   

8.
The stereochemistry of sulfate conjugation of isoproterenol (ISO) was examined with human liver, intestine, and platelets as the phenolsulfotransferase (PST) enzyme source and PAP35S as the cosubstrate. With the hepatic cytosol, two distinct sulfation reactions were identified, a high affinity reaction (Km 5 to 50 μM) and a low affinity reaction (Km 360 to 2,900 μM). The efficiency of sulfation (Vmax/Km) for both reactions was 5-fold higher for (+)- than for (?)-ISO. When the hepatic PSTs were resolved by ionexchange chromatography, it could be shown that the high affinity reaction was catalyzed by the monoamine (M) form and the low affinity reaction by the phenol (P) form of PST. Only the high affinity (M form) sulfation was detected in the jejunal cytosol with a Vmax/Km value 6.1-fold higher for (+)- than for (?)-ISO. Finally the platelet, as a potentially useful model tissue, also demonstrated only the high affinity M form reaction with a Vmax/Km value 5.7-fold higher for (+)- than for (?)-ISO. In summary, this study has shown that sulfation of ISO by PSTs in various human tissues is stereoselective and favors the inactive (+)-enantiomer over the active (?)-enantiomer by about 5-fold, a finding which should be considered in the therapeutic use of chiral drugs cleared by sulfate conjugation. © 1993 Wiley-Liss, Inc.  相似文献   

9.
A novel method to indicate the degree of chirality in polyaniline (PANI) was developed. The ( d ‐camphorsulfonic acid)‐ and (HCl)‐PANI‐based electrodes exhibited significantly different electrochemical performances in d ‐ and l ‐Alanine (Ala) aqueous solution, respectively, which can be used for the characterization the optical activity of chiral PANI. Cyclic voltammogram, tafel, and open circuit potential of PANI‐based electrodes were measured within d ‐ and l ‐Ala electrolyte solution, respectively. The open circuit potentials under different reacting conditions were analyzed by Doblhofer model formula, in which [C+]poly1/[C+]poly2 was used as a parameter to characterize the degree of chirality in chiral PANI. The results showed that [C+]poly1/[C+]poly2 can be increased with increasing concentrations of (1S)‐(+)‐ and (1R)‐(?)‐10‐camphorsulfonic acid. In addition, we detected that appropriate response time and lower temperature are necessary to improve the degree of chirality. Chirality 25:39‐42, 2013.© 2012 Wiley Periodicals, Inc.  相似文献   

10.
3H-prazosin, a new radioligand of high specific radioactivity (33 Ci/mmol) was used to characterise postsynaptic (α1) adrenoceptors in guinea pig lung membranes. Binding was saturable and of high affinity (KD 0.24 nM) with a Bmax of 54 fmol/mg protein. Adrenergic agonists competed for binding in the order (?)-epinephrine > (?)-norepinephrine ? (?)-phenyl-ephrine > (?)-isoproterenol. (+)-norepinephrine was 100x less potent than (?)-norepinephrine. α-Adrenergic antagonists competed in the order prazosin > WB 4101 > indoramin > phentolamine > haloperidol > chlorpromazine ? piperoxan > yohimbine, indicating that 3H-prazosin binding is probably to α1-adrenoceptors. Propranolol, methysergide and sulpiride inhibited binding only at high concentrations. Binding of (?)-3H-dihydroalprenolol under identical experimental conditions gave a KD of 0.93 nM and a Bmax of 870 fmol/mg protein, giving a ratio of beta : α-adrenoceptor binding sites of 16 : 1 in this lung membrane preparation. 3H-prazosin appears to be a useful ligand in studying α1-adrenoceptors.  相似文献   

11.
A (?) gamma-lactamase fromMicrobacterium hydrocarbonoxydans was purified to homogeneity by chromatography methods. SDS-PAGE showed the molecular weight of the enzyme was about 31 kDa. The purified enzyme had a specific activity of 61.3±2.5 U mg?1 for 2-azabicyclo [2.2.1] hept-5-en-3-one [(?) gamma-lactam]. The enantioselectivity factor (E) of the purified enzyme was 9.5±0.8 for unreacted (+) gamma-lactam. TheK m andV max value were 2.3±0.2 mM and 80.0±15.4 U mg?1 respectively. The highest activity was found at 30 °C and pH 8.0. ESIMS mass spectrometry analysis results and N-terminal sequence indicated the (?) gamma-lactamase might be a new enzyme.  相似文献   

12.
The in vitro aromatase activity in microsomal fractions from rat ovary and its inhibition by enantiomers of aminoglutethimide (AG), rogletimide (RG), and cyclohexylaminoglutethimide (ChAG) were studied by analysing the [3H]H2O released when [1β-3H]androstenedione was converted to estrone. Maximum velocity (Vmax) and the Michaelis-Menten constant (Km) of the microsomal aromatase enzyme were 17.40 ± 0.45 pmol/ml/mg protein/min and 1.02 ± 0.06 μM, respectively. The IC50s for the enantiomers were similar for (+)-R-AG and (?)-R-ChAG (0.86 ± 0.06 and 0.89 ± 0.15 μM, respectively). (+)S-ChA'G was most potent with IC50 of 0.075 ± 0.003 μM. The IC50s for (?)-S-AG, (+)-R-RG, and (?)-S-RG were in the same range (23.15 ± 2.74, 24.58 ± 2.46, and 24.43 ± 2.20 μM, respectively). © 1994 Wiley-Liss, Inc.  相似文献   

13.
An inducible enzyme catalysing the hydrolysis of (+)-usnic acid to (+)-2-desacetylusnic acid and acetic acid has been purified 150-fold from the mycelium of Mortierella isabellina grown in the presence of (+)-usnic acid. Purification was achieved by treatment with protamine sulfate, (NH4)2SO4 fractionation, negative adsorption on alumina Cγ gel and hydroxylapatite followed by chromatography on DEAE-cellulose and Sephadex G-200. The elution pattern from a Sephadex G-200 column indicated a MW of ca 7.6 × 104 for the enzyme. The apparent Km value for (+)-usnic acid at the pH optimum (pH 7) was 4.0 × 10?5 M. The enzyme was specific for (+)-usnic acid and inactive towards (?)-usnic acid, (+)-isousnic acid or certain phloracetophenone derivatives. Its activity was enhanced in the presence of divalent metal ions such as Co2+, Ni2+, Mn2+, Mg2+ and Zn2+.  相似文献   

14.
In the dioldehydrase reaction, L(+) propanediol reacts at a faster rate than the D(?)-isomer in competitive reactions, but when the diols are run independently the rate of the D(?)-isomer is 2.3 times greater than that of the L(+)-isomer. It is assumed that this reversal is due to a difference in the Michaelis-Menten constants. A ratio of KmD(?)/KmL(+) = 3.2 ± 0.3 at 25° is calculated from a series of competitive reactions. It is concluded that the binding site is such that while both isomers are accomodated readily, the D(?)-isomer is hindered to some degree in its approach to the binding site. However, once the enzyme-substrate complex is formed, the D(?)-isomer is in a more favorable configuration for further reaction.  相似文献   

15.
A (?)-[3H]norepinephrine binding site was identified in a crude synaptosomal fraction isolated from bovine hypothalamus which bound norepinephrine rapidly, reversibly, and stereospecifically. The results were most consistent with binding of (-)-[3H]norepinephrine to the carrier molecule used to translocate biogenic amines into synaptic vesicles. The binding studies indicated that specific binding of (?)-[3H]norepinephrine to the crude synaptosomal fraction was greatly enhanced by 1 mM MgCl2 and 1 mM ATP. The increased binding of (?)-[3H]norepinephrine also occured in the presence of MgCl2 and GTP, but AMP, adenosine and adenyl-5′-yl imidodiphosphate would not substitute for ATP. Neither CaCl2 nor ZnSO4 could be substituted for the MgCl2. In the presence of MgCl2 and ATP, the dissociation constant for (?)-[3H]norepinephrine was 280 nM with a specific binding site density of 4.8 pmol/mg protein. Binding was stereospecific with ratios of 15, 4, and 6.5 for the affinities of (?)-isomers to (+)-isomers for norepinephrine, epinephrine and isoproterenol, respectively. Drug competition studies, conducted in the presence of Mg2+ and ATP, indicated that (?)-epinephrine, (?)-norepinephrine, dopamine and serotonin had inhibitory constants ranging from 0.25 to 0.8 μM with (?)-isoproterenol and tyramine having inhibitory constants around 2 μM. Reserpine was the most potent inhibitor having an inhibition constant of 8.6 ± 0.3 nM. The binding data were not consistent with the specific site being the α- or β-receptors for norepinephrine, the Uptake1 Site for norepinephrine into synaptosomes or the metabolizing enzymes for norepinephrine.  相似文献   

16.
The chiral inversion and pharmacokinetics of two enantiomers of trantinterol, a new β2 agonist, were studied in rats dosed (+)‐ or (?)‐trantinterol separately. Plasma concentrations of (+)‐ and (?)‐trantinterol were measured by chiral stationary phase liquid chromatography tandem mass spectroscopy (LC‐MS/MS). The apparent inversion ratio was calculated as the ratio of AUC0‐t of (?)‐trantinterol or (+)‐trantinterol inverted from their antipodes to the sum of the AUC0‐t of (?)‐ and (+)‐trantinterol. Following single intravenous administration, both given enantiomers declined in similar plasma concentrations, suggesting that the two enantiomers have approximately the same disposition kinetics by the route of intravenous administration. However, after single oral administration, plasma concentrations of uninverted (?)‐trantinterol at many timepoints were significantly higher than those of uninverted (+)‐trantinterol, suggesting that the two enantiomers undergo apparently different absorption or metabolism after oral administration. Significant bidirectional chiral inversion occurred after intravenous and oral administration of (+)‐ or (?)‐trantinterol. After dosing with optically pure enantiomer, the concentration of the administered enantiomer predominated in vivo. The AUC0‐36 of (+)‐trantinterol after intravenous and oral dosing of (?)‐trantinterol were 16.6 ± 5.2 and 33.3 ± 16%, respectively of those of total [(+) + (?)] trantinterol. The AUC0‐36 of (?)‐trantinterol after intravenous and oral dosing of (+)‐trantinterol were 19.6 ± 8.8 and 37.9 ± 4.5%, respectively, of those of total [(?) + (+)] trantinterol. After intravenous administration of (+)‐ and (?)‐trantinterol the chiral inversion ratios of the two enantiomers were not significantly different and similar results were found for oral administration. The extent of chiral inversion after intravenous administration was apparently lower, indicating that the bidirectional chiral inversion was not only systemic but also presystemic. Chirality 25:934–938, 2013.© 2013 Wiley Periodicals, Inc.  相似文献   

17.
A rapid and sensitive liquid chromatography–tandem mass spectrometry assay (LC–MS/MS) with electrospray ionization was developed and validated for the quantitative determination of the concentration of methotrexate (MTX) enantiomers in intracellular and extracellular fluids of HepG2 cells. The analytes were extracted from homogenates using organic solvent to precipitate proteins. The extracted samples were analyzed by LC–MS/MS, operating in multiple reactions monitoring (MRM) mode. The condition of HPLC included the following: Gemini column (3 μm, 3.0 × 75 mm) with chromatographic column was used, and the mobile phase consisting of gradient elution utilized 0.1 % formic acid as solvent A and acetonitrile as solvent B at a flow rate of 0.4 mL min?1. The gradient was as follows: 0–7.0 min 10–90 % B, 7.0–10 min 90 % B followed by 3 min. The column temperature was maintained at 40 °C. The condition of MS included using electrospray ionization source; MRM mode with the transitions of m/z 455.2 → m/z 308.1 was used to quantify MTX enantiomers. The linear calibration curve was obtained in the concentration range of 10.0 to 10,000 ng mL?1 for MTX enantiomers in intracellular and extracellular fluids. The inter- and intraday precision was less than 15 %. The mean recovery of (+)-MTX and (?)-MTX in the extracellular fluid of HepG2 cells were 95.30 and 96.53 %, respectively, and the mean recovery of (+)-MTX and (?)-MTX in the intracellular fluid of HepG2 cells were 93.53 and 94.12 %, respectively. This method was successfully used to detect the concentration of MTX enantiomers in the intracellular and extracellular fluids of HepG2 cells and that the concentration of (+)-MTX in intracellular fluid was twice higher than the concentration of (?)-MTX in intracellular fluid. The inhibitory effect of (+)-MTX and (?)-MTX was (+)-MTX > (?)-MTX. It is a simple, precise method that can effectively explain the difference in pharamocological effect of MTX enantiomers in vitro.  相似文献   

18.
3H-Yohimbine, a potent and selective pharmacological antagonist of α2-adrenergic receptors, labeled human platelet membrane α2-receptors with high affinity. Binding was rapid and reversible at 25°C. Both saturation and kinetic experiments indicated a single order of binding sites, with an equilibrium KD value of 1.0–1.5 nM. Low Mg2+ concentrations increased the KD for 3H-yohimbine without altering the Bmax. The 3H-yohimbine site exhibited α2-receptor specificity: (?)-norepinephrine and (?)-isoproterenol were 4.8 and 330 times less potent than (?)-epinephrine; (?)-catecholamines were 17–35 times more potent than corresponding (+)-catecholamines; the selective α1-antagonist prazosin was 340 times less potent than yohimbine. Catecholamine agonists exhibited shallow curves in inhibiting 3H-yohimbine binding, with pseudo-Hill coefficients (nH) of less than 1.0, whereas the nH of antagonists was 1.0. No specific binding of 3H-prazosin to platelet membranes was observed, indicating the absence of α1-receptors. 3H-Yohimbine labeled fewer platelet sites than did 3H-dihydroergocryptine under identical conditions (80 vs 130 receptors/ cell), and may be a more specific and useful antagonist probe of platelet α2-receptors than 3H-dihydroergocryptine.  相似文献   

19.
Methods are described for the synthesis of [3H]chloramphenicol and derivatives labeled on carbon 1 of the propanediol side chain, with a specific activity of about 2 mCi/μmol. The labeling step involves the reduction of the I-oxo derivative of N-acetyl chloramphenicol base by KB3H4 to produce a mixture of the d (?) threo- and d (?) erythro-diastereoisomers, since carbon 1 is an asymmetric carbon atom. The two isomers were separated by thin-layer chromatography after acetylation of the two free hydroxyls. After hydrolysis of the three acetyl groups, the biologically active d (?) threo-[1 ? 3H]chloramphenicol base was converted to chloramphenicol. Modification of the above procedures allows the rapid and simple preparation of the mixed d (?) threo- and d (?) erythro-isomers of [1 ? 3H]chloramphenicol. This mixture can be used where the presence of the inactive d (?) erythro-isomer of chloramphenicol is not important. The modified procedure also allows the preparation of the mixed isomers of [1 ? 3H]chloramphenicol base and of chloramphenicol analogs. Attempts to prepare a 3-aldehyde derivative of chloramphenicol were not successful. If this could be done, reduction of this derivative with KB3H4 would produce the correct isomer of chloramphenicol since carbon atom 3 on the propanediol side chain is not an asymmetric carbon atom.  相似文献   

20.
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   

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