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.
A new type of molecular arrangement for dipeptides is observed in the crystal structure of l ‐phenyl­alanyl‐l ‐alanine dihydrate, C12H16N2O3·2H2O. Two l ‐Phe and two l ‐Ala side chains aggregate into large hydro­phobic columns within a three‐dimensional hydrogen‐bond network.  相似文献   

2.
The title compound, C20H21N3O3·0.75H2O, crystallizes as exceedingly thin fibers. The crystal packing arrangement is related to those of other hydro­phobic dipeptides with phenyl­alanine residues, but the structure has pseudo‐tetra­gonal symmetry in an ortho­rhom­bic space group with four peptide mol­ecules and three water mol­ecules in the asymmetric unit.  相似文献   

3.
The asymmetric unit in the crystal structure of the title compound, C15H22N2O3·0.88H2O, contains two peptide mol­ecules with completely different conformations. The structure is divided into hydro­phobic and hydro­philic layers, with channels of water mol­ecules at the layer interface.  相似文献   

4.
The structures of the title dipeptides, C9H18N2O4·0.33H2O, C12H16N2O4 and C8H16N2O4S·0.34H2O, complete a series of investigations focused on l ‐Xaa‐l ‐serine peptides, where Xaa is a hydro­phobic residue. All three structures are divided into hydro­philic and hydro­phobic layers. The hydro­philic layers are thin for l ‐phenyl­alanyl‐l ‐serine, rendered possible by an unusual peptide conformation, and thick for l ‐isoleucyl‐l ‐serine and l ‐methionyl‐l ‐serine, which include cocrystallized water mol­ecules on the twofold axes.  相似文献   

5.
The low‐temperature crystal and mol­ecular structure analyses of two modifications of l ‐alanyl‐l ‐tyrosyl‐l ‐alanine with water, C15H21N3O5·2.63H2O [(I), at 9 K], and ethanol, C15H21N3O5·C2H5O [(II), at 20 K], solvent mol­ecules in the crystal lattice show that the overall conformations of both modifications of the title tripeptide are practically the same. Moreover, despite the presence of different solvent mol­ecules in the crystal lattice, the specific inter­molecular inter­actions characteristic for individual tripeptide mol­ecules of (I) and (II) are conserved. The crystal packing of the two modifications of Ala‐Tyr‐Ala differ from each other only in the solvent region. The tight arrangements of tripeptide mol­ecules seem to be responsible for similar displacement parameters for all non‐H atoms, despite the different distances from the mol­ecular centre of mass. Comparison of the displacement parameters between the room‐ and low‐temperature structures shows that an average Ueq value decrease of about 80% takes place at 9 K [for (I)] and 20 K [for (II)] with respect to room temperature.  相似文献   

6.
One of the amino H atoms of l ‐phenyl­alanyl‐l ‐valine, C14H20N2O3, participates in a rare secondary interaction in being accepted by the aromatic ring of the phenyl­alanine side chain. The phenyl group is also a donor in a weak hydrogen bond to the peptide carbonyl group.  相似文献   

7.
The title amidino‐amino acids (a‐Hpro), C6H11N3O3·H2O, (I), and a‐Met, C6H13N3O2S·H2O, (II), respectively, exist in the form of zwitterions. The five‐membered pyrrolidine ring in (I) adopts an envelope conformation, with the Cγ atom out of the plane defined by the rest of the ring atoms, and with the hydroxyl and carboxyl­ate groups in a trans configuration relative to the ring plane. The two crystallographically independent zwitterions in (II) reveal quite different conformations of their side chains and a slightly different orientation of the guanidine moiety with respect to the carboxyl­ate group. The crystal structures of both (I) and (II) are stabilized by extensive networks of O—H·O, N—H·O and C—H·O hydrogen bonds, the network being three‐dimensional in (I) and two‐dimensional in (II).  相似文献   

8.
The crystal structures of the four dipeptides l ‐seryl‐l ‐asparagine monohydrate, C7H13N3O5·H2O, l ‐seryl‐l ‐tyrosine monohydrate, C12H16N2O5·H2O, l ‐tryptophanyl‐l ‐serine monohydrate, C14H17N3O4·H2O, and l ‐tyrosyl‐l ‐tryptophan monohydrate, C20H21N3O4·H2O, are dominated by extensive hydrogen‐bonding networks that include cocrystallized solvent water molecules. Side‐chain conformations are discussed on the basis of previous observations in dipeptides. These four dipeptide structures greatly expand our knowledge on dipeptides incorporating polar residues such as serine, asparagine, threonine, tyrosine and tryptophan.  相似文献   

9.
The valine side chains in the crystal structure of the title compound [systematic name: 2‐(2‐ammonio‐3‐methyl­butan­amido)‐3‐hydroxy­propano­ate tri­hydrate], C8H16N2O4·3H2O, stack along an a axis of 4.77 Å to form hydro­phobic columns surrounded by remarkable water/hydroxyl shells. The peptide main chains are connected by hydrogen bonds in two‐dimensional layers. The peptide mol­ecules in each layer are related only by translation, and generate a very rare pattern. This is rendered possible through the formation of the shortest Cα—H·O(carboxyl­ate) inter­action ever recorded.  相似文献   

10.
The peptide bond in the crystal structure of the title compound, C8H16N2O4, deviates substantially from planarity in the same manner as in other l ‐Ser‐l ‐Xaa dipeptides, where Xaa is a hydro­phobic residue.  相似文献   

11.
The Rose Bengal‐sensitized photooxidations of the dipeptides l ‐tryptophyl‐l ‐phenylalanine (Trp‐Phe), l ‐tryptophyl‐l ‐tyrosine (Trp‐Tyr) and l ‐tryptophyl‐l ‐tryptophan (Trp‐Trp) have been studied in pH 7 water solution using static photolysis and time‐resolved methods. Kinetic results indicate that the tryptophan (Trp) moiety interacts with singlet molecular oxygen (O2(1Δg)) both through chemical reaction and through physical quenching, and that the photooxidations can be compared with those of equimolecular mixtures of the corresponding free amino acids, with minimum, if any, influence of the peptide bond on the chemical reaction. This is not a common behavior in other di‐ and polypeptides of photooxidizable amino acids. The ratio between chemical (kr) and overall (kt) rate constants for the interaction O2(1Δg)‐dipeptide indicates that Trp‐Phe and Trp‐Trp are good candidates to suffer photodynamic action, with krlkt values of 0.72 and 0.60, respectively (0.65 for free Trp). In the case of Trp‐Tyr, a lower krlkt value (0.18) has been found, likely as a result of the high component of physical deactivation of O2(1Δg) by the tyrosine moiety. The analysis of the photooxidation products shows that the main target for O2(1Δg) attack is the Trp group and suggests a much lower accumulation of kynurenine‐type products, as compared with free Trp. This is possibly because of the occurrence of another accepted alternative pathway of oxidation that gives rise to 3a‐oxidized hydrogenated pyrrolo[2,3‐b]indoles.  相似文献   

12.
The structure of the title compound, C15H22N2O3·2H2O, was derived from data collected on a very thin twinned needle. The peptide mol­ecule is in a rare conformation normally associated with hydro­phobic dipeptides that form nanotubes. Nevertheless, the present structure is divided into hydro­phobic and hydro­philic layers.  相似文献   

13.
The Cα—C′—N—Cα (ω) torsion angle of the peptide bond in the crystal structure of the title compound, C8H16N2O4, is 157.37 (15)°. This is the second‐largest deviation from planarity observed for a small linear peptide.  相似文献   

14.
The crystal structure of methyl 4‐O‐β‐l ‐fuco­pyran­osyl α‐d ‐gluco­pyran­oside hemihydrate C13H24O10·0.5H2O is organized in sheets with antiparallel strands, where hydro­phobic interaction accounts for partial stabilization. Infinite hydrogen‐bonding networks are observed within each layer as well as between layers; some of these hydrogen bonds are mediated by water mol­ecules. The conformation of the disaccharide is described by the glycosidic torsion angles: ?H = ?6.1° and ψH = 34.3°. The global energy minimum conformation as calculated by molecular mechanics in vacuo has ?H = ?58° and ψH = ?20°. Thus, quite substantial changes are observed between the in vacuo structure and the crystal structure with its infinite hydrogen‐bonding networks.  相似文献   

15.
Ethane‐1,1,2‐trisphosphonic acid crystallizes as a hemihydrate, C2H9O9P3·0.5H2O, in which the water O atom lies on an inversion centre in the space group P21/c. The acid component, which contains a short but noncentred O—H...O hydrogen bond, adopts a gauche conformation. The acid components are linked by an extensive series of O—H...O hydrogen bonds to form layers, which are linked into pairs by the water molecules.  相似文献   

16.
The crystal structure of N‐(l ‐2‐amino­butyryl)‐l ‐alanine, C7H14N2O3, is closely related to the structure of l ‐alanyl‐l ‐alanine, both being tetragonal, while the retro‐analogue 2‐(l ‐alanyl­amino)‐l ‐butyric acid 0.33‐hydrate, C7H14N2O3·­0.33H2O, forms a new type of molecular columnar structure with three peptide mol­ecules in the asymmetric unit.  相似文献   

17.
The X‐ray crystal structure of the title compound, C8H15N3O4·H2O, at 20 K (space group P21) reveals that the mol­ecular conformation of the tripeptide is remarkably different from the water‐free form (space group P212121) reported previously [Padiyar & Seshadri (1996), Acta Cryst. C 52 , 1693–1695].  相似文献   

18.
Pd@CeO2 core–shell nanostructures with a tunable Pd core size, shape, and nanostructure as well as a tunable CeO2 sheath thickness were obtained by a biomolecule‐assisted method. The synthetic process is simple and green, as it involves only the heating of a mixture of Ce(NO3)3, l ‐arginine, and preformed Pd seeds in water without additives. Importantly, the synthesis is free of thiol groups and halide ions, thus providing a possible solution to the problem of secondary pollution by Pd nanoparticles in the sheath‐coating process. The Pd/CeO2 nanostructures can be composited well with γ‐Al2O3 to create a heterogeneous catalyst. In subsequent tests of catalytic NO reduction by CO, Pd@CeO2/Al2O3 samples based on Pd cubes (6, 10, and 18 nm), Pd octahedra (6 nm), and Pd cuboctahedra (9 nm) as well as a simply loaded Pd cube (6 nm)–CeO2/Al2O3 sample were used as catalysts to investigate the effects of the Pd core size and shape and the hybrid nanostructure on the catalytic performance.  相似文献   

19.
l ‐Valine, l ‐leucine, l ‐isoleucine, l ‐phenylalanine, and l ‐tyrosine are important proposed biomarkers for the early detection and diagnosis of type 2 diabetes. A simple and selective hydrophilic interaction chromatography with tandem mass spectrometry method was developed for the simultaneous determination of these amino acids in human serum, using stable isotope‐labeled amino acids as internal standards. Chromatographic separation was carried out on a Syncronis HILIC column (150 mm × 2.1 mm, 5 μm) with the column temperature of 35°C and a mobile phase consisted of acetonitrile/120 mM ammonium acetate (89:11, v/v), and the run time was 11.0 min. The mass spectrometric analysis was performed using a QTRAP 5500 mass spectrometer coupled with an electrospray ionization source in positive ion mode. As these five amino acids are endogenous compounds in serum, we used the corresponding stable isotope‐labeled amino acids to evaluate the matrix effect and recovery in serum. The matrix effect was 98.7–107.3%, and the recovery was 92.7–102.3%. Calibration curves spiked unlabeled amino acids in water were linear over the range of 0.200–100 μg/mL. The accuracy, inter‐, and intraday precision were below 10.2%. Analytes were stable during the study. This assay method has been validated and applied to the early diagnosis research of type 2 diabetes.  相似文献   

20.
This work reports the one‐pot enzymatic cascade that completely converts l ‐arabinose to l ‐ribulose using four reactions catalyzed by pyranose 2‐oxidase (P2O), xylose reductase, formate dehydrogenase, and catalase. As wild‐type P2O is specific for the oxidation of six‐carbon sugars, a pool of P2O variants was generated based on rational design to change the specificity of the enzyme towards the oxidation of l ‐arabinose at the C2‐position. The variant T169G was identified as the best candidate, and this had an approximately 40‐fold higher rate constant for the flavin reduction (sugar oxidation) step, as compared to the wild‐type enzyme. Computational calculations using quantum mechanics/molecular mechanics (QM/MM) molecular dynamics (MD) showed that this improvement is due to a decrease in the steric effects at the axial C4‐OH of l ‐arabinose, which allows a reduction in the distance between the C2‐H and flavin N5, facilitating hydride transfer and enabling flavin reduction.  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