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.
Nine analogues of methyl acarviosin (1), the core structure of acarbose and its homologues, the 6-hydroxy-(2), 6-azido-(3), 6-amino- (4), 6-acetamido-(5), 6-methoxy-(6), 6-hydroxy-2-O-methyl-(8), and 6-hydroxy-3-O-methyl derivatives (9), including the 5-methoxycarbonyl analogue (7) and 3,6-anhydro derivative (10) of 2, were synthesized by chemical modification of the sugar part of 2 derived by condensation of methyl 3,4-anhydro-alpha-D-galactopyranoside (17) and 4,7:5,6-di-O-isopropylidenevalienamine (26) or by direct coupling between 26 and the 6-substituted methyl 3,4-anhydro-alpha-D-galactopyranoside derivatives. Compounds 2 and 8 show notable inhibitory activity against yeast alpha-D-glucosidase almost comparable to that of 1. Introduction of a polar substituent at C-6 of 1 decreases the inhibitory activity. Interestingly, inversion of the conformation of the sugar part of 1 by introduction of the 3,6-anhydro bridge elicits almost no effect on the inhibitory activity.  相似文献   

2.
The synthesis and stereochemistry (configuration and preferred solute conformation) of some 4-alkyl (methyl, n-propyl, isobutyl)-4-(3-hydroxy-phenyl)-1-methylpiperidines and corresponding 3-methyl diastereoisomeric pairs are reported, together with their in vivo and in vitro activities as opioid ligands. All potent agonists exhibit a preference for axial 4-aryl chair conformations when protonated, and stereochemical analogies with rigid opioids of the benzomorphan class are discussed. Antagonist properties are found in compounds with preference for equatorial 4-aryl chairs, notably the cis 3,4-dimethyl derivative.  相似文献   

3.
Addition of the elements of phthalimide to methyl 2,3-anhydro-4,6-O-benzylidene-alpha-D-mannopyranoside (1) under fusion conditions has yielded methyl 4,6-O-benzylidene-3-deoxy-3-phthalimido-alpha-D-altropyranoside (2). The conformation of the pyranose ring of 2 has been shown to be non-chair by 1H NMR spectroscopy, in contrast to the conformations of related derivatives having smaller substituents at C-3. Molecular dynamics simulations of 2 in explicit chloroform-d solvent have indicated four principal conformational possibilities. Of these, the 7C5/1S5 chair/skew boat form 2d has the lowest potential energy, and is largely consistent with the observed vicinal 1H-1H NMR coupling constants.  相似文献   

4.
Single-crystal X-ray diffraction and high-resolution (1)H and (13)C NMR spectral data for methyl 3,4-di-O-acetyl-1,5-anhydro-2-deoxy-D-arabino-hex-1-enopyranuronate are reported. The (5)H(4) conformation was found to be the preferred form for this glycal, both in the crystal lattice and in solution. The factors determining the (4)H(5)<==>(5)H(4) conformational equilibrium for acetylated glycals are discussed.  相似文献   

5.
It has been proposed that the membrane allows a much more efficient binding of certain small or medium-sized amphiphilic messenger molecules to their receptor, not only by accumulation of the drug, but also by induction of orientations and conformations that are much more favorable for receptor docking than structures adopted in isotropic phases. A series of eight amphiphilic cyclic peptides containing lipophilic (L-alpha-aminodecanoic acid = Ada, L-alpha-aminohexadecanoic acid = Ahd, Nhdg = N-hexadecylglycine) and hydrophilic (Lys, Asp) amino acids were synthesized and examined by means of NMR spectroscopy and molecular dynamics (MD) simulations in isotropic (CDCl3) and membrane-mimicking anisotropic (SDS/H2O) solvents to study the influence of the environment on their individual conformations. NMR data of cyclo(-Gly1-D-Asp2-Ahd3-Ahd4-Asp5-Gly6+ ++-) (C4), cyclo(-Lys1-D-Pro2-Lys3-Ada4-Pro5-Ada6-) (C5) and cyclo(-Lys1-Pro2-Lys3-Ada4-D-Pro5-Ada6-) (C6) clearly indicate that those compounds are too rigid to perform a conformational change upon transition from an isotropic to an anisotropic environment. On the other hand, the experimental data of cyclo (-Gly1-Asp2-Ahd3-Ahd4-Asp5-Gly6-) (C1), cyclo(-Asp1-Ala2-Nhdg3-Ala4-D-Asp5-) (C7), and cyclo(-D-Asp1-Ala2-Nhdg3-Ala4-Asp5-) (C8) suggest highly flexible unstructured molecules in both environments. However, for cyclo(-Asp1-Asp2-Gly3-Ahd4-Ahd5-Gly6-) (C2) we observed a structure inducing effect of a membrane-like environment. The compound populates three different conformations in SDS/H2O, whereas in CDCI3 no preferred conformation can be detected. cyclo(-D-Asp1-Asp2-Gly3-Ahd4-Ahd5-Gly6-) (C3) clearly exhibits two different conformations with a shifted beta,beta-turn motif in CDCI3 and SDS/H2O solutions. The conformational change could be reproduced in a restraint-free MD simulation using the biphasic membrane mimetic CCl4/H2O. Our results give clear evidence that membrane interactions may not only lead to structure inductions, but can also induce major conformational changes in compounds already exhibiting a defined structure in isotropic solution.  相似文献   

6.
The P-chiral stereo-defined phosphorothioate groups have been introduced into all of the four internucleotide positions of d(T(PS1)C(PS2)C(PS3)C(PS4)C) (PSn = phosphorothioate group), and among the 16 possible diastereomers of PS-d(TC(4)), 10 stereomers have been synthesized to investigate the effects of the sense of the P-chirality upon the structure and stability of the i-motif structure. The temperature dependence of circular dichroism spectra showed that the melting temperature (T(m)) of the [all R(p)]-PS-d(TC(4)) i-motifs was 31 degrees C, identical to that of the parent oligomer, PO-d(TC(4)), while that of the [all S(p)]-PS-d(TC(4)) i-motif was largely decreased by 11 degrees C. Single substitution of R(p) with S(p) caused a decrease of T(m) by 3-4 degrees C at positions of PS1, PS2, and PS3 and by 1 degrees C at that of PS4, showing the additive property of the T(m) suppression. The comparison of the NOESY spectra between [all R(p)]-PS-, [all S(p)]-PS-, and PO-d(TC(4)) showed that intraresidual H6-H3' and H2' '-H4' NOE cross-peaks of the all S(p) isomer are weaker than those of the all R(p) isomer and PO-d(TC(4)), indicating the change in the C3'-endo conformation and glycosidic bond angle. The structural alternation for the i-motif formed by [all S(p)]-PS-d(TC(4)) is also suggested by the chemical shift differences of C2/C3/C4 H2'and H4' protons from those of [all R(p)]-PS-d(TC(4)) and PO-d(TC(4)). These results suggest that the S(p) configuration at phosphorus of the phosphorothioate linkage changes the sugar-phosphorothioate conformation and intermolecular interaction in the narrow groove, leading to the destabilization of the i-motif structure.  相似文献   

7.
The CDX2 and CDX1 homeobox genes have respectively a tumour suppressor and proliferative role in the intestinal epithelium. We analyzed DNA methylation and histones modifications associated with CDX2 and CDX1 promoters in two human colon cancer cell lines expressing differentially these genes, Caco2/TC7 [CDX2 positive-CDX1 negative] and HT29 [CDX2 negative-CDX1 negative] cells. Chromatin immunoprecipitation experiments indicated that CDX2 and CDX1 gene expression correlated with a histone modifications pattern characterizing active chromatin (H3K4 trimethylated and H3 acetylated). Bisulfite DNA sequencing and methylation-specific PCR showed that CDX2 and CDX1 promoters display no methylation in HT29 cells even though both genes are not expressed. In contrast, the CDX1 promoter is methylated in Caco2/TC7. DNA demethylation by 5aza-dC or the combination of 5aza-dC plus SAHA, an inhibitor of histone deacetylases, restored CDX1 expression in Caco2/TC7 cells but these treatments were inefficient on both CDX2 and CDX1 in HT29 cells. Thus, in colon cancer cells the changes in chromatin conformation are heterogeneous and repression of CDX2 and CDX1 in HT29 cells is not due to epigenetic mechanisms. In vivo, dietary deprivation of methyl groups in rats upregulated CDX1 mRNA and downregulated to a lesser extent CDX2 mRNA expression. Moreover, methyl group deprivation downregulated CDX2 protein by changing its phosphorylation pattern. The changes in CDX2 and CDX1 expression determined by methyl group deprivation may constitute one of the mechanisms sustaining the protective role attributed to folate in colon cancer.  相似文献   

8.
T Yamazaki  K Nunami  M Goodman 《Biopolymers》1991,31(13):1513-1528
The conformations of cis and trans cyclic retro-inverso dipeptides--2-[(4-hydroxy)benzyl]-5-benzyl-4,6(1H,2H,3H,5H)-pyrimidinedi one (c[mTyr-gPhe]), and 2-benzyl-5-amino-5-[(4-hydroxy)benzyl]-4,6(1H,2H,3H,5H)-pyrimidinedione (c[mTyr-gPhe]), and 2-benzyl-5-amino-5-[(4-hydroxy)benzyl]-4,6(1H,2H,3H,5H)-pyrimidinedione (c[(alpha-amino)mTyr-gPhe])--and the parent cyclic dipeptides--c[tyrosyl-phenylalanine] (cis-c[L-Tyr-L-Phe]) and c[tyrosyl-D-phenylalanine] (trans-c[L-Tyr-D-Phe])--were studied by using 1H-nmr spectroscopy and semiempirical energy calculations. In the cis compounds of all the cyclic retro-inverso and parent dipeptides, the most stable conformer has both aromatic side chains sharing the space over the backbone ring in a "face-to-face" fashion. All the trans compounds predominantly assume a "sandwich" conformation in which the two aromatic rings are folded back over the backbone ring on opposite sides. However, different conformational preferences were observed for the backbones between the retro-inverso and parent cyclic dipeptides. The parent cyclic dipeptide trans-c[L-Tyr-D-Phe] adopts two types of boat structures with different side-chain orientations in almost equal amounts: one with the Tyr side chain in a pseudoaxial position and the Phe side chain in a pseudoequatorial position, the other with the Tyr side chain in a pseudoequatorial position and the Phe side chain in a pseudoaxial position. On the other hand, the cyclic retro-inverso dipeptides trans-c[mPhe-gTyr] and trans c[mTyr-gPhe] assume only one type of boat structure in which the malonyl side chain is in a pseudoequatorial and the gem-diamino side chain is in a pseudoaxial position. In addition to the preferred conformations, the conformational energies of the C alpha--C beta bonds in the malonyl and gem-diamino residues were estimated from the temperature variation of vicinal 1H--1H coupling constants for the H--C alpha--C beta--H groupings observed for the trans isomers of cyclic retro-inverso dipeptides. The energies were evaluated to be 1.1 and 1.8 kcal mol-1 for the malonyl and gem-diamino residues, respectively. Applying these energies to the parent cyclic dipeptide trans-c[L-Tyr-D-Phe], the observed fractions of three side-chain conformations are reasonably reproduced. The conformational energies as well as conformational properties of the molecules estimated in this investigation may be useful to refine force constants for both parent and retro-inverso peptides with aromatic side chains.  相似文献   

9.
The conformational behavior of the synthetic hexa- and heptasaccharide methyl beta-glycosides alpha-D-Manp-(1 --> 6)-[alpha-D-Manp-(1 --> 3)-][beta-D-Xylp-(1 --> 2)-]beta-D-Manp-(1 --> 4)-beta-D-GlcpNAc-(1 --> 4)-beta-D-GlcpNAc-(1 --> OMe and alpha-D-Manp-(1 --> 6)-[alpha-D-Manp-(1 --> 3)-][beta-D-Xylp-(1 --> 2)-]beta-D-Manp-(1 --> 4)-beta-D-GlcpNAc-(1 --> 4)-[alpha-L-Fucp-(1 --> 6)-]beta-D-GlcpNAc-(1 --> OMe, representing the xylosylated and the xylosylated alpha-(1 --> 6)-fucosylated core structures of N-glycans in alpha(D)-hemocyanin of the snail Helix pomatia, respectively, were investigated by 1H NMR spectroscopy in combination with molecular dynamics (MD) simulations in water. 1H and 13C chemical shifts of the oligosaccharides were assigned using 1H-(1)H COSY, TOCSY, and NOESY, and 1H-(13)C HMQC techniques. Experimental 2D 1H cross-peak intensities from one series of NOESY and one series of ROESY experiments of the two oligosaccharides were compared with calculated values derived from MD trajectories using the CROSREL program, yielding information about the conformation of each glycosidic linkage of the methyl glycosides. The flexibility of the linkages was described by generalized order parameters and internal rotation correlation times. Analysis of the data indicated that several conformations are likely to exist for the alpha-D-Man-(1 --> 6)-beta-D-Man, the alpha-L-Fuc-(1 --> 6)-beta-D-GlcNAc, and the alpha-D-Man-(1 --> 3)-beta-D-Man linkage, whereas the beta-D-Xyl-(1 --> 2)-beta-D-Man-(1 --> 4)-beta-D-GlcNAc-(1 --> 4)-beta-D-GlcNAc fragment occurs in one rigid conformation. No significant differences were found between the corresponding structural elements in both methyl glycosides. NOESY and ROESY experiments proved to be suitable for providing the experimental data required, however, due to more overlap within the ROESY spectra, reducing the accuracy of the analysis, NOESY spectral analysis is preferred.  相似文献   

10.
Lactones of methyl 3-O-[(R)- and (S)-1-carboxyethyl]-alpha-D-gluco-, galacto- and manno-pyranoside were prepared by treatment of the sugar derivatives in acetic acid. The lactones were formed between the 1-carboxyethyl substituent and 2-OH or 4-OH in different proportions depending on the stereochemistry of the parent compounds. Relative formation rates in acetic acid-d4 and hydrolysis rates in buffered D2O solutions at pD 2.4, 4.6 and 7.4 were estimated. Hydrolysis of the formed lactones is relatively slow in D2O at pD 4.6, which permitted characterization of the lactones by 1H and 13C NMR spectroscopy in buffered D2O solutions. Hydrolysis of the lactones in 1 M aqueous NaOH at 80 degrees C gave no detectable isomerization of the alpha-carbon. The set of lactones formed from the 1-carboxyethyl substituted methyl glycosides used in this study showed large similarities in the NMR shifts (delta delta values). Deviations from the observed shift pattern were found for two lactones. Our findings strongly suggest that those two lactones differ from the rest by adopting a boat-like conformation, whereas the others adopt pseudo-chair conformations.  相似文献   

11.
Calculated and experimental (1)H, (13)C and (19)F chemical shifts were compared in BKM-824, a cyclic bradykinin antagonist mimic, c[Ava(1)-Igl(2)-Ser(3)-DF5F(4)-Oic(5)-Arg(6)] (Ava=5-aminovaleric acid, Igl=alpha-(2-indanyl)glycine, DF5F=pentafluorophenylalanine, Oic=(2S,3aS,7aS)-octahydroindole-2-carboxylic acid). The conformation of BKM-824 has been studied earlier by NMR spectroscopy (M. Miskolzie et al., J. Biomolec. Struct. Dyn. 17, 947-955 (2000)). All NMR structures have qualitatively the same backbone structure but there is considerable variation in the side chain conformations. We have carried out quantum mechanical optimization for three representative NMR structures at the B3LYP/6-31G* level, constraining the backbone dihedral angles at their NMR structure values, followed by NMR chemical shift calculations at the optimized structures with the 6-311G** basis set. There is an intramolecular hydrogen bond at Ser(3) in the optimized structures. The experimental (13)C chemical shifts at five C(alpha) positions as well as at the Cbeta, Cgamma and Cdelta position of Ava(1), which forms part of the backbone, are well reproduced by the calculations, confirming the NMR backbone structure. A comparison between the calculated and experimental H(beta) chemical shifts in Igl(2) shows that the dominant conformation at this residue is gauche. Changes of proton chemical shifts with the scan of the chi(1) angle in DF5F(4) suggest that chi(1)180 degrees. The calculated (1)H and (13)C chemical shifts are in good agreement with experiment at the rigid residue Oic(5). None of the models gives accurate results for Arg(6), presumably because of its positive charge. Our study indicates that calculated NMR shifts can be used as additional constraints in conjunction with NMR data to determine protein conformations. However, to be computationally effective, a database of chemical shifts in small peptide fragments should be precalculated.  相似文献   

12.
D-ring-fused dioxaphosphorinanes (4-6) in the estrone series were synthetized as epimeric pairs and investigated by NMR and computational methods in order to determine their stereostructures and predominant conformations. The study was performed to evaluate the influence of the rigid sterane framework on the geometry of the condensed hetero ring, with regard to the possible steric effect of the angular methyl group at position 13. Additionally, the steric and electronic effects of the P-substituents on the conformational equilibrium were examined. The distorted-boat conformation of the hetero ring of dioxaphosphorinoestrone 3-methyl ether 4a was confirmed by single-crystal X-ray analysis. This is in good agreement with the observation in solution that, in the case of the boat conformation, the anisotropic shielding effect of the phenyl group of cyclic phosphonate 4a generates an upfield shift for 17-H, as compared with the corresponding chemical shift for epimer 4b. A similar boat conformation was substantiated for derivatives 4b, 5a, 5b and 6b on the basis of the J(H, H) and J(H, P) coupling constants and also ab initio calculations, regardless of the P-configuration. At the same time, the hetero ring of 6a seems to tilt towards a chair-like conformation due to the strong equatorial preference of the N-bis(2-chloroethyl) group.  相似文献   

13.
The metabolism of 1-fluoronaphthalene by Cunninghamella elegans ATCC 36112 was studied. The metabolites were isolated by reverse-phase high-pressure liquid chromatography and characterized by the application of UV absorption, 1H nuclear magnetic resonance, and mass spectral techniques. C. elegans oxidized 1-fluoronaphthalene predominantly at the 3,4- and 5,6-positions to form trans-3,4-dihydroxy-3,4-dihydro-1-fluoronaphthalene and trans-5,6-dihydroxy-5,6-dihydro-1-fluoronaphthalene. In addition, 1-fluoro-8-hydroxy-5-tetralone, 5-hydroxy-1-fluoronaphthalene, and 4-hydroxy-1-fluoronaphthalene as well as glucoside, sulfate, and glucuronic acid conjugates of these phenols were formed. Circular dichroism spectra of the trans-3,4- and trans-5,6-dihydrodiols formed from 1-fluoronaphthalene indicated that the major enantiomers of the dihydrodiols have S,S absolute stereochemistries. In contrast, the trans-5,6-dihydrodiol formed from 1-fluoronaphthalene from 3-methylcholanthrene-treated rats had Cotton effects that are opposite in sign (R,R) to those formed by C. elegans. The results indicate that the fungal monooxygenase-epoxide hydrolase systems are highly stereoselective in the metabolism of 1-fluoronaphthalene and that a fluoro substituent blocks epoxidation at the fluoro-substituted double bond, decreases oxidation at the aromatic double bond that is peri to the fluoro substituent, and enhances metabolism at the 3,4- and 5,6-positions of 1-fluoronaphthalene.  相似文献   

14.
Diethylglycine (Deg) residues incorporated into peptides can stabilize fully extended (C5) or helical conformations. The conformations of three tetrapeptides Boc-Xxx-Deg-Xxx-Deg-OMe (Xxx=Gly, GD4; Leu, LD4 and Pro, PD4) have been investigated by NMR. In the Gly and Leu peptides, NOE data suggest that the local conformations at the Deg residues are fully extended. Low temperature coefficients for the Deg(2) and Deg(4) NH groups are consistent with their inaccessibility to solvent, in a C5 conformation. NMR evidence supports a folded beta-turn conformation involving Deg(2)-Gly(3), stabilized by a 4-->1 intramolecular hydrogen bond between Pro(1) CO and Deg(4) NH in the proline containing peptide (PD4). The crystal structure of GD4 reveals a hydrated multiple turn conformation with Gly(1)-Deg(2) adopting a distorted type II/II' conformation, while the Deg(2)-Pro(3) segment adopts a type III/III' structure. A lone water molecule is inserted into the potential 4-->1 hydrogen bond of the Gly(1)-Deg(2) beta-turn.  相似文献   

15.
The X-ray crystal structure of 1d-1,4-di-O-[(S)-O-acetylmandeloyl]-2,3:5,6-di-O-isopropylidene-myo-inositol is described. Both inositol ring and OAM (O-acetylmandeloyl) moiety deviate from their respective ideal conformations. Inositol ring adopts a flattened chair conformation while OAM adopts an ap (antiperiplanar) conformation. A comparison of its conformation in solution with that in solid was made by the use of NOESY and anisotropic shielding effect in (1)H NMR. This conformational study revealed that the title compound adopts similar conformations in both the states.  相似文献   

16.
Exhaustive conformational analysis of the 5'-deoxyadenylic acid molecule, has been carried out by the quantum-mechanical density functional theory method at the MP2/6-311++G(d,p)//DFT B3LYP/6-31G(d,p) theory level. As many as 726 of its conformations have been revealed with the relative gas phase Gibbs energies under standard conditions from 0 to 12.1 kcal/mole. It has been shown, that the energetically most favorable conformation has north sugar puckering and synorientation of the nitrogenous base and is stabilized by intramolecular O(p1)H(p1)-N3 and O3'H-O(p) hydrogen bonds. Four conformations have been shown to have their geometry similar to that of AI-DNA and four - of BI-DNA. One conformer of the 5'-deoxyadenylic acid molecule is similar to its sodium salt hexahydrate structure in crystalline state resolved by the X-ray diffraction method and taken from literature. It is shown that effective charges of C4' and C5' atoms are the most sensitive to the molecule conformation ones. The role of the intramolecular OH-N hydrogen bonds in formation of the 5'-deoxyadenylic acid molecule structure has been demonstrated.  相似文献   

17.
Transferred nuclear Overhauser enhancement spectroscopy (TRNOE) was used to observe changes in a ligand's conformation upon binding to its specific antibody. The ligands studied were methyl O-beta-D-galactopyranosyl(1----6)-4-deoxy-4-fluoro-beta-D-galactopyra nos ide (me4FGal2) and its selectively deuteriated analogue, methyl O-beta-D-galactopyranosyl(1----6)-4-deoxy-2-deuterio-4-fluoro-beta -D- galactopyranoside (me4F2dGal2). The monoclonal antibody was mouse IgA X24. The solution conformation of the free ligand me4F2dGal2 was inferred from measurements of vicinal 1H-1H coupling constants, long-range 1H-13C coupling constants, and NOE cross-peak intensities. For free ligand, both galactosyl residues adopt a regular chair conformation, but the NMR spectra are incompatible with a single unique conformation of the glycosidic linkage. Analysis of 1H-1H and 1H-13C constants indicates that the major conformer has an extended conformation: phi = -120 degrees; psi = 180 degrees; and omega = 75 degrees. TRNOE measurements on me4FGal2 and me4F2dGal2 in the presence of the specific antibody indicate that the pyranose ring pucker of each galactose ring remains unchanged, but rotations about the glycosidic linkage occur upon binding to X24. Computer calculations indicate that there are two sets of torsion angles that satisfy the observed NMR constraints, namely, phi = -152 +/- 9 degrees; psi = -128 +/- 7 degrees; and omega = -158 +/- 6 degrees; and a conformer with phi = -53 +/- 6 degrees; psi = 154 +/- 10 degrees; and omega = -173 +/- 6 degrees. Neither conformation is similar to any of the observed conformations of the free disaccharide.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Conformational properties of three novel zidovudine analogs, namely 3'-azido-3'-deoxy-5'-O-isonicotinoylthymidine (AZT-Iso, 2), (-)-trans-(5S,6S)-5-bromo-6, 5'-epoxy-5,6-dihydro-3'-azido-3'-deoxythymidine (3) and (+)-trans-(5R,6R)-5-bromo-6,5'-epoxy-5,6-dihydro-3'-azido-3'-deoxythymidine (4), have been investigated by AM1 calculations and NMR studies, and compared with those of the parent nucleoside (AZT, 1). Based on the results obtained the following correlation may be established, a) AZT and AZT-Iso exhibit a conformational behavior analog to other pyrimidinic nucleosides, displaying a dynamic equilibrium in solution where the two conformers (North and South) undergo a constant transformation. b) Compounds 3 and 4 show a different conformational profile. The estimate of the pseudorotation phase angle reveals the rigid structures of the latter compounds, which do not evidence conformational equilibrium in solution; the azide group being the only group free to rotate. c) Diastereoisomers 3 and 4 exhibit an extra conformational parameter compared with other pyrimidinic nucleosides: the chair or boat conformation in the third ring formed between the sugar and the base. In all cases, a reasonable correlation was obtained between theoretical and NMR spectroscopic data.  相似文献   

19.
Two galactosyl derivatives of [DMet2,Pro5] enkephalin-amide (compound 1), namely [DMet2,Pro5] enkephalin [N1.5-beta-D-galactopyranosyl] amide (compound 2) and O1.5-(beta-D-galactopyranosyl) [DMet2,Hyp5] enkephalin-amide (compound 3) have been synthesized. Such glycosylpeptides have been shown to be extremely potent analgesic agonists. The conformational analysis of these three compounds in DMSO-d6 solution has been carried out using two-dimensional NMR methods. Both the parent compound (1) and the beta N-galactosyl derivative (2) show similar NMR parameters which are consistent with fairly rigid beta-strands at both the N-terminus and C-terminus, connected by a glycine residue that displays a mixture between multiple conformational states. Thus, although the beta N-galactosyl derivative (2) has been shown to be significantly more potent than the parent compound (1) in the tail immersion and paw pressure tests of analgesia, no correlation can be established between the conformation of (1) and (2) in DMSO and the difference in analgesic activity. In contrast, important conformational differences with respect to (1) and (2) have been detected in the beta O-galactosyl derivative (3). In this case, only one of the likely conformations for (1) and (2) are consistent with the experimental data. These data show that the position of the galactose residue in compound (3) causes Gly3 to loose flexibility leading to a more rigid folded conformation. Such a change in conformation could be related to the difference in analgesic activity between (2) and (3).  相似文献   

20.
The 5,6-dimethylbenzimidazole moiety of vitamin B12 is formed from riboflavin in aerobic and some aerotolerant bacteria. Thereby C1' of riboflavin is transformed into C2 of the vitamin B12 base. In the present publication a study on this transformation with riboflavin 2H-labeled in the 1'R or 1'S position is described. This study was undertaken in order to find out if one of the two hydrogens at C1' is transferred to C2 of 5,6-dimethylbenzimidazole. The 2H-labeled riboflavin samples were synthesized starting from unlabeled or 1-2H-labeled ribose and 3,4-dimethylaniline yielding N-beta-D-ribopyranosyl-3,4-dimethylaniline. The unlabeled riboside was reduced to N-D-ribityl-3,4-dimethylaniline with sodium cyanoborotrideuteride, the 2H-labeled riboside with sodium cyanoborohydride. The ribityl derivatives were transformed into N-D-ribityl-2-phenylazo-4,5-dimethylaniline, and condensed with barbituric acid to riboflavin. The reduction of the ribosyl compound to the ribityl derivative is only partially stereospecific. Thus the riboflavin synthesized from unlabeled ribose had a 2H ratio of 3/1 (1'R/1'S), the riboflavin obtained from D-[1-2H1]ribose of 1/3 (1'R/1'S). The 2H content in these positions was determined from the 1H-NMR spectra. These spectra showed also that 1 mol 2H/mol riboflavin was present in position 1'. The deuterated riboflavin samples were incubated under aerobic conditions with broken cell preparations of Propionibacterium shermanii. The deuterium content of the 5,6-dimethylbenzimidazole isolated was determined by mass spectrometry and by 1H NMR. These measurements revealed that the hydrogen in the pro-S position at C1' of riboflavin is retained during 5,6-dimethylbenzimidazole formation, and is thus found at C2 of this base.  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