首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The crystal structure of potassium hydroxide complexed amylose, obtained by heterogeneous deacetylation of amylose triacetate, has been determined through a combined stereochemical structure-refinement and X-ray diffraction-analysis. The structure crystallizes in an orthorhombic unit-cell with parameters a  8.84, b  12.31, and c (fiber repeat)  22.41 Å, and with P212121 symmetry. The conformation of the amylose chain is a distorted, left-handed helix with 6 d-glucose residues per turn. Each three-residue asymmetric unit is complexed with one molecule of potassium hydroxide and three molecules of water. The K+ ion coordinates with four oxygen atoms of the amylose chain and with two other oxygen atoms, and this coordination is probably the cause for the more-extended amylose chain-conformation than would be predicted from a φ, ψ map. The distortions in the chain are primarily manifested by different O-6 rotations and by slightly different bridge and φ, ψ angles for the individual residues. The structure is extensively hydrogen bonded, although largely through water molecules, which accounts for its ready water solubility. The left-handed conformation of the chain in this structure is consistent with the conformations of amylose triacetate and V-amylose, both of which are left-handed.  相似文献   

2.
The local conformation of RNA molecules is an important factor in determining their catalytic and binding properties. The analysis of such conformations is particularly difficult due to the large number of degrees of freedom, such as the measured torsion angles per residue and the interatomic distances among interacting residues. In this work, we use a nearest-neighbor search method based on the statistical mechanical Potts model to find clusters in the RNA conformational space. The proposed technique is mostly automatic and may be applied to problems, where there is no prior knowledge on the structure of the data space in contrast to many other clustering techniques. Results are reported for both single residue conformations, where the parameter set of the data space includes four to seven torsional angles, and base pair geometries, where the data space is reduced to two dimensions. Moreover, new results are reported for base stacking geometries. For the first two cases, i.e., single residue conformations and base pair geometries, we get a very good match between the results of the proposed clustering method and the known classifications with only few exceptions. For the case of base stacking geometries, we validate our classification with respect to geometrical constraints and describe the content, and the geometry of the new clusters.  相似文献   

3.
The substitution pattern of the water-soluble l-arabino-(4-O-methyl-d-glucurono)-d-xylan from redwood (Sequoia sempervirens) has been studied by enzymic degradation. Exhaustive hydrolysis by an endo-xylanase (EC 3.2.1.8) from a Basidiomycete Sporotrichum dimorphosporum left a residue accounting for 20% of the original d-xylan. In the dialyzable material, oligosaccharides having arabinose or 4-O-methylglucuronic acid residues attached to the non-reducing d-xylosyl end-group of xylobiose or xylotriose, respectively, were the smallest branched oligomers released. Action of the xylanase appears to involve a region of the polysaccharide backbone having three xylosyl residues. A mode of action is proposed that requires unsubstituted hydroxyl groups at C-2, C-3, and C-2′ of a xylobiosyl residue. The binding site seems to correspond to a shallow cavity. The composition and structure of the final residue of attack shows that the enzyme has no action when the xylosyl residues branched through O-2 are separated by only one, unsubstituted xylose residue. This pattern of action, the nature of the dialyzable products, and the production of a final residue in which the substituents are accumulated, suggest that the arabinosyl and glucosyl-uronic groups are irregularly distributed on the main chain of the xylan from redwood and that in some regions they are in close vicinity when not actually on adjacent xylosyl residues.  相似文献   

4.
《Carbohydrate research》1988,172(1):97-112
The complete structure of the acidic, extracellular, capsular polysaccharide of Rhizobium trifolii 843 has been elucidated by a combination of chemical, enzymic, and spectroscopic methods, confirming an earlier proposed sugar sequence and assigning the locations of the acyl substituents. The polysaccharide was depolymerized by a lyase into octasaccharide units which were uniform in carbohydrate composition and linkage. These units also contained a uniform distribution of acetyl and pyruvic acetal [O-(1-carboxyethylidene)] groups, and half of them were further acylated with d-3-hydroxybutanoyl groups. A much smaller proportion (<5%) of the oligomers was further acylated by a second d-3-hydroxy-butanoyl group. The locations of the substituents were determined chemically and by J-correlated, 1H-n.m.r. spectroscopy, proton nuclear Overhauser effect (n.O.e.)_ measurements, doubie-resonance 1H-n.m.r. spectroscopy, and 13C-n.m.r. spectroscopy. The composition and structure of the carbohydrate chain were determined by methylation analysis using g.l.c.-m.s. fast-atom-bombardment mass spectrometry, and n.m.r. studies on the reduced, deacylated oligomer. Structural studies were supplemented by n.m.r. analyses on the original polymer. The oligosaccharides were found to be branched octasaccharides with four sugar residues in each branch, and the carbohydrate sequence agreed well with that expected from earlier work. In the abbreviated sequence and structure (1a), the sugar residues are labelled “a” through “h”. The main chain (a–d) is composed of a 4-deoxy-α-l-threo-hex-4-enopyranosyluronic acid group (a) that is linked to O-4 of a 3-O-acetyl-d-glucosyluronic acid residue (b) which is β-linked to O-4 of a d-glucosyl residue (c). Residue c is β-linked to O-4 of the branching d-linked to O-4 of a d-glucosyl residue (d). The side chain consists of a substituted d-galactosyl group (h) which is β-linked to O-3 of residue 9 of a β-(1→4)-linked d-glucose trisaccharide (fragment e–f–g). The reducing end of the resulting tetrasaccharide (e–f–g–h) is β-linked to O-6 of the branching d-glucose residue (d). In the native polymer, this branching residue is α-linked to O-4 of the modified d-glucuronic acid residue (a) which is the unsaturated sugar in the oligomer. A small proportion of the O-2 atoms of the acetylated d-glucosyluronic acid residues is acetylated because of ester migration. The two terminal sugars (g and h) of the branch chain bear 4,6-O-(1-carboxyethylidene) groups. The d-galactosyl groups of half of the oligomers are acylated by d-3-hydroxybutanoyl groups at O-3. About 5% of the oligomers bear a second d-3-hydroxybutanoyl group at O-2 of the d-galactosyl group (h).  相似文献   

5.
The structure of the O-specific side-chains of the Shigella sonnei phase I lipopolysaccharide has been investigated. The side chains are composed of disaccharide repeating-units containing two uncommon sugar components, one of witch, 2-amino-2-deoxy-L-altruronic acid, has been identified previously. The other has now been identified as 2-acetamido-4-amino-2,4,6-trideoxy-D-galactose. The uronic acid, as N-acetylated α-pyranosyl residues, is linked through O-4, and the diamino sugar, as β-pyranosyl residues, is linked through O-3. The pyranosyluronic acid residue assumes the 4C1 conformation in the polymer, with the carboxyl group in the axial position.  相似文献   

6.
The suitability of the reductive-cleavage method for analysis of the linkage positions in d-fructofuranosyl residues of d-fructans was examined by using sucrose, chicory-root inulin, and Aerobacter levanicum levan as models. Permethylation, and reductive cleavage with triethylsilane in the presence of either boron trifluoride etherate or trimethylsilyl trifluoromethanesulfonate, gave the expected methylated derivatives of 2,5-anhydro-d-mannitol and 2,5-anhydro-d-glucitol. With either catalyst, nonreducing (terminal) d-fructofuranosyl groups and d-fructofuranosyl residues linked at O-1 gave derivatives having the manno configuration as the major product, whereas d-fructofuranosyl residues linked at O-6, and at both O-1 and O-6, gave derivatives having the gluco configuration as the major product. The independent synthesis, and n.m.r.- and mass-spectral characterization, of the methylated 2,5-anhydro-d-mannitol and 2,5-anhydro-d-glucitol derivatives formed from these residues by reductive cleavage are reported.  相似文献   

7.
The crystal structure of the B-polymorph of amylose appears to be based on double-stranded helices. The individual strands are in a right-handed six-fold helical conformation repeating in 20.8 Å and are wound parallel around each other. The steric disposition of O-6 is gt. The double helices pack in a hexagonal unit-cell (ab  18.50 Å, c (fiber repeat)  10.40 Å, γ  120°), with two helices (12 d-glucose residues) per cell. The helices are packed antiparallel and leave an open channel within a hexagonal array that is filled with water molecules. The reliability of the structure analysis is indicated by R  0.22. The structure of B-amylose is consistent with the diffraction diagrams of B-starches and accounts for the physical properties of such starches.  相似文献   

8.
A previous investigation of the structure of the extracellular polysaccharide gum from the nitrogen-fixing Rhizobium strain cb744 (a member of the slow-growing Cowpea group) indicated that there were two β-(1→4)-linked d-glucopyranosyl residues for each α-(1→4)-linked d-mannopyranosyl residue, and that each mannose was substituted at O-6 by a β-d-galactopyranosyl residue having 71% of the galactose present as 4-O-methylgalactose. The present study shows that, although the gum appeared to have a simple tetrasaccharide repeating unit, it is composed of two closely associated components. One is a (1→4)-linked α-d-mannan substituted at each O-6 by a β-d-galactopyranosyl residue (71% 4-O-methylated). The second component is a (1→4)-linked β-d-glucan. The existence of the two polysaccharides was established by separation of the β-d-galactosidase-treated gum on a column of concanavalin A-Sepharose 4B. The d configurations were determined and the anomeric attribution of the linkages confirmed by the use of enzymes. The interaction between the two gum components is discussed.  相似文献   

9.
Vacuolar (H+)-ATPase (V-ATPase) is a proton pump present in several compartments of eukaryotic cells to regulate physiological processes. From biochemical studies it is known that the interaction between arginine 735 present in the seventh transmembrane (TM7) segment from subunit a and specific glutamic acid residues in the subunit c assembly plays an essential role in proton translocation. To provide more detailed structural information about this protein domain, a peptide resembling TM7 (denoted peptide MTM7) from Saccharomyces cerevisiae (yeast) V-ATPase was synthesized and dissolved in two membrane-mimicking solvents: DMSO and SDS. For the first time the secondary structure of the putative TM7 segment from subunit a is obtained by the combined use of CD and NMR spectroscopy. SDS micelles reveal an α-helical conformation for peptide MTM7 and in DMSO three α-helical regions are identified by 2D 1H-NMR. Based on these conformational findings a new structural model is proposed for the putative TM7 in its natural environment. It is composed of 32 amino acid residues that span the membrane in an α-helical conformation. It starts at the cytoplasmic side at residue T719 and ends at the luminal side at residue W751. Both the luminal and cytoplasmatic regions of TM7 are stabilized by the neighboring hydrophobic transmembrane segments of subunit a and the subunit c assembly from V-ATPase.  相似文献   

10.
The nonbonded interaction energy of disaccharides, mannobiose and galactobiose and polysaccharides mannan and galactan have been computed as a function of dihedral angles (?,ψ). The conformation (40°, ?20°) has been preferred for the mannan chain from nonbonded interaction energy considerations. The O5…O3′ type of intramolecular hydrogen bond has been found to be possible in the above conformation. Comparison of the allowed region of mannan with those of cellulose and xylan indicates that the monomer unit, in mannan chain has slightly higher freedom of rotation than that of cellulose and less than that of xylan. As in cellulose and mannan, the freedom of rotation of the monomer units in β-1,4′ galactan is highly restricted. Unlike mannan (which prefers an extended conformation) the β-1,4′ galactan prefers a helical conformation similar to amylose. Just as in amylose the O2…O3′ type hydrogen bond between contiguous residues is also possible in β-1,4′ galactan.  相似文献   

11.
The structures of two crystal forms of the RNA 16-mer with the sequence GUGGUCUGAUGAGGCC, grown in the presence of a high concentration of sulphate ions, have been determined using synchrotron radiation at 1.4- and 2.0-Å resolution. RNA with this sequence is known as one of the two strands of the noncleavable form of the hammerhead ribozyme. In both crystal structures, two G(syn)–G(anti) noncanonical base pairs are observed in the middle of a 14 base-pair (bp) duplex having 5′-dangling GU residues. Both structures contain sulphate anions interacting with the G–G bp stabilizing G in its syn conformation and bridging the two RNA strands. In both cases the interactions take place in the major groove, although the anions are accommodated within different helix geometries, most pronounced in the changing width of the major groove. In one structure, where a single sulphate spans both G–G pairs, the major groove is closed around the anion, while in the other structure, where each of the two G–G pairs is associated with a separate sulphate, the groove is open. This work provides the first examples of a G–G pair in syn-anti conformation, which minimizes the purine–purine clash in the center of the duplex, while utilizing its residual hydrogen bonding potential in specific interactions with sulphate anions.  相似文献   

12.
Conformation in aqueous solution at pH 6.95 of tripeptides having cyclic dipeptide backbones, cyclo[l-Glu(l-Leu-OBzl)-l-His] and cyclo[l-Glu(l-Leu-OH)-l-His], was investigated by u.v., c.d. and n.m.r. spectroscopy and by the lanthanide probe method. In the major conformation of cyclo[l-Glu(l-Leu-OBzl)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechain of the l-His residue is nearly parallel with the backbone plane and the sidechain of the l-Glu residue protrudes outside the backbone plane. In the major conformation of cyclo[l-Glu(l-Leu-OH)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechains of the l-His and l-Glu residues are accommodated in the same side of the backbone plane so that the imidazolyl sidechain of l-His residue is twisted slightly. Tripeptides were not found to change the conformation when metal salts or ammonium salts such as Cl?H3N?(CH2)11 COOEt, Gly-OEt-HCl, dl-Val-OEt-HCl and l-Leu-OEt-HCl were added, but a significant conformation change occurred upon adding d-Leu-OEt·HCl. If the same situation holds with the addition of α-amino acid p-nitrophenyl ester hydrochlorides, the previously reported enantiomer-selective catalysis by the tripeptides which hydrolysed d-Leu-OPh(NO2·HCl faster than l-Leu-OPh(NO2)·HCl can be explained; that is, the tripeptides change the conformation only when d-Leu-OPh(NO2)·HCl is bound and consequently the intramolecular reaction is facilitated. This phenomenon may be compared with that of ‘induced fit’ in enzyme catalysis.  相似文献   

13.
From φ, ψ data on eleven proteins, a 20 × 20 × 20 Table of tripeptides has been computed to evaluate the influence of nearest neighbors (n ? 1) and (n + 1) on the φ, ψ angles of amino acid (n). From this Table, having removed values for horse cytochrome c and using the sequences of 18 cytochromes c and the procedure of Kabat &; Wu (1972), an attempt was made to select a set of φ, ψ angles for positions 2 to 103 of cytochrome c and compare them with the values obtained from the atomic co-ordinates. Agreement was good for 56, intermediate for 29 and poor for 17 residues. Eleven of the 17 with poor agreement were residues contacting the heme or adjacent to a contacting residue. Moreover, 6 of the 17 poor values were in regions of the φ, ψ plot for which no occurrences in the ten known proteins were reported, and for four others known values were minimal so that no basis for selection existed. Frequency distributions on the Ramachandran plot (Ramachandran &; Sasisekharan, 1968) of all φ, ψ values in the eleven known proteins are given as well as a contour plot for such frequencies. The uses and limitations of the procedure are discussed and the need for obtaining accurate estimates of errors in experimentally determined φ, ψ angles is emphasised.  相似文献   

14.
The structure of a dextran produced by Leuconostoc mesenteroides NRRL B-1397 has been investigated in relation to its immunological properties. The methylated dextran yielded on acid hydrolysis 2,3,4,6-tetra-, 2,3,4-tri-, 3,4,-di-, and 2,4-di-O-methyl-d-glucose, in the molar ratio of 1.0:3.1:0.7:0.2, together with a trace of 2,4,6-tri-O-methyl-dglucose, indicating that the branches occur mainly at O-2 and the remainder at O-3. A carboxyl-dextran, obtained by catalytic oxidation of the dextran to convert the terminal, non-reducing d-glucose residues d-glucuronic acid residues, was partially hydrolyzed with acid. Fractionation gave 2-O-(α-d-glucopyranosyluronic acid) d-glucose (major), 6-O-(α-d-glucopyranosyluronic acid)-d-glucose, and mixtures of aldotri-, aldotetra-, and aldopentaouronic acid that contain both (1 → 6)- and (1 → 2)-d-glucosidic linkages. It is concluded that the branches at O-2 are mainly single d-glucose units, whereas those occurring at O-3 may be longer than two glucose units, forming a highly branched structure having an average repeating- unit of 5 sugar residues.  相似文献   

15.
A new program, TALOS-N, is introduced for predicting protein backbone torsion angles from NMR chemical shifts. The program relies far more extensively on the use of trained artificial neural networks than its predecessor, TALOS+. Validation on an independent set of proteins indicates that backbone torsion angles can be predicted for a larger, ≥90 % fraction of the residues, with an error rate smaller than ca 3.5 %, using an acceptance criterion that is nearly two-fold tighter than that used previously, and a root mean square difference between predicted and crystallographically observed (?, ψ) torsion angles of ca 12º. TALOS-N also reports sidechain χ1 rotameric states for about 50 % of the residues, and a consistency with reference structures of 89 %. The program includes a neural network trained to identify secondary structure from residue sequence and chemical shifts.  相似文献   

16.
Hylaseptin-4 (HSP-4, GIGDILKNLAKAAGKAALHAVGESL-NH2) is an antimicrobial peptide originally isolated from Hypsiboas punctatus tree frog. The peptide has been chemically synthetized for structural investigations by CD and NMR spectroscopies. CD experiments reveal the high helical content of HSP-4 in biomimetic media. Interestingly, the aggregation process seems to occur at high peptide concentrations either in aqueous solution or in presence of biomimetic membranes, indicating an increase in the propensity of the peptide for adopting a helical conformation. High-resolution NMR structures determined in presence of DPC-d38 micelles show a highly ordered α-helix from amino acid residues I2 to S24 and a smooth bend near G14. A large separation between hydrophobic and hydrophilic residues occurs up to the A16 residue, from which a shift in the amphipathicity is noticed. Oriented solid-state NMR spectroscopy show a roughly parallel orientation of the helical structure along the POPC lipid bilayer surface, with an insertion of the hydrophobic N-terminus into the bilayer core. Moreover, a noticeable pH dependence of the aggregation process in both aqueous and in biomimetic membrane environments is attributed to a single histidine residue (H19). The protonation degree of the imidazole side-chain might help in modulating the peptide-peptide or peptide-lipid interactions. Finally, molecular dynamics simulations confirm the orientation and preferential helical conformation and in addition, show that HSP-4 tends to self-aggregate in order to stabilize its active conformation in aqueous or phospholipid bilayer environments.  相似文献   

17.
Two l-arabino-d-galactan-containing glycoproteins having a potent inhibitory activity against eel anti-H agglutinin were isolated from the hot saline extracts of mature radish leaves and characterized to have a similar monosaccharide composition that consists of l-arabinose, d-galactose, l-fucose, 4-O-methyl-d-glucuronic acid, and d-glucuronic acid residues. The chemical structure features of the carbohydrate components were investigated by carboxyl group reduction, methylation, periodate oxidation, partial acid hydrolysis, and digestion with exo- and endo-glycosidases, which indicated a backbone chain of (1→3)-linked β-d-galactosyl residues, to which side chains consisting of α-(1→6)-linked d-galactosyl residues were attached. The α-l-arabinofuranosyl residues were attached as single nonreducing groups and as O-2- or O-3-linked residues to O-3 of the β-d-galactosyl residues of the side chains. Single α-l-fucopyranosyl end groups were linked to O-2 of the l-arabinofuranosyl residues, and the 4-O-methyl-β-d-glucopyranosyluronic acid end groups were linked to d-galactosyl residues. The O-α-l-fucopyranosyl-(1→2)-α-l-arabinofuranosyl end-groups were shown to be responsible for the serological, H-like activity of the l-arabino-d-galactan glycoproteins. Reductive alkaline degradation of the glycoconjugates showed that a large proportion of the polysaccharide chains is conjugated with the polypeptide backbone through a 3-O-d-galactosylserine linkage.  相似文献   

18.
A statistical mechanical model of protein conformation with medium-range interactions between theith and (i+k)th residues (k<-4) is presented. Two two-state models, an α-helix-coil and an extended-structure-coil model, are formulated using the same form of the partition function, but the two models are applied independently to predict the locations of α-helical, extended, and coil segments; in the relatively few cases (<2%) where the predictions from the two models are in conflict, the prediction is scored as an incorrect one. Two independent sets of statistical weights (one set for each model) are derived to describe the interactions between the 20 amino acid residues for each range of interactionk; they are evaluated by minimizing an objective function so that the probability profiles for the α-helix or extended structure, respectively, in proteins computed from these statistical weights correlate optimally with the experimentally observed native conformations of these proteins. Examination of the resulting statistical weights shows that those for the interactions between hydrophobic residues and between a hydrophobic and a hydrophilic residue have reasonable magnitudes compared to what would be expected from the spatial arrangements of the side chains in the α-helix and the extended structure, and that those for the α-helix-coil model correlate well with experimentally determined values of the Zimm-Bragg parameterss and σ of the helix-coil transition theory. From the point of view of a method to predict the conformational states (i.e., α-helix, extended structure, and coil) of each residue, the statistical weights (as inall empirical prediction schemes) depend very much on the proteins used for the data base, since the presently available set of proteins of known structure is still too small for very high predictability; as a result, the correctness of the prediction is not very good for proteins not included in the data base. However, the correctness of the prediction, at least for the 37 proteins utilized as the data base in this study, is 91% and 87% for the α-helix-coil and the extended-structure-coil models, respectively; further, 79% of all the residues are predicted correctly when both the α-helix-coil and extended-structure-coil models are applied independently.  相似文献   

19.
An Wang 《Carbohydrate research》2010,345(9):1216-7185
We describe here the synthesis of two oligosaccharide fragments of the tumor associated carbohydrate antigen LeaLex. While the linear lacto-N-triose I: β-d-Galp-(1→4)-β-d-GlcNAcp-(1→3)-β-d-Galp-OMe is a known compound, this is the first reported preparation of the branched tetrasaccharide β-d-GlcNAcp-(1→3)-β-d-Galp-(1→4)-[α-l-Fucp-(1→3)]-β-d-GlcNAcp-OMe. Our synthetic schemes involved using an N-trichloroacetylated trichloroacetimidate glucosaminyl donor activated with excess TMSOTf at 0 °C for glycosylation at O-3 of galactosyl residues and that of trichloroacetimidate galactosyl donors activated with excess BF3·OEt2 to glycosylate either O-3 or O-4 of glucosamine residues. The fucosylation at O-3 of the glucosamine acceptor was accomplished using a thiofucoside donor activated with copper(II) bromide and tetrabutylammonium bromide. Thus, syntheses of the protected tri- and tetrasaccharides were achieved easily and efficiently using known building blocks. Of particular interest, we also report that these protected oligosaccharides were submitted to dissolving metal conditions (Na-NH3) to provide in one single step the corresponding deprotected compounds. Under these conditions all protecting groups (O-acyl, benzylidene, benzyl, and N-trichloroacetyl) were efficiently cleaved. The work-up procedure for such reactions usually involves quenching with excess methanol and then neutralization with acetic acid. In our work the neutralization was carried out using acetic anhydride rather than acetic acid to ensure N-acetylation of the glucosamine residue. Both fully deprotected compounds were then simply purified and desalted by gel permeation chromatography on a Biogel P2 column eluted with water.  相似文献   

20.
The structures of the two title C-glycopyranosylarene nucleosides have been determined by X-ray diffraction. The aim of this work was to relate the conformation around the extracyclic C-1C-7 bond to steric hindrance between the pyranose and benzene rings. The torsion angles observed in the two compounds (O-5C-1C-7C-8: +61,7° for 1, ?13,4° for 2) signify of a C-2 configurational modification. Moreover, the interaction between O-5 and an o-phenyl hydrogen could explain the particular conformation of the aryl substituent in 2.  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