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.
Igor Tvaro&#x;ka 《Biopolymers》1982,21(9):1887-1897
The conformational equilibria of four β-maltose conformers have been studied theoretically in 12 solvents. The stability of the conformers in dilute solution has been compared by using the continuum reaction field method. The solvation energy consists of electrostatic, dispersion, and cavity terms, which have been determined from the calculated properties of β-maltose and physicochemical properties of solvents. The calculated population of four conformers significantly depends on the solvent (e.g., in dioxane, M1:M2:M3:M4 = 53.3: 20.3:17.7:8.7; in dimethyl sulfoxide, 40.1:21.8:22.8:15.3; and in water, 25.7:17.5:26.3:30.5) and was found to correlate with experimentally observed data. The results obtained indicate that the conformation adopted by β-maltose in the crystalline form is not the one preferred in solution. The roles of the individual contributions to the solvation energy have been analyzed. Based on the determined abundance of conformers, averaged residual optical activity and vicinal carbon–proton coupling constants have been calculated and discussed from the point of view of the solution behavior of β-maltose.  相似文献   

2.
Igor Tvaro&#x;ka 《Biopolymers》1984,23(10):1951-1960
The conformational equilibria of five β-cellobiose conformers have been studied theoretically in 10 solvents. The stability of the conformers in dilute solution has been compared by using the method that has already been tested for 2-methoxytetrahydropyran, β-maltose, and D -glucose. The solvation energy consists of electrostatic, dispersion, and cavity terms which have been determined from the properties of the solute calculated by the PCILO quantum-chemical method and physicochemical properties of the solvents. The calculated abundance of conformers depends on the solvent (e.g., in dioxane C1:C2:C3:C4:C5 = 60.0:34.1:2.9:2.0:1.0; in dimethylsulfoxide, 75.5:22.1:1.8:0.5:0.2; and in water, 82.2:16.2:1.3:0.2:0.1). The results obtained indicate that the preponderant conformer in the aqueous solution is similar to the one adopted by β-cellobiose in the crystalline form. The role of individual contributions to the solvation energy have been analyzed. Based on the determined abundance of conformers, averaged residual optical activity and nmr parameters have been calculated and compared with observable properties. The marked differences observed between solvent-induced conformational changes for β-cellobiose and β-maltose have been discussed from the viewpoint of the solubility of the cellulose.  相似文献   

3.
Benzylidenation of β-maltose monohydrate with α,α-dimethoxytoluene in N,N-dimethylformamide in the presence of p-toluenesulfonic acid gave, in 70% yield, 4′,6′-O-benzylidenemaltose, which was acetylated to afford, 1,2,3,6,2′,3′-hexa-O-acetyl-4′,6′-O-benzylidene-β-maltose (4). Removal of the benzylidene group of 4 gave 1,2,3,6,2′,3′-hexa-O-acetyl-β-maltose (5), which was transformed into 1,2,3,6,2′,3′,4′-hepta-O-acetyl-6′-O-p-tolylsulfonyl-β-maltose (8). Several 6′-substituted β-maltose heptaacetates were synthesized by displacement reactions of 8 with various nucleophiles. Condensation of 5 with 2,3,4,6-tetra-O-benzyl-α-d-glucopyranosyl bromide, under catalysis by halide ion, followed by removal of protecting groups, furnished panose in good yield.  相似文献   

4.
Arabinoxylan preparations from sugar cane show temperature-induced shifts of optical rotation in aqueous methyl sulphoxide solution, the sign and magnitude of which depend on the content of arabinofuranose side-groups. The shift has a sigmoidal form and shows a distinct hysteresis loop on heating and cooling. This evidence, together with the sign and magnitude, is interpreted in terms of a conformation change of the backbone of β-(1→4)-linked D-xylose residues, from the random coil to an ordered, ribbon-like conformation similar to that which is known to exist in the solid state. The driving force for this change is not intramolecular but derives from an aggregation that occurs simultaneously and can be detected by other methods. The arabinofuranose side-groups can be incorporated into the ordered assembly and contribute to the optical rotation shift, and they therefore have an unusual role compared with other polysaccharide side-chains. We conclude that the optical rotation shifts show a “melting” and re-formation of ordered associations which may imitate the natural biological cohesion between hemicellulose chains.  相似文献   

5.
Selective, double irradiation allows the assignment of most 13C-n.m.r. signals in a series of per-O-acetyl disaccharides composed of two D-glucose residues linked α-(1→3), β-(1→3), α-(1→4), β-(1→4), α-(1→6), β-(1→6), and α,α-(1→1). The main influences that affect the chemical shifts are discussed and the spectra of β-cellobiose octaacetate and β-maltose octaacetate are compared to those of cellulose and amylose triacetate, respectively, to show the possibilities and limitations of a disaccharide model for the interpretation of the 13C-spectrum of a polymer.  相似文献   

6.
6-Aminohexyl glycosides covalently linked to solid matrices are effective reagents for the isolation of proteins that bind to carbohydrates [Schnaar and Lee, Biochemistry, 14 (1975) 1535–1541], and for the study of interactions between intact cells and immobilized carbohydrates [Weigel et al., J. Biol. Chem., 253 (1978) 330–333]. The preparation of the 6-aminohexyl glycosides of the following D-pyranoses is now described: β-glucose, β-galactose, 2-acetamido-2-deoxy-β-glucose, α-mannose, β-maltose, β-melibiose, β-lactose, and β-cellobiose. These glycosides were prepared by glycosylation of 6-(trifluoroacetamido)hexanol with the appropriate acetylated glycosyl halide in 1:1 (v/v) benzene-nitromethane, with mercuric cyanide as the catalyst. Deacylation of the glycosides was achieved in two steps: use of sodium methoxide for O-deacetylation, and of an anion-exchange resin for N-de(trifluoroacetyl)ation.  相似文献   

7.
In a study undertaken to illustrate the inadequacy of the familiar concept of carbohydrases as hydrolases, crystalline α-amylases from six different sources, as well as crude salivary amylase, were examined and found to catalyze the synthesis of maltose and maltosaccharides from α-d-glucopyranosyl fluoride, a stereoanalog of α-d-glucopyranose. These syntheses apparently involve initial formation of maltosyl fluoride and higher maltosaccharide 1-fluorides, traces of which were found in digests with certain α-amylases. That the reactions are due to the α-amylases themselves and not to some accompanying enzyme(s) appears certain from the purity and diversity of the preparations; their failure (with one exception) to attack α- or β-maltose; the correspondence of the synthesized products with the known specificity of α-amylases for α-1,4-d-glucosidic linkages (and capacity of different α-amylases to hydrolyze saccharides of different sizes). The “saccharifying” α-amylase of B. sublilis var amylosacchariticus was unique in producing maltosaccharides from both α- and β-maltose (i.e., by α-d-glucosyl transfer). However, the entire group of α-amylases had the capacity to promote α-d-glucosyl transfer from α-d-glucosyl fluoride to C4-carbinol sites, demonstrating for the first time that the catalytic range of α-amylase extends beyond hydrolysis and its reversal. Indeed, all transferred the glucosyl group of α-d-glycosyl fluoride preferentially to C4-carbinols rather than water—a finding neither anticipated nor explained by the representation of α-amylases as hydrolases.  相似文献   

8.
Minimal basis set SCF-MO computations on conformations of α-D-gluco-pyranose, β-D-glucopyranose, and β-maltose resulting from empirical energy minimisation reproduce known trends in relative energy. Analysis of electron population and molecular orbital valency-state energy leads to separation of atoms into classes, two for oxygen, three for carbon, and two for hydrogen, the classes being correlated with the chemical environments of the atoms. Partitioning of the total energy into two-centre terms gives a quantification of non-bonded interactions, leading to potential energy curves for interactions of all types of atom present. Peculiar details of the electronic structure at and around the anomeric carbon atoms are noted: C(1s) chemical shift is insensitive to configuration, but depends on conformation.  相似文献   

9.
Anomeric forms of glucose and maltose produced from phenyl, p-nitrophenyl, p-tert-butylphenyl, p-ethylphenyl and p-chlorophenyl α-maltosides and maltopentaose by α- and β-amylases were determined quantitatively by a gas-liquid chromatographic method. All of the three kinds of α-amylases tested, B. subtilis saccharifying α-amylase, Taka-amylase A, and porcine pancreas α-amylase, were found to produce only α-maltose from the maltosides. Sweet potato and barley β-amylases produced β-maltose from maltopentaose.

Saccharifying α-amylase from B. subtilis also released α-maltose from all the maltosides mentioned above, contrary to the report by Shibaoka et al. that the enzyme released β-maltose from maltosides other than phenyl α-maltoside: FEBS Lett., 16, 33 (1971); J. Biochem., 77, 1215 (1975). It appears unlikely that the α-amylase releases β-maltose, depending on the kind of substrate.  相似文献   

10.
The peptidoglycan recognition protein PGRP-S is an innate immunity molecule that specifically interacts with microbial peptidoglycans and other pathogen-associated molecular patterns. We report here two structures of the unique tetrameric camel PGRP-S (CPGRP-S) complexed with (i) muramyl dipeptide (MDP) at 2.5 Å resolution and (ii) GlcNAc and β-maltose at 1.7Å resolution. The binding studies carried out using surface plasmon resonance indicated that CPGRP-S binds to MDP with a dissociation constant of 10−7 m, whereas the binding affinities for GlcNAc and β-maltose separately are in the range of 10−4 m to 10−5 m, whereas the dissociation constant for the mixture of GlcNAc and maltose was estimated to be 10−6 m. The data from bacterial suspension culture experiments showed a significant inhibition of the growth of Staphylococcus aureus cells when CPGRP-S was added to culture medium. The ELISA experiment showed that the amount of MDP-induced production of TNF-α and IL-6 decreased considerably after the introduction of CPGRP-S. The crystal structure determinations of (i) a binary complex with MDP and (ii) a ternary complex with GlcNAc and β-maltose revealed that MDP, GlcNAc, and β-maltose bound to CPGRP-S in the ligand binding cleft, which is situated at the interface of molecules C and D of the homotetramer formed by four protein molecules A, B, C, and D. In the binary complex, the muramyl moiety of MDP is observed at the C-D interface, whereas the peptide chain protrudes into the center of tetramer. In the ternary complex, GlcNAc and β-maltose occupy distinct non-overlapping positions belonging to different subsites.  相似文献   

11.
The mucilage extracted from the cladodes (modified stems) of Opuntia ficus-indica contains residues of d-galactose, d-xylose, l-arabinose, l-rhamnose, and d-galacturonic acid. Seasonal variation in the sugar composition of the mucilage has been investigated. Fractionation studies indicate that the mucilage is essentially homogeneous. The rate of release of the constituent sugars and the change in optical rotation on mild hydrolysis coupled with the results of chromic acid oxidation suggest that the mucilage contains α-arabinofuranosyl, β-xylopyranosyl, β-rhamnopyranosyl, β-galactopyranosyl, and α-galactopyranosyluronic acid residues. The results also suggest a core containing galacturonic acid, rhamnose, and galactose, to which xylose and arabinose are attached in peripheral positions.  相似文献   

12.
The conformational properties of xanthans with partially hydrolyzed side chains were in vestigated by optical rotation, CD, and differential scanning calorimetry (DSC). All variants displayed the well-known temperature-driven, cooperative order–disorder transition, and both optical rotation and DSC showed that the transition temperature was essentially independent of the content of terminal β-mannose. It was found that up to 80% of the changes in the specific optical rotation accompanying the transition reflects conformational changes linked to the terminal β-mannose in the side chains. Modification of the sidechains also affected the CD when xanthan was in the ordered state, but in this case the data suggest that the glucuronic acid is the major component determining the magnitude of the CD signal. DSC measurements showed that the transition enthalpy (ΔHcal) increased linearly with the fraction of β-mannose, again indicating that a significant part (up to 80%) of ΔHcal reflects conformational changes in the side chains. The conformational transition of the xanthan variants generally showed a higher degree of cooperativity (sharper transition) than unmodified, pyruvated xanthan. Calculation of the cooperativity parameter σ by means of the Zimm–Bragg theory (OR data) or from the ratio between ΔHcal and the van't Hoff enthalpy (ΔHvH) using DSC data showed a correlation between σ and the content of β-mannose, but the two methods gave different results when the content of β-mannose approached 100%. The ionic strength dependence of the transition temperature, expressed as d (log I)/d(T?1m), was nearly identical for intact xanthan and a sample containing only 6% of the terminal β-mannose. Application of the Manning polyelectrolyte theory does not readily account for the observed ΔHcal values, neither does it provide new information on the nature of the ordered and disordered conformations in xanthan. © 1993 John Wiley & Sons, Inc.  相似文献   

13.
Medium chain β-1-alkylglycosides show interesting mild detergent properties. Therefore, their synthesis and purification have been investigated and improved so as to permit preparation of 50–100 g amounts. Preparatory methods are presented for the already known compounds β-1-octyl-, β-1-nonyl and β-1-decyl-glucose and for the new compounds β-1-undecylglucose and β-1-dodecylmaltose. Some relevant properties such as melting point, optical rotation, critical micelle concentration and NMR-spectra have been determined. They illustrate the suitability of this class of detergents for membrane research.  相似文献   

14.
A trihydroxy derivative of phenylalanine was synthesized from S-methyl-L-cysteine and pyrogallol by the crystalline tyrosine phenollyase (L-tyrosine phenol-lyase (deaminating) EC 4. I. 99.2 formerly known as β-tyrosinase) from Escherichia intermedia. The product was isolated as its N-acetyl-triacetoxymethylester and identified as 2,3,4-trihydroxy-L-phenylalanine by the analyses of NMR, MS spectra and optical rotation.  相似文献   

15.
A chiroptical method of conformational analysis is applied to linear (1 → 3)-β-D -glucans and the dimeric analogues β- and α-laminaribioside. The method is based on a recently developed semiempirical calculational model for saccharide optical activity. We conclude that disaccharide conformational energy maps in the literature represent the effective potential energy surface in aqueous solution well. The positive optical rotation observed with long chains in dilute alkaline solution is not characteristic of any single–chain conformation, and must reflect chain association.  相似文献   

16.
The effects of solute and hydrochloric acid concentrations on optical rotation were studied using 20 naturally occuring amino acids.

There appeared to be no common factor among the amino acids as far as the inclination of optical rotation was concerned. Lutz-Jirgenson’s rule could be applied to few amino acids in the cationic form. Therefore, in the determination of the optical rotation, the concentration of the solute, nature of solvent and temperature must be rigorously controlled. The optical conditions of measurement and the specific rotation of 20 amino acids were recommended based on this work.  相似文献   

17.
We have synthesized (Pro-Pro-βAla)n as a model for collagen. The synthetic polytripeptide, mol wt 6500, exhibits a large negative optical rotation with a very strong negative Cotton effect centered at 216 nm. The optical rotatory dispersion of (Pro-Pro-βAla)n followed a single-term Drude equation and the λc was 195 nm. The rotation decreased markedly on heating with the midpoint of the broad transition at 55°C. Preliminary studies also showed loss of structure in guadinine HCl. The circular dichroism spectrum of the polymer exhibited a deep trough at 190 nm. The marked similarities of solution properties of (Pro-Pro-βAla)n to (Pro-Pro-Gly)n suggest that β-alanine can replace glycine in generating collagen-like helix in solution.  相似文献   

18.
This paper concerns the 13C NMR signal assignment in the DMSO of a neutral polysaccharide, scleroglucan. The previously proposed chemical structure is confirmed. The 13C NMR spectrum shows that scleroglucan is a regular poly (A, B, C, D) type glucan. The relaxation times of the different series of carbon atoms demonstrate that a single, pendant glucose group is attached to each third monomer along the main chain of what is a β(1 → 3)-glucan. Partial acid hydrolysis gives a spectrum analogous to that of the β(1 → 3)-d-glucan, curdlan, and confirms the structure of the polymer backbone.In aqueous solution, no signal has been obtained due to the existence of a rigid, ordered conformation as demonstrated by optical rotation; in the presence of sodium hydroxide, a conformational transition is produced just as with curdlan. The conclusion is that the behaviour of scleroglucan in solution is similar to that of other β(1 → 3)-d-glucans even though it is more soluble.  相似文献   

19.
The use of optical dielectrophoresis (ODEP) to manipulate microparticles and biological cells has become increasingly popular due to its tremendous flexibility in providing reconfigurable electrode patterns and flow channels. ODEP enables the parallel and free manipulation of small particles on a photoconductive surface on which light is projected, thus eliminating the need for complex electrode design and fabrication processes. In this paper, we demonstrate that mouse cells comprising melan-a cells, RAW 267.4 macrophage cells, peripheral white blood cells and lymphocytes, can be manipulated in an opto-electrokinetics (OEK) device with appropriate DEP parameters. Our OEK device generates a non-rotating electric field and exerts a localized DEP force on optical electrodes. Hitherto, we are the first group to report that among all the cells investigated, melan-a cells, lymphocytes and white blood cells were found to undergo self-rotation in the device in the presence of a DEP force. The rotational speed of the cells depended on the voltage and frequency applied and the cells'' distance from the optical center. We discuss a possible mechanism for explaining this new observation of induced self-rotation based on the physical properties of cells. We believe that this rotation phenomenon can be used to identify cell type and to elucidate the dielectric and physical properties of cells.  相似文献   

20.
Menthols were converted to Δ3-menthone enol acetate (VII) via menthones having only one asymmetric carbon atom. It was shown that the optical rotation of menthone enol acetate was proportional to the optical purity of starting menthols. Optical purity of original menthol could, therefore, be determined by optical rotation of menthone enol acetate derived from.  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