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 78 毫秒
1.
Kinetic and thermodynamic studies are presented showing that the cofactor activity of fibrin I (polymerized des-A fibrinogen) in the alpha-thrombin-catalyzed proteolysis of activation peptide (AP) from plasma factor XIII can be attributed to formation of a fibrin I-plasma factor XIII complex (Kd = 65 nM), which is processed by alpha-thrombin more efficiently (kcat/Km = 1.2 x 10(7) M-1 s-1) than free, uncomplexed plasma factor XIII (kcat/Km = 1.4 x 10(5) M-1 s-1). The increase in the specificity constant (kcat/Km) is shown to be largely due to an increase in the apparent affinity of alpha-thrombin for the complex of plasma factor XIII and fibrin I, as reflected by the 30-fold decrease in the Michaelis constant observed for fibrin I bound plasma factor XIII relative to that for uncomplexed plasma factor XIII. Analysis of the initial rates of alpha-thrombin-catalyzed hydrolysis of fibrinopeptide B (FPB) from fibrin I polymer in the presence of plasma factor XIII indicated that alpha-thrombin bound to fibrin I in the ternary complex of alpha-thrombin, plasma factor XIII, and fibrin I polymer is competent to catalyze cleavage of both FPB from fibrin I and AP from plasma factor XIII. This observation is consistent with the view that alpha-thrombin within the ternary complex is anchored to fibrin I polymer through a binding site distinct from the active site (an exosite) and that the active site is alternatively complexed with the AP moiety of plasma factor XIII or the FPB moiety of fibrin I. This conclusion is supported by the observation that a 12-residue peptide, which binds to an exosite of alpha-thrombin and blocks the interaction of alpha-thrombin with fibrinogen and fibrin, competitively inhibits alpha-thrombin-catalyzed release of both FPB and AP from the fibrin I-plasma factor XIII complex.  相似文献   

2.
Steady-state kinetic parameters were determined for the action of human alpha-thrombin on human fibrin I polymer, an intermediate in the alpha-thrombin-catalyzed conversion of fibrinogen to the fibrin matrix of blood clots during the terminal phase of the blood clotting cascade. Values of 49 s-1 and 7.5 microM were determined (at 37 degrees C, pH 7.4, gamma/2 0.17) for kcat and Km, respectively. Studies of the effect of fibrin I on alpha-thrombin-catalyzed hydrolysis of the fluorogenic substrate N-p-Tos-Gly-L-Pro-L-Arg-7-amido-4-methylcoumarin (tos-GPR-amc) and the effect of fibrin I on the reaction of alpha-thrombin with antithrombin III (AT) were presented which indicate that the active site of alpha-thrombin is accessible while it is bound to its substrate fibrin I. Fibrin I inhibited alpha-thrombin-catalyzed hydrolysis of tos-GPR-amc in a manner inconsistent with the pure competitive inhibition expected for an alternative substrate, whereas fibrinogen, an alpha-thrombin substrate, behaved as a pure competitive inhibitor of the alpha-thrombin-catalyzed hydrolysis of tos-GPR-amc. The effect of fibrin I on alpha-thrombin-catalyzed hydrolysis of tos-GPR-amc was shown to be consistent with alpha-thrombin binding to fibrin I in alternative orientations. In one orientation both the active site and a site distinct from the active site (an exosite) of alpha-thrombin are occupied by fibrin I. In the other orientation only the exosite of alpha-thrombin is occupied and the active site is freely accessible to other substrates. The values of both kcat (21 s-1) and Km (less than 0.23 microM) determined for fibrin I-bound alpha-thrombin acting on tos-GPR-amc were decreased relative to the values of kcat (180 s-1) and Km (7.3 microM) observed for the action of uncomplexed alpha-thrombin on tos-GPR-amc. This observation suggests that the active site of alpha-thrombin is altered in fibrin I-bound alpha-thrombin. Studies of the effect of fibrin I on the reaction of AT with alpha-thrombin (at 37 degrees C, pH 7.4, gamma/2 0.17) indicated that when alpha-thrombin is bound to fibrin I in an orientation where the active site of alpha-thrombin is accessible, AT reacts with alpha-thrombin with a rate constant (greater than 4.2 x 10(4) M-1 s-1) that is greater than the rate constant (1.5 x 10(4) M-1 s-1) for reaction of AT with the free enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
Promotion of thrombin-catalyzed activation of factor XIII by fibrinogen   总被引:10,自引:0,他引:10  
T J Janus  S D Lewis  L Lorand  J A Shafer 《Biochemistry》1983,22(26):6269-6272
High-performance liquid chromatography was used to analyze the kinetics of the thrombin-catalyzed release of the activation peptide from the factor XIII zymogen (fibrin-stabilizing factor). The specificity constant (kcat/Km) for this reaction, measured at factor XIII concentrations much below Km, was (0.13-0.16) X 10(6) M-1 s-1 at pH 7.4, mu = 0.15, and 37 degrees C. Separate estimates, obtained from the dependence of the initial rates of release of the activation peptide on the concentration of factor XIII, gave values of 10 (+/- 3) s-1 for kcat and 84 (+/- 30) microM for Km, in terms of ab protomers of the zymogen. The thrombin-mediated release of the activation peptide was dramatically enhanced in the presence of fibrinogen. Furthermore, the time course of release, in relation to that of fibrinopeptide A, suggested that some des-A-fibrinogen species (e.g., alpha 2B beta 2 gamma 2) may be the true activator for promoting the cleavage of the Arg-36 peptide bonds in the a subunits of factor XIII. This observation suggests that generation of factor XIIIa and its substrate (fibrin) is coordinated so that thrombin-mediated zymogen activation proceeds efficiently only after the process of clotting has been initiated by the removal of fibrinopeptide A from fibrinogen.  相似文献   

4.
The time dependence of the release of fibrinopeptides from fibrinogen was studied as a function of the concentration of fibrinogen, thrombin, and Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization. The release of fibrinopeptides during fibrin assembly was shown to be a highly ordered process. Rate constants for individual steps in the formation of fibrin were evaluated at pH 7.4, 37 degrees C, gamma/2 = 0.15. The initial event, thrombin-catalyzed proteolysis at Arg-A alpha 16 to release fibrinopeptide A (kcat/Km = 1.09 X 10(7) M-1s-1) was followed by association of the resulting fibrin I monomers. Association of fibrin I was found to be a reversible process with rate constants of 1 X 10(6) M-1s-1 and 0.064 s-1 for association and dissociation, respectively. Assuming random polymerization of fibrin I monomer, the equilibrium constant for fibrin I association (1.56 X 10(7) M-1) indicates that greater than 80% of the fibrin I protofibrils should contain more than 10 monomeric units at 37 degrees C, pH 7.4, when the fibrin I concentration is 1.0 mg/ml. Association of fibrin I monomers was shown to result in a 6.5-fold increase in the susceptibility of Arg-B beta 14 to thrombin-mediated proteolysis. The 6.5-fold increase in the observed specificity constant from 6.5 X 10(5) M-1s-1 to 4.2 X 10(6) M-1s-1 upon association of fibrin I monomers and the rate constant for fibrin association indicates that most of the fibrinopeptide B is released after association of fibrin I monomers. The interaction between a pair of polymerization sites in fibrin I dimer was found to be weaker than the interaction of fibrin I with Gly-Pro-Arg-Pro and weaker than the interaction of fibrin I with fibrinogen.  相似文献   

5.
Hirudin, a potent 65-residue polypeptide inhibitor of alpha-thrombin found in the saliva of the leech Hirudo medicinalis, and fragments thereof are potentially useful as antithrombotic agents. Hirugen, the synthetic N-acetylated COOH-terminal dodecapeptide (Ac-Asn-Gly-Asp-Phe-Glu-Glu-Ile-Pro-Glu-Glu-Tyr(SO3)-Leu) of hirudin was shown in the present study to behave as a pure competitive inhibitor (Ki = 0.54 microM) of human alpha-thrombin-catalyzed release of fibrinopeptide A from human fibrinogen. In contrast to this inhibitory activity, hirugen slightly enhanced (increased kcat/Km 1.6-fold) alpha-thrombin-catalyzed hydrolysis of the fluorogenic tripeptide substrate N-p-Tosyl-Gly-Pro-Arg-7-amino-4-methylcoumarin. These observations indicate that hirugen binds to alpha-thrombin at an exosite distinct from the active site, and that interaction with this exosite is a major determinant of the competence of alpha-thrombin to bind fibrinogen. Consistent with this view, hirugen blocked binding of fibrin II to alpha-thrombin. Studies of the effect of hirugen on the rate of inactivation of alpha-thrombin by antithrombin III (AT), the major plasma inhibitor of alpha-thrombin, indicated that binding of hirugen to alpha-thrombin results in less than a 2.5-fold decrease in the rate of inactivation of alpha-thrombin by AT, both in the absence and presence of heparin. This behavior is distinct from that of active site-directed competitive inhibitors of alpha-thrombin which bind to alpha-thrombin and block both conversion of fibrinogen to fibrin and inactivation of alpha-thrombin by AT. Hirugen, an exosite-directed competitive inhibitor, blocks the interaction of alpha-thrombin with fibrinogen while leaving alpha-thrombin competent to react with AT. Thus, unlike active site-directed competitive inhibitors, hirugen should act in concert with AT and heparin to reduce the amount of fibrinogen that is processed during the lifetime of alpha-thrombin in plasma.  相似文献   

6.
Fibrinogen, purified from a recently identified case of dysfibrinogenaemia, fibrinogen Sydney I, was shown by thrombin digestion, high-performance liquid chromatography (HPLC) and amino acid analysis to be a heterozygous case of an A alpha Arg-16----His substitution. Kinetic studies have been carried out on the thrombin-induced release of fibrinopeptide A (FPA), fibrinopeptide B (FPB) and the variant peptide [His16]FPA. When thrombin was added to fibrinogen Sydney I at a concentration of 0.2 U/ml release of FPA was rapid and there was a 79-fold reduced rate of release of [His16]FPA, but the rate of release of FPB was not appreciably reduced. In contrast, at lower thrombin concentrations the rate of FPB release was reduced in proportion to the rate of total FPA release, supporting the view that release of fibrinopeptides is a sequential process. The second-order kinetic constant kcat/Km for hydrolysis of the abnormal A alpha chain by thrombin was calculated from Lineweaver-Burk plots to be 16-30-fold less than that for the normal A alpha chain. Molecular modelling studies, using a refined model of the trypsin-pancreatic-trypsin-inhibitor complex have been used to suggest how the histidine at the P1 site can be accommodated within the enzyme hydrophobic active-site pocket.  相似文献   

7.
The vampire bat salivary plasminogen activator (BatPA) is virtually inactive toward Glu-plasminogen in the absence of a fibrin-like cofactor, unlike human tissue-type plasminogen activator (tPA) (the kcat/Km values were 4 and 470 M-1 s-1, respectively). In the presence of fibrin II, tPA and BatPA activated Glu-plasminogen with comparable catalytic efficiencies (158,000 and 174,000 M-1 s-1, respectively). BatPA's cofactor requirement was partially satisfied by polymeric fibrin I (54,000 M-1 s-1), but monomeric fibrin I was virtually ineffective (970 M-1 s-1). By comparison, a variety of monomeric and polymeric fibrin-like species markedly enhanced tPA-mediated activation of Glu-plasminogen. Fragment X polymer was 2-fold better but 9-fold worse as cofactor for tPA and BatPA, respectively, relative to fibrin II. Fibrinogen, devoid of plasminogen, was a 10-fold better cofactor for tPA than fibrinogen rigorously depleted of plasminogen, Factor XIII, and fibronectin; the enhanced stimulatory effect of the less-purified fibrinogen was apparently due to the presence of Factor XIII. By contrast, the two fibrinogen preparations were equally poor cofactors of BatPA-mediated activation of Glu-plasminogen. BatPA possessed only 23 and 4% of the catalytic efficiencies of tPA and two-chain tPA, respectively, in hydrolyzing the chromogenic substrate Spectrozyme tPA. However in the presence of fibrin II, BatPA and tPA exhibited similar kcat/Km values for the hydrolysis of Spectrozyme tPA. Our data revealed that BatPA, unlike tPA, displayed a strict and fastidious requirement for polymeric fibrin I or II. Consequently, BatPA may preferentially promote plasmin generation during a narrow temporal window of fibrin formation and dissolution.  相似文献   

8.
The kinetics of activation of platelet factor XIII, an a-subunit dimer, were characterized by determining rate constants for activation peptide (AP) release, generation of activity, and exposure of the active-site thiol group. The specificity constant (kappacat/Km) for alpha-thrombin-catalyzed AP release, 1.2 x 10(5) M-1s-1, was found to be similar to that for AP release from the tetramer plasma factor XIII (a2b2) [Janus, T.J., Lewis, S. D., Lorand, L., & Shafer, J. A. (1983) Biochemistry 22, 6269-6272], implying that the b subunits of plasma factor XIII do not hinder alpha-thrombin-catalyzed cleavage of AP from the a subunit. Platelet factor XIIIa activity was generated at a rate approximately twice the rate of AP release. This difference in rates was shown to be consistent with a reaction pathway for activation of platelet factor XIII wherein full factor XIIIa activity is generated when one AP is removed from the dimeric zymogen so that removal of the second AP has no detectable effect on catalytic activity. In accord with this conclusion, the rate constant for exposure of the active-site thiol group, as measured by the incorporation of [1-14C]-iodoacetamide, was about twice that observed for the removal of AP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Enzymic and nonenzymic properties of human beta-thrombin   总被引:4,自引:0,他引:4  
Autolysis or tryptic hydrolysis converts human alpha-thrombin to its beta-derivative and subsequently to gamma-thrombin. Human beta-thrombin was obtained by tryptic digestion of alpha-thrombin and isolated by BioRex chromatography. The kinetic parameters for human alpha- and beta-thrombins with H-D-phenylalanyl-L-pipecolyl-L-arginine-para-nitroanilide were similar, as well as the rate of inactivation by tosyl-lysine chloromethyl ketone. By contrast, the rate of inactivation by diisopropyl fluorophosphate was reduced by half, and the inhibition constant for benzamidine was increased 2.5-fold. Moreover, the beta cleavages induced a drastic reduction in reactivity toward protein C, affinity for thrombomodulin, and fibrinogen clotting activity. Unlike alpha-thrombin, beta-thrombin was not protected from inhibition by diisopropyl fluorophosphate in the presence of fibrinogen and failed to bind to fibrin-Sepharose. Our results indicate that the beta cleavages induce multiple defects in the functions of human thrombin. Although the three catalytic residues remain in an active configuration, subtle changes are induced in the microenvironment of the active serine. However, the drastic reduction of fibrinogen clotting activity should rather be ascribed to major alterations observed in both the fibrinopeptide groove and the fibrin recognition site. These observations provide further evidence for a double-site mechanism in the interaction of fibrinogen with thrombin.  相似文献   

10.
Thrombin preferentially cleaves fibrinopeptides A (FPA) from fibrinogen resulting in the formation of desAA-fibrin from which most of the fibrinopeptides B (FPB) are then released with an enhanced rate. Kinetics of fibrinopeptide release from normal and dysfunctional fibrinogens were investigated in order to further characterize the mechanism of accelerated FPB release during desAA-fibrin polymerization. Dysfunctional fibrinogens London I and Ashford, exhibiting primary polymerization abnormalities (i.e., an abnormality present when all fibrinopeptides have been cleaved), which in the case of fibrinogen London I is believed to be caused by a defect in the D-domain, were shown to exhibit a decreased rate of FPB release compared with normal fibrinogen. While Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization, was shown to decrease the rate of FPB release from normal fibrinogen by a factor of 5, normal fragment D1, although inhibiting clot formation of normal fibrinogen, did not influence the acceleration of FPB release. On the other hand, the presence of fragment D1 did not enhance FPB release from fibrinogen London I, suggesting that interaction of D-domains in functional isolation with desAA-fibrin E-domains is not sufficient to enhance FPB release. Although clot formation was inhibited by the concentrations of fragment D1 used, the formation of small desAA-fibrin oligomers was hardly affected. Thus, small fibrin polymers, but not desAA-fibrin monomers, act as optimal substrates for the release of FPB by thrombin.  相似文献   

11.
In this study we report a kinetic model for the alpha-thrombin-catalyzed production of fibrin I and fibrin II at pH 7.4, 37 degrees C, gamma/2 0.17. The fibrin is produced by the action of human alpha-thrombin on plasma levels of human fibrinogen in the presence of the major inhibitor of alpha-thrombin in plasma, antithrombin III (AT). This model quantitatively accounts for the time dependence of alpha-thrombin-catalyzed release of fibrinopeptides A and B concurrent with the inactivation of alpha-thrombin by AT and delineates the concerted interactions of alpha-thrombin, fibrin(ogen), and AT during the production of a fibrin clot. The model also provides a method for estimating the concentration of alpha-thrombin required to produce a clot of known composition and predicts a direct relationship between the plasma concentration of fibrinogen and the amount of fibrin produced by a bolus of alpha-thrombin. The predicted relationship between the concentration of fibrinogen and the amount of fibrin produced in plasma provides a plausible explanation for the observed linkage between plasma concentrations of fibrinogen and the risk for ischemic heart disease.  相似文献   

12.
Anion-binding exosite of human alpha-thrombin and fibrin(ogen) recognition   总被引:6,自引:0,他引:6  
Activation of prothrombin to alpha-thrombin generates not only the catalytic site and associated regions but also an independent site (an exosite) which binds anionic substances, such as Amberlite CG-50 resin [cross-linked poly(methylacrylic acid)]. Like human alpha-thrombin with high fibrinogen clotting activity (peak elution at I = 0.40 +/- 0.01 M, pH 7.4, approximately 23 degrees C), catalytically inactivated forms (e.g., i-Pr2P-alpha- and D-Phe-Pro-Arg-CH2-alpha-thrombins) were eluted with only slightly lower salt concentrations (I = 0.36-0.39 M), while gamma-thrombin with very low clotting activity was eluted with much lower concentrations (I = 0.29 M) and the hirudin complex of alpha-thrombin was not retained by the resin. In a similar manner, hirudin complexes of alpha-, i-Pr2P-alpha-, and gamma-thrombin were not retained by nonpolymerized fibrin-agarose resin. Moreover, the ionic strengths for the elution from the CG-50 resin of seven thrombin forms were directly correlated with those from the fibrin resin (y = 0.15 + 0.96x, r = 0.95). In other experiments, the 17 through 27 synthetic peptide of the human fibrinogen A alpha chain was not an inhibitor of alpha-thrombin, while the NH2-terminal disulfide knot (NDSK) fragment was a simple competitive inhibitor of alpha-thrombin with a Ki approximately 3 microM (0.15 M NaCl, pH 7.3, approximately 23 degrees C). These data suggest that alpha-thrombin recognizes fibrin(ogen) by a negatively charged surface, noncontiguous with the A alpha cleavage site but found within the NDSK fragment. Such interaction involving an anion-binding exosite may explain the exceptional specificity of alpha-thrombin for the A alpha cleavage in fibrinogen and alpha-thrombin incorporation into fibrin clots.  相似文献   

13.
The kinetic parameters for release of fibrinopeptide A (FPA) from human fibrinogen by thrombin are: Km = 2.3 X 10(-6)M and Vmax. = 1.1 X 10(-10)mol of FPA/s per unit of thrombin; for fibrin formation, Km is similar to that for FPA release, but, the conditions of the present study, Vmax. was approximately half of that for FPA release. The formation of fibrin polymer before the sol-gel transition was studied by gel-permeation chromatography combined with effluent analysis for fibrinogen antigen and residual FPA. Polymer formation in purified fibrinogen incubated with thrombin proceeded as a bimolecular association of exposed sites in a manner predicted by probability calculations and assuming random FPA cleavage. Each oligomer consisted of n molecules of fibrin monomer and two fibrinogen molecules, each of the latter lacking one FPA molecule, i.e. each oligomer, regardless of molecular size, retains two FPA molecules. The addition of 5 mM-CaCl2 to the reaction mixture changed the rate of polymer formation, so that dimer was no longer the prevalent oligomer; in the presence of Ca2+, the trimer was the oligomer in highest concentration. The polymers formed in the presence of calcium were similar in composition to those without, i.e. 2 mol of FPA/mol of oligomer. EDTA-treated plasma samples incubated for short periods of time, 30s or less, with thrombin ranging in concentration up to 1 N.I.H. unit/ml did not form clots during the 10-15 min period of observation until they were applied to the column, though a large proportion of the available FPA was cleaved (maximum 45%). The soluble polymers in plasma were mostly of the high-Mr variety (tetramer and greater); these high-Mr polymers contained less than 2 mol of FPA/mol of polymer, whereas dimer and trimer in plasma were similar to those in the purified systems, i.e. 2 mol of FPA/mol.  相似文献   

14.
Human neutrophil cathepsin G or bovine chymotrypsin proteolytically cleaved human alpha-thrombin at the B-chain Trp148-Thr149 bond generating a new form, zeta-thrombin. While incubation of alpha-thrombin with cathepsin G at pH 7.4 and 37 degrees C resulted in a partial loss of fibrinogen clotting activity, 86 +/- 13% of the clotting activity and 99 +/- 16% of the active sites titratable with p-nitrophenyl p-guanidinobenzoate were retained upon controlled passage of alpha-thrombin through chymotrypsin-Sepharose 4B at pH 6.2 or 7.4 and 24 degrees C (n = 15). Kinetic parameters for H-D-hexahydrotyrosyl-Ala-Arg p-nitroanilide were Km = 1.52 +/- 0.60 vs 1.32 +/- 0.18 microM and kcat = 51.9 +/- 2.9 vs 35.8 +/- 6.4 s-1 with alpha-thrombin vs chymotrypsin-prepared zeta-thrombin (n = 4 vs 3), respectively (I = 0.15 M, pH 7.4, and 24 degrees C). Some 95% of the clotting activity was lost when zeta-thrombin was passed through trypsin-Sepharose 4B under conditions for converting alpha- to nonclotting beta- and subsequently gamma-thrombin. The resulting gamma-like thrombins eluted bimodally with 260 and 310 mM NaCl when applied to Amberlite CG-50 resin [cross-linked poly(methylacrylic acid)] developed with a linear salt gradient in 50 mM Tris at pH 7.4 and 24 degrees C. These elution peaks correspond to 240, 330, and 350 mM NaCl for gamma-, alpha-, and zeta-thrombin, respectfully, implying that the anion-binding exosite is partially destroyed in gamma-like thrombins but is intact in zeta-thrombin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Human and bovine alpha-thrombin cleaved at the B-chain by chymotrypsin generates catalytically competent zeta-thrombins, which are comprised of two noncovalently linked fragments: a 36-(human) or 49-(bovine) residue A-chain linked by a disulfide to B-chain residues B1-148 (zeta 1-thrombin) and B-chain residues B149-259 (zeta 2-thrombin). Human and bovine D-Phe-Pro-Arg-CH2-zeta- and PhMeSO2-zeta-thrombins were prepared by reaction of the active-site histidine (H-B43) and serine (S-B205) with PPACK and PMSF, respectively. Unfolding and dissociation of the noncovalently linked polypeptide chains of either human or bovine D-Phe-Pro-Arg-CH2-zeta- and PhMeSO2-zeta-thrombins in 4.5 M guanidine-HCl and refolding upon 30-fold dilution in 50 mM sodium phosphate buffer pH 6.5, 750 mM NaCl, 0.1% PEG resulted in biphasic generation of catalytic activity. The slow phase was eliminated in the presence of the competitive inhibitor benzamidine-HCl. Unfolding and refolding mixtures of the appropriate inactive precursors generated the active chimeric thrombins bovine zeta 1-thrombin:human zeta 2-thrombin and human zeta 1-thrombin:bovine zeta 2-thrombin. Human zeta 1-thrombin and zeta 2-thrombin were isolated, and, upon recombining, the isolated fragments refolded to generate catalytically competent zeta-thrombin with an active-site content, specific activity toward Chromozym-TH, and a specificity constant (kcat/Km) for FPA release from fibrinogen that were all within 60% of those of native alpha-thrombin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

17.
Multiple factors affect the thrombin-catalyzed conversion of fibrinogen to fibrin, including: fibrinopeptide (FPA and FPB) release leading to exposure of two types of polymerization domains ("A" and "B," respectively) in the central portion of the molecule, and exposure of a noncatalytic "secondary" thrombin-binding site in fibrin. Fibrinogen containing the FPA sequence but lacking the B beta 1-42 sequence ("des-(B beta 1-42)-fibrinogen"), was compared to native fibrinogen (containing both FPA and FPB) to investigate the role played by B beta 1-42 in the polymerization of alpha-fibrin (i.e. fibrin lacking FPA), to compare reptilase and thrombin cleavage of FPA from fibrinogen, and to explore the location and function of the secondary thrombin-binding site. Electron microscopy of evolving polymer structures (mu, 0.14; pH 7.4) plus turbidity measurements, showed that early thin fibril formation as well as subsequent lateral fibril associations were impaired in des-(B beta 1-42)-alpha-fibrin, thus indicating that the B beta 1-42 sequence contributes to the A polymerization site. Reptilase-activated des-(B beta 1-42)-alpha-fibrin polymerized even more slowly than thrombin-activated des-(B beta 1-42)-alpha-fibrin, differences that disappeared when repolymerization of preformed fibrin monomers was carried out. Since existing data indicate that thrombin releases FPA in a concerted manner, resulting in relatively rapid evolution of fully functional divalent alpha-fibrin monomers, it can be inferred that delayed fibrin assembly of reptilase fibrin is due to slower formation of divalent alpha-fibrin monomers. Thrombin-activated des-(B beta 1-42)-alpha-fibrin polymerized more rapidly at low ionic strength (mu, 0.04) than did native alpha,beta-fibrin, a reversal of their behavior at physiological ionic strength (mu, 0.14). Concomitant measurement of FPA release revealed modest slowing of release at low ionic strength from des-(B beta 1-42)-fibrinogen (t1/2, 36.5 versus 21.5 min) and marked slowing from native fibrinogen (t1/2, 138 versus 22.2 min). This behavior correlated with increased thrombin binding to native alpha,beta-fibrin at low ionic strength, coupled with weak thrombin binding to des-(B beta 1-42)-alpha-fibrin, and indicates that secondary thrombin binding plays an important role in regulating thrombin diffusion and catalytic activity. Des-(B beta 1-42)-fibrinogen lacks or has a markedly defective secondary thrombin-binding site, from which we conclude that the B beta 15-42 sequence in fibrin plays a major role in forming or providing this site.  相似文献   

18.
Nieman MT  Schmaier AH 《Biochemistry》2007,46(29):8603-8610
Investigations determined the critical amino acids for alpha-thrombin's interaction with protease-activated receptors 1 and 4 (PAR1 and PAR4, respectively) at the thrombin cleavage site. Recombinant PAR1 wild-type (wt) exodomain was cleaved by alpha-thrombin with a Km of 28 microM, a kcat of 340 s-1, and a kcat/Km of 1.2 x 10(7). When the P4 or P2 position was mutated to alanine, PAR1-L38A or PAR1-P40A, respectively, the Km was unchanged, 29 or 23 microM, respectively; however, the kcat and kcat/Km were reduced in each case. In contrast, when Asp39 at P3 was mutated to alanine, PAR1-D39A, Km and kcat were both reduced approximately 3-fold, making the kcat/Km the same as that of PAR1-wt exodomain. Recombinant PAR4-wt exodomain was cleaved by alpha-thrombin with a Km of 61 microM, a kcat of 17 s-1, and a kcat/Km of 2.8 x 10(5). When the P5 or P4 position was mutated to alanine, PAR4-L43A or PAR4-P44A, respectively, there was no change in the Km (69 or 56 microM, respectively); however, the kcat was lowered in each case (9.7 or 7.7 s-1, respectively). Mutation of the P2 position (PAR4-P46A) also had no effect on the Km but markedly lowered the kcat and kcat/Km approximately 35-fold. PAR1-wt exodomain and P4 and P3 mutants were noncompetitive inhibitors of alpha-thrombin hydrolyzing Sar-Pro-Arg-pNA. However, PAR1-P40A displayed a mixed type of inhibition. Mutation of P4, P3, or P2 had no effect on the Ki. All PAR4 exodomains were competitive inhibitors of alpha-thrombin. Mutation of P5, P4, or P2 had no effect on the Ki. These investigations show that Leu at P4 in PAR1 or P5 in PAR4 critically influences the kinetics of alpha-thrombin binding and cleavage of PAR1 and PAR4 exodomains. It also implies that factors other than the hirudin-like binding region on PAR1 exodomain predominate in influencing PAR1 cleavage on cells.  相似文献   

19.
The activation kinetics of single chain urinary-type plasminogen activator (scu-PA) by plasmin have been studied in detail. Nonstandard Michaelis-Menten kinetics were observed. To explain our results, we propose a model in which plasmin can exist in two conformations of lower activity (kcat/Km = 1.4 x 10(6) M-1 s-1) or higher activity (kcat/Km = 16.7 x 10(6) M-1 s-1) depending on whether a lysine binding site is occupied or free, respectively. These kinetic studies demonstrate that scu-PA interacts at this binding site (KD approximately 30 nM) and so is able to act as both a substrate and effector in this reaction. Binding was also demonstrated between scu-PA and Glu- or Lys-plasminogen at a high affinity site (KD approximately 65 nM), sensitive to the presence of lysine analogs. This suggests that scu-PA may be almost completely bound to plasminogen in plasma under normal physiological conditions and provides a possible explanation for the fibrin specificity of this activator, as discussed.  相似文献   

20.
The present study represents detailed investigations into the nature of interactions between an intracellular "tissue" transglutaminase and a plasma protein, fibrinogen. We demonstrate a specific, saturable, and reversible binding of transglutaminase to fibrin(ogen). The binding was time- and temperature-dependent, was independent of divalent metal ions, did not require the release of either fibrinopeptide A or B, and was partially inhibited by the presence of sodium chloride or plasma proteins, properties similar to Factor XIII binding to fibrin(ogen). Both Factor XIII and liver transglutaminase also shared similar binding sites on fibrinogen, the A alpha- and the B beta-chains. The binding characteristics of liver transglutaminase were thus similar to Factor XIII binding to fibrin, but there were also important differences. Scatchard analyses of the binding data indicated that the affinity of liver transglutaminase (Kd = 4.17 x 10(-7) M) was at least 40-fold weaker compared with the affinity of Factor XIII to fibrinogen. Consequently, a 20-fold molar excess of Factor XIII a-chains specifically and completely inhibited the binding of liver transglutaminase to des-A-fibrinogen. The association between liver transglutaminase and fibrin(ogen) was also critically controlled by the conformational states of the two proteins. Substances capable of altering the conformation of either transglutaminase (such as guanosine 5'-triphosphate) or of fibrinogen (such as the tetrapeptide Gly-Pro-Arg-Pro and Fragment D) disrupted binding. Excess CaCl2 was able to counteract the effects of guanosine 5'-triphosphate on transglutaminase binding to fibrin. In contrast, Factor XIII binding to fibrin was unaffected by either guanosine 5'-triphosphate, CaCl2, or Gly-Pro-Arg-Pro, suggesting a more stable association between the two proteins. The physiologic implications of transglutaminase-fibrin(ogen) interactions are discussed.  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