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 interaction of the isomers of verapamil with sites on the calcium channel and alpha 1-adrenergic receptor has been examined. The inhibitory potency of these enantiomers differ with respect to the agonist. KCl- or clonidine-induced contractions of rabbit aortic rings were inhibited in a stereoselective manner by the enantiomers of verapamil with the (-)-isomer being more potent than the (+)-isomer. Similarly, (-)-verapamil was also more potent at displacing (-)-[N-methyl-3H]desmethoxyverapamil than was the (+)-isomer. In contrast, the inhibition of norepinephrine- or phenylephrine-induced aortic contractions was not stereoselective. Differences in enantiomer potency were also observed in vivo. The ability of clonidine to increase blood pressure in the anesthetized rat was blocked in a stereoselective manner by the verapamil enantiomers, while inhibition of the pressor actions of phenylephrine was not. In summary, for agents that rely heavily on calcium channel function (KCl, clonidine), stereoselective inhibition was observed. Stereoselective inhibition was not observed against high efficacy alpha 1-agonists. This difference in stereochemistry argues that verapamil does not act at the same site when inhibiting clonidine or KCl action when compared with norepinephrine or phenylephrine.  相似文献   

2.
The intravenous (0.5 mg/kg) and oral (5 mg/kg) dose kinetics of verapamil were studied in 6 dogs during steady-state oral verapamil dosing (5 mg/kg every 8 h for 3 days). Racemic verapamil and norverapamil, a metabolite of verapamil, were quantitated in plasma by HPLC-fluorescence detection. The verapamil peaks eluting off the column were collected and rechromatographed on an Ultron-OVM column, which resolved the two verapamil enantiomers. After intravenous administration, the systemic clearance and apparent volume of distribution of (?)-(S)-verapamil were nearly twice that of the (+)-(R)-isomer. There was no difference in the elimination half-lives between the two isomers. After oral administration, the oral clearance of (?)-(S)-verapamil was 20 times that of the (+)-(R)-isomer. The apparent bioavailability of (+)-(R)-verapamil was over 14 times that of (?)-(S)-verapamil. The plasma protein binding of the (+)-(R)-isomer was slightly higher by 5% than (?)-(S)-verapamil; however, this effect was not enough to account for the difference between the apparent volume of distribution of the enantiomers, indicating that the tissue binding of (?)-(S)-verapamil was greater than that of the (+)-(R)-isomer. This data on the disposition of the enantiomers of verapamil in the dog is similar to that reported for man and demonstrates that the dog may be an appropriate animal model for man in future studies on the disposition of the enantiomers of verapamil. © 1993 Wiley-Liss, Inc.  相似文献   

3.
dl-2-Haloacid dehalogenase from Pseudomonas sp. 113 is a unique enzyme because it acts on the chiral carbons of both enantiomers, although its amino acid sequence is similar only to that of d-2-haloacid dehalogenase from Pseudomonas putida AJ1 that specifically acts on (R)-(+)-2-haloalkanoic acids. Furthermore, the catalyzed dehalogenation proceeds without formation of an ester intermediate; instead, a water molecule directly attacks the alpha-carbon of the 2-haloalkanoic acid. We have studied solvent deuterium and chlorine kinetic isotope effects for both stereoisomeric reactants. We have found that chlorine kinetic isotope effects are different: 1.0105 +/- 0.0001 for (S)-(-)-2-chloropropionate and 1.0082 +/- 0.0005 for the (R)-(+)-isomer. Together with solvent deuterium isotope effects on V(max)/K(M), 0.78 +/- 0.09 for (S)-(-)-2-chloropropionate and 0.90 +/- 0.13 for the (R)-(+)-isomer, these values indicate that in the case of the (R)-(+)-reactant another step preceding the dehalogenation is partly rate-limiting. Under the V(max) conditions, the corresponding solvent deuterium isotope effects are 1.48 +/- 0.10 and 0.87 +/- 0.27, respectively. These results indicate that the overall reaction rates are controlled by different steps in the catalysis of (S)-(-)- and (R)-(+)-reactants.  相似文献   

4.
Recently we synthesized a naphthalene analog of medetomidine, 4-[1-(1-naphthyl)ethyl]-1H-imidazole hydrochloride (1), and found it to be highly potent in adrenergic systems. The separation of optical isomers of this naphthalene analog was achieved by using the isomers of tartaric acid. The optical purities of the isomers were determined by HPLC using a chiral column. Using X-ray analysis the (+)-isomer was determined to have the S absolute configuration. It has been reported that the (+)-isomer of medetomidine (2) is the most potent enantiomer on alpha 2-adrenergic receptors. There were both qualitative and quantitative differences in biological activities of the optical isomers of 1 in alpha 1- and alpha 2-adrenergic receptor systems of guinea pig ileum and human platelets. (+)-(S)-1, but not (-)-(R)-1 was a selective agonist of alpha 2-mediated responses in ileum whereas (-)-(R)-1 was more potent than (+)-(S)-1 as an inhibitor of alpha 2-mediated platelet aggregation.  相似文献   

5.
The (−)- and (+)-clausenamide (CLA) enantiomers have different pharmacokinetic effects in animals, but their association with putative stereoselective regulation of P-glycoprotein (P-gp) remains unclear. Using three cells expressing P-gp—Caco-2, KBv and rat brain microvessel endothelial cells(RBMEC), this study investigated the association of CLA enantiomers with P-gp. The results showed that the rhodamine 123 (Rh123) accumulation, an indicator of P-gp activity, in Caco-2, KBv and RBMECs was increased by (−)CLA (1 or 5 μmol/L) at 8.2%–28.5%, but reduced by (+)CLA at 11.7%–25.9%, showing stereoselectivity in their regulation of P-gp activity. Following co-treatment of these cells with each CLA enantiomer and verapamil as a P-gp inhibitor, the (+)-isomer clearly antagonized the inhibitory effects of verapamil on P-gp efflux, whereas the (−)-isomer had slightly synergistic or additive effects. When higher concentrations (5 or 10 μmol/L) of CLA enantiomers were added, the stimulatory effects of the (+)-isomer were converted into inhibitory ones, leading to an enhanced intracellular uptake of Rh123 by 24.5%–58.2%; but (−)-isomer kept its inhibition to P-gp activity, causing 30.0%–63.0% increase in the Rh123 uptake. The biphasic effects of (+)CLA were confirmed by CLA uptake in the Caco-2 cells. (+)CLA at 1 μmol/L had significantly lower intracellular uptake than (−)CLA with a ratio[(−)/(+)] of 2.593, which was decreased to 2.167 and 1.893 after CLA concentrations increased to 2.5 and 5 μmol/L. Besides, in the non-induced KB cells, (+)CLA(5 μmol/L) upregulated P-gp expression at 54.5% relative to vehicle control, and decreased Rh123 accumulation by 28.2%, while (−)CLA(5 μmol/L) downregulated P-gp expression at 15.9% and increased Rh123 accumulation by 18.0%. These results suggested that (−)CLA could be a P-gp inhibitor and (+)CLA could be a modulator with concentration-dependent biphasic effects on P-gp activity, which may result in drug—drug interactions when combined with other P-gp substrate drugs.  相似文献   

6.
Tang K  Yi J  Huang K  Zhang G 《Chirality》2009,21(3):390-395
This article reports a new chiral separation method-biphasic recognition chiral extraction for the separation of mandelic acid enantiomers. Distribution behavior of mandelic acid enantiomers was studied in the extraction system with O,O'-di-benzoyl-(2S,3S)-4-toluoyl-tartaric acid (D-(+)-DTTA) in organic phase and beta-CD derivatives in aqueous phase, and the influence of the types and concentrations of extractants and pH on extraction efficiency was investigated. Hydroxypropyl-beta-cyclodextrin (HP-beta-CD), hydroxyethyl-beta-cyclodextrin (HE-beta-CD), and methyl-beta-cyclodextrin (Me-beta-CD) have stronger recognition abilities for S-mandelic acid than those for R-mandelic acid, among which HP-beta-CD has the strongest ability. D-(+)-DTTA preferentially recognizes R-mandelic acid. pH and the concentrations of extractants have great effects on chiral separation ability. A high enantioseparation efficiency with a maximum enantioselectivity of 1.527 is obtained at pH of 2.7 and the ratio of 2:1 of [D-(+)-DTTA] to [HP-beta-CD]. The obtained results indicate that the biphasic recognition chiral extraction is of stronger chiral separation ability than the monophasic recognition chiral extraction. It may be very helpful to optimize the extraction systems and realize the large-scale production of pure enantiomers.  相似文献   

7.
Intraperitoneally administered R-(?)- and S-(+)- enantiomers of 2,5-dimethoxy-4-bromoamphetamine were evaluated for their ability to induce head-body shake, limb flick and abortive grooming behaviors in cats. The R-(?)-enantiomer was consistently more effective than the S-(+)-isomer in all three behavioral measures. Dose-response relationships were evident for head-body shakes and limb flicks for both enantiomers, but reliable abortive grooming responses appeared only after the higher doses of R-(?)-DOB. Cinanserin and methysergide pretreatments effectively antagonized the induction of head-body shakes and limb flicks by 0.1 mg/kg R-(?)-DOB. In addition, haloperidol pretreatment significantly antagonized the appearance of these behaviors suggesting that dopaminergic as well as serotonergic stimulation is involved in the elicitation of these cat behaviors by R-(?)-DOB.  相似文献   

8.
Zhu CJ  Zhang JT 《Chirality》2009,21(3):402-406
Stereoselective differences in pharmacokinetics between clausenamide (CLA) enantiomers have been found after intravenous and oral administration of each enantiomer to rats. The differences could be associated with protein binding of CLA enantiomers. By equilibrium dialysis methods, the binding of CLA enantiomers to rat plasma protein was investigated. The results showed that mean percentages of (-) and (+)CLA in the bound form were 28.5% and 38.0%, respectively, indicating that the unbound fraction of (-)CLA was higher than that of (+)CLA, which provided an explanation for stereoselective pharmacokinetics of CLA enantiomers in rats. The results also showed that there were species differences in plasma protein binding of (-)-isomer between rats (28.5%) and rabbits (47.2%). Furthermore, effects of plasma protein binding on the distribution of CLA enantiomers to their possible target tissues were observed. The amount of (-)CLA in brain was greater than that of (+)CLA 15 min after administration of each enantiomer to rats. But the results were reverse at 4 h postdose. Further studies in distributional kinetics showed that (-)CLA had a more rapid absorption and distribution to hippocampus, cortex, and cerebellum than (+) CLA. (+)CLA had greater values for T(max), t(1/2) (beta), and AUC(0) (-->infinity), and smaller ones for CL/F and V(d)/F than its antipode. The data indicated that the distribution of (-) and (+)CLA in their target tissues was stereoselective. The stereoselective distribution might be involved in the metabolism and transport of two enantiomers in the central nerve system.  相似文献   

9.
Enantiospecific disposition of pranoprofen in beagle dogs and rats   总被引:1,自引:0,他引:1  
Imai T  Nomura T  Aso M  Otagiri M 《Chirality》2003,15(4):312-317
The pharmacokinetic characteristics of pranoprofen enantiomer were examined and compared with the disposition of the corresponding isomer after the administration of racemic pranoprofen to beagle dogs and rats. The plasma levels of (+)-(S)-isomer were significantly higher than those of (-)-(R)-isomer in dogs and rats by either intravenous or oral administration. Although the oral bioavailability and absorption rate constant between the (-)-(R)- and (+)-(S)-form was the same, the elimination rate constant of the (+)-(S)-form was significantly lower than that of the (-)-(R)-form in both dogs and rats. This discrepancy can be explained on the basis of differences in protein binding and the metabolism of the two enantiomers. The (-)-(R)-isomer was predominantly conjugated depending on its higher free plasma level and its faster metabolic rate than the (+)-(S)-form, and thus was excreted more rapidly in the urine and bile in the form of pranoprofen glucuronide. Furthermore, a (-)-(R)- to (+)-(S)-inversion occurred to the extent of 14% in beagle dogs, but not in rats. This chiral inversion might be an important factor in the slow elimination of the (+)-(S)-form in dogs. The most efficient organ for chiral inversion was the liver, followed by kidney and intestine.  相似文献   

10.
A selective, accurate and reproducible high-performance liquid chromatographic (HPLC) method for the separation of individual enantiomers of DRF 2725 [R(+)-DRF 2725 and S(-)-DRF 2725 or ragaglitazar] was obtained on a chiral HPLC column (Chiralpak). During method optimization, the separation of enantiomers of DRF 2725 was investigated to determine whether mobile phase composition, flow-rate and column temperature could be varied to yield the base line separation of the enantiomers. Following liquid-liquid extraction, separation of enantiomers of DRF 2725 and internal standard (I.S., desmethyl diazepam) was achieved using an amylose based chiral column (Chiralpak AD) with the mobile phase, n-hexane-propanol-ethanol-trifluoro acetic acid (TFA) in the ratio of 89.5:4:6:0.5 (v/v). Baseline separation of DRF 2725 enantiomers and I.S., free from endogenous interferences, was achieved in less than 25 min. The eluate was monitored using an UV detector set at 240 nm. Ratio of peak area of each enantiomer to I.S. was used for quantification of plasma samples. Nominal retention times of R(+)-DRF 2725, S(-)-DRF 2725 and I.S. were 15.8, 17.7 and 22.4 min, respectively. The standard curves for DRF 2725 enantiomers were linear (R(2) > 0.999) in the concentration range 0.3-50 microg/ml for each enantiomer. Absolute recovery, when compared to neat standards, was 70-85% for DRF 2725 enantiomers and 96% for I.S. from rat plasma. The lower limit of quantification (LLOQ) for each enantiomers of DRF 2725 was 0.3 microg/ml. The inter-day precisions were in the range of 1.71-4.60% and 3.77-5.91% for R(+)-DRF 2725, S(-)-DRF 2725, respectively. The intra-day precisions were in the range of 1.06-11.5% and 0.58-12.7% for R(+)-DRF 2725, S(-)-DRF 2725, respectively. Accuracy in the measurement of quality control (QC) samples was in the range 83.4-113% and 83.3-113% for R(+)-DRF 2725, S(-)-DRF 2725, respectively. Both enantiomers and I.S. were stable in the battery of stability studies viz., bench-top (up to 6 h), auto-sampler (up to 12 h) and freeze/thaw cycles (n = 3). Stability of DRF 2725 enantiomers was established for 15 days at -20 degrees C. The application of the assay to a pharmacokinetic study of ragaglitazar [S(-)-DRF 2725] in rats is described. It was unequivocally demonstrated that ragaglitazar does not undergo chiral inversion to its antipode in vivo in rat plasma.  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.
Probenecid-induced changes in the clearance of pranoprofen enantiomers   总被引:1,自引:0,他引:1  
Imai T  Nomura T  Otagiri M 《Chirality》2003,15(4):318-323
Probenecid is known to inhibit the elimination of several acidic drugs. Its influence on the pharmacokinetics of pranoprofen was investigated in rabbit after a single intravenous injection of racemic mixture (5 mg/kg). Levels of (-)-(R)- and (+)-(S)-pranoprofen and their glucuronide (after hydrolysis with sodium hydroxide) were determined in plasma, urine, and several tissues. The plasma concentration of the (+)-(S)-isomer was higher than that of the (-)-(R)-form. Oral coadministered probenecid (100 mg/kg) resulted in an increased plasma concentration of both enantiomers. Probenecid reduced the apparent total clearance and excretion of pranoprofen enantiomers in urine. It had a slight effect on the tissue distribution of pranoprofen at the dose used, but significantly reduced the formation of glucuronide for both enantiomers to the same extent in kidney microsomes. The differences caused by probenecid were significant with respect to its ability to inhibit glucuronidation in the kidney and subsequent excretion into urine, but enantioselective effects were negligible.  相似文献   

13.
Compounds 2a and 3a-e are racemic 2-[(acylamino)ethyl]-1,4-benzodiazepines, tifluadom analogs, with high affinity and selectivity towards the kappa-opioid receptor. We describe the enantiomeric separation of all compounds through liquid chromatography with chiral stationary phases, as well as the resolution of the enantiomers of the most interesting compounds, 2a and 3a, by the semipreparative column Chiralpak AD. The configuration of the resolved enantiomers was investigated: the comparative study of CD and (1)H NMR spectra shows that compounds (-)-2a and (-)-3a have the same absolute configuration of (+)-(S)-tifluadom. A study on the stereoselective interaction with opiate receptors is reported.  相似文献   

14.
1) (R)-2-Hydroxyglutaryl-1-CoA was synthesised starting from (R)-5-oxotetrahydrofuran-2-carboxylic acid (gamma-lactone of (R)-2-hydroxyglutarate) which was converted to the acylchloride and condensed with N-capryloylcysteamine. The lactone ring of the resulting thiolester was opened by acid hydrolysis and the CoA derivative was obtained by transesterification. 2) Pure glutaconate CoA-transferase from Acidaminococcus fermentans catalysed the formation of the 1- and the 5-isomer of (R)-2-hydroxyglutaryl-CoA from acetyl-CoA and (R)-2-hydroxyglutarate. The isomers were separated by HPLC and characterised by their reaction with acetate under the catalysis of the CoA-transferase. V/Km for the 1-isomer was 80 times higher than that for the 5-isomer. 3) Studies with cell-free extracts from A. fermentans showed that only (R)-2-hydroxyglutaryl-1-CoA but not its 5-isomer was dehydrated to glutaconyl-1-CoA. The data indicate that (R)-2-hydroxyglutaryl-5-CoA is an erroneous product of glutaconate CoA-transferase which only occurs in vitro.  相似文献   

15.
An aqueous solution of the (+)-monoethyl ester of N-(l′-hydroxymethyl-)propyl-α-aminobenzylphosphonic acid has been proposed as a suitable chiral eluent for enantiomeric analysis of amino acids by ligand-exchange chromatography. Asymmetric synthesis of the chiral selector using (−)-(R)-2-aminobutan-1-ol as a starting reactant is described. The dependence of the parameters of separation of valine enantiomers on concentration of the complexing ion, pH, and temperature has been investigated. It is shown that the order in which enantiomers are eluted from a column depends on the concentration of the complexing ion and pH. © 1996 Wiley-Liss, Inc.  相似文献   

16.
The enantiomers of 4-tert-butyl-3-isopropyl-2,6,7-trioxa-1-phosphabicyclo[2.2.2 ]octane 1-sulfide (TBIPPS) were prepared in nine steps from diethyl tert-butylmalonate, and their abilities to compete with [3H]1-(4-ethynylphenyl)-4-n-propyl-2,6,7-trioxabicyclo[2.2.2 ]octane (EBOB), a noncompetitive antagonist of ionotropic gamma-aminobutyric acid (GABA) receptors, at their binding site were investigated using rat brain and housefly head membranes. The (S)-(-)-isomer of TBIPPS (IC50 = 398 nM) was more potent than was the (R)-(+)-isomer of TBIPPS (IC50 = 1220 nM) in rat receptors, while the potencies of (S)-TBIPPS 104 nM) and (R)-TBIPPS (IC50 = 94.4 nM) in housefly receptors were almost the same. The different enantiospecificities of rat and housefly receptors indicate that the three-dimensional structure of the binding site might be different between these receptors. In a region of the rat binding site there might be a steric bulk that interacts less favorably with (R)-TBIPPS than with (S)-TBIPPS, while in the corresponding region of the housefly binding site there might not be such a steric bulk that leads to specificity for these compounds.  相似文献   

17.
The role of calcium in interleukin- (IL) 8-, IL-1 alpha- and IL-1 beta-induced lymphocyte migration has been investigated by using the calcium channel antagonists, verapamil, nifedipine, diltiazem (IL-8) and the optical isomers of the dihydropyridine analogue SDZ 202-791 (IL-8, IL-1 alpha and IL-1 beta). Potent inhibition of IL-8-induced migration was observed in response to nifedipine (IC50 = 10 nM), verapamil (IC50 = 60 nM) and diltiazem (IC50 = 10 nM). The (+)-isomer of SDZ 202-791 was without effect on any of the agonists tested, however, the (-)-isomer induced dose-related inhibition of stimulated migration, IC50 values being 0.1 nM, 10 pM and 1.0 nM, for IL-8-, IL-1 alpha- and IL-1 beta-induced migration, respectively. Reversal of the inhibitory effects of the (-)-isomer was obtained in the presence of increasing concentrations of (+)-isomer. The induction of lymphocyte migration by IL-8, IL-1 alpha and IL-1 beta therefore appears to be a process dependent on calcium channel activation.  相似文献   

18.
R(-)-Nipecotic acid was a more potent inhibitor than the S(+)-isomer of the uptake of GABA, (+)-nipecotic acid, and β-alanine in rat brain slices. (-)-Nipecotic acid was an order of magnitude more potent as an inhibitor of GABA uptake than as an inhibitor of β-alanine uptake, whereas the (+)-isomer was less selective. (–)-Nipecotic acid was a weak inhibitor of L-proline uptake and of rat brain acetylcholinesterase activity. Kinetic studies showed that both isomers of nipecotic acid were competitive inhibitors of GABA uptake when added at the same time as GABA, but non-competitive inhibitors when preincubated with the tissue for 15 min before addition of GABA. The apparent slope inhibition constants, which were not influenced by preincubation, indicated that (–)-nipecotic acid has an affinity for the carrier some 5 times higher than that for (+)-nipecotic acid. (–)-Nipecotic acid stimulated the release of preloaded radioactive GABA from rat brain slices. These observations indicate that (–)-nipecotic acid is a substrate-competitive inhibitor of GABA which combines with the GABA carrier and is taken up. (?)-Nipecotic acid and (+)-2,4-diaminobutyric acid, on the basis of their absolute structures and inhibition kinetics, are proposed to interact in a similar way with the GABA transport system.  相似文献   

19.
The new chiral derivatizing agent (CDA), alpha-cyano-alpha-fluoro(2-naphthyl)-acetic acid (2-CFNA) 1 was prepared in optically pure form by chiral HPLC separation of racemic 2-CFNA methyl ester (2-CFNA Me ester) (+/-)-2. The ester was obtained by fluorination of methyl alpha-cyano(2-naphthyl)acetate with FClO3. 2-CFNA 1 has proven to be a significantly superior CDA for determination of enantiomeric excess (ee) of a primary alcohol when compared to alpha-methoxy-alpha-trifluoromethylphenylacetic acid (MTPA, Mosher's agent) and alpha-cyano-alpha-fluoro(p-tolyl)acetic acid (CFTA). The ee of (-)-3-acetoxy-2-fluoro-2-(hexadecyloxymethyl)propan-1-ol (-)-9, a fluorinated analog of anticancer active ether lipids, was determined using (+)-2-CFNA (+)-1.  相似文献   

20.
Racemic 2-(5,6-dichloro-3-indolyl)propionic acid (5,6-Cl2-2-IPA) was synthesized from 5,6-dichloroindole-3-acetic acid (5,6-Cl2-IAA) by successive esterification, methoxycarbonylation, methylation, and double hydrolysis. The racemate was converted to the diastereomeric esters of (S)-(-)-1-phenylethyl alcohol. These were separated by HPLC into two optically active diastereomers and then hydrolyzed with p-TsOH to the optically active enantiomers of 5,6-Cl2-2-IPA. The absolute configurations of both the 5,6-Cl2-2-IPA enantiomers were determined by comparing the 1H-NMR spectra of their diastereomeric (S)-(-)-1-phenylethyl esters with those of the diastereomeric (S)-(-)-1-phenylethyl esters of 2-(3-indolyl)propionic acid (2-IPA) whose absolute configurations are already known. There was no essential difference between (S)-(+)- and (R)-(-)-5,6-Cl2-2-IPA in hypocotyl growth-inhibiting activity toward Chinese cabbage, but their inhibitory activities were stronger than that of the potent mother auxin, 5,6-Cl2-IAA. No essential difference in the coleoptile elongating activity of Avena sativa was apparent for the enantiomers, this activity being about one-third that of 5,6-Cl2-IAA.  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