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 0 毫秒
1.
The kinetics of α-methyl-d-glucoside accumulation by rat kidney cortex slices under conditions of varying extracellular pH are compared with values obtained at pH 7.4. At pH below 7.4 there is a diminished initial uptake and reduced influx of the sugar which results in a decrease in the steady-state intracellular pool. This was associated with a decrease in the V of the entry process without affecting the apparent Km of transport. At pH 8 there is no change in the rate of glucoside entry. The efflux of the glucoside, however, is impaired and the steady-state concentration gradient becomes greater than that observed at pH 7.4.  相似文献   

2.
The uptake of α-methyl-d-glucoside was stimulated in slices of rat kidney cortex by pretreatment with reduced glutathione. Diamide, an oxidizing agent with high specificity for GSH, caused an inhibition of α-methyl-d-glucoside uptake. These effects appeared to be related specifically to GSH, since dithiothreitol and mercaptoethanol did not increase α-methyl-d-glucoside uptake, and were not as effective as GSH in reversing the effects of diamide. GSH and diamide had no effect on the uptake of another sugar analog, 3-O-methylglucose, which is not actively transported. Kinetic studies indicated that GSH increased the apparent V without affecting Km. The results are discussed in terms of the possible role of GSH in the process of sugar transport.  相似文献   

3.
α-Methyl-d-glucoside has been shown to be a non-metabolizable sugar which is accumulated against a concentration gradient by a Na+-dependent and phlorizin inhibited process by adult rat renal cortical slices incubatedin vitro at 37 °C. (2) The velocity of accumulation increased linearly with substrate concentrations up to 1.5 mM, but at higher concentrations obeyed saturable kinetics with an apparentKm of about 6 mM. (3) Uptake was enhanced as Na+ was increased from 0 to 100 mequiv/l. Higher Na+ concentrations caused no further effect. (4) A pH maximum of transport occurred between 7.35 and 8.0. (5) Glucoside uptake was inhibited byd-glucose,d-galactose,d-fructose,d-mannose andd-ribose. The inhibition byd-glucose andd-galactose was competitive with apparentKt of 24 and 53 mM, respectively. (6) Bothd-glucose andd-galactose accelerated the efflux of α-methyl-d-glucoside from preloaded cells. (7) Kidney cortex slices from 1-day-old rats were unable to accumulate α-methyl-d-glucoside to form a concentration gradient. The ability to concentrate the glucoside increased progressively after birth, reaching near normal in tissue from 15-day-old animals. The data indicate that the transport process in the newborn is rudimentary, failing also to display accelerated efflux phenomenon. (8) α-Methyl-d-glucoside is transported in rat kidney cortex by a mechanism similar in many ways to that ofd-galactose.  相似文献   

4.
Cyclic adenosine 3′,5′-monophosphate and N6-2′-O-dibutyryl cyclic adenosine 3′,5′-monophosphate increase the accumulation of α-methyl-d-glucoside by cortical slices from rat, rabbit, dog and human kidney. The characteristics of the effect have been studied in rat tissue. At least 90 min of exposure of the tissue to cyclic nucleotide prior to onset of glucoside accumulation is required as well as presence of the cyclic nucleotide during the accumulation phase. Inhibition of protein synthesis does not abolish the effect of N6-2′-O-dibutyryl cyclic adenosine 3′,5′-monophosphate. The cyclic nucleotide causes an increase in the initial entry rate of α-methyl-d-glucoside into cells and an increase in the intracellular steady state concentration. The cyclic nucleotide does not affect the apparent Km of the glucoside entry process but increases the maximum velocity of accumulation.  相似文献   

5.
Dizene dicarboxylic acid bis-(N,N-dimethylamide), commonly called diamide, is known to oxidize stoichiometrically intracellular pools of reduced glutathione and inhibit the accumulation of sugars and amino acids by rat kidney slices. Incubation of rat renal cortical slices in diamide also leads to a significant decrease in the level of endogenous protein kinase activity. The inhibition of sugar and amino acid transport and protein kinase activity by diamide is partially reversible by the addition of exogenous glutathione or other thiols. A comparison of protein kinase activity with amino acid and sugar transport at various concentrations of diamide indicates that there is a high degree of correlation between these two processes.  相似文献   

6.
1-O-Tosyl-d-glucopyranose derivatives having a nonparticipating benzyl group at O-2 have been shown to react rapidly in various solvents with low concentrations of alcohols, either methanol or methyl 2,3,4-tri-O-benzyl-α-d-glucopyranoside. The stereospecificity of the glucoside-forming reaction could be varied from 80% of β to 100% of α anomer by changing the solvent or modifying the substituents on the 1-O-tosyl-d-glucopyranose derivative. 2,3,4-Tri-O-benzyl-6-O-(N-phenylcarbamoyl)-1-O-tosyl-α-d-glucopyranose in diethyl ether gave a high yield of α-d-glucoside. Kinetic measurements of reaction with various alcohols (methanol, 2-propanol, and cyclohexanol) show a high rate even at low concentrations of alcohol, and give some insight into the reaction mechanism. The high rate and stereoselectivity of their reaction suggest that the 1-O-tosyl-d-glucopyranose derivatives may be used as reagents for oligosaccharide synthesis.  相似文献   

7.
The intestinal transport of three actively transported sugars has been studied in order to determine mechanistic features that, (a) can be attributed to stereospecific affinity and (b) are common.The apparent affinity constants at the brush-border indicate that sugars are selected in the order, β-methyl glucose >d-galactose > 3-O-methyl glucose, (the Km values are 1.23, 5.0 and 18.1 mM, respectively.) At low substrate concentrations the Kt values for Na+ activation of sugar entry across the brush-border are: 27.25, and 140 mequiv. for β-methyl glucose, galactose and 3-O-methyl glucose, respectively. These kinetic parameters suggest that Na+, water, sugar and membrane-binding groups are all factors which determine selective affinity.In spite of these differences in operational affinity, all three sugars show a reciprocal change in brush-border entry and exit permeability as Ringer [Na] or [sugar] is increased. Estimates of the changes in convective velocity and in the diffusive velocity when the sugar concentration in the Ringer is raised reveal that with all three sugars, the fractional reduction in convective velocity is approximately equal to the (reduction of diffusive velocity)2. This is consistent with the view that the sugars move via pores in the brush-border by convective diffusion.Theophylline reduces the serosal border permeability to β-methyl glucose and to 3-O-methyl glucose relatively by the same extent and consequently, increases the intracellular accumulation of these sugars.The permeability of the serosal border to β-methyl glucose entry is lower than permeability of the serosal border to β-methyl glucose exit, which suggests that β-methyl glucose may be convected out of the cell across the lateral serosal border.  相似文献   

8.
Methyl α-d-mannopyranoside (1 mole) reacts with 2,2-dimethoxypropane (1 mole), to give the 4,6-O-isopropylidene derivative (2) which rearranges to the 2,3-O-isopropylidene derivative (4). Compound4 can also be prepared by graded hydrolysis of methyl 2,3:4,6-di-O-isopropylidene-α-d-mannopyranoside. Successive benzoylation, oxidation, and reduction of4 provides a useful route to a number ofd-talopyranoside compounds. Methyl α-d-mannofuranoside (1 mole) reacts with 1–2 moles of 2,2-dimethoxypropane to give the 5,6-O-isopropylidene derivative (16) in 90% yield.  相似文献   

9.
Reaction of 2,3,5-tri-O-benzyl-d-ribofuranosyl bromide with mercuric cyanide afforded an anomeric mixture of cyanides (3) and 1,4-anhydro-2,3,5-tri-O-benzyl-d-erythro-pent-1-enitol (6). Reduction of 3 with lithium aluminum hydride gave a pair of epimeric amines (4 and 5), which were separated by chromatography and characterized by conversion into the known 2,5-anhydro-3,4,6-tri-O-benzyl-1-deoxy-1-ureido-d-allitol (7) and its epimer, 2,5-anhydro-3,4,6-tri-O-benzyl-1-deoxy-1-ureido-d-altritol (8). Compound 8 and its precursor were used for the synthesis of various “α-homonucleosides”.  相似文献   

10.
The rate constants for the hydrolysis of six alkyl and four aryl β-d-xylofuranosides in aqueous perchloric acid at various temperatures have been measured. The effects of varying the aglycon structure on the hydrolysis rate are interpreted in terms of two concurrent reactions. Either, the substrate, protonated on the glycosidic oxygen atom, undergoes a rate-limiting heterolysis to form a cyclic oxocarbonium ion, or, an initial rapid protonation of the ring oxygen is followed by a unimolecular cleavage of the five-membered ring, all subsequent reactions being fast. It is suggested that xylofuranosides having strongly electron-attracting aglycon groups react mainly by the former pathway, whereas the latter is more favourable for substrates having electron-repelling aglycon groups. The negative entropies of activation obtained with the latter compounds are attributed to the rate-limiting opening of the five-membered ring. The rate variations of the hydrolyses of alkyl β-d-xylofuranosides in aqueous perchloric acid-methyl sulfoxide mixtures are interpreted as lending further support for the suggested chance in mechanism.  相似文献   

11.
The formation of 1,6-anhydro-β-d-glucopyranose and several d-glucosyl oligosaccharides has been observed during the action of a purified, fungal glucosyltransferase (EC 2.4.1.24) on maltose. Such products are synthesized by a transglucosylation mechanism involving the formation of a d-glucosyl-enzyme complex and the displacement of the d-glucosyl group by appropriate acceptor-substrates. The formation of the 1,6-anhydro bond is a novel type of transfer reaction and occurs by displacement of the enzyme from the d-glucosyl-enzyme complex by the proton of the primary hydroxyl group of the same glucosyl group. This reaction is characterized by inversion of configuration at the position of glucosidic bond-cleavage of the substrate. Synthesis of the d-glucosyl oligosaccharides occurs by displacement of the d-glucosyl groups from the enzyme by suitable acceptor-substrates. In these cases, the reactions are characterized by retention of configuration of the d-glucosidic bonds of the substrate. The list of oligosaccharides produced from maltose includes nigerose, kojibiose, isomaltose, maltotriose, panose, isomaltotriose, and 6-O-d-glucosyl-panose. The identity of these compounds has been established by methylation analysis and enzymic hydrolysis. d-Glucose is also a product of the reaction and arises from both the reducing and the non-reducing groups of maltose.  相似文献   

12.
Reaction of 1,2-O-cyclopentylidene-α-d-glucofuranurono-6,3-lactone (2) with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl bromide (1) gave 1,2-O-cyclopentylidene- 5-O-(2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl)-α-d-glucofuranurono-6,3-lactone (3, 45%) and 1,2-O-cyclopentylidene-5-O-(2,3,4,6-tetra-O-acetyl-β-d-glucopyranosyl)-α-d-glucofuranurono-6,3-lactone (4, 38%). Reduction of 3 and 4 with lithium aluminium hydride, followed by removal of the cyclopentylidene group, afforded 5-O-α-(9) and -β-d-glucopyranosyl-d-glucofuranose (12), respectively. Base-catalysed isomerization of 9 yielded crystalline 5-O-α-d-glucopyranosyl-d-fructopyranose (leucrose, 53%).  相似文献   

13.
d-xylose is a highly anomer-specific, powerful and competitive inhibitor of plant α-galactosidases. The apparent inhibition constant (Ki) for the interaction of d-xylose with chick pea (Cicer arietinum) α-galactosidase and the apparent number of inhibitor molecules (n) bound per enzyme molecule, using p-nitrophenyl-α-d-galactopyranoside as substrate, were found to be 0.4 × 10?2M and 0.8, respectively.  相似文献   

14.
15.
The effect of phenols on the hydrolysis of substituted phenyl β-d-gluco- and β-d-xylo-pyranosides by β-d-glucosidase from Stachybotrys atra has been investigated. Depending on the glycon part of the substrate and on the phenol substituent, the hydrolysis is either inhibited or activated. With aryl β-d-xylopyranosides, transfer of the xylosyl residue to the phenol, with the formation of new phenyl β-d-xylopyranosides, is observed. With aryl β-d-glucopyranosides, such transfer does not occur when phenols are used as acceptors, but it does occur with anilines. A two-step mechanism, in which the first step is partially reversible, is proposed to explain these observations. A qualitative analysis of the various factors determining the overall effect of the phenol is given.  相似文献   

16.
Maltitol, crystallised from aqueous solution, has m.p. 146.5–147°, [α]d + 106.5° (water), and is orthorhombic with the space group P212121 and Z = 4, and with cell dimensions a = 8.166(5), b = 12.721(9), and c = 13.629(6) Å. The molecule shows a fully extended conformation with no intramolecular hydrogen-bonds. All nine hydroxyl groups are involved in intermolecular hydrogen-bond networks and in bifurcated, finite chains. The d-glucopyranosyl moiety has the 4C1 conformation, and the conformation about the C-5–C-6 bond is gauche-gauche. The d-glucitol residue has the bent [ap, Psc, Psc (APP)] conformation. The empirical formula for the solubility in water is C = 119.1 + 1.204 T + 4.137 × 10?2 T2 ? 7.137 × 10?4 T3 + 7.978 × 10?6 T4. The thermal properties are as follows: ΔHf = 13.5 kcal.mol?1, and Q = ?5.57 kcal.mol?1.  相似文献   

17.
The preparation of 2,3-di-O-benzoyl-4,6-O-benzylidene-α-d-glucopyranosyl-2-O-benzoyl-4,6-O-benzylidene-α-d-ribo-hexopyranosid-3-ulose (3) from 4,6:4′,6′-di-O-benzylidene-α,α-trehalose (1) via the 2,3,2′-tribenzoate 2 has been improved. Reduction of 3 with sodium borohydride gave 2-O-benzoyl-4,6-O-benzylidene-α-d-allopyranosyl 2,3-di-O-benzoyl-4,6-O-benzylidene-α-d-glucopyranoside (4), which was converted into the methanesulfonate 5 and trifluoromethanesulfonate 6. Displacement of the sulfonic ester group in 6 with lithium azide was very facile and afforded a high yield of 3-azido-2-O-benzoyl-4,6-O-benzylidene-3-deoxy-α-d-glucopyranosyl 2,3-di-O-benzoyl-4,6-O-benzylidene-α-d-glycopyranoside (7), whereas similar displacement in 5 proceeded sluggishly, giving a lower yield of 7 together with an unsaturated disaccharide (8). The azido sugar 7 was converted by conventional reactions into the analogous 2,3,2′-triacetate 9, the corresponding 2,3,2′-triol 10, and deprotected 3-azido-3-deoxy-α-d-glucopyranosyl α-d-glucopyranoside (11). Hydrogenation of 11 over Adams' catalyst furnished crystalline 3-amino-3-deoxy-α,α-trehalose hydrochloride (12), the overall yield from 3 being 35%.  相似文献   

18.
Water-insoluble, non-adherent α-d-glucans have been obtained from Streptococcus salivarius HHT under two sets of conditions: from a growing culture, or synthesized enzymically by using a glucosyltransferase. In the former case, the glucan ([α]d + 197°) was shown by methylation analysis to have a slightly branched structure containing a relatively high proportion (80 %) of (1→3)-α-d-glucosidic linkages, together with small proportions of (1→6)- and (1→4)-α-d-glucosidic linkages. The enzymically synthesized glucan had a much less-branched structure, containing 88 % of (1→3)-α-d-glucosidic linkages. Both glucans, on Smith degradation (sequential periodate oxidation, borohydride reduction, and mild acid hydrolysis), gave linear, (1→3)-α-d-glucosidic polysaccharides (yields, 82-90%) that constitute the backbone chains. The presence of small proportions of glycerol, erythritol, 1-O-α-d-glucosyl-d-glycerol, and also 2-O-α-d-glucosyl-d-erythritol in the products of Smith degradation suggests that the short side-chains are attached to the backbone chain by (1→4)-, (1→6)-, and (1→3)-α-d-glucosidic linkages  相似文献   

19.
The crystal structures of α-d-glucopyranosyl β-d-psicofuranoside and α-d-galactopyranosyl β-d-psicofuranoside were determined by a single-crystal X-ray diffraction analysis, refined to R1 = 0.0307 and 0.0438, respectively. Both disaccharides have a similar molecular structure, in which psicofuranose rings adopt an intermediate form between 4E and 4T3. Unique molecular packing of the disaccharides was found in crystals, with the molecules forming a layered structure stacked along the y-axis.  相似文献   

20.
De-etherification of 6,6′-di-O-tritylsucrose hexa-acetate (2) with boiling, aqueous acetic acid caused 4→6 acetyl migration and gave a syrupy hexa-acetate 14, characterised as the 4,6′-dimethanesulphonate 15. Reaction of 2,3,3′4′,6-penta-O-acetylsucrose (5) with trityl chloride in pyridine gave a mixture containing the 1′,6′-diether 6 the 6′-ether 9, confirming the lower reactivity of HO-1′ to tritylation. Subsequent mesylation, detritylation, acetylation afforded the corresponding 4-methanesulphonate 8 1′,4-dimethanesulphonate 11. Reaction of these sulphonates with benzoate, azide, bromide, and chloride anions afforded derivatives of β-D-fructofuranosyl α-D-galactopyranoside (29) by inversion of configuration at C-4. Treatment of the 4,6′-diol 14 the 1,′4,6′-triol 5, the 4-hydroxy 1′,6′-diether 6 with sulphuryl chloride effected replacement of the free hydroxyl groups and gave the corresponding, crystalline chlorodeoxy derivatives. The same 4-chloro-4-deoxy derivative was isolated when the 4-hydroxy-1′,6′-diether 6 was treated with mesyl chloride in N,N-dimethylformamide.  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