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.
(+)-2,9 alpha-Dimethyl-5-(m-hydroxyphenyl)morphan is the only phenylmorphan analog whose affinity for opioid kappa-receptors is greater than its affinity for opioid mu-receptors. Pharmacologically, the compound is a pure opioid antagonist devoid of agonist activity in in vivo assays of antinociception. The absolute configuration of the compound has been determined to be (1R,5S,9R) from an X-ray crystallographic study of the chloride salt. Thus, the absolute configuration corresponds to that of the atypical opioid agonist (-)-phenylmorphan while the weak atypical agonist (-)-2,9 alpha-dimethyl-5-(m- hydroxyphenyl)morphan corresponds to the potent morphine-like (+)-phenylmorphan. The preferred orientations of the phenyl ring for the two stereoisomers were determined using the molecular mechanics program MM2-87 and found to vary from that of the two parent compounds. The atypical properties of the two 9 alpha-methyl analogs is consistent with an opioid ligand model which proposes that morphine-like properties require a particular range of phenyl orientations. There was good agreement between the structure obtained from X-ray crystallography and computed with the MM2-87 program.  相似文献   

2.
The absolute configurations of the enantiomers of the opiod picenadol [cis-1,3-dimethyl-4-propyl-4-propyl-4-(3-hydroxyphenyl)piperidine; cis-3-methyl, 4-propyl] have been determined by an X-ray crystallographic study of the chloride salt of the (+)-enantiomer. The agonist (+)-enantiomer and the antagonist (?)-enantiomer were found to have the 3R, 4R and 3S, 4S absolute configurations, respectively. The conformational properties of the enantiomers were also examined with MM2–87 calculations. There was good agreement between the computed global minimum and the crystallographic structure with the phenyl ring approximately bisecting the piperidine ring by both methods. This orientation of the phenyl ring differs from that of related opioids such as the phenylmorphans, prodines, meperidine, and ketobemidone in which the phenyl ring tends to eclipse one edge of the piperidine ring. Because the phenyl ring bisects the piperidine ring in picenadol, there is little difference in the three-dimensional orientations of the phenyl rings of the two enantiomers when one superimposes the piperidine rings. The agonist (+)-enantiomer is ambiguous with respect to an opioid ligand model, which suggests that agonist activity requires a specific range of dihedral angles for the phenyl ring. While the global minimum of the agonist is not consistent with the model, a second conformer that is only 1.2 kcal/mol above the global minimum is consistent. An alternative explanation is that agonist or antagonist activity is solely due to the presence of the 3-methyl group on the different edges of the piperidine ring. MM2–87 calculations were also performed on the opioid agonist des-3-methyl analog of picenadol and the closely related trans-1,3,4-trimethyl-4-(3-hydroxyphenyl)piperidines (trans-3-methyl, 4-methyl) in which both enantiomers are opioid antagonists. The conformational properties of these compounds are consistent with the ligand model. © 1995 Wiley-Liss, Inc.  相似文献   

3.
Conformational studies of enkephalins are hampered by their high flexibility which leads to mixtures of quasi-isoenergetic conformers in solution and makes NOEs very difficult to detect in NMR spectra. In order to improve the quality of the NMR data, Leu–enkephalin was synthesized with 15N-labelled uniformly on all amide nitrogens and examined in a viscous solvent medium at low temperature. HMQC NOESY spectra of the labelled Leu–enkephalin in a DMSOd6/H2O mixture at 275 K do show numerous NOEs, but these are not consistent with a single conformer and are only sufficient to describe the conformational state as a mixture of several conformers. Here a different approach to the structure–activity relationships of enkephalins is presented: it is possible to analyse the NMR data in terms of limiting canonical structures (i.e. β- and γ-turns) and finally to select only those consistent with the requirements of δ selective agonists and antagonists. This strategy results in the prediction of a family of conformers that may be useful in the design of new δ selective opioid peptides. © 1998 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
The (S)-enantiomer of 5-fluoro-8-hydroxy-2-(dipropylamino) tetralin [(S)- 2a; (S)-UH301] was the first reported 5-HT1A receptor antagonist. We now give a full account on the synthetic effort leading to the preparation of the racemate and the enantiomers of 2a. The crystal and molecular structure of 2a · HBr has been determined by X-ray diffraction and the absolute configuration has been deduced using statistical tests of the crystallographic R values. The unit cell is tetragonal (P41212) with a = b = 13.2235 (2), c = 39.560(1) Å and contains two crystallographically independent molecules in each asymmetric unit. The two solid state conformers differ in the conformation of the N-propyl groups. The pharmacological characterization of the enantiomers was done by use of in vivo biochemical and behavioural assays in rats. The (R)-enantiomer of 2a is a 5-HT1A receptor agonist of low potency while (S)- 2a does not exhibit any agonist properties at 5-HT1A receptors. As a consequence of the opposing effects of the enantiomers, the racemate, rac- 2a, does not produce any clear-cut effects in rats. The reduced efficacy of (S)- 2a as compared to the well known 5-HT1A receptor agonist 8-hydroxy-2-(dipropylamino)tetralin ( 1; 8-OH-DPAT) may be due to the fluoro-substituent induced negative potential of the aromatic ring. Chirality 8:531–544, 1996. © 1997 Wiley-Liss, Inc.  相似文献   

5.
B E Hingerty  S Broyde 《Biopolymers》1985,24(12):2279-2299
The low-energy conformations accessible to dCpdG modified at guanine N2 via trans epoxide opening by (+) and (?) 7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (anti-BPDE) have been delineated by minimized semiempirical potential-energy calculations with all torsion angles flexible. Nearly 4000 trials were made, representing a fairly thorough investigation of the conformation space of the adducts. Carcinogen–base stacked states and base–base stacked conformers were found in the low-energy regions of both enantiomers. Many ω′, ω, ψ domains accommodate the two types of conformations, with B-like backbones among the most preferred states in each case. The conformational differences between the two enantiomers on the dimer level reside in subtle distinctions in orientation of the carcinogen–base linkage.  相似文献   

6.
A modified β-cyclodextrin bearing a 2-aminomethylpyridine binding site for copper(II) (6-deoxy-6-[N-(2-methylamino)pyridine)]-β-cyclodextrin, CDampy was synthesized by C6-monofunctionalization. The acid-base properties of the new ligand in aqueous solution were investigated by potentiometry and calorimetry, and its conformations as a function of pH were studied by NMR and circular dichroism (c.d.). The formation of binary copper(II) complexes was studied by potentiometry, EPR, and c.d. The copper(II) complex was used as chiral selector for the HPLC enantiomeric separation of underivatized aromatic amino acids. Enantioselectivity in the overall stability constants of the ternary complexes with D- or L-Trp was detected by potentiometry, whereas the complexes of the Ala enantiomers did not show any difference in stability. These results were consistent with a preferred cis coordination of the amino group of the ligand and of the amino acid in the ternary complexes (“cis effect”), which leads to the inclusion of the aromatic side chain of D-Trp, but not of that of L-Trp. In Trp-containing ternary complexes, the two enantiomers showed differences in the fluorescence lifetime distribution, consistent with only one conformer of D-Trp and two conformers of L-Trp, and the latter were found to be more accessible to fluorescence quenching by acrylamide and KI. Chirality 9:341–349, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

7.
Chiral discrimination observed in high‐performance liquid chromatography (HPLC) with the novel chiral stationary phase (CSP‐18C6I) derived from (+)‐(R)‐18‐crown‐6 tetracarboxylic acid [(+)‐18C6H4] was investigated by X‐ray crystallographic analysis of the complex composed of the R‐enantiomer of 1‐(1‐naphthyl)ethylamine (1‐NEA) and (+)‐18C6H4. Mixtures of 1‐NEA (the R‐ or S‐enantiomer) and (+)‐18C6H4 were dissolved in methanol‐water (1:1) solution and allowed to stand for crystallization. The R‐enantiomer crystallized with (+)‐18C6H4 as a co‐crystal, although the S‐enantiomer did not. This result was in good agreement with the enantiomer elution order of 1‐NEA in CSP‐18C6I. The apparent binding constants (Ka) of the enantiomers to the (+)‐18C6H4 obtained from 1H‐NMR experiments also supported the above‐mentioned result. The X‐ray crystal structure of the 1:1 complex of the R‐enantiomer and (+)‐18C6H4 indicated the four sets of hydrogen bond association between the naphthylethylammonium cation and oxygen of polyether ring or carbonyl group of (+)‐18C6H4. Chirality 11:173–178, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

8.
Metal-induced sequential transitions among DNA conformations   总被引:1,自引:0,他引:1  
The action of [Co(NH3)6]Cl3 on poly(dGdC) · poly(dGdC) can lead to a series of consecutive reactions, in which B-DNA is first converted to Z-DNA [M. Behe and G. Felsenfeld (1981) Proc. Natl. Acad. Sci USA 78 , 1619–1623], which in turn is transformed into an unidentified conformer that we tentatively call “U”, and finally the highly associated anisotropic ψ-DNA is produced. Conditions are given under which the sequence B ? Z ? “U” ψ(+) can be stopped at any point in the direction from left to right. The reverse processes, from right to left, occur when ψ(+) or “U”-DNA are treated with various amounts of salt concentrations and lowering temperature. Thus it is demonstrated that four conformers of poly(dGdC) · poly(dGdC) are readily interconvertible, and that Z-DNA and “U” conformers are intermediates in the reversible transformations of B- and ψ-DNA.  相似文献   

9.
The enantiomers of 1-methyl-3-(10H-phenothiazine-10-ylmethyl)-1-azoniabicyclo[2,2,2]octane iodide ( 1 ) were prepared by chiral chromatographic resolution of the precursor mequitazine ( 2 ). The (+)-(S)-enantiomer 1b is 10-fold more potent than (?)-(R)-enantiomer 1a as a histamine antagonist, while the two enantiomers show the same antimuscarinic activity in vitro. The absolute configuration of the more active dextrorotatory isomer has been determined by X-ray analysis. Conformational analysis and molecular modeling suggest that the (+)-(S)-enantiomer can adopt a conformation similar to that attributed to the receptor binding conformers of classical antihistamines. © 1994 Wiley-Liss, Inc.  相似文献   

10.
Vincent Madison 《Biopolymers》1977,16(12):2671-2692
A survey of over 50 crystal structures indicates that both imino acid and peptide derivatives of proline populate ring conformers consistent with the torsional potentials about single bonds. In both cases, lower barriers for rotation about C? N bonds relative to those about C? C bonds favor smaller values for dihedral angles about the former bonds. In peptides a minimum in the torsional potential about C? N bonds occurs at zero dihedral angle, further favoring small angles. The pyrrolidine-ring dihedral angles of the proline compounds in the solid state obey a cyclopentane-type pseudorotation function. Thus the puckering of the five-membered ring can be quantitatively described by two parameters. Consistent with small dihedral angles about C? N bonds, Cβ and/or Cγ are puckered out of the mean plane of the ring in nearly all of the nonstrained compounds. Utilizing the consistent force-field method of Lifson and coworkers [see A. Warshel, M. Levitt, and S. Lifson (1970) J. Mol. Spectrosc. 33 , 84] the intramolecular energy of five proline peptides was minimized with respect to all internal coordinates. In addition, the energy surface near minima was explored by constraining a particular dihedral angle and reminimizing the energy with respect to all remaining variables. In linear peptides two types of pyrrolidine-ring conformers have identical predicted energies. In the cyclic dipeptide cyclo (Pro-Gly) one of the ring conformers is favored by about 3 kcal/mol, while the cyclic tripeptide cyclo(Pro-Gly-Gly) favors the other conformer by a comparable margin. In agreement with observations in the solid state and in solution, Cβ and/or Cγ are puckered in the predicted conformers. A correlation between proline Φ and the details of the puckered conformation was predicted and found to match precisely conformers observed in crystals. For the diamides N-acetyl-L -proline-N′-methyl-amide and N-acetyl-L -proline-N′,N′-dimethylamide (AcProMe2A) 30% and 60% cis acetyl peptide bonds were predicted in good agreement with observations in nonpolar solvents for the respective compounds. The conformational distributions with respect to proline Ψ are also in accord with experimental observations. For AcProMe2A, a model for a -Pro-Pro-sequence in a peptide chain, this study is the first to predict stable conformers for proline Ψ either ca. ?50° or 140° for both cis and trans peptides.  相似文献   

11.
Chiral high‐performance liquid chromatography (HPLC) separation of trans‐bis[2‐(2‐pyridyl)aminophenolato] dichlorocyclotriphosphazene 1 was achieved and the absolute configuration of (+)-1 was assigned to be S,S by single‐crystal X‐ray structural analysis. The optically pure 1,2‐diphenyl‐1,2‐ethanediolate derivatives (+)‐ 2a and (?)‐ 2b were synthesized by the reactions of (+)-1 and (-)-1 with (R,R)‐hydrobenzoin, respectively, in refluxing toluene in the presence of an excess amount of triethylamine and a catalytic amount of 4‐(dimethylamino)pyridine. The racemization of the enantiomers of 1 and the epimerization of diastereomers of 2 were not observed in refluxing toluene neither under acidic nor basic conditions. The stereochemistry of (+)-1 was confirmed by the crystal structure of (+)‐ 2a and bis[(4‐methyl‐2‐pyridyl)oxy]cyclotriphosphazene (+)-3 derived from (+)-1 . Chirality 28:556–561, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

12.
The enantioselective bioactivity against pathogens (Cercospora arachidicola, Fulvia fulva, and Phytophthora infestans) and acute toxicity to Daphnia magna of the fungicide myclobutanil enantiomers were studied. The (+)‐enantiomer in an antimicrobial activity test was about 1.79–1.96 times more active than the (–)‐enantiomer. In the toxicity assay, the calculated 24‐h LC50 values of the (–)‐form, rac‐form and (+)‐form were 16.88, 13.17, and 11.91 mg/L, and the 48‐h LC50 values were 10.15, 9.24, and 5.48 mg/L, respectively, showing that (+)‐myclobutanil was more toxic. Meanwhile, the enantioselective metabolism of myclobutanil enantiomers following a single intravenous (i.v.) administration was investigated in rabbits. Total plasma clearance value (CL) of the (+)‐enantiomer was 1.68‐fold higher than its antipode. Significant differences in pharmacokinetics parameters between the two enantiomers indicated that the high bioactive (+)‐enantiomer was preferentially metabolized and eliminated in plasma. Consistent consequences were found in the tissues (liver, brain, heart, kidney, fat, and muscle), resulting in a relative enrichment of the low‐activity (–)‐myclobutanil. These systemic assessments of the stereoisomers of myclobutanil cannot be used only to investigate environmental and biological behavior, but also have human health implications because of the long persistence of triazole fungicide and enantiomeric enrichment in mammals and humans. Chirality 26:784–789, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
Dβ (or D-iso)- and Lβ- (or iso)- aspartyl (Asp) residues are accumulated in aged lens crystallins and amyloid beta (Aβ) proteins, respectively, as a result of spontaneous, nonenzymatic isomerization of normal Lα-Asp. To explore why such uncommon Asp isomers are accumulated, the stability of Lα-, Lβ-, and Dβ-Asp was compared in view of the staggered side-chain conformers. By using cylindrin (KVKVLGD7VIEV) from αB-crystallin and Aβ17-25 (L17VFF20AED23)VG25) containing Asp isomers, the vicinal spin-spin coupling constants of Asp Hα-Hβ1 and Hα-Hβ2 were quantified by high-resolution solution 1H NMR. It was found that the trans conformer was extremely preferred in Dβ-Asp7 side-chain of cylindrin. In Aβ17–25, the side chain of Lβ-Asp23 was likely to adopt trans conformer, while gauche conformers were rather rich in Lα-Asp23. In gauche conformers, the close distance between Asp carboxylate carbon (CCOO-) and backbone nitrogen (N) next to Asp is advantageous to the intramolecular cyclization to form succinimide intermediate, followed by the conversion from α- to β-Asp. The cyclization is limited in the trans conformer because of the long distance between CCOO- and N, to keep Dβ- or Lβ-Asp stable. This would be the reason for the site specificity of Asp isomerization in proteins. The higher population of trans conformer in Asp side chain, the less isomerization of Asp as shown as Asp76 in αA-crystallin. The stability and less reactivity of normal Asp and its isomers are the potential factors to determine whether or not the abnormal accumulation is permitted in aged crystallins and Aβ.  相似文献   

14.
Myclobutanil, (RS)‐2‐(4‐chlorophenyl)‐2‐(1H‐1, 2, 4‐triazol‐1‐ylmethyl)hexanenitrile is a broad‐spectrum systemic triazole fungicide which consists of a pair of enantiomers. The stereoselective degradation of myclobutanil was investigated in rat liver microsomes. The concentrations of myclobutanil enantiomers were determined by high‐performance liquid chromatography (HPLC) with a cellulose‐tris‐(3,5‐dimethyl‐phenylcarbamate)‐based chiral stationary phase (CDMPC‐CSP) under reversed phase condition. The t1/2 of (+)‐myclobutanil is 8.49 min, while the t1/2 of (–)‐myclobutanil is 96.27 min. Such consequences clearly indicated that the degradation of myclobutanil in rat liver microsomes was stereoselective and the degradation rate of (+)‐myclobutanil was much faster than (–)‐myclobutanil. In addition, significant differences between two enantiomers were also observed in enzyme kinetic parameters. The Vmax of (+)‐myclobutanil was about 4‐fold of (–)‐myclobutanil and the CLint of (+)‐myclobutanil was three times as much as (–)‐myclobutanil after incubation in rat liver microsomes. Corresponding consequences may shed light on the environmental and ecological risk assessment for myclobutanil and may improve human health. Chirality 26:51–55, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
The crystal structure of cyclo(Pro-Gly)3 (PG3) complex with LiSCN (C22H30N7O6SLi) has been solved by x-ray diffraction. The crystals belong to the space group R3 in the hexagonal setting with unit cell parameters of a = 12.581(1), c = 29.705(3) Å, V = 4072.0 Å3, Z = 6, Mr = 527.53, Dc = 1.23 g/cm3. The crystal structure was solved by direct methods using the program SHELXS-86 and refined to an R value of 5.3% for 1645 reflections (I > 2σI). There are two conformers in the crystal structure. One conformer has three carbonyls on one side and three on the other side of the peptide plane. The other conformer has all six of the carbonyls on the same side of the peptide plane. Both of these conformers bind independently to a Li ion. Based on the conformers of the Li complex and other reported ion complexes formed by PG3, we propose a model for the transport of ions across the lipid membrane. The features of the model are as follows: (1) PG3 forms a hexameric stack in a lipid bilayer when complexing and transporting metal ions. (2) It undergoes a conformational flipping in order pass the ion along the channel. The energy required for the conformational change involved in the flipping of the PG3 molecule may be provided by the applied potential during ion transport. © 1994 John Wiley & Sons, Inc.  相似文献   

16.
Cyclic tetradepsipeptides, AM-toxin I and II, are the host-specific phytotoxins of Alternaria mali. In order to elucidate conformation-toxicity relationships, we analyzed the 270-MHz proton nmr spectra of AM-toxins and hydrogenated analogs, (D -Ala2)AM-toxin I (toxic) and (L -Ala2)AM-toxin I (not toxic), in (C2H3)2SO. These cyclic tetradepsipeptides do not contain N-substituted amino acid residues, and all the peptide and ester groups have been found to be transoid. Two conformers with very unequal populations have been found for AM-toxin I and II; the Cβ?Cα? C?O conformations of the Dha2 residues are nonplanar S-trans in the major conformer and nonplanar S-cis in the minor conformer. Only one ring conformation has been found for each of (L -Ala2) and (D -Ala2)AM-toxin I. (L -Ala2)AM-toxin I takes a C4-type ring conformation; all the C?O groups and Cα-H bonds are oriented to the same side of the ring. (D -Ala2)AM-toxin I takes a new ring conformation; the side chain and C?O group of the L -Amp1 residue are oriented to the same side of the ring. This new conformation is also found for the major conformers of AM-toxin I and II and thus appears to be required for the toxicity. The ring conformations of Tyr(OCH3)1-bearing analog tetradepsipeptides have been found to be much the same as those of Amp1-bearing depsipeptides. Furthermore, on the basis of the two distinct conformations of (D -Ala2) and (L -Ala2)AM-toxin I, an empirical rule is proposed for the stable ring conformations of cyclic tetra-D ,L -peptides, not containing N-substituted amino acid residues.  相似文献   

17.
(±)-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   

18.
In bifidobacteria, phosphoketolase (PKT) plays a key role in the central hexose fermentation pathway called “bifid shunt.” The three-dimensional structure of PKT from Bifidobacterium longum with co-enzyme thiamine diphosphate (ThDpp) was determined at 2.1 Å resolution by cryo-EM single-particle analysis using 196,147 particles to build up the structural model of a PKT octamer related by D4 symmetry. Although the cryo-EM structure of PKT was almost identical to the X-ray crystal structure previously determined at 2.2 Å resolution, several interesting structural features were observed in the cryo-EM structure. Because this structure was solved at relatively high resolution, it was observed that several amino acid residues adopt multiple conformations. Among them, Q546–D547–H548–N549 (the QN-loop) demonstrate the largest structural change, which seems to be related to the enzymatic function of PKT. The QN-loop is at the entrance to the substrate binding pocket. The minor conformer of the QN-loop is similar to the conformation of the QN-loop in the crystal structure. The major conformer is located further from ThDpp than the minor conformer. Interestingly, the major conformer in the cryo-EM structure of PKT resembles the corresponding loop structure of substrate-bound Escherichia coli transketolase. That is, the minor and major conformers may correspond to “closed” and “open” states for substrate access, respectively. Moreover, because of the high-resolution analysis, many water molecules were observed in the cryo-EM structure of PKT. Structural features of the water molecules in the cryo-EM structure are discussed and compared with water molecules observed in the crystal structure.  相似文献   

19.
  1. In US Pacific Northwest ponderosa pine forests the primary attraction order shown previously for red turpentine beetle, Dendroctonus valens (Coleoptera: Curculionidae: Scolytinae), is (−)-β-pinene+ethanol > (+)-3-carene+ethanol > (+)-α-pinene+ethanol. The monoterpenes are bicyclic C10H16 isomers containing one 6-carbon ring with one double bond. Both pinenes have a 4-carbon second ring and differ only by their endocyclic or exocyclic double bond. The (+)-3-carene second ring has 3-carbons; its double bond is endocyclic like (+)-α-pinene.
  2. Ring system and double bond influences on primary attraction were evaluated by hydrogenating (+)-3-carene and (+)-α-pinene to cis-carane and cis-pinane, respectively. Field test primary attraction strengths were (−)-β-pinene+ethanol > cis-carane+ethanol > cis-pinane+ethanol > ethanol.
  3. In combination with ethanol (i) a double bond is not required in either ring system to attract D. valens, (ii) the cis-carane bicyclic 3, 6-carbon ring system provides stronger beetle attraction than the cis-pinane 4, 6-carbon bicyclic ring system, and likely structural basis for stronger (+)-3-carene attraction over (+)-α-pinene, (iii) adding an exocyclic double bond to the 4, 6-carbon ring system elevates attraction above the 3, 6-carbon ring system with no double bond, and (iv) the 4, 6-carbon ring system is a much stronger attractant with an exocyclic rather than endocyclic double bond.
  相似文献   

20.
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   

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