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 578 毫秒
1.
3-Deoxy-d-manno-octulosonate 8-phosphate synthase (KDO8PS) catalyzes the reaction between phosphoenol pyruvate and d-arabinose 5-phosphate to generate KDO8P. This reaction is part of the biosynthetic pathway to 3-deoxy-d-manno-octulosonate, a component of the lipopolysaccharide of the Gram-negative bacterial cell wall. Two distinct groups of KDO8PSs exist, differing by the absolute requirement of a divalent metal ion. In this study Acidithiobacillus ferrooxidans KDO8PS has been expressed and purified and shown to require a divalent metal ion, with Mn2+, Co2+ and Cd2+ (in decreasing order) being able to restore activity to metal-free enzyme. Cd2+ significantly enhanced the stability of the enzyme, raising the Tm by 14 °C. d-Glucose 6-phosphate and d-erythrose 4-phosphate were not substrates for A. ferrooxidans KDO8PS, whereas 2-deoxy-d-ribose 5-phosphate was a poor substrate and there was negligible activity with d-ribose 5-phosphate. The 243AspGlyPro245 motif is absolutely conserved in the metal-independent group of synthases, but the Gly and Pro sites are variable in the metal-dependent enzymes. Substitution of the putative metal-binding Asp243 to Ala in A. ferrooxidans KDO8PS gave inactive enzyme, whereas substitutions Asp243Glu or Pro245Ala produced active enzymes with altered metal-dependency profiles. Prior studies indicated that exchange of a metal-binding Cys for Asn converts metal-dependent KDO8P synthase into a metal-independent form. Unexpectedly, this mutation in A. ferrooxidans KDO8P synthase (Cys21Asn) gave inactive enzyme. This finding, together with modest activity towards 2-deoxy-d-ribose 5-phosphate suggests similarities between the A. ferrooxidans KDO8PS and the related metal-dependent 3-deoxy-d-arabino-heptulosonate phosphate synthase, and highlights the importance of the AspGlyPro loop in positioning the substrate for effective catalysis in all KDO8P synthases.  相似文献   

2.
The tautomeric composition of a solution of ammonium 3-deoxy-d-manno-octulosonate (KDO, 1a) in D2O at 28° was assessed by means of 13 C-F.t.-n.m.r. spectroscopy. The results revealed the presence of 6?0 and 11 % of the α and β anomers of the pyranose, and 20 and 9 % of the two furanoses, and suggested, but did not unequivocally prove, that the major furanose form is the α anomer. To facilitate interpretation of the spectral results for 1, ammonium 3,5-dideoxy-d-arabino(or ribo)-octulosonate (3a) was prepared by the reaction of 5-deoxy-d-erythro-pentose with sodium oxalacetate at pH 11. A chromatographically homogeneous, noncrystalline sample of 3 was obtained by lyophilization, and characterized as its (4-nitrophenyl)hydrazone (m.p. 162-163°). The 13C-n.m.r. spectrum of a solution of 3a in D2O revealed it to be substantially all in the α-pyranose form. No signals were obtained for the possible 1,4-lactone of 3. As the 1,5-lactone and furanose forms are impossible for 3, it exhibited no signals analogous to those attributed to furanoid 1. On the basis of these results for 3, the two lactone forms of 1 were excluded from consideration, and the three pairs of 13C-n.m.r. signals observed at ≈45, 86, and 104 p.p.m. were assigned to the furanose forms of 1.  相似文献   

3.
Acid-catalysed condensation of methyl β-d-glycero-l-manno-heptopyranoside with cyclohexanone yielded an approximately 3:1 mixture of the 2,3:6,7- and 2,3:4,7-di-O-cyclohexylideneheptosides (1 and 2), which could be separated either as their benzoates (3 and 4) or as their methyl ethers (5 and 6). The latter compounds afforded the 4- and 6-methyl ethers (7 and 8) of d-glycero-l-manno-heptitol.  相似文献   

4.
2,6-Anhydro-1-deoxy-1-diazo-D-glycero-L-manno-heptitol (2) decomposes in 0.01M methanolic sodium methoxide with a half-life of approx. 18 min. Decomposition in aqueous solution is too rapid for spectrophotometric measurement. Seven products could be identified in methanolic and aqueous reaction mixtures. 2,6-Anhydro-1-deoxy-D-galacto-hept-1-enitol (6), 2,7-anhydro-1-deoxy-β-D-galacto-heptulopyranose (10), and 4-O-vinyl-D-lyxose (12) are products of rapid intramolecular reactions. The major portion consists of the direct solvolysis products 2,6-anhydro-1-O-methyl-D-glycero-L-manno-heptitol (3) and 2,6-anhydro-D-glycero-L-manno-heptitol (5).  相似文献   

5.
A series of l-pyroglutamic acid analogues from natural product lead were designed and synthesized, as well as their antifungal activities against Phytophthora infestans, neuritogenic activities, antibacterial activities and anti-inflammatory activities are described. The bioassays and SAR study showed that the majority of l-pyroglutamic acid esters have a significant antifungal activity against P. infestans, especially 2d and 2j demonstrated the best activities with EC50 values of 1.44 and 1.21?μg?mL?1, which were about seven times that of commercial azoxystrobin (7.85?μg?mL?1). Moreover, compounds 2e, 2g and 4d displayed anti-inflammatory activity against LPS-induced NO production in BV-2 microglial cells; neuritogenic activity in NGF-induced PC-12 cells is the same activity. This study demonstrates that compounds 2d and 2j are potential drugs to control P. infestans.  相似文献   

6.
Kinetic data for the oxidations of d-fructose and l-sorbose by chromium(VI) and vanadium(V) in perchloric acid medium are reported. The addition of perchloric acid and sodium perchlorate increases the pseudo-first-order rate constants. Change of the reaction medium from water to deuterium oxide appreciably affects the rates of chromium(VI) oxidations, but does not affect those of vanadium(V) oxidations. The activation parameters are ΔH3 = 46.6 ±3.4 (fructose) and 50.6 ±6.3 (sorbose) kJ.mol?1, and ΔS3 = ?105 ±11 (fructose) and ?100 ±20 (sorbose) J.deg?1.mol?1 for chromium(VI) oxidations, and, for the other reactions, ΔH3 = 53.2 ±4.2 (fructose) and 52.3 ±6.3 (sorbose) kJ.mol?1, and ΔS3 = ?139.0 ±14 (fructose) and ?137 ±20 (sorbose) J.deg?1.mol?1. The kinetics of the oxidations of ketohexoses by chromium(VI) indicate no intermediate-complex formation, whereas those for vanadium(V) indicate the formation of a 1:1 intermediate complex between ketohexoses and vanadium(V).  相似文献   

7.
Extracellular d-glucosyltransferases (GTase) and d-fructosyltransferases (FTase) were isolated from Streptococcus mutans IB (serotype c), B14 (e), and OMZ175 (f) by chromatofocusing, followed by hydroxyapatite column chromatography. The GTases isolated from serotypes c, e, and f are basic proteins (pI 7.4). The serotype c and e enzymes have two protein components having Mr 173 000 and 158 000 and the enzyme of the serotype f one component having Mr 156 000. The GTases of all the serotypes showed a Km value for sucrose of 10–14mm and an optimum pH 5.5–6.0 for enzyme activity, and their activities were enhanced by the presence of primer Dextran T10. The α-d-glucans synthesized by the purified GTases are water soluble and primarily consist of (1→6)-α-d-glucosidic linkage (41–66 mol/100 mol) and α-d-(1→3,6)-branch linkage (6–20 mol/100 mol), but significant proportions of α-d-(1→3), α-d-(1→4), and α-d-(1→3,4) linkages (11, 6, and 14 mol/100 mol, respectively) were detected in the serotype c α-d-glucan. The isolated FTases of the serotypes c, e, and f are acidic enzymes (pI 4.6) and consist of two components having Mr 84 000 and 76 000 for the serotype c enzyme, and 106 000 and 84 000 for the serotypes e and f enzymes, respectively. The Km value for sucrose was 6, 10, and 17mm for the serotypes c, e, and f enzymes, respectively, and the optimum pH of enzymic activity 5.5–6.0. Reactivity with Concanavalin A, susceptibility to acid hydrolysis, and paper chromatography of the hydrolyzates suggested that the water-soluble β-d-fructans synthesized by the purified FTases were of the inulin-type and had chemical structures somewhat different among the serotypes.  相似文献   

8.
Three-dimensional X-ray diffraction data were used to determine the crystal structure of sodium β-d-glucuronate monohydrate, a model system for investigating the factors involved in the binding of sodium ions to d-glucuronate residues of glycosaminoglycans. Crystals of the salt are monoclinic, space group P21, with a = 9.206(3) Å, b = 7.007(2) Å, c = 7.378(3) Å, β = 96.84(3)°, and Z = 2. Intensity data for 858 reflections were measured with an automated diffractometer. A trial structure, obtained by direct methods, was refined by least squares to R = 0.035. An outstanding feature of the crystal packing is the interaction of d-glucuronate anions with sodium ions. The sodium ion is coordinated to three symmetry-related d-glucuronate anions and to one water molecule. The d-glucuronate anion binds sodium cations through the three following sites: one that involves a carboxyl oxygen atom combined with ring oxygen O-5; one that includes a single carboxyl oxygen atom, and one composed of the O-3–O-4 pair of hydroxyl groups.  相似文献   

9.
When kept at 105° for 2.5 h, weakly alkaline, syrupy d-erythrose was readily converted into a mixture containing mainly d-glycero-tetrulose, the previously unknown β-d-altro-l-glycero-3-octulofuranose (2), and α-d-gluco-l-glycero-3-octulopyranose, which were isolated as the corresponding acetates. Treatment of 2 with Dowex 50 (H+) resin yielded 3,8-anhydro-β-d-altro-l-glycero-octulopyranose, identified as its acetate. Previous discrepancies in the [α]d values for d-erythrose appear partly to originate in the self-aldol reaction. The dimerisation of d-erythrose 4-phosphate is also described.  相似文献   

10.
Cyclic dipeptide cyclo(l- or d-Glu-l-His) carrying an anionic site and a nucleophilic site has been synthesized and used as a catalyst for the solvolysis of cationic esters in aqueous alcohols. In the solvolysis of 3-acyloxy-N-trimethylanilinium iodide (S+n, n = 2 and 10) and Cl?H3N+(CH2)11COOPh(NO2), no efficient nucleophilic catalysis was observed. On the other hand, in the solvolysis of Gly-OPh(NO2)·HCl, Val-OPh(NO2)·HCl and Leu-OPh(NO2)·HCl a very efficient general base-type catalysis by cyclo(l-Glu-l-His) was observed. In particular, with the latter two substrates the catalysis by cyclo(l-Glul-His) was more efficient than that by imidazole, although the catalysis was not enantiomer-selective. The diastereomeric cyclic dipeptide cyclo(d-Glu-l-His) was almost inactive under the same conditions. Confomation of cyclo(l- or d-Glu-l-His) in aqueous solution was investigated and the structure/catalysis relationship is discussed.  相似文献   

11.
d-glycero-d-manno-Heptopyranose 7-phosphate—an intermediate in the biosynthesis of nucleotide-activated heptoses—has been prepared in good overall yield from benzyl 5,6-dideoxy-2,3-O-isopropylidene-α-d-lyxo-(Z)-hept-5-enofuranoside by a short-step synthesis. Phosphitylation using the phosphoramidite procedure followed by in situ oxidation afforded the corresponding 7-O-phosphotriester derivative in high yield. Subsequent osmylation proceeded in good diastereoselectivity (4:1) to furnish the d-glycero-d-manno-configured derivative, which was separated from the l-glycero-l-gulo-isomer by chromatography. Hydrogenolysis led to simultaneous removal of the benzyl and isopropylidene groups and afforded the target compound in high yield, which serves as a substrate of bacterial heptose 7-phosphate kinases.  相似文献   

12.
Capillary g.l.c. on SE-30 of the trimethylsilylated (-)-2-butyl glycosides of d and l monosaccharides gives multiple peak patterns, which can be used for the assignment of the absolute configurations. (-)-2-Butyl glycosides can be prepared from monosaccharides or their methyl glycosides; consequently, for the analysis of oligo- or poly-saccharides, hydrolysis as well as methanolysis can be applied. Provided that the peaks of the (-)-2-butyl glycosides do not completely overlap, mixtures of monosaccharides can be analysed directly, as illustrated for the constituents of the cell-wall lipopolysaccharide from Salmonella typhimurium LT-2.  相似文献   

13.
An economically viable procedure for the isolation and purification of d-mannose from palm kernel was developed in this research. The palm kernel was catalytically hydrolyzed with sulfuric acid at 100 °C and then fermented by mannan-degrading enzymes. The solution after fermentation underwent filtration in a silica gel column, desalination by ion-exchange resin, and crystallization in ethanol to produce pure d-mannose in a total yield of 48.4% (based on the weight of the palm kernel). Different enzymes were investigated, and the results indicated that endo-β-mannanase was the best enzyme to promote the hydrolysis of the oligosaccharides isolated from the palm kernel. The pure d-mannose sample was characterized by FTIR, 1H NMR, and 13C NMR spectra.  相似文献   

14.
The hydrothermolysis of cellobiose in the range 180–249° has been studied. Kinetic analysis of the reaction showed that 60% of the cellobiose is converted into d-glucose, and 40% into other products. The rate (k1) of cellobiose disintegration is approximately eight times that (k2) of d-glucose. Thus, hydrothermolysis differs from acidic hydrolysis. Hydrothermolysis is not dependent on pH, at least in the range 3–7.  相似文献   

15.
The benzothiazoline (1, R1 = R2 = H) formed by the reaction of d-galactose with o-aminobenzenethiol gives bis[o-(α-d-galactofuranosylamino)benzenethiol]-mercury(II) (2, R = H) on treatment with mercury(II) acetate in refluxing acetic acid. O-Acetylation of the chelate occurs smoothly, and demercuration of the product with hydrogen sulphide gives the thiol (3, R1 - Ac, R2 = R3 = H) which, with catalytic acid or when kept in chloroform solution, isomerises to the thiazoline compound (1, R1 = Ac, R2 = H). Under mild acetylating conditions, this product (and the starting material) gives diastereoisomeric 2,3,4,5,6-penta-acetates (1, R1 = R2 = Ac), but appreciable reversion to thiol occurs with acyl chlorides, with the consequence that thioesters (3, R1 = R2 = Ac, R3 = H; R1 = Ac, R2 = Bz, R3 = H) were major products. The value of the tetraester (1, R1 = Ac; R2 = H) as a means of obtaining galactose derivatives specifically modified at C-4 is therefore limited.  相似文献   

16.
2,6-Anhydro-d-glycero-l-manno-heptose (1) is converted by the cyanohydrin reaction into crystalline d-threo-l-talo-octononitrile (3), which shows mutarotation in water. The equilibrium mixture, as measured by 13C-n.m.r. spectroscopy, contains about equal amounts of 3 and its epimer, d-threo-l-galacto-octononitrile. On evaporation of the aqueous mixture, pure, crystalline 3 is again obtained. Labelling experiments in 3H2O proved that epimerization proceeds through reversible deprotonation. Stabilization of 3 in the solid state is explained by intramolecular hydrogen-bonding. In pyridine, rapid isomerization of 3 occurs. When acetylation of 3 is conducted in this solvent, the yield of 2,4,5,6,8-penta-O-acetyl-3,7-anhydro-d-threo-l-talo-octono-nitrile (4) depends strongly on the conditions of acetylation. Acetylation after equilibration produces an equimolar mixture of 4 and its isomer 2,4,5,6,8-penta-O-acetyl-3,7-anhydro-d-threo-l-galacto-octononitrile. Structural assignment for both was achieved by 360-Mhz, 1H- and 13C-n.m.r. spectroscopy. Reduction of 4 in pyridine-acetic acid-water in the presence of N,N-diphenylethylenediamine yields a 1:2.36 mixture of 2,4,5,6,8-penta-O-acetyl-3,7-anhydro-d-threo-l-talo-octose N,N-diphenylimidazolidine (6) and 2,4,5,6,8-penta-O-acetyl-3,7-anhydro-d-threo-l-galacto-octose N,N-diphenylimidazolidine (8). Compounds 6 and 8 could be separated and obtained as crystalline solids, and their structure proved by 1H- and 13C-n.m.r. spectroscopy. Hydrolysis of 6 and 8 gave 2,4,5,6,8-penta-O-acetyl-3,7-anhydro-d-threo-l-galacto-octose and -d-threo-l-talo-octose.  相似文献   

17.
The binding of substrate and product analogs to phenylalanine ammonia-lyase (EC 4.3.1.5) from maize has been studied by a protection method. The ligand dissociation constants, KL, were estimated from the variation with [L] of the pseudo-first-order rate constants for enzyme inactivation by nitromethane. The phenylalanine analogs d- and l-2-aminooxy-3-phenylpropionic acid showed KL, values over 20,000-fold lower than the Km for l-phenylalanine. From these and other KL values it is deduced that when the enzyme binds l-phenylalanine the structural free energy stored in the protein is higher than when it binds the superinhibitors. Models for binding d- and l-phenylalanine and the superinhibitors are described. The enantiomeric pairs are considered to have similar KL values because they pack into the active site in a mirror-image relationship. If the elimination reaction approximates to the least-motion course deduced on stereoelectronic grounds, the mirror-image packing of the superinhibitors into the active site mimics the conformation inferred for a transition state in the elimination. It appears, therefore, that structural changes take place in the enzyme as the transition state conformation is approached causing stored free energy to be released. This lowers the activation free energy for the elimination reaction and accounts for the strong binding by the above analogs.  相似文献   

18.
An electron diffraction study was carried out on thin single micro-crystals of l-type and dl-type dipalmitoyl lecithins grown in xylene suspensions and fine net patterns were obtained and the mechanism of the thermotropic phase transitions of them was clarified.From the apparent structure of diffraction patterns in low temperature, it is confirmed that the two dimensional lattices have p mm symmetry in l-type and in dl-type lecithins. Lattice parameters from the [001] projection are d100 = 9.9 A? and d010 = 8.8 A? in l-type, and d100 = 17.2 A? and d010 = 8.9 A? in dl-type.With anisotropic variation of dimensions along a and b axes, i.e. contraction for a and expansion for b, induced by temperature rise by electron irradiation during the observation, these diffraction patterns of the lattices of l-type and dl-type were transformed into those characterized by the six diffraction spots having nearly the same spacings. Four of them are observed on slightly outer and two are slightly inner positions as compared with their mean spacings of about (4.1 Å)?1 in l-type and about (4.2 Å)?1 in dl-type. The changes in the patterns observed indicate that at low temperatures the hydrocarbon chains are nearly perpendicular to the layer in dl-type lipid, and tilted with a more complicated packing in l-type ones. The dimension along a in dl-type is twice as large as that in l-type.  相似文献   

19.
NG-Monoethyl-l-arginine, a putative in vivo product after administration of the potent hepatocarcinogen l-ethionine to rats, has been chemically synthesized by coupling N-ethyl, S-methylthiopseudouronium iodide with α-amino-blocked l-ornithine. The structure of the compound as NG-monoethyl-l-arginine was confirmed by 13C NMR. Its elution time on an automatic amino acid analyzer, Rf values using thin-layer chromatography, and isoelectric point have been compared with those of NG-monomethyl-l-arginine.  相似文献   

20.
Transport of l-proline into Saccharomyces cerevisiae K is mediated by two systems, one with a KT of 31 μM and Jmax of 40 nmol · s?1 · (g dry wt.)?1, the other with KT > 2.5 mM and Jmax of 150–165 nmol · s?1 · (g dry wt.)?1, The kinetic properties of the high-affinity system were studied in detail. It proved to be highly specific, the only potent competitive inhibitors being (i) l-proline and its analogs l-azetidine-2-carboxylic acid, sarcosine, d-proline and 3,4-dehydro-dl-proline, and (ii) l-alanine. The other amino acids tested behaved as noncompetitive inhibitors. The high-affinity system is active, has a sharp pH optimum at 5.8–5.9 and, in an Arrhenius plot, exhibits two inflection points at 15°C and 20–21°C. It is trans-inhibited by most amino acids (but probably only the natural substrates act in a trans-noncompetitive manner) and its activity depends to a considerable extent on growth conditions. In cells grown in a rich medium with yeast extract maximum activity is attained during the stationary phase, on a poor medium it is maximal during the early exponential phase. Some 50–60% of accumulated l-proline can leave cells in 90 min (and more if washing is done repeatedly), the efflux being insensitive to 0.5 mM 2,4-dinitrophenol and uranyl ions, to pH between 3 and 7.3, as well as to the presence of 10–100 mM unlabeled l-proline in the outside medium. Its rate and extent are increased by 1% d-glucose and by 10 μg nystatin per ml.  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