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 interaction of heparin with human α-thrombin was investigated in the present report. Hydrolysis of synthetic tripeptide anilide substrates by thrombin was enhanced in the presence of heparin. With both N-α-benzoyl-l-phenylalanyl-l-valyl-l-arginine-p-nitroanilide (BzPheValArgNaN) and N-α-p-tosyl-l-glycyl-l-prolyl-l-arginine-p-nitroanilide (TosGlyProArgNaN), saturating concentrations of heparin enhanced the binding of substrate two-to threefold as determined by a decrease in the apparent Michaelis constant value, while having a marginal inhibitory effect on V. Substrate inhibition was observed with BzPheValArgNaN, which was enhanced in the presence of heparin. The enhancing effect of heparin on the binding of TosGlyProArgNaN was used to determine a dissociation constant value of 1.7 × 10?9m for the heparin · thrombin complex. This value is nearly two orders of magnitude lower than the dissociation constant value determined for the heparin · antithrombin III complex (B. Nordenman and I. Bjork, 1978, Biochemistry17, 3339–3344), suggesting strongly that heparin must bind to thrombin to account for the enhancing effect of heparin on the antithrombin III/thrombin reaction. Heparin also enhanced the rate of inactivation of thrombin by 1-chloro-3-tosylamido-7-amino-l-2-hepatonone, but had little effect on the inactivation rate with phenylmethanesulfonyl fluoride.  相似文献   

2.
The following new compounds were prepared and characterized: N-benzyl-oxycarbonyl-O-(tetra-O-acetyl-β-D-glucopyranosyl)-N-glycyl-L-serine methyl ester (1) and L-threonine methyl ester (2), N-benzyloxycarbonyl-O-(β-D-glucopyranosyl)-N-glycyl-L-serine amide (3), N-benzyloxycarbonyl-O-(β-D-glucopyranosyl)-N-glycyl-L-threonine methyl ester (4) and L-threonine amide (5), N-benzyloxycarbonyl-O-(tri-O-acetyl-2-deoxy-2-trifluoroacetamido-β-D-glucopyranosyl)-N-glycyl-L-serine methyl ester (6), and N-benzyloxycarbonyl-O-(2-deoxy-2-trifluoroacetamido-β-D-glucopyranosyl)-N-glycyl-L-serine amide (7). Although various modifications of the Koenigs-Knorr synthesis were used, the best, over-all yields of the deacetylated dipeptide derivatives were only 5–10%. Although the products are alkali-labile, deacetylation was accomplished with methanolic ammonia. Of the deacetylated products, the threonine derivatives (4 and 5) were more rapidly hydrolyzed by acids than phenyl β-D-glucopyranoside, which in turn was more rapidly cleaved than the serine derivatives (3 and 7). The stabilities of 3, 4, 5, and 7 to sodium hydroxide and sodium borohydride were similar, and essentially complete β-elimination of the glycosyl residue occurred for the amide derivatives (3, 5, and 7). For the ester derivative 4, pH 9 was optimal; above this pH, ester hydrolysis was more rapid than β-elimination, and the resulting carboxyl derivatives did not undergo β-elimination. Under optimal conditions with sodium borohydride, the β-elimination reaction was complete, but the corresponding alanine and α-aminobutyric acid residues were not formed; presumably reductions to the amino alcohols occurred. A mechanism for the β-elimination is proposed.  相似文献   

3.
The following ethers, of potential value for the synthesis of α-D-galactopyranosides, were prepared: 2-O-benzyl-D-galactose, 2,6-di-O-benzyl-D-galactose, and 2,3-di-O-benzyl-D-galactose. Isopropylidenation of methyl α-D-galactopyranoside in the presence of phosphorus pentaoxide gave its 3,4-, and 4,6-O-isopropylidene derivatives. Treatment of the 3,4-acetal with trityl chloride in pyridine produced the 6-trityl ether, which was benzylated with benzyl chloride and sodium hydride in N,N-dimethylformamide to yield the 2-benzyl ether. Acid hydrolysis of this product gave 2-O-benzyl-D-galactose. Benzylation of methyl 3,4-O-isopropylidene-α-D-galactopyranoside, followed by hydrolysis, gave 2,6-di-O-benzyl-D-galactose. Similarly, 2,3-di-O-benzyl-D-galactose was obtained by acid hydrolysis of methyl 2,3-di-O-benzyl-4,6-O-isopropylidene-α-D-galactopyranoside and of methyl 2,3-di-O-benzyl-4,6-O-benzylidene-β-D-galactopyranoside.  相似文献   

4.
Data are presented which demonstrate that the α-N-benzoyl-l-argine ethyl ester rate assay procedure, based on a burst titration with N-benzyloxy-carbonyl-l-tyrosine p-nitrophenyl ester as previously desribed (1), is an accurate and reliable method for determining the normality of papain in solution.  相似文献   

5.
Sequential digestion of human thrombin and antithrombin with neuraminidase, βgalactosidase, β-N-acetylglucosaminidase, and endo-β-N-acetylglucosaminidase D resulted in the successive removal of sialic acid, galactose, N-acetylglucosamine, and mannose and more N-acetylglucosamine residues. The products obtained after each stage of deglycosylation had electrophoretic mobilites that were consistent with the calculated change in mass expected from the cleavage of the sugar moieties. The modified thrombins did not lose fibrinogen-clotting activity, amidolytic activity, nor the ability to form complexes with antithrombin. In addition, asialothrombin and asialoagalactothrombin caused the same extent of platelet release as did control thrombin. The products obtained after removal of sugars from antithrombin retained thrombin-neutralizing activity. In the presence of heparin the inhibition of thrombin as well as factor Xa was enhanced. Thus, the sugar residues of thrombin and antithrombin are not required for the formation of enzyme-inhibitor complexes or for the other activities that were measured.  相似文献   

6.
《Carbohydrate research》1985,144(1):77-86
3-Amino-3-deoxy-α-d-mannopyranosyl α-d-mannopyranoside was synthesized from known 2-O-benzyl-4,6-O-benzylidene-α-d-altropyranosyl 3-O-benzyl-4,6-O-benzylidene-α-d-glucopyranoside, which is available in four steps from commercial α,α-trehalose. The 3,2′-ditriflate of the blocked disaccharide was first treated with sodium azide under phase-transfer conditions, which effected regioselective displacement of the 3-triflyloxy group, and subsequent reaction with sodium benzoate in N,N-dimethylformamide displaced the 2′-triflyloxy group. The blocked, 3-azido-2′-O-benzoyl derivative of α-d-mannopyranosyl α-d-mannopyranoside so obtained was conventionally debenzoylated and debenzylidenated, and subsequent, palladium-catalyzed transfer hydrogenation with formic acid effected reduction of the azido group and cleavage of the benzyl protecting groups, to give the title disaccharide in 13% over-all yield.  相似文献   

7.
Easily deprotoned hydroxyl groups of isatine 3-oximes were glycosylated in high yields by α-D-glucosaminyl chloride peracetate in the solid potassium carbonate-acetonitrile phase transfer system. It was found that catalytic amounts of 15-crown-5 supported a twofold acceleration of the process. The resulting β-D-glucosaminides were identified by 1H NMR spectroscopy. Specific features of the NMR spectra of the synthesized compounds are discussed in comparison with those of other l-O-derivatives of N-acetylglucosamine. Biological activities of oximes with different substituents in the isatin residue were studied by the bacterial luminescence inhibition test with marine luminescent bacteria Photobacterium leiognathi Sh1. The relationship of the structures of the isatin N-substituent and the 5-indolyl substituent and the glycoside capacity to suppress bacterial luminescence was analyzed.  相似文献   

8.
Diacyl and alkylacyl glycerophosphoserines were synthesized by condensation of the respective diradylglycerophosphates with N-t-butoxycarbonyl-L-serinebenzhydryl ester in the presence of 2,4,6-triisopropylbenzenesulfonyl chloride in pyridine at room temperature. The protective groups were simultaneously removed from the resultant N-t-butoxycarbonyl-phosphatidylserinebenzhydryl esters by treatment with HCl in chloroform at 0°C. Products were obtained in 60–65% yield after purification by medium pressure liquid chromatography (MPLC). Dipalmitoyl-N-methylphosphatidylserine was synthesized by an analogous procedure.  相似文献   

9.
It is known that the enzymatic activity of papain (EC 3.4.22.2) toward α-N-benzoyl-l-arginine p-nitroanilide can be substantially increased by hydroxynitrobenzylation of Trp-177 through reaction of the enzyme with the active site-directed reagent, 2-chloromethyl-4-nitrophenyl (N-carbobenzoxy)glycinate (S.-M. T. Chang and H. R. Horton, 1979, Biochemistry18, 1559–1563). To determine the effect of such hydroxynitrobenzylation on the nucleophilicity of the essential thiol group at the active site of the enzyme, rates of inactivation by SN2 reactions of Cys-25 with chloroacetamide and chloroacetate and by Michael addition of Cys-25 to N-ethylmaleimide were monitored. The kinetics revealed that, at pH 6.5, the reactivities of the sulfhydryl group of hydroxynitrobenzylated papain with chloroacetamide and with N-ethylmaleimide are 24 and 27% greater than those of the sulfhydryl group of native papain. At pH 7.1, the rate enhancements are 34 and 39%, respectively. These increases in reactivity of Cys-25 as an attacking nucleophile appear to account for the increased catalytic activity of hydroxnitrobenzyl-papain toward an oligopeptide substrate, α-N-benzoyl-l-phenylalanyl-l-valyl-l-arginine p-nitroanilide, and toward an ester substrate, N-carbobenzoxyglycine p-nitrophenyl ester. However, the presence of the hydroxynitrobenzyl reporter group provides substantially greater improvement (250%) in enzymatic efficiency toward α-N-benzoyl-l-arginine p-nitroanilide, apparently by blocking nonproductive binding of this substrate to the enzyme. Fluorescence changes accompanying the various chemical modifications are interpreted in terms of a charge-transfer interaction between the imidazolium ion of His-159 and the indole moiety of Trp-177 in the active form of native papain, which should help to stabilize the catalytically essential mercaptide-imidazolium ion-pair (Cys-25, His-159).  相似文献   

10.
Hydroxylation of 6-N-trimethyl-l-lysine(lys(Me3)) to 3-hydroxy-6-N-trimethyl-l-lysine(3-HO-lys(Me3)) by several rat tissues has been examined and compared. The kidney enzyme, which previously was shown to require molecular oxygen and α-ketoglutarate as cosubstrates, ferrous iron and ascorbate as cofactors, and to be stimulated by catalase, has a broad pH optimum ranging between 6.5 to 7.5 at 37 °C. As determined with crude tissue extracts from kidney, liver, heart, and skeletal muscle, similar apparent Km values were obtained for substrate, cosubstrates, and cofactors. In view of similar kinetic parameters among the several lys(Me3) hydroxylases examined in rat tissues, and the fact that the level of skeletal muscle lys(Me3) hydroxylase activity is comparable to that of heart, liver, and kidney, because of its large total mass, skeletal muscle may contribute significantly to the biosynthesis of l-carnitine from lys(Me3). The most effective inhibitors found, competitive with lys(Me3), were 2-N-acetyl-6-N-trimethyl-l-lysine, 6-N-monomethyl-l-lysine, and 6-N-dimethyl-l-lysine. l-2-Amino-6-N-trimethylammonium-4-hexynoate, d-2-amino-6-N-trimethylammonium-4-hexynoate, and dl2-amino-6-N-trimethylammonium-cis-4-hexenoate, also inhibited hydroxylase activity but by a yet undetermined mechanism. Oxalacetate, succinate, and citrate inhibited the hydroxylation reaction by competing with α-ketoglutarate. The binding of ferrous iron to the enzyme was competitively inhibited by ions of “soft metals” (e.g., Cd2+, Zn2+) but not by those of “hard metals” (e.g., Ca2+, Mg2+). Preincubation of the crude kidney enzyme for 15 min at 37 °C with mercuriphenylsulfonate, N-ethylmaleimide, iodoacetate, or iodoacetamide resulted in considerable inhibition of 3-HO-lys(Me3) formation. The degree of inhibition by N-ethylmaleimide could be reduced by including Zn (II) during preincubation of the enzyme. The effects of “soft” metals and sulfhydryl reagents on the enzyme suggest that sulfhydryl groups are required for ferrous iron binding in the active site.  相似文献   

11.
β-N-Acetylaminoglucohydrolase (β-2-acetylamino-2-deoxy-D-glucoside acetylaminodeoxyglucohydrolase, EC 3.2.1.30) was extracted from malted barley and purified. The partially purified preparation was free from α-and β-glucosidase, α- and β-galactosidase, α-mannosidase and β-mannosidase. This preparation was free from α-mannosidase only after affinity chromatography with p-amino-N-acetyl-β-D-glucosaminidine coupled to Sepharose. The enzyme was active between pH 3 and 6.5 and had a pH optimum at pH 5. A MW of 92000 was obtained by sodium dodecyl sulfate-acrylamide gel electrophoresis and a sedimentation coefficient of 4.65 was obtained from sedimentation velocity experiments. β-N-Acetylaminoglucohydrolase had a Km of 2.5 × 10?4 M using the p-nitrophenyl N-acetyl β-D-glucosaminidine as the substrate.  相似文献   

12.
The highest antithrombogenic activity was achieved by the sulphation of partially N-deacetylated O-(carboxymethyl)chitin among variously modified chitin derivatives. It was also suggested that the distribution of N-sulphate and N-acetyl groups on the C-2 position might be essential to the selective binding by antithrombin-III to inhibit thrombin activity. Kinetic evaluations demonstrated the non-competitive inhibition on direct interaction with thrombin (Ki = 9.26 × 10−8m) and the competitive inhibition with antithrombin-III (Ki = 3.33 × 10−8) m as well as with heparin. 6-O-Carboxymethyl groups were found, from the data of intravenous injection in mice, to suppress the toxicity of chitin heparinoids.  相似文献   

13.
2,3,4,6-Tetra-O-acetyl-β-d-mannopyranosyl chloride (2) was obtained in 70% yield by the action of lithium chloride on 2,3,4,6-tetra-O-acetyl-α-d-mannopyranosyl bromide (1) in hexamethylphosphoric triamide. p-Nitrobenzenethiol reacted with 1 and 2 as well as with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl bromide (9) or its β-d-chloro analog (10), giving exclusively and in good yield the corresponding p-nitrophenyl 1-thioglycosides of inverted anomeric configuration. The 1,2-cis-d-manno and -glucop-nitrophenylglycosides were likewise prepared. α-d-Glucopyranosyl 1-thio-α-d-glucopyranoside was similarly obtained by the action of the sodium salt of 1-thio-α-d-glucopyranose on the β-chloride 10 in hexamethylphosphoric triamide, or by treatment of 10 with sodium sulfide, with subsequent deacetylation. Analogous procedures allowed the preparation of β-d-mannopyranosyl 1-thio-β-d-mann opyranoside, the corresponding α,β anomer and α-d-glucopyranosyl 1-thio-α-d-mannopyranoside, starting from bromide 1, 1-thio-α-d-mannopyranose (8),and chloride 10, respectively. When acetone was used as solvent, the reaction between 1 and 8 led instead to the α,α anomer. The thio disaccharides that are interglycosidic 4-thio analogs of methyl 4-O-(β-d-galactopyranosyl)-α-d-galactopyranoside, methyl α-cellobioside, and methyl α-maltoside, respectively, were obtained by way of the peracetates of methyl 4-thio-α-d-galactopyranoside and -glucopyranoside by reaction of the corresponding thiolates with tetra-O-acetyl-α-d-galactopyranosyl bromide, bromide 9, or chloride 10, respectively, in hexamethylphosphoric triamide. These 1-thioglycosides, and (1→1)- and (1→4)-thiodisaccharides, were characterized by 1H- and 1 3C-n.m.r. spectroscopy. Correlations were established between the polarity of the sulfur atom and certain proton and carbon chemical-shifts in the 1-thioglycosides in comparison with the O-glycosyl analogs; these correlations permitted in particular the unambigous attribution of anomeric configuration.  相似文献   

14.
The use of crown ethers for a phase transfer-catalyzed synthesis of heteroaromatic glycosides of N-acetylglucosamine was studied. The solid-liquid system and catalysis by 15-crown-5 were found to provide for both the 100% conversion of α-D-glucosaminyl chloride peracetate and a high reaction rate. The interaction of α-D-glucosaminyl chloride peracetate and oxadiazole and triazole mercapto derivatives capable of thiol-thione tautomerism carried out at room temperature in acetonitrile in the presence of anhydrous potassium carbonate and crown ethers was shown to lead to both S- and N-glucosides. The structures of the compounds synthesized were confirmed by X-ray analysis and 13C and 1H NMR spectroscopy.  相似文献   

15.
6-N-[3-3H]Trimethyl-dl-lysine was synthesized from 6-N-acetyl-l-lysine by the following chemical scheme: 6-N-acetyl-l-lysine → 2-keto-6-N-acetylcaproic acid → 2-[3-3H]keto-6-N-acetylcaproic acid → 2-[3-3H]keto-6-N-acetylcaproic acid oxime → 6-N-[3-3H]acetyl-dl-lysine → dl-[3-3H]lysine → 2-N-[3-3H]formyl-dl-lysine → 2-[3-3H]formyl-6-N-trimethyl-dl-lysine → 6-N-[3-3H]trimethyl-dl-lysine. Using a 70% ammonium sulfate fraction obtained from a high-speed rat kidney supernatant, the cosubstrate and cofactor requirements for 6-N-trimethyl-l-lysine hydroxylase activity as measured by tritium release from 6-N-[3-3H]trimethyl-dl-lysine were: α-ketoglutarate, ferrous ions, l-ascorbate, and oxygen, with added catalase showing a slight but distinct stimulatory effect. On incubation with the crude rat kidney preparation, the release of tritium from 6-N-[3-3H]trimethyl-dl-lysine was linear with both time of incubation and protein concentration. Hydroxylation of 6-N-trimethyl-l-lysine, as measured by tritium release from the labeled substrate, was examined in rat kidney, heart, liver, and skeletal muscle tissues, and found to be most active in the kidney.  相似文献   

16.
The unidirectional fluxes of Na+ and Cl- were measured across the isolated gastric mucosa of the bullfrog (R. catesbiana). The addition of strophanthidin, a cardiac aglycone, resulted in marked reductions of the spontaneous potential and short-circuit current. Associated with these changes, the isolated gastric mucosa ceased secreting chloride and hydrogen ion. Although the active component of chloride transfer was inhibited, the exchange diffusion component seemed to increase. No significant changes in membrane conductance or sodium flux were noted. Possible mechanisms of strophanthidin inhibition were discussed in view of its effect on chloride transport across the gastric mucosa and on sodium and potassium transfer in other tissues. It was concluded that the cardiac glycosides may not be specific inhibitors of sodium and potassium transport. This non-specific inhibition suggests that active chloride transport is affected by strophanthidin directly and/or anion secretion is dependent upon normal functioning of cation transport systems in the tissue.  相似文献   

17.
The addition of ammonium ions to the external medium results in an inhibition of the sodium influx and net uptake in Carassius auratus, while intraperitoneal injection of ammonium produces the opposite effect. The simultaneous chloride balance is not significantly affected by these treatments. The addition of bicarbonate ions to the external medium results in a reduction of the influx and net flux of chloride, while injection of bicarbonate produces the opposite effect. The simultaneous sodium balance is not significantly altered. The effects of the external additions are reversible after elimination of the excess ammonium or bicarbonate ions by rinsing. Inhibition of carbonic anhydrase in the gill by injection of acetazoleamide produces a simultaneous inhibition of both sodium and chloride exchanges. These results confirm the hypothesis of an exchange of sodium for ammonium, and of bicarbonate for chloride across the gill. A tentative schematic representation of the ionic absorption mechanisms in the branchial cell of the fresh-water teleosts is given. Similarities with other biological membranes and especially with the renal tubule are pointed out.  相似文献   

18.
2-[4-(p-Toluenesulfonamido)phenyl]ethyl 2,3,4-tri-O-benzyl-α-D-glucopyranoside was condensed with 2,3,4-tri-O-benzyl-6-O-(N-phenylcarbamoyl)-1-O-tosyl-D-glucopyranose to give 2-[4-(p-toluenesulfonamido)phenyl]ethyl 2,3,4,2′,3′,4′-hexa-O-benzyl-6′-O-(N-phenylcarbamoyl)α-isomaltoside. The disaccharide was decarbanilated in ethanol with sodium ethoxide. The sequence of coupling with the 1-O-tosyl-glucose derivative followed by decarbanilation was repeated to form the tri- and tetra-saccharide derivatives. The di-, tri-, and tetra-oligo-saccharides, were deblocked with sodium in liquid ammonia to give the 2-(4-aminophenyl)ethyl α-isomalto-oligosaccharides, which were diazotized with sodium nitrite in acid, and then coupled to bovine serum albumin and edestin to give the protein conjugates.  相似文献   

19.
Proteins in any solution with a pH value that differs from their isoelectric point exert both an electric Donnan effect (DE) and colloid osmotic pressure. While the former alters the distribution of ions, the latter forces water diffusion. In cells with highly Cl--permeable membranes, the resting potential is more dependent on the cytoplasmic pH value, which alters the Donnan effect of cell proteins, than on the current action of Na/K pumps. Any weak (positive or negative) electric disturbances of their resting potential are quickly corrected by chloride shifts. In many excitable cells, the spreading of action potentials is mediated through fast, voltage-gated sodium channels. Tissue cells share similar concentrations of cytoplasmic proteins and almost the same exposure to the interstitial fluid (IF) chloride concentration. The consequence is that similar intra- and extra-cellular chloride concentrations make these cells share the same Nernst value for Cl-. Further extrapolation indicates that cells with the same chloride Nernst value and high chloride permeability should have similar resting membrane potentials, more negative than -80 mV. Fast sodium channels require potassium levels >20 times higher inside the cell than around it, while the concentration of Cl- ions needs to be >20 times higher outside the cell. When osmotic forces, electroneutrality and other ions are all taken into account, the overall osmolarity needs to be near 280 to 300 mosm/L to reach the required resting potential in excitable cells. High plasma protein concentrations keep the IF chloride concentration stable, which is important in keeping the resting membrane potential similar in all chloride-permeable cells. Probable consequences of this concept for neuron excitability, erythrocyte membrane permeability and several features of circulation design are briefly discussed.  相似文献   

20.
Specific spectrophotometric assays for cathepsin B1.   总被引:6,自引:0,他引:6  
Cathepsin B1 from bovine spleen was partially purified by acetone fractionation and by chromatography on Sephadex G-150 and DEAE Sephadex A-50. The enzyme was shown to catalyze the hydrolysis of p-nitrophenyl benzyloxycarbonylglycinate and p-nitrophenyl α-N-benzyloxycarbonyl-l-lysinate. Under the assay conditions, cathepsin B1 is the major enzyme present in bovine spleen homogenates hydrolyzing these substrates. The kinetic parameters for the hydrolysis of p-nitrophenyl benzyloxycarbonylglycinate and p-nitrophenyl α-N-benzyloxycarbonyl-l-lysinate were measured and compared with those obtained for other cathepsin B1 substrates. These results form the basis of an improved spectrophotometric assay for this enzyme in which the liberation of p-nitrophenol from either the N-benzyloxycarbonyl glycine or lysine p-nitrophenyl ester is monitored continuously at 326 nm.  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