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.
The stereochemistry and mechanism in the reduction of the C–C double bond of carvone by the cultured cells of Astasia longa, a nonchlorophyllous cell line classified in Euglenales, was studied. The reduction of the C–C double bond of carvone with the cultured cells involved the anti-addition of hydrogen atom from the si face at the -position and the re face at the β-position of carbonyl group. Two different enone reductases were isolated from the cultured cells of A. longa. Both reductases catalyzed stereospecifically the anti-addition of hydrogen atoms from the si face at C-1 and the re face at C-6. However, one of the reductases participated in a hydrogen transfer of the pro-4R hydrogen of NADH to C-6 position of carvone and the other used the pro-4S hydrogen of NADH.  相似文献   

2.
Using a conditioning paradigm, the olfactory sensitivity of six CD-1 mice for the enantiomers of carvone and of limonene as well as for their racemic mixtures was investigated. With all six stimuli, the animals significantly discriminated concentrations 相似文献   

3.
Using DECODER (direction exchange with correlation for orientation distribution evaluation and reconstruction) NMR, we probe the orientations of carbonyl carbons in [1-(13)C]glycine-labeled dragline silk under conditions of varying strain and fiber draw rate. A model-specific reconstruction of the molecular orientation distribution incorporating beta sheets and polyglycine II helices indicates that the structures' alignment along the fiber can be described by a pair of Gaussian distributions with full width at half-maxima of 20 and 68 degrees and approximately 45 and approximately 55% relative contributions to the signal intensity. The alignment along the fiber was found to change appreciably when the drawing tension on the fiber was relaxed in a sample drawn at 4 cm/s while little change was observed in a sample drawn at 2 cm/s. The degree of alignment along the fiber was found to increase with fiber draw rate.  相似文献   

4.
Degradation of, and measurement of isotope ratios in, (?)-carvone that had been biosynthesized in Mentha spicata from 3H- and 14C-labelled geraniol and mevalonate indicate that (a) oxidation of limonene or its biogenetic equivalent to form carvone involves shift of the endocyclic double bond; (b) (+)-limonene and (?)-carvone are biogenetically related and are probably formed on divergent pathways from a common intermediate; and (c) the exocyclic double bond of carvone is not formed regiospecifically. These results enable the mechanisms for the introduction of the carbonyl group and for the formation of the isopropenyl side-chain to be delimited.  相似文献   

5.
The feeding deterrent effect of carvone on the slug Arion lusitanicus was investigated. Carvone, a natural compound from caraway seeds, was incorporated into mulch to reduce its inherent volatility. In a laboratory choice experiment, boxes were filled on one side with carvone‐treated mulch and on the other side with untreated mulch. At carvone concentrations ranging from 0.03–0.75 ml litre?1 mulch, slugs ate significantly more lettuce on the untreated side. In a laboratory based no‐choice experiment, carvone concentrations of 0.25 and 0.75 ml litre?1 mulch significantly reduced slug feeding in comparison with the untreated control. Moreover at the highest concentration of carvone (0.75 ml litre?1 mulch) 50% mortality was recorded over a period of 5 days, indicating a clear molluscicidal effect. Due to its volatility carvone did not decrease plant defoliation by A. lusitanicus when applied directly onto lettuce. Subsequent field evaluation showed 0.75 ml litre?1 mulch to partially reduce slug feeding damage. However, this effect was not sufficient to significantly increase lettuce yield. The incorporation of a higher carvone concentration into mulch is still to be tested to confirm whether carvone‐treated mulch can be recommended as an effective alternative approach to chemical slug control.  相似文献   

6.
Abstract:  The monoterpenoid carvone ( 1 ) has been shown to have strong antifeedant effects on Hylobius spp. However, because of the high volatility of carvone, long-time protection of conifer seedlings in the field using this compound has not been possible. We demonstrate, in several bioassay steps, that less-volatile, heavier analogues retain their pre-ingestive feeding inhibition activity in the large pine weevil, Hylobius abietis (L.) for a longer time. The first step in the evaluation of the biological activity of 12 carvone analogues was a micro-assay, a choice test lasting 4 h. Compounds active at 100 nmol/cm2 were further dose–response tested to give the effective dose needed to inhibit feeding by 50% (ED50). Of the 14 compounds tested, including both carvone enantiomers, seven heavier analogues were active at low doses (had low ED50 values). As expected from their lower vapour pressure compared with carvone, the heavier analogues proved more resistant to evaporation before testing. Thus, whereas the effect of 8-hydroxy- p -menth-en-6-one 4 declined after 2 days, some of the compounds with high molar masses, such as the alkylhydroxymenthenones 6 and 8 , retained a stable activity for 4 days. The retained activity at 4 days was strongly correlated to molecular mass and boiling point. When tested on natural material (host Pinus sylvestris L. twig sections for 48 h), the heavier analogues showed a rather low activity. Probably, the activity of the more volatile compound carvone ( 1 ) is due to a repellent effect (olfactory mode) rather than the deterrent effects (gustatory mode) of the heavier compounds. In agreement with the relatively low activity on twigs in the laboratory, the hydroxymenthenone 4 was not active in the field when tested for 2 months as a 1 : 9 mixture with a polar wax.  相似文献   

7.
A Blume  W Hübner  G Messner 《Biochemistry》1988,27(21):8239-8249
Fourier transform infrared spectroscopy has been used to characterize the carbonyl stretching vibration of DMPC, DMPE, DMPG, and DMPA, all labeled with 13C at the carbonyl group of the sn-2 chain. Due to the vibrational isotope effect, the 13C = O and the 12C = O vibrational bands are separated by ca. 40-43 cm-1. This frequency difference does not change when the labeling is reversed with the 13C = O group at the sn-1 chain. For lipids in organic solvents possible conformational differences between the sn-1 and sn-2 ester groups have no effect on the vibrational frequency of the C = O groups. In aqueous dispersion unlabeled phospholipids always show a superposition of two bands for the C = O vibration located at ca. 1740 and 1727 cm-1. These two bands have previously been assigned to the sn-1 and sn-2 C = O groups. FT-IR spectra of 13C-labeled phospholipids show that the vibrational bands of both, the sn-1 as well as the sn-2 C = O group, are clearly superpositions of at least two underlying components of different frequency and intensity. Band frequencies were determined by Fourier self-deconvolution and second-derivative spectroscopy. The difference between the component bands is ca. 11-17 cm-1. Again, the conformational effect as shown by reversed labeling is negligible with only 1-2 cm-1. The splitting of the C = O vibrational bands in H2O and D2O is caused by hydrogen bonding of water molecules to both C = O groups as shown by a comparison with spectra of model ester compounds in different solvents. To extract quantitative information about changes in hydration, band profiles were stimulated with Gaussian-Lorentzian functions. The chemical nature of the head group and its electronic charge have distinctive effects on the extent of hydration of the carbonyl groups. In the gel and liquid-crystalline phase of DMPC the sn-2 C = O group is more hydrated than the sn-1 C = O. This is accord with the conformation determined by X-ray analysis. In DMPG the sn-1 C = O group seems to be more accessible to water, indicating a different conformation of the glycerol backbone.  相似文献   

8.
Aplysiatoxin (ATX) is a naturally occurring tumor promoter isolated from a sea hare and cyanobacteria. ATX binds to, and activates, protein kinase C (PKC) isozymes and shows anti-proliferative activity against human cancer cell lines. Recently, ATX has attracted attention as a lead compound for the development of novel anticancer drugs. In order to predict the binding mode between ATX and protein kinase Cδ (PKCδ) C1B domain, we carried out molecular docking simulation, atomistic molecular dynamics simulation in phospholipid membrane environment, and structure–activity study on a simple acyclic analog of ATX. These studies provided the binding model where the carbonyl group at position 27, the hydroxyl group at position 30, and the phenolic hydroxyl group at position 20 of ATX were involved in intermolecular hydrogen bonding with the PKCδ C1B domain, which would be useful for the rational design of ATX derivatives as anticancer lead compounds.  相似文献   

9.
Fourier transform infrared spectroscopy was used to characterize the lamellar phases of 1,3-dipalmitoyl-sn-glycero-2-phosphocholine (1,3-DPPC), a positional isomer of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (1,2-DPPC). The molecule exists in three distinct phases over the temperature interval 0-70 degrees C. In the low-temperature (LC) phase, the spectra are indicative of acyl chains packed in an orthorhombic subcell, while the carbonyl groups and phosphate ester at the head group show evidence of only partial hydration. The transition from the low-temperature (LC) phase to the intermediate-temperature (L beta) phase at 25 degrees C corresponds to a temperature-induced head-group hydration in which the hydration of the phosphate and carbonyl ester groups results in the reorganization of the hydrocarbon chain-packing subcell from orthorhombic to hexagonal. The transition from the intermediate (L beta) to the high-temperature (L alpha) phase at 37 degrees C is a gel-to-liquid-crystalline phase transition analogous to the 41.5 degrees C transition of 1,2-DPPC. The spectra of the acyl-chain carbonyl groups show evidence of significant differences in molecular conformation at the carbonyl esters in the LC phase. In the L beta and L alpha phases, the carbonyl band contour becomes much more symmetric. However, two components are clearly present in the spectra indicating that the sn-1 and sn-3 carbonyls experience slightly different environments. The observed differences are likely due to a preferred conformation of the phosphocholine group relative to the glycerol backbone. Indications from the infrared spectra of differences in the structure of the C = O groups provide a possible explanation for the selection of the sn-1 chain of 1,3-DPPC by phospholipase A2 on the basis of a preferred head group conformation.  相似文献   

10.
A novel molecular modeling study, involving inhibitors bound to the iron of cytochrome P450 heme, is described for nonsteroidal inhibitors of aromatase (AR). Study of compounds such as aminoglutethimide (AG) suggests that it utilizes hydrogen bonding group(s) at the active site which would usually H-bond to the steroid C(17) carbonyl group. Interaction between AG's carbonyl groups and the area of the active site corresponding to the substrate C(3)==O group is not possible due to steric interaction. Possible reasons for the difference in activity of enantiomers of alternative inhibitors is also suggested, as well as the mode of action of the new AR inhibitor, Arimidex-whose inhibitory activity previously has not been rationalized. The present study proposes that it is able to use hydrogen bonding groups at the active site corresponding to the steroid C(17)==O and C(3)==O area, contradicting a previous study where it is postulated that azole-type compounds only use polar groups at the active site corresponding to the steroid D ring. Using the hypotheses of the modeling study, we designed and synthesized a number of novel (enantiomerically pure) inhibitors, which upon biochemical evaluation were found to be good inhibitors; the N-nonyl derivative of the S-enantiomer was found to possess 39% inhibition at 100 microM inhibitor concentration (using androstenedione as the substrate), under similar conditions, and AG possessed 20% inhibition.  相似文献   

11.
Oriented multilayers of 1-myristoyl-2(1-13C)-myristoyl-sn-glycero-3-phosphatidylcholine (2[1-13C]DMPC) and 1-palmitoyl-2(1-13C)-palmitoyl-sn-glycero-3-phosphatidylcholine (2[1-13C]DPPC) were investigated by use of attenuated total reflection infrared spectroscopy with polarized light. Experiments were performed with the aim to determine the orientation of the two ester groups in these phospholipids in the solid state and in the hydrated state at temperatures below and above the respective gel to liquid-crystalline phase transitions. Substitution of the naturally occurring 12C carbonyl carbon atom by 13C in the ester group of the sn-2 chain of DMPC and DPPC shifts the infrared absorption of the carbonyl double bond stretching vibration to lower frequency. This results in two well-resolved ester C=O bands which can be assigned unequivocally to the sn-1 and sn-2 chains as they are separated by more than 40 cm-1. The two ester CO-O single bond stretching vibrations of the molecular fragments-CH2CO-OC-are also affected and the corresponding infrared absorption band shifts by 20 cm-1 on 13C-labeling of the carbonyl carbon atom. From the dichroic ratios of the individual ester bands in 2(1-13C)DMPC and 2(1-13C)DPPC we were able to demonstrate that the sn-1 and sn-2 ester C=O groups are similarly oriented with respect to the bilayer plane, with an angle greater than or equal to 60 degrees relative to the bilayer normal. The two CO-O single bonds on the other hand have very different orientations. The CH2CO-OC fragment of the sn-1 chain is oriented along the direction of the all-trans methylene chain, whereas the same molecular segment of the sn-2 carbon chain is directed toward the bilayer plane. This orientation of the ester groups is retained in the liquid-crystalline phase. The tilt angle of the hydrocarbon all-trans chains, relative to the membrane normal, is 25 degrees in the solid state of DMPC and DPPC multibilayers. In the hydrated gel state this angle varies between 26 degrees and 30 degrees, depending on temperature. Neither the orientation of the phosphate group, nor that of the choline group varies significantly in the different physical states of these phospholipids.  相似文献   

12.
Carvone has previously been found to highly inhibit its own production at concentrations above 50 mM during conversion of a diastereomeric mixture of (−)-carveol by whole cells of Rhodococcus erythropolis. Adaptation of the cells to the presence of increasing concentrations of carveol and carvone in n-dodecane prior to biotransformation proved successful in overcoming carvone inhibition. By adapting R. erythropolis cells for 197 h, an 8.3-fold increase in carvone production rate compared to non-adapted cells was achieved in an air-driven column reactor. After an incubation period of 268 h, a final carvone concentration of 1.03 M could be attained, together with high productivity [0.19 mg carvone h−1 (ml organic phase)−1] and high yield (0.96 g carvone g carveol−1).  相似文献   

13.
A porphyrin-based photoexcited system has been revealed to be an efficient catalyst for d-limonene biotransformation under mild conditions and using molecular oxygen or/and H2O2 as oxidants. The influence of the oxidant, the wavelength of visible light, and the photoexcitation time on the catalytic system were studied for limonene oxidation with 5,10,15,20-tetraphenylporphyrin (H2TPP) as a catalyst. This porphyrin-catalyzed oxidation of limonene to three main products identified as carvone, an unknown product with a verbenone-like mass spectrum (1), and a (1S,4R)-p-mentha-2,8-diene 1-hydroperoxide (2). The highest conversion yield of these products was achieved at a very high molar ratio of H2TPP to limonene. The dependence of the biotransformation yield on the kind of solvent with different acceptor/donor electron properties was also investigated. Ethyl alcohol proved to be the best among the considered additives used for the reaction. Limonene photooxidation was not significantly dependent on wavelengths of visible light. It was concluded by UV-vis experiments that the reaction proceeds via a free-radical or/and molecular mechanism. Additional evidence for its radical nature was obtained from reactivity investigations. Maximal yield of carvone was obtained in the medium containing 90% of the substrate, within the period of 18 to 36 h of exposition to sunlight.  相似文献   

14.
The rates of thermal transformation of organic molecules containing carbon, hydrogen, and oxygen were systematically examined in order to identify the kinetic constraints that governed origin-of-life organic chemistry under mild aqueous conditions. Arrhenius plots of the kinetic data were used to estimate the reaction of half-lifes at 50 degrees C. This survey showed that hydrocarbons and organic substances containing a single oxygenated group were kinetically the most stable; whereas organic substances containing two oxygenated groups in which one group was an alpha- or beta-positioned carbonyl group were the most reactive. Compounds with an alpha- or beta-positioned carbonyl group (aldehyde or ketone) had rates of reaction that were up to 10(24)-times faster than rates of similar molecules lacking the carbonyl group. This survey of organic reactivity, together with estimates of the molecular containment properties of lipid vesicles and liquid spherules, indicates that an origins process in a small domain that used C,H,O-intermediates had to be catalytic and use the most reactive organic molecules to prevent escape of its reaction intermediates.  相似文献   

15.
Yang J  Hao Q  Liu X  Ba C  Cao A 《Biomacromolecules》2004,5(1):209-218
This study presents chemical synthesis, structural, and physical characterization of novel biodegradable aliphatic poly(butylene succinate-co-cyclic carbonate)s P(BS-co-CC) bearing functional carbonate building blocks. First, five kinds of six-membered cyclic carbonate monomers, namely, trimethylene carbonate (TMC), 1-methyl-1,3-trimethylene carbonate (MTMC), 2,2-dimethyl-1,3-trimethylene carbonate (DMTMC), 5-benzyloxytrimethylene carbonate (BTMC), and 5-ethyl-5-benzyloxymethyl trimethylene carbonate (EBTMC), were well prepared from ethyl chloroformate and corresponding diols at 0 degrees C in THF solution with our modified synthetic strategies. Then, a series of new P(BS-co-CC)s were synthesized at 210 degrees C through a simple combination of poly-condensation and ring-opening-polymerization (ROP) of hydroxyl capped PBS macromers and the prepared carbonate monomers, and titanium tetra-isopropoxide Ti(i-OPr)4 was used as a more suitable catalyst of 5 candidate catalysts which could concurrently catalyze poly-condensation and ROP. By means of NMR, GPC, FTIR, and thermal analytical instruments, macromolecular structures and physical properties have been characterized for these aliphatic poly(ester carbonate)s. The experimental results indicated that novel biodegradable P(BS-co-CC)s were successfully synthesized with number average molecular weight Mn ranging from 24.3 to 99.6 KDa and various CC molar contents without any detectable decarboxylation and that the more bulky side group was attached to a cyclic carbonate monomer, the lower reactivity for its copolymerization would be observed. The occurrences of 13C NMR signal splitting of succinyl carbonyl attributed to the BS building blocks could be proposed due to the randomized sequences of BS and CC building blocks. FTIR characterization indicated two distinct absorption bands at 1716 and 1733 approximately 1735 cm(-1), respectively, stemming from carbonyl stretching modes for corresponding BS and CC units. With regard to their thermal properties, it is seen that the synthesized P(BS-co-CC)s exhibited thermal degradation temperatures 10 approximately 20 degrees C higher than that of PBS. On the basis of the synthesized P(BS-co-BTMC)s, new aliphatic poly(butylene succinate-co-5-hydroxy trimethylene carbonate)s were further synthesized, bearing hydrophilic hydroxyl pendant functional groups through an optimized Pd/C catalyzed hydrogenation. These semi-crystalline new biodegradable aliphatic copolymers with tunable physical properties and functional carbonate building blocks might be expected as potential new biomaterials.  相似文献   

16.
Nuclear magnetic resonance has been used to study the binding of [13C]carbamyl-P (90% enriched) to the catalytic subunit of Escherichia coli aspartate transcarbamylase. Upon forming a binary complex, there is a small change in the chemical shift of the carbonyl carbon resonance, 2 Hz upfield at pH 7.0, indicating that the environments of the carbonyl group in the active site and in water are similar. When succinate, an analog of L-aspartate, is added to form a ternary complex, there is a large downfield change in the chemical shift for carbamyl-P, consistent with interaction between the carbonyl group and a proton donor of the enzyme. The change might also be caused by a ring current froma nearby aromatic amino acid residue. From the pH dependence of this downfield change and from the effects of L-aspartate analogs other than succinate, the form of the enzyme involved is proposed to be an isomerized ternary complex, previously observed in temperature jump and proton NMR studies. The downfield change to chemical shift for carbamyl-P bound to the isomerized complex is 17.7 +/- 1.0 Hz. Using this value, the relative ability of other four-carbon dicarboxylic acids to form isomerized ternary complexes with the enzyme and carbamyl-P has been evaluated quantitatively. The 13C peak for the transition state analog N-(phosphonacetyl)-L-aspartate (PALA), 90% enriched specifically at the amide carbonyl group, is shifted 20 Hz downfield of the peak for free PALA upon binding to the catalytic subunit at pH 7.0. In contrast, the peak for [1-13C] phosphonaceatmide shifts upfield by about 6 Hz upon binding. Since PALA induces isomerization of the enzyme and phosphonacetamide does not, these data provide further evidence consistent with protonation of the carbonyl group only upon isomerization. The degrees of protonation is strong acids of the carbonyl groups of PALA, phosphonacetamide and urethan (a model for the labile carbamyl-P) have been determined, as have the chemical shifts for these compounds upon full protonation. From these data it is calculated that the amide carbonyl groups of carbamyl-P and PALA might be protonated to a maximum of about 20% in the isomerized complexes at pH 7.0. The change in conformation of the enzyme-carbamyl-P complex upon binding L-aspartate, previously proposed to aid catalysis by compressing the two substrates together in the active site, may be accompanied by polarization of the C=O bond, making this ordinarily unreactive group a much better electrophile. A keto analog of PALA, 4,5-dicarboxy-2-ketopentyl phosphonate, also binds tightly to the catalytic subunit and induces a very similar conformational change, whereas an alcohol analog, 4,5-dicarboxy-2-hydroxypentyl phosphonate, does not bind tightly, indicating the critical importance of an unhindered carbonyl group with trigonal geometry.  相似文献   

17.
T B Woolf  B Roux 《Biophysical journal》1997,72(5):1930-1945
The location of the main binding site for sodium in the gramicidin A (GA) channel was investigated with molecular dynamics simulations, using an atomic model of the channel embedded in a fully hydrated dimyristoyl phosphatidycholine (DMPC) bilayer. Twenty-four separate simulations in which a sodium was restrained at different locations along the channel axis were generated. The results are compared with carbonyl 13C chemical shift anisotropy solid-state NMR experimental data previously obtained with oriented GA:DMPC samples. Predictions are made for other solid-state NMR properties that could be observed experimentally. The combined information from experiment and simulation strongly suggests that the main binding sites for sodium are near the channel's mouth, approximately 9.2 A from the center of the dimer channel. The 13C chemical shift anisotropy of Leu10 is the most affected by the presence of a sodium ion in the binding site. In the binding site, the sodium ion is lying off-axis, making contact with two carbonyl oxygens and two single-file water molecules. The main channel ligand is provided by the carbonyl group of the Leu10-Trp11 peptide linkage, which exhibits the largest deviation from the ion-free channel structure. Transient contacts with the carbonyl group of Val8 and Trp15 are also present. The influence of the tryptophan side chains on the channel conductance is examined based on the current information about the binding site.  相似文献   

18.
In this study we examined the effect of polychlorinated biphenyls (PCBs) on biomass production of a PCB-degrading Pseudomonas stutzeri, and on the fatty acid profile of its major membrane lipids. Growth based on biomass weight was stimulated when PCBs were added at the time of inoculation, but PCB addition three days after inoculation led to a significant decrease in biomass. Simultaneous addition of PCBs plus biphenyl or PCBs plus carvone negatively affected P. stutzeri biomass (addition of biphenyl or carvone at the time of inoculation and PCBs to three-day-old culture). In the presence of PCBs alone the amount of the prevalent fatty acids C16:0 and C17-cyclopropyl fatty acid (C17-CP) of P. stutzeri in total and neutral lipids was significantly reduced. When PCBs were added together with carvone (carvone at the time of inoculation and PCBs after three days) a significant reduction of these fatty acids was obtained, but, in addition, oleic, cis-vaccenic, and cyclononadecanic (C19-CP) acids were increased. When PCBs were combined to biphenyl the prevalent fatty acids were reduced and oleic, cis-vaccenic, and cyclononadecanic acids were increased in total and neutral lipids. Addition of 3-chlorobenzoic acid led to a significant growth inhibition and to the production of oleic and cis-vaccenic acids in the membrane fraction phosphatidylcholine.  相似文献   

19.
P O Quist 《Biophysical journal》1998,75(5):2478-2488
The natural-abundance 13C NMR spectrum of gramicidin A in a lipid membrane was acquired under magic-angle spinning conditions. With fast sample spinning (15 kHz) at approximately 65 degrees C the peaks from several of the aliphatic, beta-, alpha-, aromatic, and carbonyl carbons in the peptide could be resolved. The resolution in the 13C spectrum was superior that observed with 1H NMR under similar conditions. The 13C linewidths were in the range 30-100 Hz, except for the alpha- and beta-carbons, the widths of which were approximately 350 Hz. The beta-sheet-like local structure of gramicidin A was observed as an upfield shift of the gramicidin alpha and carbonyl resonances. Under slow sample spinning (500 Hz), the intensity of the spinning sidebands from 13C in the backbone carbonyls was used to determine the residual chemical shift tensor. As expected, the elements of the residual chemical shift tensor were consistent with the single-stranded, right-handed beta6.3 helix structure proposed for gramicidin A in lipid membranes.  相似文献   

20.
A new, semirigid, nicotinic agonist ( ± )-octahydro-2-methyl-trans-5 (1H)-isoquinolone methiodide was synthesized. The disposition of this agonist's nitrogen and carbonyl group conforms well to the prevailing notion of a pharmacophore for the nicotinic receptor. Comparing its structure and electrostatic potential surfaces, we predicted that its activity would be similar to that of carbamylcholine at the frog neuromuscular junction. Instead, the potency of the isoquinolone was only 0.015 times as potent as (+) carbamylcholine. We conclude, after eliminating other possibilities, that the vicinity of the carbonyl group of an agonist must be planar to fit a confined space within the receptor's recognition site. The isoquinolone is a weak agonist because its methylene group β to the carbonyl intrudes on this space.  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