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.
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   

2.
Chiral column liquid chromatography and enantiospecific enzymatic hydrolysis were utilized to separate the enantiomers of α- and β-dihydrotetrabenazine and α-9-O-desmethyldihydrotetrabenazine, three benzo[a]quinolizines derived from the amine-depleting drug tetrabenazine. An X-ray crystal structure analysis of (−)-α-9-O-desmethyldihydrotetrabenazine gave an absolute structure of that compound as the 2S, 3S, 11bS isomer. Therefore, (−)-α-dihydrotetrabenazine also has the 2S, 3S, 11bS absolute configuration. (+)-α-Dihydrotetrabenazine, the single biologically active isomer from the metabolic reduction of tetrabenazine, thus has the absolute configuration of 2R, 3R, 11bR. For further in vitro and in vivo studies of the vesicular monoamine transporter, it is now possible to use the single enantiomer of radiolabeled α-dihydrotetrabenazine. Chirality 9:59–62, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

3.
The enantioselectivity of 4‐hydroxylation of bunitrolol (BTL), a β‐adrenoceptor blocking drug, was studied in microsomes from human liver, human hepatoma (Hep G2) cells expressing CYP2D6, and lymphoblastoid cells expressing CYP2D6. Kinetics in human liver microsomes showed that the Vmax value for (+)‐BTL was 2.1‐fold that of (−)‐BTL, and that the Km value for (+)‐BTL was lower than that for the (−)‐antipode, resulting in the intrinsic clearance (Vmax/Km) of (+)‐BTL being 2.1‐fold over its (−)‐antipode. CYP2D6 (CYP2D6‐met) expressed in Hep G2 cells had a methionine residue at position 373 of the amino acid sequence and a rat‐type N‐terminal peptide (MELLNGTGLWSM) instead of the human‐type (MGLEALVPLAVIV), and showed enantioselectivity of [(+)‐BTL < (−)‐BTL] for the rate of BTL 4‐hydroxylation. In contrast, enantioselectivity [(+)‐BTL > (−)‐BTL] for Hep G2‐CYP2D6 (CYP2D6‐val) with a human‐type N‐terminal peptide that had a valine residue at 374, which corresponds to the methionine of the CYP2D6‐met variant, was the same as that for human liver microsomes. We further confirmed that CYP2D6‐met and CYP2D6‐val expressed in human lymphoblastoid cells, both of which have methionine and valine, respectively, at position 374 and a human‐type N‐terminal peptide, exhibited the same enantioselectivities as those obtained from CYP2D6‐met and CYP2D6‐val expressed in the Hep G2 cell system. These results indicate that the amino acid at 374 of CYP2D6 is one of the key factors influencing the enantioselectivity of BTL 4‐hydroxylation. Chirality 11:1–9, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

4.
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   

5.
(±)-Pantoprazole ((±)-PAN), (±)-5-(difluoromethoxy)-2-[[(3.4-dimethoxy-2-pyridinyl)methyl]sulfinyl]-1H-benzimidazole is a chiral sulfoxide that is used clinically as a racemic mixture. The disposition kinetics of (+)-PAN and (−)-PAN given separately has been studied in rats. Serum levels of (+)- and (−)-PAN and its metabolites, pantoprazole sulfone (PAN-SO2), pantoprazole sulfide (PAN-S), 4′-O-demethyl pantoprazole sulfone (DMPAN-SO2), and 4′-O-demethyl pantoprazole sulfide (DMPAN-S) were measured by HPLC. Following single intravenous or oral administration, both enantiomers were rapidly absorbed and metabolized, resulting in similar serum concentrations, suggesting that the two enantiomers have approximately the same disposition kinetics. The major metabolite of both (+)- and (−)-PAN was PAN-SO2, while DMPAN-SO2 was also detected as a minor metabolite. Serum levels of PAN-S and DMPAN-S could not be quantified after intravenous or oral administration of either enantiomer. Significant chiral inversion occurred after intravenous and oral administration of (+)-PAN. The AUCs of (−)-PAN after intravenous and oral dosing of (+)-PAN were 36.3 and 28.1%, respectively of those of total [(+) + (−)] PAN. In contrast, the serum levels of (+)-PAN were below quantitation limits after intravenous or oral administration of (−)-PAN. Therefore, chiral inversion was observed only after administration of (+)-PAN, supporting the hypothesis that stereoselective inversion from (+)-PAN to (−)-PAN occurs in rats. Chirality 10:747–753, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

6.
The N-oxides of (−)-hyoscyamine and (−)-hyoscine have been prepared and characterized. Two N-oxides of hyoscyamine have been isolated from species of Atropa,Datura,Hyoscyamus,Scopolia andMandragora,andoneN-oxideofhyoscinehasbeenisolatedfromspeciesofthefirstfourgenera.  相似文献   

7.
N‐[1‐(4‐(4‐fluorophenyl)‐2,6‐dioxocyclohexylidene)ethyl] (Fde) protected amino acids have been prepared and applied in solid‐phase peptide synthesis monitored by gel‐phase 19F NMR spectroscopy. The Fde protective group could be cleaved with 2% hydrazine or 5% hydroxylamine solution in DMF as determined with gel‐phase 19F NMR spectroscopy. The dipeptide Ac‐L ‐Val‐L ‐Val‐NH2 12 was constructed using Fde‐L ‐Val‐OH and no noticeable racemization took place during the amino acid coupling with N,N′‐diisopropylcarbodiimide and 1‐hydroxy‐7‐azabenzotriazole or Fde deblocking. To extend the scope of Fde protection, the hydrophobic nonapeptide LLLLTVLTV from the signal sequence of mucin MUC1 was successfully prepared using Fde‐L ‐Leu‐OH at diagnostic positions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Two preparations of the enantiomers of 2 are described. The first makes use of the chromatographic separation of the diastereomeric amides 6a and 6b. Standard hydrolysis of these amides caused racemization, so a milder sequence was developed which utilized carbonyldiimidazole and 1 equivalent of 1 N LiOH. The second preparation involved classical resolution of 9 with (?)-cinchonidine. Subsequent transformations of this substrate involved ester formation, Friedel–Crafts acylation, and ester hydrolysis, all without racemization. The most notable of these reactions was the use of EtAlCl2 in the Friedel–Crafts step, which provided a mild acylation of 10. This second preparation affords a high yield, mild process for the potential preparation of kilogram quantities of (?)-(R)-2b.  相似文献   

9.
Lipase from Candida antarctica, commercially available immobilised on acrylic resin as Novozym® 435, allows for enantioselective esterification of rac-indoprofen (±)-1, with methanol in a dioxane-toluene solvent system. A double esterification process affords methyl ester (−)-(R)- 2 in 85% e.e. and enantiopure (+)-(S)- 1 , both in good chemical yield. Chirality 10:321–324, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
《Chirality》2017,29(2):70-79
The chirality issues in the essential oils (EOs) of leaves and flowers from Bubonium graveolens were addressed by chiral high‐performance liquid chromatography (HPLC) with polarimetric detection and vibrational circular dichroism (VCD). The chemical compositions of the crude oils of three samples were established by gas chromatography / mass spectrometry (GC/MS). The well‐known cis ‐chrysanthenyl acetate ( 1 ), oxocyclonerolidol ( 2 ), and the recently disclosed cis ‐acetyloxychrysanthenyl acetate ( 3 ), the three major chiral compounds, were isolated by preparative HPLC. The naturally occurring oxocycloneroledol ( 2 ), mostly found in the leaf oil (49.4–55.6%), presents a (+) sign in the mobile phase during HPLC on a chiral stationary phase (CSP) with a Jasco polarimetric detection. The naturally occurring cis ‐chrysanthenyl acetate ( 1 ) and cis ‐acetyloxychrysanthenyl acetate ( 3 ), mostly found in the flower EO (35.9–74.9% and 10.0–34.3%, respectively), both present a (−) sign. HPLC on a CSP with polarimetric detection is an unprecedented approach to readily differentiate the flower and leaf EOs according to their chiral signature. The comparison of the experimental and calculated VCD spectra of pure isolated 1 , 2, and 3 provided their absolute configuration as being (1S ,5R ,6S )‐(−)‐2,7,7‐trimethylbicyclo[3.1.1]hept‐2‐en‐6‐yl acetate 1 , (2R ,6R )‐(+)‐6‐ethenyl‐2,6‐dimethyl‐2‐(4‐methylpent‐3‐en‐1‐yl)dihydro‐2H‐pyran‐3(4H)‐one) 2 and (1S ,5R ,6R ,7S )‐(−)‐7‐(acetyloxy)‐2,6‐dimethylbicyclo[3.1.1]hept‐2‐en‐6‐yl]methyl acetate 3 . Compounds 1 , 2, and 3 were already known in B. graveolens but this is the first report of the absolute configuration of (+)‐ 2 and (−)‐ 3 . The VCD chiral signatures of the crude oils were also recorded.  相似文献   

11.
  1. Lure attraction strength for red turpentine beetle, Dendroctonus valens (Coleoptera: Curculionidae: Scolytinae) observed previously in US Pacific Northwest ponderosa pine forests is (−)-β-pinene+ethanol > (+)-3-carene+ethanol, but untested elsewhere in its western US range. Thus, both were tested with (−)-β-pinene, (+)-3-carene, ethanol, and a blank in Oregon and California sites burned by wildfire, whereas in Arizona the first four lures were tested in a thinned-unburned site.
  2. The D. valens responses in burned Oregon and California sites were similar, (−)-β-pinene+ethanol > (−)-β-pinene > 3-carene = 3-carene+ethanol > ethanol > blank, whereas in the cut-unburned Arizona site it was 3-carene+ethanol > 3-carene = (−)-β-pinene+ethanol > (−)-β-pinene. Whether this variation was influenced by beetle genetic differences, or chemical and physical parameters in the different environments and remaining stressed host resources 1-year post disturbance warrants additional study.
  3. Responses to (−)-β-pinene varied, from a stronger attractant than (+)-3-carene in Oregon and California, to a weaker lure than (+)-3-carene in Arizona. This (−)-β-pinene variability was minimized when released in combination with ethanol, making (−)-β-pinene+ethanol the most consistent attractant of those tested across the three states, and a reliable lure for detection, monitoring, and management projects for D. valens in western US pine forests.
  相似文献   

12.
More effective use of optical resolution processes can be obtained by increasing the overall yields after development of methods for inversion of the chiral centre of the unwanted isomer. The configuration of some optically active amines can be inverted in a three-step synthesis via the N,N-ditosylimides and a subsequent nucleophilic substitution by the azide ion. The azide product is reduced by hydrogenolysis. Low stereoselectivity caused by racemization to some extent was at first observed for the inversion of the benzylic substrate, (S)-α-methylbenzylamine ( 5a ). However, modified reaction conditions allowed increased stereoselectivity, a more rapid and almost complete inversion of this substate as well. © 1994 Wiley-Liss, Inc.  相似文献   

13.
The synthesized 7-aryl derivatives of (7R,7′S,8S,8′S)-(+)-verrucosin were applied to growth inhibitory activity test against ryegrass at 1 mM. 7-(3-Ethoxy-4-hydroxyphenyl) derivative 12 and 7-(2-hydroxyphenyl) derivative 4 showed comparable activity to those of (+)-verrucosin against the root (−95%) and the shoot (−60%), respectively. The growth inhibitory activity test against lettuce using synthesized 7-aryl derivatives of (7S,7′R,8R,8′R)-(−)-verrucosin at 1 mM showed that the activities of 7-(3-hydroxyphenyl) derivative 20 and 7-(3-ethoxy-4-hydroxyphenyl) derivative 28 are similar to that of (−)-verrucosin against the root (−95%). Against the shoot, 7-(3-hydroxyphenyl) derivative 20 showed higher activity (−80%) than that of (−)-verrucosin (−60%). As the next step, (7S,7′R,8R,8′R)-7-(3-hydroxyphenyl)-7′-aryl-(−)-verrucosin derivatives, in which the most effective 3-hydroxyphenyl group is employed as 7-aromatic ring, were synthesized for the assay against lettuce. In this experiment, 7′-(2-hydroxyphenyl) derivative 37 and 7′-(3-hydroxyphenyl) derivative 38 showed similar activity to that of derivative 20. The effect of 7- and 7′-aryl structures of 7,7′-epoxylignanes on the plant growth inhibitory activity was clarified. The 7- and 7′-aryl structures were simplified to show comparable activity to or higher activity than that of (−)-verrucosin. The plant growth inhibitory activity of a nutmeg component, (+)-fragransin C3b, was estimated as −80% inhibition at 1 mM against ryegrass roots.  相似文献   

14.
Thermostable N-acylamino acid recemase from Amycolatopsis sp. TS-1-60, a rare actinomycete strain selected for its ability to grow on agar plates incubated at 40° C, was purified to homogeneity and characterized. The relative molecular mass (M r) of the native enzyme and the subunit was estimated to be 300 000 and 40 000 on gel filtration chromatography and sodium dodecyl sulfate-polyacrylamide gel electrophoresis respectively. The isoelectric point (pI) of the enzyme was 4.2. The optimum temperature and pH were 50° C and 7.5 respectively. The enzyme was stable at 55° C for 30 min. The enzyme catalyzed the racemization of optically active N-acylamino acids such as N-acetyl-l-or d-methionine, N-acetyl-l-valine, N-acetyl-l-tyrosine and N-chloroacetyl-l-valine. In addition, the enzyme also catalyzed the recemization of the dipeptide l-alanyl-l-methionine. By contrast, the optically active amino acids, N-alkyl-amino acids and methyl and athyl ester derivatives of N-acetyl-d- and l-methionine were not racemized. The apparent K m values for N-acetyl-l-methionine and N-acetyl-d-methionine were calculated to be 18.5 mM and 11.3 mM respectively. The enzyme activity was markedly enhanced by the addition of divalent metal ions such as Co2+, Mn2+ and Fe2+ and was inhibited by addition of EDTA and P-chloromercuribenzoic acid. The similarity between the NH2-terminal amino acid sequence of the enzyme and that of Streptomyces atratus Y-53 [Tokuyama et al. (1994) Appl Microbiol Biotechnol 40:835–840] was above 80%.  相似文献   

15.
Two new compounds, (S)-(−)-N-[2-(3-Hydroxy-2-oxo-2,3-dihydro-1H-indol-3-yl)-ethyl]-acetamide (1) and 6-formyl-5-isopropyl-3-hydroxymethyl-7-methyl-1H-indene (2), were isolated from the leaves of Selaginella pulvinata. Their structures were determined by spectroscopic methods. Additionally, compound 1 could inhibit the growth of SK-mel-110 cells and induce cell apoptosis in vitro through up-regulating the expression of inhibitor of growth family member 4.  相似文献   

16.
Hydrazides of five N-acylamino acids have been used alone as substrates for papain catalysis to yield N1,N2-diacylhydrazines. With the exception of N-(benzyloxycarbonyl)(Z)- -alanine hydrazide, they were very effective as both acylating agents of the enzyme and nucleophiles in attacking the enzyme-substrate intermediate. Although Z- -alanine hydrazide was a minimal acylating agent, it was a satisfactory nucleophile. The most favorable reaction involved Z- -alanine hydrazide in producing N1,N2-bis(Z- -alanyl)hydrazine. When Z- -alanine hydrazide was the substrate, this same chiral diacylhydrazine was formed along with meso N1-(Z- -alanyl)-N2-(Z- -alanyl)hydrazine. For the acylation step, the enzyme displayed powerful, essentially stereospecific, bias toward the enantiomer. Once the thioester intermediate was formed, little preference was detected for attack by the enantiomers as nucleophiles. The most direct procedure for synthesis of substrates was conversion of Z-amino acids to their esters by means of dry HCl in an absolute alcohol. Treatment with hydrazine produced the hydrazides in excellent yield.  相似文献   

17.
《Chirality》2017,29(11):726-736
Pyricularia grisea has been identified as a foliar pathogen on buffelgrass (Cenchrus ciliaris ) in North America and was studied as a potential source of phytotoxins for buffelgrass control. Two monosubstituted hex‐4‐ene‐2,3‐diols, named pyriculins A and B, were isolated from its culture filtrate organic extract together with (10S ,11S )‐(−)‐epipyriculol, trans ‐3,4‐dihydro‐3,4,8‐trihydroxy‐1(2H )‐napthalenone, and (4S )‐(+)‐isosclerone. Pyriculins A and B were characterized by spectroscopic (essentially nuclear magnetic resonance [NMR], High‐resolution electrospray ionization mass spectrometry [HRESIMS]) and chemical methods such as (4E )‐1‐(4‐hydroxy‐1,3‐dihydroisobenzofuran‐1‐yl)hex‐4‐ene‐2,3‐diols. The relative and absolute configuration of these compounds was determined by a combination of spectroscopic (NMR, electronic circular dichroism [ECD]) and computational tools. When bioassayed in a buffelgrass coleoptile and radicle elongation test, (10S ,11S )‐(−)‐epipyriculol proved to be the most toxic compound. Seed germination was much reduced and slowed with respect to the control and radicles failed to elongate. All five compounds delayed germination, but only (10S ,11S )‐(−)‐epipyriculol was able to prevent radicle development of buffelgrass seedlings. It had no effect on coleoptile elongation, while the other four compounds caused significantly increased coleoptile development relative to the control.  相似文献   

18.
  • 1.1. In the narrow Mediterranean area of Il Rogiolo, the acetates of the secondary alcohols of β-chamigrene (−)-1b, branched lauroxepane (−)-2b, and isopimarane (+)-3b are found in the dictyoceratid sponge Spongia zimocca, whereas the corresponding free alcohols (+)-1a, (−)-2a, and probably also (+)-3a, are found in the red seaweed Laurencia microcladia.
  • 2.2. No one of these acetylated metabolites could be detected either in this seaweed or in the global algal mass of Il Rogiolo.
  • 3.3. Terpenes bearing tertiary alcoholic functionalities of either L. microcladia (−)-4E, (−)-4Z or another red seaweed of the same area, Sphaerococcus coronopifolius [(+)-5, (−)-6], are found unaltered also in S. zimocca.
  • 4.4. These findings imply transfer of the hydroxyl-bearing metabolites from the seaweeds to the sponge and it is tempting to speculate that the secondary alcoholic function is acetylated in the sponge while tertiary alcohols and the secondary alcohol (+)-7 escape acetylation, the first ones as sterically-hindered alcohols and the latter one as a sponge metabolite residing in a special cell compartment.
  相似文献   

19.
In a series of 5-acyl-6-phenyl-2,4-substituted-3(2H)-pyridazinones the derivative 1a , with a sulfur stereogenic center, had the most potent activity as human platelet aggregation inhibitor. The resolution of rac- 1a was successfully performed by chiral chromatography on Chiralcel OD-R, OD-H, and Chiralpak AD columns and scaled up to a preparative level. The absolute configuration of (−)-(S)- 1a was determined by X-ray crystallographic analysis. In vitro human platelet aggregation inhibitory activity was evaluated. Both the enantiomers showed IC50 values in the same micromolar range, but the (−)-(S) isomer was slightly more potent [(S)/(R) potency ratio was 4/1]. Chirality 9:681–685, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

20.
Improved radiopharmaceuticals for imaging cerebral acetylcholinesterase (AChE) are needed for the diagnosis of Alzheimer’s disease (AD). Thus, 11C-labeled (−)-galanthamine and its enantiomers were synthesized as novel agents for imaging the localization and activity of AChE by positron emission tomography (PET). C-11 was incorporated into (−)- and (+)-[11C]galanthamine by N-methylation of norgalanthamines with [11C]methyl triflate. Simple accumulation of 11C in the brain was measured in an in vivo biodistribution study using mice, whilst donepezil was used as a blocking agent in analogous in vivo blocking studies. In vitro autoradiography of rat brain tissue was performed to investigate the distribution of (−)-[11C]galanthamine, and confirmed the results of PET studies in mice. The radiochemical yields of N-methylation of (−)- and (+)-norgalanthamines were 13.7% and 14.4%, respectively. The highest level of accumulation of 11C in the brains of mice was observed at 10 min after administration (2.1% ID/g). Intravenous pretreatment with donepezil resulted in a 30% decrease in accumulation of (−)-[11C]galanthamine in the striatum; however, levels in the cerebellum were unchanged. In contrast, use of (+)-[11C]galanthamine led to accumulation of radioactivity in the striatum equal to that in the cerebellum, and these levels were unaffected by pretreatment with donepezil. In in vitro autoradiography of regional radioactive signals of brain sections showed that pretreatment with either (−)-galanthamine or donepezil blocked the binding of (−)-[11C]galanthamine to the striatum, while sagittal PET imaging revealed accumulation of (−)-[11C]galanthamine in the brain. These results indicate that (−)-[11C]galanthamine showed specific binding to AChE, whereas (+)-[11C]-galanthamine accumulated in brain tissue by non-specific binding. Thus, optically pure (−)-[11C]galanthamine could be a useful PET tracer for imaging cerebral AChE.  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