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.
Some proline-containing tripeptides with the general formulas R0CO-L -Pro-X-NHR3 (X = Gly,Sar,L -Ala,D -Ala) and R0CO-X-L -Pro-NHR3 (X = Gly,L -Ala,D -Ala) have been investigated in solution by ir and 1H-nmr spectroscopies. Their favored conformational states depend mainly on both the primary structure and the chiral sequence of the molecules. In inert solvents the βII-folding mode is the most favored conformation for the L -Pro-D -Ala and L -Pro-Gly tripeptides, while the βII′-turn is largely preferred by D -Ala-L -Pro derivatives. Under the same conditions only about one-third of the whole conformers of L -Pro-L -Ala molecules adopts the βI-folding mode. Semiopened C7C5 and C5C7 conformations are appreciably populated in the L -Pro-L -Ala sequence, on the one hand, and in the Gly-L -Pro and L -Ala-L -Pro derivatives, on the other hand. In L -Pro-Sar and X-L -Pro models, the cistrans isomerism around the middle tertiary amide function is observed. Thus cis L -Pro-Sar and L -Ala-L -Pro conformers are folded by an intramolecular i + 3 → i hydrogen bond, whereas cis D -Ala-L -Pro and Gly-L -Pro molecules accommodate an open conformation. In dimethylsulfoxide the βII- and βII′-folding modes are not essentially destabilized, as contrasted with the βI conformation, which is less populated. In water solution all the above-mentioned conformations, with the possible exception of the βII′-folding mode for D -Ala-L -Pro molecules, seem to vanish. Solute conformations are also compared with the crystal structures of four proline-containing tripeptides.  相似文献   

2.
Cyclo(L -Pro-Sar)n (n = 2–4) with moderate flexibility and hydrophobicity of molecular structure was synthesized, and the characteristics of these cyclic peptides and their metal complexes in acetonitrile were investigated in connection with the residual properties using 13C-nmr measurements. The cyclic tetrapeptide cyclo(L -Pro-Sar)2 showed a sterically hindered phenomenon in acetonitrile in which the amide backbone adopted a cis-trans-cis-trans sequence. The cyclic hexapeptide cyclo(L -Pro-Sar)3 existed as a mixture of several conformers whose interconversion is slow on the nmr time scale, including cis-cis-trans and/or cis-trans-trans arrangement of the Sar-Pro bond. Finally, it was demonstrated that the cyclic octapeptide cyclo(L -Pro-Sar)4 behaved as a mixture of multiple conformers which allowed for cis-trans isomerism about the Pro-Sar peptide bond, of which 20–30% had the all-cis Sar-Pro bond isomer and the remaining 70–80% had one (or more) cis Sar-Pro bond isomer. 13C-nmr spectra also demonstrated that cyclo(L -Pro-Sar)n (n = 3,4) formed a 1:1 ion complex whose conformation was characterized by an all-trans peptide bond in the presence of excess metal salt. Cation binding studies, using CD measurements, established that the ion selectivity of cyclo(L -Pro-Sar)4 in acetonitrile decreased in the order, Ba2+ > Ca2+ > Na+ > Mg2+ > Li+.  相似文献   

3.
Five N-protected tetrapeptide esters of the structure Gly-Pro-X-X*-O-methyl were synthesized in such a way that one of the two variable amino acid residues (X) was isotopically enriched in 15N (denoted by*). The variable amino acids are glycine, alanine, leucine, valine, and phenylalanine. For the natural abundance 15N-nmr spectra of these tetrapeptide derivatives in methylene chloride only the signals of the Gly-Pro trans isomer were found. In a 2:1 mixture of acetone and dimethylsulfoxide, signals for both the cis and trans isomers were observed. Three of the five tetrapeptide derivatives show cis/trans splitting of all four nitrogen signals. The 15N-nmr spectra of Z-Pro-Pro-OH and of (D ,L -proline)n were measured in a 2:1 mixture of acetone and dimethylsulfoxide as well as in water. The effects of solvents and neighboring residues and the influence of the cis/trans isomerism on the nmr spectra are discussed. The determination of the cis/trans equilibria and the assignment of the 15N-nmr signals of all oligopeptides were achieved by selective isotopic enrichment and by means of 13C-nmr spectra.  相似文献   

4.
1H-Nmr was used to measure the rate of cistrans interconversion of X-Pro bonds in linear and cyclic oligopeptides. k(cis → trans) = 2.5 × 10?3 s?1 at 25°C was found for the zwitterionic form of H-Ala-Pro-OH, in good agreement with earlier measurements. Replacement of Ala by Phe, Tyr, or Trp resulted in a 10-fold slower interconversion rate, whereas after substitution of Ala by His or Glu, the rate decreased only slightly. Independent of the residues X, the interconversion rate was increased by a factor of ca. 20 when the peptide chain was elongated by addition of Ala to the C-terminal Pro. An additional increase by a factor of 6 was observed when going from the protected linear peptide CF3CO-Gly-Gly-Pro-Ala-OCH3 to the closely related cyclic compound c[-Gly-Gly-Pro-Gly-Ala-]. These data are evaluated with regard to their possible use in future studies on the role of X-Pro cistrans isomerization in the kinetics of protein folding.  相似文献   

5.
Empirical conformational energy calculations with the use of ECEPP energy functions have been carried out for linear dipeptides H-X-L -Pro-OH, with X = Gly, L -Ala, D -Ala, L -Leu, D -Leu, L -Phe, and D -Phe, in different states of protonation of the end groups. The results of these calculations are compared with the previously reported experimental equilibrium populations for the cis and trans isomers of the X-Pro bond in the different species. For all the protonation states of the seven dipeptides, the calculated nonbonded interactions and the conformational entropy term lead to a preference of the trans forms over the cis isomers by at least 1 kcal/mol. The electrostatic interactions stabilize the cis conformations in all species except the cationic forms of the D ,L -peptides, and it could further be shown that only the carbonyl group of X and the two end groups contribute significantly to the total electrostatic energy. One of the principal results of the experimental studies, i.e., the occurrence of 5–15% cis-proline in all the peptides with an uncharged C-terminus, was corroborated by our investigation of the cationic species. A detailed assessment of the electrostatic contribution to the total energy of the different conformations of H-Gly-L -Pro-OH indicates that the standard ECEPP parameters tend to overestimate the electrostatic interactions in aqueous solutions of the X-Pro dipeptides.  相似文献   

6.
The crystal structures and molecular conformations of two tetraproline derivatives with alternating configurations Boc(D -Pro,L -Pro)2OH and Boc(D -Pro,L -Pro)2OCH3 are investigated in connection with the ability of the homologous polymer to selectively increase (as an ion channel) the ion permeability across bilayer membranes. Both tetramers are characterized by the cis-trans alternating conformation of the peptide bonds, which formally transforms in a turn of the poly-D ,L -proline channel after a cis-trans change of the central peptide residue.  相似文献   

7.
Combinations of L - and D -proline residues are useful compounds for finding new structures and properties of cyclic peptides. This is demonstrated with one striking example, the cyclic tetrapeptide c(D -Pro-L -Pro-D -Pro-L -Pro). For this molecule composed of strictly alternating D - and L -configurated residues, a highly symmetrical structure is expected, which should be an optically inactive meso-form. Cyclization of the enantiomeric pure linear precursor D -Pro-L -Pro-D -Pro-L -Pro, however, yields a racemic mixture of two enantiomeric cyclotetrapeptides, both with twofold symmetry and a cistranscistrans sequence of the peptide bonds. Remarkably, this formation of a racemate was not caused by racemization, but by cis/trans isomerization of all peptide bonds in the ring. This process may occur in the linear precursor during the ring formation (cyclization of conformers with transcistrans or cistranscis arrangement of the amide bonds) as well as in the enantiomeric pure cyclic tetrapeptide at higher temperature. In the latter case, an all-cis structure should exist as the intermediate, which can form a cistranscistrans sequence in two equivalent ways, leading finally to two enantiomeric cyclotetrapeptides. In the first one, the cis peptide bonds are attributed to the L -residues and the trans peptide bonds to the D -residues; in the second one, the cis bonds belong to the D and the trans bonds to the L -residues. The mixture of these two enantiomers does not crystallize in the racemic form, but in enantiomeric pure separate crystals. The structural properties could be proved by 1H- and 13C-nmr spectroscopy and x-ray analysis. The cis/trans isomerization process was confirmed by optical rotation measurements and CD spectroscopy, as well as DREIDING model studies. Calorimetric measurements in the solid state suggest the existence of the expected all-cis intermediate. The backbone conformation of the 12-membered medium-sized ring shows only slight deviations—up to 6° —from the planarity of the peptide bonds. On the other hand, the four pyrrolidine rings show different types of puckering of the Cγ or the Cβ atoms.  相似文献   

8.
The conformations of the phytotoxic cyclic tetrapeptide tentoxin [cyclo-(L -MeAla1-L -Leu2-MePhe[(Z)Δ]3-Gly4 )] have been studied in aqueous solution by two-dimensional proton nmr at various temperatures. Contrary to what is observed in chloroform, tentoxin exhibits multiple exchanging conformations in water. Aggregation phenomena were also observed. Four conformations with different proportions (51, 37, 8, and 4%) were observed at ?5°C. Models were constructed from nmr parameters and restrained molecular dynamics simulations. All the models exhibit cis-trans-cis-trans conformation of the amide bond sequence. The conversion from one form to another is accomplished by a conformational peptide flip consisting of a 180° rotation of a nonmethylated peptide bond. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
Molecular dynamics of the cyclic dipeptides cyclo(Gly-L -Pro), cyclo-(L-Pro-L -Pro), and cyclo(L-Pro-D-Pro) and the linear dipeptides L-Pro-Gly and cis and trans Gly-L -Pro were studied in neutral aqueous solution by 13C nuclear magnetic resonance. Spinlattice relaxation times (T1) were determined for each individual carbon atom. The correlation times, τ, were derived from a semiquantitative analysis of the T1 data. The correlation times of the proline ring carbons, β, γ, and δ suggest that the cyclic dipeptides have more restriction of conformational freedom in the proline ring than the linear dipeptides. This effect is most pronounced on the γ carbon.  相似文献   

10.
Cyclic hexapeptides of the type cyclo(L -Xxx-L -Pro-D -Yyy)2 or cyclo(L -Xxx-L -Pro-Gly)2 exist in solution predominantly in two forms of C2 average symmetry, one with all-trans peptide bonds and generally well-established conformation, and another with both Xxx-Pro peptide bonds cis. We have been measuring the thermodynamic parameters of this equilibrium using carbon and proton nmr spectroscopy. Data have been obtained for peptides in which Yyy = Gly, D -Ala, or D -Phe, and Xxx = Gly, L -Ala, L -Leu, and L -Val. In a given solvent, stability of the all-trans form decreases (ΔG0 increases) as Xxx is changed through the series Gly, L -Ala-, L -Leu, and L -Val, consistent with expected increasing repulsion between the Xxx side chain and the proline δ methylene across the trnas Xxx-Pro bond. Also, for a given set of side chains, the stability of the all-trnas form increases as the polarity of the solvent decreases, consistent with models in which all C?O and N? H groups are accessible for solvation in the two-cis form, but two C?O and two N? H groups are somewhat sequestered in the all-trans form. With the available data it is not possible to identify pure intramolecular (solvent-independent) or pure peptide-bond solvation (side chain-independent) terms in ΔH° or ΔS°, although trends are discernible.  相似文献   

11.
Conformations of the cyclic tetrapeptide cyclo(L -Pro-Sar)2 in solution were studied by 1H- and 13C-nmr spectrometry and model building. The nmr data provide definite evidence that this cyclic peptide exists chiefly in two conformations, namely, a C2-symmetric conformation and an asymmetric structure. The former was demonstrated to be predominant in polar solvents (100% in Me2SO-d6). This structure contains all cis-peptide bond linkages and all trans′ Pro Cα?CO bonds. It represents the first cyclic tetrapeptide in which all four peptide bonds have been found in the cis-conformation. As the polarity of the solvent decreases, the population of C2-symmetric conformers decreases (88% in CD3CN and 65% in CDCl3). At the same time, a minor asymmetric conformer, characterized by cis-cis-cis-trans peptide bond sequences (two cis Sar-Pro bonds, one cis Pro-Sar bond, and one trans Pro-Sar bond), is seen to increase (9% in CD3CN and 30% in CDCl3). A proposed predominant conformation in solution for cyclo(L -Pro-Sar)2 was compared with a crystal structure, as reported in an accompanying paper. Both structures show striking overall similarities.  相似文献   

12.
P R Andrews 《Biopolymers》1971,10(11):2253-2267
The molecular orbital method PCILO is applied to eight. N-monsubstituted amides. Experimentally known geometric properties are reasonably predicted by minimization of total energy with respect to molecular geometry. The same procedure shows that molecular deformations during rotation around the peptide bond significantly lower calculated barriers. Experimental heats of activation and the free-energy changes associated with cis–trans isomerism are in good agreement with those calculated, which include qualitative estimates of configurational entropy contributions to the isomerism energies. Both the calculations and revised infrared data indicate that N-phenylurethane, which has been used as a model for the cis peptide bond, should be predominantly trans. However the variations in rotational barriers and cis–trans isomerism energies among the N-monosubstituted amides provide no reason to suppose that the cis peptide bond should be excluded from stable protein conformations.  相似文献   

13.
The molecular conformations of the linear oligopeptides H-(L -Ala)n-L -Pro-OH, with n = 1,2 and 3, have been investigated. 13C nmr observation of the equilibrium between the cis and trans forms of the Ala-Pro peptide bond indicated the occurrence of nonrandom conformations in solutions of these flexible peptides. The formation of the nonrandom species containing the cis form of the Ala-Pro bond was found to depend on the deprotonation of the carboxylic acid group of proline, the solvent, and the ionic strength in aqueous solution. The influence of intramolecular hydrogen bonding on the relative conformational energies of the species containing the cis and trans Ala-Pro peptide bond was studied by comparison of the peptides H-(Ala)n-Pro-OH with analogous molecules where hydrogen bond formation was excluded by the covalent structure. In earlier work a hydrogen bond between the protonated terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue had been suggested to stabilize conformations including trans proline. For the systems described here this hypothesis can be ruled out, since the cis:trans ratio is identical for molecules with methyl ester protected and free protonated terminal carboxylic acid groups of proline. Direct evidence for hydrogen bond formation between the deprotonated terminal carboxylic acid group and the amide proton of the penultimate amino acid residue in the molecular species containing cis proline was obtained from 1H nmr studies. However, the cis:trans ratio of the Ala-Pro bond was not affected by N-methylation of the penultimate amino acid residue, which prevents formation of this hydrogen bond. Overall the experimental observations lead to the conclusion that the relative energies of the peptide conformations including cis or trans proline are mainly determined by intramolecular electrostatic interactions, whereas in the molecules considered, intramolecular hydrogen bonding is a consequence of specific peptide backbone conformations rather than a cause for the occurrence of energetically favored species. Independent support for this conclusion was obtained from model consideration which indicated that electrostatic interactions between the terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue could indeed account for the observed relative conformational energies of the species containing cis and trans proline, respectively.  相似文献   

14.
Microbial hydrolysis of the acetates of unsaturated cyclic terpene alcohols by Pseudomonas sp. NOF-5 isolated from soil was investigated. (±)-trans-Carveyl acetate ((±)-trans-3) was enantio-selectively hydrolyzed with NOF-5 strain to give ( – )-trans-carveol (( – )-trans-2 of 86.6% optical purity). However, the hydrolysis of (±)-cis-3 was less enantioselective, while (±)-piperitylacetate ((±)-6, a cis and trans mixture) was hydrolyzed to give the ( – )-trans- and ( – )-cis-piperitols (( – )- trans-5 and ( – )-cis-5) in a poor optical yield. In this case, other tert-alcohols, ( + )-trans- and ( – )- ds-2-p-menthen-1-ols ((±)-trans-7 and ( – )-cis-7), were also produced. Furthermore, microbial and enzymic allyl rearrangements of ( + )-trans-6 and ( – )-trans-verbenylacetate (( – )-trans-11) were studied. Biological treatment of (+)-trans-6 and ( – )-trans-11 with NOF-5 or its esterase gave (+)-trans- and (-)-cis-1 and ( + )-cis-3-pinen-2-ol (( + )-cis-12), respectively.  相似文献   

15.
M Sisido  Y Imanishi  T Higashimura 《Biopolymers》1973,12(10):2375-2380
Cis-trans isomerism was investigated with N-acetyl and N-propionyl, N-alkylglycine dimethylamides as model compounds for poly(N-alkylglycine dimethlamides as model compounds for (N-alkylglycine)s using n.m.r. spectroscopy. The population of the cis isomer measured in benzene and methylene chloride solutions did not show any marked dependence on the bulkiness of N-alkyl substituents. This contrasts with polyN-alkylglycine)s, whose cis isomer population increased with the introduction of bulky N-alkyl groups. Kinetics of the Cis-trans isomerization was also investigated with N-acetyldimethylamides of sacrosine, N-n-propylglycine, and N-isopropylglycine in tetrachloroethane solution. The δG? values for Cis-trans isomerization in these amides were 18 ~ 19 kcal/mole, which were virtually the same as that of polysacrosine.  相似文献   

16.
Eubacterium lentum (33 strains) isomerized the 12-cis double bond of C18 fatty acids with cis double bonds at C-9 and C-12 into an 11-trans double bond before reduction of the 9-cis double bond. The 14-cis double bond of homo-γ-linolenic acid was isomerized by 29 strains into a 13-trans double bond. The same strains isomerized the 14-cis double bond of arachidonic acid into a 13-trans double bond and then isomerized the 8-cis double bond into a 7-trans double bond; the 13-cis double bond of 10-cis, 13-cis-nonadecadienoic acid was isomerized into a 12-trans double bond. None of these isomerization products was further reduced. Studies with resting cells showed optimal isomerization velocity at a linoleic acid concentration of 37.5 μM; higher concentrations were inhibitory. The pH optimum for isomerization was 7.5 to 8.5. The isomerase was inhibited by the sulfhydryl reagents iodoacetamide, bromoacetate, and N-ethylmaleimide and by the chelators EDTA and 1,10-phenanthroline.  相似文献   

17.
Abstract

Conformational preferences of the hypermodified nucleic acid bases N6-(Δ2 -cis-hydroxyisopentenyl)adenine, cis-io6Ade also known as cis-zeatin, and N6-(Δ2 -trans-hydroxyisopentenyl)adenine, trans-io6ade or trans-zeatin, and 2-methylthio derivatives of these cis-ms2io6Ade or cis-ms2zeatin, and trans-ms2io6Ade or trans-ms2zeatin have been investigated theoretically by the quantum chemical Perturbative Configuration Interaction with Localized Orbitals (PCILO) method. Automated geometry optimization using quantum chemical MNDO, AMI and PM3 methods has also been made to compare the salient features. The predicted most stable conformation of cis-io6Ade, trans-io6Ade, cis-ms2io6Ade and trans-ms2io6Ade are such that in each of these molecules the isopentenyl substituent spreads away (has “dista” conformation) from the five membered ring imidazole moiety of the adenine. The atoms N(6), C(10) and C(11) remain coplanar with the adenine ring in the predicted preferred conformation for each of these molecules. In cis-io6Ade as well as cis- ms2io6Ade the hydroxyl oxygen may participate in intramolecular hydrogen bonding with the H-C(10)-H group. In trans-io6Ade the hydroxyl group is oriented towards the H-C(2) instead. This orientation is retained in trans-ms2io6Ade, possible O-H…S hydrogen bonding may be a stabilizing factor. In all these four modified adenines C(11)-H is favourably placed to participate in intramolecular hydrogen bonding with N(1). In cis-ms2io6Ade as well as trans-ms2io6Ade the 2-methylthio group preferentially orients on the same side as C(2)-N(3) bond, due to this nonobstrusive placing, orientation of the hydroxyisopentenyl substituent remains unaffected by 2-methylthiolation. Thus the N(1) site remains shielded irrespective of the 2-methylthiolation status in these various cis-and trans-zeatin analogs alike. Firmly held orientation of hydroxyisopentenyl substituent in zeatin isomers and derivatives, in contrast to adaptable orientation of isopentenyl substituent in i6Ade and ms2i6Ade, may account for the increased efficiency of suppressor tRNA and reduced codon context sensitivity accompanied with the occurrence of ms2-zeatin (ms2io6Ade) modification.  相似文献   

18.
The main component of Japanese Ho-leaf oil has been shown to be (?)-linalool (80~90%), and the following twenty minor constituents newly have been identified; methyl vinyl ketone, methyl isobutyl ketone, mesityl oxide, β-pinene, myrcene, (+)-limonene, cis- and trans-ocimene, n-hexanol, cis-3-hexenol, cis- and trans-linalool oxide, (?)-1-terpinen-4-ol, (+)-cis and (+)-trans-2,6,6-trimethyl-2-vinyl-5-hydroxytetrahydropyran, citronellol, nerol, (+)-β-selinene, (+)-tagetonol and (?)-trans-hotrienol. (+)-Tagetonol and (?)-trans-hotrienol have been demonstrated to be (+)-3,7-dimethyl-3-hydroxy-1-octen-5-one (III) and (3R)-(?)-trans-3,7-dimethyl-3-hydroxy-1,5,7-octatriene (IX), respectively.  相似文献   

19.
A new method of high-performance liquid chromatography (HPLC) analysis to quantify isomers of retinol, retinal and retinoic acid simultaneously was established. The HPLC system consisted of a silica gel absorption column and a linear gradient with two kinds of solvents containing n-Hexane, 2-propanol, and glacial acetic acid in different ratios. It separated six retinoic acid isomers (13-cis, 9-cis, all-trans, all-trans-4-oxo, 9-cis-4-oxo, 13-cis-4-oxo), three retinal isomers (13-cis-, 9-cis-, and all-trans) and two retinol isomers (13-cis- and all-trans). Human serum samples were subjected to this HPLC analysis and at least, all-trans retinol, 13-cis retinol, and all-trans retinoic acid were detectable. This HPLC system is useful for evaluating retinoic acid formation from retinol via a two-step oxidation pathway. Moreover, it could be applied to monitoring the concentrations of various retinoids, including all-trans retinoic acid in human sera.  相似文献   

20.
All-trans-retinoic acid (all-trans-RA) and 13-cis-retinoic acid (13-cis-RA), due to their effects on cell differentiation, proliferation and angiogenesis, improved treatment results in some malignancies. Pharmacokinetic studies of all-trans-RA and 13-cis-RA along with monitoring of retinoic acid metabolites may help to optimize retinoic acid therapy and to develop new effective strategies for the use of retinoic acids in cancer treatment. Therefore, we developed a HPLC method for the simultaneous determination in human plasma of the physiologically important retinoic acid isomers, all-trans-, 13-cis- and 9-cis-retinoic acid, their 4-oxo metabolites, 13-cis-4-oxoretinoic acid (13-cis-4-oxo-RA) and all-trans-4-oxoretinoic acid (all-trans-4-oxo-RA), and vitamin A (all-trans-retinol). Analysis performed on a silica gel column with UV detection at 350 nm using a binary multistep gradient composed on n-hexane, 2-propanolol and glacial acetic acid. For liquid-liquid extraction a mixture of n-hexane, dichloromethane and 2-propanolol was used. The limits of detection were 0.5 ng/ml for retinoic acids and 10 ng/ml for all-trans-retinol. The method showed good reproducibility for all components (within-day C.V.: 3.02–11.70%; day-to-day C.V.: 0.01–11.34%. Furthermore, 9-cis-4-oxoretinoic acid (9-cis-4-oxo-RA) is separated from all-trans-4-oxo-RA and 13-cis-4-oxo-RA. In case of clinical use of 9-cis-retinoic acid (9-cis-RA) the pharmacokinetics and metabolism of this retinoic acid isomer can also be examined.  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