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.
Seven fungi, which are found to reduce ethyl 3-oxobutanoate in high yields, were tested for their reducing ability for ethyl 2-methyl 3-oxobutanoate. We obtained some interesting findings. In particular, Penicillium purpurogenum reduced ethyl 2-methyl 3-oxobutanoate to the corresponding alcohols with the diastereomer (anti/syn) ratio of 93/7 with the enantiomeric excess of anti-(2S,3S)- and syn-(2S,3R)- hydroxy esters of 90 and >99 ee%, respectively.  相似文献   

2.
Chlorella pyrenoidosa Chick reduced ethyl 2-methyl 3-oxobutanoate to the corresponding alcohols with the diastereomer (anti/syn) ratio of 53/47. The enantiomer excesses of anti-(2S, 3S)- and syn-(2S, 3R)-hydroxy esters were 89 and > 99ee% respectively. C. vulgaris and C. regularis afforded predominantly the syn-isomer, contrary to C. pyrenoidosa. The differences in the activity of reducing ethyl 2-methyl 3-oxobutanoate were observed among three strains of Chlorella. Addition of 2% metal salts slightly increased the chemical yield of the hydroxy ester.  相似文献   

3.
The synthesis of 10 analogues of pepstatin modified so that statine is replaced by 4-amino-3-hydroxy-3,6-dimethylheptanoic acid (Me3Sta) or 4-amino-3-hydroxy-3-methyl-5-phenylpentanoic acid (Me3AHPPA) residues is reported. Both the 3S,4S and 3R,4S diastereomers of each analogue were tested as inhibitors of the aspartic proteases, porcine pepsin, cathepsin D, and penicillopepsin. In all cases the 3R,4S diastereomer (rather than the 3S,4S diastereomer) of the Me3Sta and Me3AHPPA derivatives was found to be the more potent inhibitor of the aspartic protease (Ki = 1.5-10 nM for the best inhibitors), in contrast to the results obtained with statine (Sta) or AHPPA derivatives, where the 3S,4S diastereomer is the more potent inhibitor for each diastereomeric pair of analogues. The Me3Sta- and Me3AHPPA-containing analogues are only about 10-fold less potent than the corresponding statine and AHPPA analogues and 100-1000-fold more potent than the corresponding inhibitors lacking the C-3 hydroxyl group. Difference NMR spectroscopy indicates that the (3R,4S)-Me3Sta derivative induces conformational changes in porcine pepsin comparable to those induced by the binding of pepstatin and that the (3S,4S)-Me3Sta derivatives do not induce the difference NMR spectrum. These results require that the C-3 methylated analogues of statine-containing peptides must inhibit enzymes by a different mechanism than the corresponding statine peptides. It is proposed that pepstatin and (3S)-statine-containing peptides inhibit aspartic proteases by a collected-substrate inhibition mechanism. The enzyme-inhibitor complex is stabilized, relative to pepstatin analogues lacking the C-3 hydroxyl groups, by the favorable entropy derived when enzyme-bound water is returned to bulk solvent.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
5.
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   

6.
Six analogs of tryptophanyl-adenylate, which is an important intermediate in the enzymatic synthesis of Trp-tRNATrp, have been prepared. Four compounds, tryptophanyl-8-bromoadenylate, tryptophanyl-2-chloroadenylate, tryptophanyl-7-deazaadenylate and tryptophanyl-(N6-methyl)adenylate, contain modifications in the nucleobase moiety, while tryptophanyl-2′ deoxyadenylate and tryptophanyl-3′-deoxyadenylate were modified in the carbohydrate part of the molecule. Three of these analogs (2-chloro, 7-deaza, 2′-deoxy analogs) as well as ATP analogs with the same modifications were substrates in the aminoacylation reaction; three analogs (8-bromo, N6-methyl, 3′-deoxy analogs) were inactive as well as the corresponding ATP analogs. In contrast, in the ATPPPi pyrophosphate exchange in the absence of tRNA all ATP analogs except 8-bromo-ATP were substrates. However, the presence of tRNA reduced the number of ATP analogs being substrates to that number of substrates observed in the aminoacylation. Therefore, it can be concluded that the presence of tRNA is responsible for an increase of specificity. The diastereomers of adenosine 5′-O-(3-thiotriphosphate) (ATPαS), adenosine 5′-O-(2-thiotriphosphate) (ATPβS), and adenosine 5′-O-(3-thiotriphosphate) (ATPγS) were tested with various divalent metals as substrates in the pyrophosphate exchange reaction. The Sp diastereomer of ATPαS is a substrate with Mg2+, whereas the Rp diastereomer is inactive. Both diastereomers are inactive in the presence of Zn2+. Since Zn2+ binds preferentially to the sulfur atom, an explanation of these results is that the Mg2+ ion is not bound to the α-phosphate. Only the Sp isomer of the diastereomers of ATPβS acts as substrate in the presence of Mg2+. The stereospecificity becomes reversed in the presence of Zn2+. ATPγS acts as substrate with both Mg2+ and Zn2+. These results suggest that the Δ isomer of the β,γ-bidentate ATP-Mg2+ complex is the substrate for this enzyme. From these results a molecular model of the ATP-Mg2+ complex in the active site can be derived in which the nucleotide is attached to the enzyme by interactions in which the 3′-OH and 6-NH2 group, one oxygen atom of the α-phosphorus atom, and the coordinated magnesium cation are all involved.  相似文献   

7.
JB Parker  CT Walsh 《Biochemistry》2012,51(28):5622-5632
The first four enzymes of the bacilysin antibiotic pathway, BacABGF, convert prephenate to a tetrahydrotyrosine (H(4)Tyr) diastereomer on the way to the anticapsin warhead of the dipeptide antibiotic. BacB takes the BacA product endocyclic-Δ(4),Δ(8)-7R-dihydrohydroxyphenylpyruvate (en-H(2)HPP) and generates a mixture of 3E- and 3Z-olefins of the exocyclic-Δ(3),Δ(5)-dihydrohydroxyphenylpyruvate (ex-H(2)HPP). The NADH-utilizing BacG then catalyzes a conjugate reduction, adding a pro-S hydride equivalent to C(4) to yield tetrahydrohydroxyphenylpyruvate (H(4)HPP), a transamination away (via BacF) from 2S-H(4)Tyr. Incubations of the pathway enzymes in D(2)O yield deuterium incorporation at C(8) from BacA and then C(9) from BacB action. By (1)H NMR analysis of samples of H(4)Tyr, the stereochemistry at C(4), C(8), and C(9) can be assigned. BacG (followed by BacF) converts 3E-ex-H(2)HPP to 2S,4R,7R-H(4)Tyr. The 3Z isomer is instead reduced and transaminated to the opposite diastereomer at C(4), 2S,4S,7R-H(4)Tyr. Given that bacilysin has the 2S,4S stereochemistry in its anticapsin moiety, it is likely that the 2S,4S-H(4)Tyr is the diastereomer "on pathway". NMR determination of the stereochemistry of the CHD samples at C(8) and C(9) allows assignment of all stereogenic centers (except C(3)) in this unusual tetrahydro-aromatic amino acid building block, giving insights into and constraints on the BacA, BacB, and BacG mechanisms.  相似文献   

8.
We have examined the electronic and molecular structure of 3,7-diaminophenothiazin-5-ium dye (thionine) in the electronic ground state and in the lowest excited states. The electronic structure was calculated using a combination of density functional theory and multi-reference configuration interaction (DFT/MRCI). Equilibrium geometries were optimized employing (time-dependent) density functional theory (B3LYP functional) combined with the TZVP basis set. Solvent effects were estimated using the COSMO model and micro-hydration with up to five explicit water molecules. Our calculated electronic energies are in good agreement with experimental data. We find the lowest excited singlet and triplet states at the ground state geometry to be of π→π* (S(1), S(2), T(1), T(2)) and n→π* (S(3), T(3)) character. This order changes when the molecular structure in the electronically excited states is relaxed. Geometry relaxation has almost no effect on the energy of the S(1) and T(1) states (~0.02 eV). The relaxation effects on the energies of S(2) and T(2) are moderate (0.14-0.20 eV). The very small emission energy results in a very low fluorescence rate. While we were not able to locate the energetic minimum of the S(3) state, we found a non-planar minimum for the T(3) state with an energy which is very close to the energy of the S(1) minimum in the gas phase (0.04 eV above). When hydration effects are taken into account, the n→π* states S(3) and T(3) are strongly blueshifted (0.33 and 0.46 eV), while the π→π* states are only slightly affected (<0.06 eV).  相似文献   

9.
Metabolism of (+)-, (-)-, and (+/-)-trans-3,4-dihydroxy-3, 4-dihydrobenzo[c]phenanthrenes by liver microsomes from rats and mice and by a purified monooxygenase system reconstituted with cytochrome P-450c has been examined. Bay-region 3,4-diol 1,2-epoxides are minor metabolites of both enantiomers of the 3,4-dihydrodiol with liver microsomes from 3-methylcholanthrene-treated rats or with the reconstituted system (less than 10% of total metabolites). Microsomes from control and phenobarbital-treated rats and from control mice form higher percentages of these diol epoxides (13-36% of total metabolites). Microsomes from 3-methylcholanthrene-treated rats and cytochrome P-450c in the reconstituted system form exclusively the diol expoxide-1 diastereomer, in which the benzylic hydroxyl group and oxirane oxygen are cis to each other, from the (+)-(3S,4S)-dihydrodiol. The same enzymes selectively form the diol expoxide-2 diastereomer, with its oxirane oxygen and benzylic hydroxyl groups trans to each other, from the (-)-(3R,4R)-dihydrodiol (77% of the total diol epoxides). Liver microsomes from control rats show similar stereoselectivity whereas liver microsomes from phenobarbital-treated rats and from control mice are less stereoselective. Three bis-dihydrodiols and three phenolic dihydrodiols are also formed from the enantiomeric 3,4-dihydrodiols of benzo[c]phenanthrene. A single diastereomer of one of these bis-dihydrodiols with the newly introduced dihydrodiol group at the 7,8-position accounts for 79-88% of the total metabolites of the (-)-(3R,4R)-dihydrodiol formed by liver microsomes from 3-methylcholanthrene-treated rats or by the reconstituted system containing epoxide hydrolase. In contrast, the (+)-(3S,4S)-dihydrodiol is metabolized to two diastereomers of this bis-dihydrodiol, a third bis-dihydrodiol, and two phenolic dihydrodiols.  相似文献   

10.
Dehydrogenation of (2S)-[3-13C]isobutyryl-CoA was carried out in vitro using 2-methyl-branched-chain acyl-CoA dehydrogenase purified from rat liver mitochondria. The product was subsequently hydrated by the addition of bovine crotonase. The resulting 3-hydroxyisobutyric acid was predominantly enriched with 13C at the beta-hydroxy position as determined as a methyl ester using electron ionization-gas chromatography-mass spectroscopy. This finding indicates that isobutyryl-CoA is dehydrogenated stereospecifically at the (pro-2S)-methyl and alpha-methine groups to form methacrylyl-CoA, which is later hydrated with the addition of hydrogen on the same side of the molecule from which it was subtracted to produce 3-hydroxyisobutyryl-CoA.  相似文献   

11.
The preferred conformations of deoxyribo and ribonucleoside 3'-methylphosphonates are analysed by minimizing the conformational energy as a function of all the major parameters including the sugar ring for both the S- and R-isomers. The results show that neither the substitution nor the nature of the diastereomer affects significantly the preferred conformations compared to the naturally occurring nucleoside 3'-phosphates. The preferred range of C3'-O3' bond torsions or the phase angles of pseudorotation (P) of the sugar are unaffected. The chiral substitution on the phosphate always adopts a conformation distal to the secondary C3' carbon atom in the minimum energy conformational state. Further, it introduces certain restrictions on the preferred range of P-O3' torsions depending on the methylphosphonate configuration. Methylphosphonate, especially the S-isomer, renders the normal gauche- range of P-O3' bond torsions responsible for the stacked helical duplexes to be energetically unfavourable besides introducing a high energy barrier between trans and gauche conformations. Therefore it is suggested that duplexes with S-methylphosphonate may favour extended phosphodiester conformations. These factors explain the observed lower melting temperature as well as the downfield shifts in the 31P signals in duplexes containing the S-isomer.  相似文献   

12.
The interdependence of the lateral distribution of molecules which are embedded in a membrane (such as integral membrane proteins) and the shape of a cell with no internal structure (such as phospholipid vesicles or mammalian erythrocytes) has been studied. The coupling of the lateral distribution of the molecules and the cell shape is introduced by considering that the energy of the membrane embedded molecule at a given site of the membrane depends on the curvature of the membrane at that site. Direct interactions between embedded molecules are not considered. A simple expression for the interaction of the membrane embedded molecule with the local membrane curvature is proposed. Starting from this interaction, the consistently related expressions for the free energy and for the distribution function of the embedded molecules are derived. The equilibrium cell shape and the corresponding lateral distribution of the membrane embedded molecules are determined by minimization of the membrane free energy which includes the free energy of the membrane embedded molecules and the membrane elastic energy. The resulting inhomogeneous distribution of the membrane embedded molecules affects the cell shape in a nontrivial manner. In particular, it is shown that the shape corresponding to the absolute energy minimum at given cell volume and membrane area may be elliptically non-axisymmetric, in contrast to the case of a laterally homogeneous membrane where it is axisymmetric.  相似文献   

13.
Utilizing a new method for modeling furanose pseudorotation (D. A. Pearlman and S.-H. Kim, J. Biomol. Struct. Dyn. 3, 85 (1985)) and the empirical multiple correlations between nucleic acid torsion angles we derived in the previous report (D. A. Pearlman and S.-H. Kim, previous paper in this issue), we have made an energetic examination of the entire conformational spaces available to two nucleic acid oligonucleotides: d(ApApApA) and ApApApA. The energies are calculated using a semi-empirical potential function. From the resulting body of data, energy contour map pairs (one for the DNA molecule, one for the RNA structure) have been created for each of the 21 possible torsion angle pairs in a nucleotide repeating unit. Of the 21 pairs, 15 have not been reported previously. The contour plots are different from those made earlier in that for each point in a particular angle-angle plot, the remaining five variable torsion angles are rotated to the values which give a minimum energy at this point. The contour maps are overall quite consistent with the experimental distribution of oligonucleotide data. A number of these maps are of particular interest: delta (C5'-C4'-C3'-O3')-chi (O4'-C1'-N9-C4), where the energetic basis for an approximately linear delta-chi correlation can be seen: zeta (C3'-O3'-P-O5')-delta, in which the experimentally observed linear correlation between zeta and delta in DNA(220 degrees less than zeta less than 280 degrees) is clearly predicted; zeta-epsilon (C4'-C3'-O3'-P), which shows that epsilon increases with decreasing zeta less than 260 degrees; alpha (O3'-P-O5'-C5')-gamma (O5'-C5'-C4'-C3') where a clear linear correlation between these angles is also apparent, consistent with experiment; and several others. For the DNA molecule studied here, the sugar torsion delta is predicted to be the most flexible, while for the RNA molecule, the greatest amount of flexibility is expected to reside in alpha and gamma. Both the DNA and RNA molecules are predicted to be highly polymorphic. Complete energy minimization has been performed on each of the minima found in the energy searches and the results further support this prediction. Possible pathways for B-form to A-form DNA interconversion suggested by the results of this study are discussed. The results of these calculations support use of the new sugar modeling technique and torsion angle correlations in future conformational studies of nucleic acids.  相似文献   

14.
Although the hydrophobic effect is generally considered to be one of the most important forces in stabilizing the folded structure of a globular protein molecule, there is a lack of consensus on the precise magnitude of this effect. The magnitude of the hydrophobic effect is most directly measured by observing the change in stability of a protein molecule when an internal hydrophobic residue is mutated to another of smaller size. Results of such measurements have, however, been confusing because they vary greatly and are generally considerably larger than expected from the transfer free energies of corresponding small molecules. In this article, a thermodynamic argument is presented to show (1) that the variation is mainly due to that in the flexibility of the protein molecule at the site of mutation, (2) that the maximum destabilization occurs when the protein at the site of mutation is rigid, in which case the value of the destabilization is approximately given by the work of cavity formation in water, and (3) that the transfer free energy approximately gives the minimum of the range of variations. The best numerical agreements between the small molecule and the protein systems are obtained when the data from the small molecule system are expressed as the molarity-based standard free energies without other corrections.  相似文献   

15.
The preferred conformations of L-α-dipalmitoyl-lecithin (DPL) have been refined using a steepest descent procedure throughout non-bonded potential energy calculations. The results indicate that energy differences between the conformers is very low and the energy parameters are quite constant around the minimum, suggesting a large degree of flexibility of this molecule.The molecular packing energy calculations have been performed by separating the two hydrocarbon chains from the polar head groups of the molecule. From the energy parameters for the packing of the aliphatic chains it results that for distances between adjacent chains up to 4·4 Å the intermolecular forces allow the maximum degree of freedom. This suggests that hydrocarbon chains do not play the main role in the packing process of DPL molecule. Therefore the energy parameters of the polar segment have been calculated, assuming that the C2 asymmetric carbon atom represents the points of the hexagonal lattice and the rotation centre for each molecule. For the internal symmetry of this segment of the molecule two non-equivalent conformers have been selected over all sets of allowed conformations (the GGG and GGG1 for the α2, α3 and α5 torsion angles). The energy packing calculation has been carried out for two independent sets of data, with and without the electrostatic contributions. In the first case a unique topological situation is allowed with the P-N dipole lying parallel to the lattice plane. In the second case different situations including that with the P-N dipole lying orthogonal to the plane are allowed. These data are discussed in relation to different physical conditions.  相似文献   

16.
The chemical and physical properties of a biosurfactant synthesized by hexadecane-grown Rhodococcus species H13-A are described. The biosurfactant is an anionic glycolipid consisting of 1 major and 10 minor components. The hydrophilic portion of the molecule is trehalose, which is acylated with normal C(10) to C(22) saturated and unsaturated fatty acids, C(35) to C(40) mycolic acids, hexanedioic and dodecanedioic acids, and 10-methyl hexadecanoic and 10-methyl octadecanoic acids. The major glycolipid species was identified as 2,3,4,6,2',3',4',6'-octaacyltrehalose, plus minor glycolipid species of di-, tetra- and hexa-acyltrehalose derivatives. The glycolipid exhibited a critical micelle concentration of 1.5?mg/mL and minimum interfacial tension value of 2?×?10(-2)?mN/m against decane, with a further reduction in interfacial tension to 6?×?10(-5)?mN/m in the presence of the cosurfactant pentanol. The phase behavior of the glycolipid indicates the formation of a surfactant-rich, "middle-phase" microemulsion containing liquid crystals, both of which are associated with surfactant systems having ultralow interfacial tension values. Key words: trehalose lipids, glycolipids, biosurfactants.  相似文献   

17.
Gastric juice was collected from gastric pouches in dogs stimulated with histamine. 15(R)-15-methyl PGE2, methyl ester inhibited gastric secretion in dogs when given orally, but was almost inactive when given intravenously, whereas 15(S)-15-methyl PGE2 methyl ester was active by both routes. When given directly into the small intestine (intrajejunally), the 15(S) was active and the 15(R) was inactive. The 15(R), diluted in acid and administered intrajejunally, became active in inhibiting gastric secretion. When the 15(S) was diluted in acid and administered intrajejunally, it lost half of its activity. When each analog was incubated in an acid medium, each was epimerized to give approximately a 1:1 mixture of both 15(R) and 15(S). Incubation of the 15(R) in pH 3 buffer resulted in only a trace of formation of 15(S). These results explain why the 15(R) is active orally but not intrajejunally. When given orally, the low pH of gastric secretion epimerizes much of the 15(R) into the 15(S),which is active by any route. The degree of acidity of gastric contents may determine whether the 15(R) will exert an antisecretory effect.  相似文献   

18.
Vasil'kov SL 《Biofizika》2003,48(1):27-34
Formulae were obtained for the quantitative analysis of the following parameters of excitons in DNA: 1) the lifetime of electronic excitation; 2) the numbers of exciton runs along the DNA sequence; 3) the energy loss by an exciton for one run; 4) the maximum length of the DNA sequence capable of deactivating an exciton for one run. The maximum and minimum ranges for the constant of electronic excitation migration was determined to meet the requirement of inductive-resonance energy transfer for the case of strong interaction. The constant of exciton energy migration was shown to depend on the activation energy, which is equal to the energy of absorbed quantum. An analytical formula was derived to determine the number of quanta the DNA molecule is able to absorb, depending on its length, without nonlinear effects (without overlapping of spatial areas of electronic excitation). By this formula, DNA sequences consisting of only identical AT and GC nucleotide pairs and aggregate AT + GC (in the ratio 1:1) DNA sequences ranging from 1 up to 10(10) base pairs were analyzed. The results of the analysis suggest that the overlapping of spatial areas of electronic excitation induced by a single ultraviolet quantum occurs in short DNA sequences characteristic of prokaryotes. To achieve the same effects on long DNA sequences specific for eukaryotes, DNA must synchronously absorb a great number of ultraviolet quanta. Based on the above results, the following conclusions were made: 1) disturbances in the normal activity of DNA and RNA polymerases may be due to electromagnetic field, which is caused by the oscillatory relaxation of vibronic levels of nucleotides. The energy enters the vibronic levels of nucleotides from an exciton running along the DNA sequence; 2) the increase in the noncoding DNA sequences in eukaryotes due to evolution is a way of DNA protection from undesirable mutations; 3) prokaryotes must possess a greater potentiality and a higher rate of mutagenesis in comparison with eukaryotes, which is proved by their greater diversity in nature.  相似文献   

19.
Rate constants for binding of five inhibitors of human immunodeficiency virus (HIV) protease were determined by stopped-flow spectrofluorometry. The two isomers of quinoline-2-carbonyl-Asn-Phe psi-[CH(OH)CH2N]Pro-O-t-Bu (R diastereomer = 1R; S diastereomer = 1S) quenched the protein fluorescence of HIV protease and thus provided a spectrofluorometric method to determine their binding rate constants. The dissociation rate constants for acetyl-Thr-Ile-Leu psi(CH2NH)Leu-Gln-Arg-NH2 (2), (carbobenzyloxy)-Phe psi[CH(OH)CH2N]Pro-O-t-Bu (3), and pepstatin were determined by trapping free enzyme with 1R as 2, 3, and pepstatin dissociated from the respective enzyme.inhibitor complex. Association rate constants of 1R, 2, and pepstatin were calculated from the time-dependent inhibition of protease-catalyzed hydrolysis of the fluorescent substrate (2-aminobenzoyl)-Thr-Ile-Nle-Phe(NO2)-Gln-Arg-NH2 (4). The kinetic data for binding of 1S to the protease fit a two-step mechanism. Kd values for these inhibitors were calculated from the rate constants for binding and were similar to the respective steady-state Ki values.  相似文献   

20.
This paper reports the conformation energy (CE) calculations on PGE2, PGE2 alpha and 15-(s)-methyl PGE2 alpha on the basis of empirical potential energy functions for the simultaneous rotations around C7-C8 (psi), C12-C13 (theta) and C14-C15 (phi) bonds. The variation of the minimum conformation energy E for each isoenergy map in the psi theta plane with respect to phi gives two minima around 90 degrees and 240 degrees in PGE2, 60 degrees and 245 degrees in PGF2 alpha, and 60 degrees and 150 degrees in 15-(s)-methyl PGF2 alpha. The latter two forms also have a small dip at 270 degrees. The pattern of allowed low energy conformations of PGF2 alpha and 15-(s)-methyl PGF2 alpha is quite similar and is characterized by the existence of six low energy regions.  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