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 875 毫秒
1.
Heparin and dermatan sulfate activate heparin cofactor II (HCII) comparably, presumably by liberating the amino terminus of HCII to bind to exosite I of thrombin. To explore this model of activation, we systematically substituted basic residues in the glycosaminoglycan-binding domain of HCII with neutral amino acids and measured the rates of thrombin inactivation by the mutants. Mutant D, with changes at Arg(184), Lys(185), Arg(189), Arg(192), Arg(193), demonstrated a approximately 130-fold increased rate of thrombin inactivation that was unaffected by the presence of glycosaminoglycans. The increased rate reflects displacement of the amino terminus of mutant D because (a) mutant D inactivates gamma-thrombin at a 65-fold slower rate than alpha-thrombin, (b) hirudin-(54-65) decreases the rate of thrombin inactivation, and (c) deletion of the amino terminus of mutant D reduces the rate of thrombin inactivation approximately 100-fold. We also examined the contribution of glycosaminoglycan-mediated bridging of thrombin to HCII to the inhibitory process. Whereas activation of HCII by heparin was chain-length dependent, stimulation by dermatan sulfate was not, suggesting that dermatan sulfate does not utilize a template mechanism to accelerate the inhibitory process. Fluorescence spectroscopy revealed that dermatan sulfate evokes greater conformational changes in HCII than heparin, suggesting that dermatan sulfate stimulates HCII by producing more effective displacement of the amino terminus.  相似文献   

2.
Thrombomodulin (TM) slows down the interaction rate between thrombin and plasminogen activator inhibitor 1 (PAI-1). We now show that the 12-fold reduced inhibition rate in the presence of TM does not result from an altered distribution between PAI-1 cleavage and irreversible complex formation. Surface plasmon resonance (SPR) revealed an over 200-fold reduced affinity of TM for thrombin-VR1tPA as compared to thrombin, demonstrating the importance of the VR1 loop in the interaction of thrombin with both TM and PAI-1. Furthermore, in contrast to ATIII, PAI-1 was not able to bind the thrombin/TM complex demonstrating complete competitive binding between PAI-1 and TM. Kinetic modeling on the inhibitory effect of TM confirms a mechanism that involves complete steric blocking of the thrombin/PAI-1 interaction. Also, it accurately decribes the biphasic inhibition profile resulting from the substantial reduction of the extremely fast rate of reversible Michaelis complex formation, which is essential for efficient inhibition of thrombin by PAI-1. Vitronectin (VN) is shown to partially relieve TM inhibitory action only by vastly increasing the initial rate of interaction between free thrombin and PAI-1. In addition, SPR established that solution-phase PAI-1/VN complexes and non-native VN (extracellular matrix form) bind TM directly via the chondroitin sulphate moiety of TM. Collectively, these results show that VR1 is a subsite of exosite 1 on thrombin's surface, which regulates exclusive binding of either PAI-1 or TM. This competition will be physiologically significant in controlling the mitogenic activity of thrombin during vascular disease.  相似文献   

3.
Assembly of ternary thrombin-heparin-fibrin complexes, formed when fibrin binds to exosite 1 on thrombin and fibrin-bound heparin binds to exosite 2, produces a 58- and 247-fold reduction in the heparin-catalyzed rate of thrombin inhibition by antithrombin and heparin cofactor II, respectively. The greater reduction for heparin cofactor II reflects its requirement for access to exosite 1 during the inhibitory process. Protection from inhibition by antithrombin and heparin cofactor II requires ligation of both exosites 1 and 2 because minimal protection is seen when exosite 1 variants (gamma-thrombin and thrombin Quick 1) or an exosite 2 variant (Arg93 --> Ala, Arg97 --> Ala, and Arg101 --> Ala thrombin) is substituted for thrombin. Likewise, the rate of thrombin inhibition by the heparin-independent inhibitor, alpha1-antitrypsin Met358 --> Arg, is decreased less than 2-fold in the presence of soluble fibrin and heparin. In contrast, thrombin is protected from inhibition by a covalent antithrombin-heparin complex, suggesting that access of heparin to exosite 2 of thrombin is hampered when ternary complex formation occurs. These results reveal the importance of exosites 1 and 2 of thrombin in assembly of the ternary complex and the subsequent protection of thrombin from inhibition by heparin-catalyzed inhibitors.  相似文献   

4.
Bothrojaracin, a 27-kDa C-type lectin from Bothrops jararaca venom, is a selective and potent thrombin inhibitor (K(d) = 0.6 nM) which interacts with the two thrombin anion-binding exosites (I and II) but not with its catalytic site. In the present study, we analyzed the allosteric effects produced in the catalytic site by bothrojaracin binding to thrombin exosites. Opposite effects were observed with alpha-thrombin, which possesses both exosites I and II, and with gamma-thrombin, which lacks exosite I. On the one hand, bothrojaracin altered both kinetic parameters K(m) and k(cat) of alpha-thrombin for small synthetic substrates, resulting in an increased efficiency of alpha-thrombin catalytic activity. This effect was similar to that produced by hirugen, a peptide based on the C-terminal hirudin sequence (residues 54-65) which interacts exclusively with exosite I. On the other hand, bothrojaracin decreased the amidolytic activity of gamma-thrombin toward chromogenic substrates, although this effect was observed with higher concentrations of bothrojaracin than those used with alpha-thrombin. In agreement with these observaions, bothrojaracin produced opposite effects on the fluorescence intensity of alpha- and gamma-thrombin derivatives labeled at the active site with fluorescein-Phe-Pro-Arg-chloromethylketone. These observations support the conclusion that bothrojaracin binding to thrombin produces two different structural changes in its active site, depending on whether it interacts exclusively with exosite II, as seen with gamma-thrombin, or with exosite I (or both I and II) as observed with alpha-thrombin. The ability of bothrojaracin to evoke distinct modifications in the thrombin catalytic site environment when interacting with exosites I and II make this molecule an interesting tool for the study of allosteric changes in the thrombin molecule.  相似文献   

5.
Exosite 1 on thrombin mediates low affinity binding to sites on the NH2 termini of the alpha- and beta-chains of fibrin. A subpopulation of fibrin molecules (gammaA/gamma'-fibrin) has an alternate COOH terminus of the normal gamma-chain (gammaA/gammaA-fibrin) that binds thrombin with high affinity. To determine the roles of exosites 1 and 2 in the high affinity interaction of thrombin with gammaA/gamma'-fibrin, binding studies were done with thrombin variants and exosite 1- or 2-directed ligands. alpha-Thrombin bound gammaA/gamma'-fibrin via high and low affinity binding sites. A peptide analog of the COOH terminus of the gamma'-chain that binds alpha-thrombin via exosite 2 blocked the high affinity binding of alpha-thrombin to gammaA/gamma'-fibrin, suggesting that the interaction of alpha-thrombin with the gamma'-chain is exosite 2-mediated. In support of this concept, (a) gamma-thrombin, which lacks a functional exosite 1, bound to gammaA/gamma'-fibrin, but not to gammaA/gammaA-fibrin; (b) thrombin R93A/R97A/R101A, an exosite 2-defective variant, bound only to gammaA/gamma'-fibrin via low affinity sites; and (c) exosite 2-directed ligands reduced alpha-thrombin binding to gammaA/gamma'-fibrin. However, several lines of evidence indicate that exosite 1 contributes to the high affinity interaction of thrombin with gammaA/gamma'-fibrin. First, the affinity of gamma-thrombin for gammaA/gamma'-fibrin was lower than that of alpha-thrombin. Second, removal of a low affinity binding site on the beta-chain of gammaA/gamma'-fibrin reduced its affinity for alpha-thrombin. Third, exosite 1-directed ligands reduced alpha-thrombin binding to gammaA/gamma'-fibrin. Taken together, these data suggest that, although exosite 2 mediates the interaction of thrombin with the gamma'-chain of gammaA/gamma'-fibrin, simultaneous ligation of exosite 1 by low affinity binding sites is essential for the high affinity interaction of thrombin with gammaA/gamma'-fibrin.  相似文献   

6.
Thrombin (T) inactivation by the serpin, heparin cofactor II (HCII), is accelerated by the glycosaminoglycans (GAGs) dermatan sulfate (DS) and heparin (H). Equilibrium binding and thrombin inactivation kinetics at pH 7.8 and ionic strength (I) 0.125 m demonstrated that DS and heparin bound much tighter to thrombin (K(T(DS)) 1-5.8 microm; K(T(H)) 0.02-0.2 microm) than to HCII (K(HCII(DS)) 236-291 microm; K(HCII(H)) 25-35 microm), favoring formation of T.GAG over HCII.GAG complexes as intermediates for T.GAG.HCII complex assembly. At [GAG] < K(HCII(GAG)) the GAG and HCII concentration dependences of the first-order inactivation rate constants (k(app)) were hyperbolic, reflecting saturation of T.GAG complex and formation of the T.GAG.HCII complex from T.GAG and free HCII, respectively. At [GAG] > K(HCII(GAG)), HCII.GAG complex formation caused a decrease in k(app). The bell-shaped logarithmic GAG dependences fit an obligatory template mechanism in which free HCII binds GAG in the T.GAG complex. DS and heparin bound fluorescently labeled meizothrombin(des-fragment 1) (MzT(-F1)) with K(MzT(-F1)(GAG)) 10 and 20 microm, respectively, demonstrating a binding site outside of exosite II. Exosite II ligands did not attenuate the DS-accelerated thrombin inactivation markedly, but DS displaced thrombin from heparin-Sepharose, suggesting that DS and heparin share a restricted binding site in or nearby exosite II, in addition to binding outside exosite II. Both T.DS and MzT(-F1).DS interactions were saturable at DS concentrations substantially below K(HCII(DS)), consistent with DS bridging T.DS and free HCII. The results suggest that GAG template action facilitates ternary complex formation and accommodates HCII binding to GAG and thrombin exosite I in the ternary complex.  相似文献   

7.
Heparin cofactor II (HCII) is a glycoprotein in human plasma that inhibits thrombin and chymotrypsin. Inhibition occurs when the protease attacks the reactive site peptide bond in HCII (Leu444-Ser445) and becomes trapped as a covalent 1:1 complex. Dermatan sulfate and heparin increase the rate of inhibition of thrombin, but not of chymotrypsin, greater than 1000-fold. The N-terminal portion of HCII contains two acidic repeats (Glu56-Asp-Asp-Asp-Tyr-Leu-Asp and Glu69-Asp-Asp-Asp-Tyr-Ile-Asp) that may bind to anion-binding exosite I of thrombin to facilitate covalent complex formation. To examine the importance of the acidic domain, we have constructed a series of 5' deletions in the HCII cDNA and expressed the recombinant HCII (rHCII) in Escherichia coli. Apparent second-order rate constants (k2) for inhibition of alpha-thrombin and chymotrypsin by each variant were determined. Deletion of amino acid residues 1-74 had no effect on the rate of inhibition of alpha-thrombin or chymotrypsin in the absence of a glycosaminoglycan. Similarly, the rate of inhibition of alpha-thrombin in the presence of a glycosaminoglycan was unaffected by deletion of residues 1-52. However, deletion of residues 1-67 (first acidic repeat) or 1-74 (first and second acidic repeats) greatly decreased the rate of inhibition of alpha-thrombin in the presence of heparin, dermatan sulfate, or a dermatan sulfate hexasaccharide that comprises the minimum high-affinity binding site for HCII. Deletion of one or both of the acidic repeats increased the apparent affinity of rHCII for heparin-Sepharose, suggesting that the acidic domain may interact with the glycosaminoglycan-binding site of native rHCII. The stimulatory effect of glycosaminoglycans on native rHCII was decreased by a C-terminal hirudin peptide which binds to anion-binding exosite I of alpha-thrombin. Furthermore, the ability of native rHCII to inhibit gamma-thrombin, which lacks the binding site for hirudin, was stimulated weakly by glycosaminoglycans. These results support a model in which the stimulatory effect of glycosaminoglycans on the inhibition of alpha-thrombin is mediated, in part, by the N-terminal acidic domain of HCII.  相似文献   

8.
We previously showed that the alpha-thrombin-antithrombin III complex causes antigenic change in vitronectin as monitored by the monoclonal anti-vitronectin antibody 8E6 (Tomasini & Mosher, 1988). We have extended these studies to other protease-serpin complexes and to gamma-thrombin, a proteolytic derivative of alpha-thrombin. In the presence of heparin, recognition of vitronectin by 8E6 was increased 64- or 52-fold by interaction with the complex of alpha-thrombin and heparin cofactor II or the Pittsburgh mutant (Met358----Arg) of alpha 1-protease inhibitor, respectively. This was comparable to the value obtained with the alpha-thrombin-antithrombin III complex. Factor Xa-serpin complexes were approximately 4-fold less effective than the corresponding thrombin complexes. alpha-Thrombin-serpin complexes but not Xa-serpin complexes formed disulfide-bonded complexes with vitronectin. Antigenic changes and disulfide-bonded complexes were not detected when trypsin- or chymotrypsin-serpin complexes were incubated with vitronectin. gamma-Thrombin caused 7- and 34-fold increases in recognition of vitronectin by MaVN 8E6 in the absence and presence of heparin, respectively. In contrast, alpha-thrombin by itself had no effect. The antigenic change induced by gamma-thrombin was maximal when gamma-thrombin and vitronectin were equimolar, was not dependent on cleavage of vitronectin, and was abolished by inhibition of gamma-thrombin with Phe-Pro-Arg-chloromethyl ketone but not with diisopropyl fluorophosphate. These data indicate that alpha-thrombin is the component in alpha-thrombin-serpin complexes that induces the antigenic change in vitronectin, probably via a region that is preferentially exposed in gamma-thrombin.  相似文献   

9.
Recently a thrombin receptor with a unique mechanism of activation was cloned from a megakaryocyte-like cell line (Vu et al., Cell 64:1057-1068, 1991). Thrombin cleaves a portion of this receptor creating a new N-terminus that acts as a "tethered-ligand" to activate the receptor. A thrombin receptor activating peptide (SFLLRNPNDKYEPF) homologous to the new N-terminus was shown to activate platelets. We synthesized this peptide and demonstrated that it desensitized platelets to activation by low concentrations of alpha-thrombin but not gamma-thrombin. We also synthesized a thrombin exosite inhibitor (BMS 180742) that inhibited platelet aggregation induced by low, but not high, concentrations of alpha-thrombin. In contrast, a thrombin active site inhibitor, N alpha-(2-naphthylsulfonyl-glycyl)-D,L-amidinophenylalanylpiperi dide, competitively inhibited thrombin-induced platelet aggregation. We conclude that thrombin-induced platelet activation is mediated by at least two pathways: one activated by low concentrations of alpha-thrombin and blocked by a thrombin exosite inhibitor that appears to be coupled to the "tethered-ligand" thrombin receptor, and another that is stimulated by higher concentrations of alpha-thrombin and by gamma-thrombin and does not require the thrombin exosite for activation. Both pathways are blocked by a thrombin active site inhibitor.  相似文献   

10.
Alpha-thrombin has two separate electropositive binding exosites (anion binding exosite I, ABE-I and anion binding exosite II, ABE-II) that are involved in substrate tethering necessary for efficient catalysis. Alpha-thrombin catalyzes the activation of factor V and factor VIII following discrete proteolytic cleavages. Requirement for both anion binding exosites of the enzyme has been suggested for the activation of both procofactors by alpha-thrombin. We have used plasma-derived alpha-thrombin, beta-thrombin (a thrombin molecule that has only ABE-II available), and a recombinant prothrombin molecule rMZ-II (R155A/R284A/R271A) that can only be cleaved at Arg(320) (resulting in an enzymatically active molecule that has only ABE-I exposed, rMZ-IIa) to ascertain the role of each exosite for procofactor activation. We have also employed a synthetic sulfated pentapeptide (DY(SO(3)(-))DY(SO(3)(-))Q, designated D5Q1,2) as an exosite-directed inhibitor of thrombin. The clotting time obtained with beta-thrombin was increased by approximately 8-fold, whereas rMZ-IIa was 4-fold less efficient in promoting clotting than alpha-thrombin under similar experimental conditions. Alpha-thrombin readily activated factor V following cleavages at Arg(709), Arg(1018), and Arg(1545) and factor VIII following proteolysis at Arg(372), Arg(740), and Arg(1689). Cleavage of both procofactors by alpha-thrombin was significantly inhibited by D5Q1,2. In contrast, beta-thrombin was unable to cleave factor V at Arg(1545) and factor VIII at both Arg(372) and Arg(1689). The former is required for light chain formation and expression of optimum factor Va cofactor activity, whereas the latter two cleavages are a prerequisite for expression of factor VIIIa cofactor activity. Beta-thrombin was found to cleave factor V at Arg(709) and factor VIII at Arg(740), albeit less efficiently than alpha-thrombin. The sulfated pentapeptide inhibited moderately both cleavages by beta-thrombin. Under similar experimental conditions, membrane-bound rMZ-IIa cleaved and activated both procofactor molecules. Activation of the two procofactors by membrane-bound rMZ-IIa was severely impaired by D5Q1,2. Overall the data demonstrate that ABE-I alone of alpha-thrombin can account for the interaction of both procofactors with alpha-thrombin resulting in their timely and efficient activation. Because formation of meizothrombin precedes that of alpha-thrombin, our findings also imply that meizothrombin may be the physiological activator of both procofactors in vivo in the presence of a procoagulant membrane surface during the early stages of coagulation.  相似文献   

11.
Anophelin: kinetics and mechanism of thrombin inhibition   总被引:5,自引:0,他引:5  
Anophelin is a 6.5-kDa peptide isolated from the salivary gland of Anopheles albimanus that behaves as an alpha-thrombin inhibitor. In this paper, kinetic analyses and the study of mechanism of alpha-thrombin inhibition by anophelin were performed. Anophelin was determined to be a reversible, slow, tight-binding inhibitor of alpha-thrombin, displaying a competitive type of inhibition. The binding of anophelin to alpha-thrombin is stoichiometric with a dissociation constant (K(i)) of 5.87 +/- 1.46 pM, a calculated association rate constant (k(1)) of 2.11 +/- 0.06 x 10(8) M(-1) s(-1), and a dissociation rate constant (k(-1)) of 4.05 +/- 0.97 x 10(-4) s(-1). In the presence of 0.15 and 0.4 M NaCl, a 17.6- and 207-fold increase in the K(i) of anophelin-alpha-thrombin complex was observed, respectively, indicating that ionic interactions are important in anophelin-alpha-thrombin complex formation. Incubation of alpha-thrombin with C-terminal hirudin fragment 54-65 that binds to alpha-thrombin anion binding exosite 1 (TABE1) attenuates alpha-thrombin inhibition by anophelin; anophelin also blocks TABE1-dependent trypsin-mediated proteolysis of alpha-thrombin. Using gamma-thrombin, an alpha-thrombin derivative where the anion binding exosite has been disrupted, anophelin behaves as a fast and classical competitive inhibitor of gamma-thrombin hydrolysis of small chromogenic substrate (K(i) = 0. 694 +/- 0.063 nM). In addition, anophelin-gamma-thrombin complex formation is prevented by treatment of the enzyme with D-Phe-Pro-Arg-chloromethyl ketone (PPACK), a reagent that irreversibly blocks the catalytic site of thrombin. It is concluded that anophelin is a potent dual inhibitor of alpha-thrombin because it binds both to TABE1 and to the catalytic site, optimal binding being dependent on the availability of both domains. Finally, anophelin inhibits clot-bound alpha-thrombin with an IC(50) of 45 nM and increases the lag phase that precedes explosive in vitro alpha-thrombin generation after activation of intrinsic pathway of blood coagulation. Because of its unique primary sequence, anophelin may be used as a novel reagent to study the structure and function of alpha-thrombin.  相似文献   

12.
Secondary structure and enzymatic properties of human a-thrombin and its gamma-form (obtaining during autolysis of the native enzyme) have been studied by differential scanning calorimetry (DSC) and circular dichroism (CD). According to DSC-data both alpha-thrombin and gamma-thrombin contained only one thermal transition peak at 58.5 and 53.3 degrees C, respectively. A comparison of these values suggested that gamma-form is less stable than initial a-thrombin. In contrast to that the thermogram of DIP-a-thrombin had two peaks (57.5 and 64.5 degrees C). CD spectra showed that conversion a- to gamma-thrombin influenced the secondary structure of the enzyme slightly. The study of the inhibitory effect of such polyanions as ATP and dextran sulfate (DS) upon thrombin-catalyzed cleavages of fibrinogen has shown that the growth of the negative charge of the polyanion molecule resulted in the increase of its inhibitory activity. The catalytically non-active DIP-alpha-thrombin, which retained the native anion-binding exosite 1, was shown to decrease the inhibitory power of the dextran sulfate. It was explained by competition of DS with the exosite 1 of both alpha- and DIP-alpha -thrombin. In contrast to that DIP-gamma-thrombin having exosite 1 destroyed neither competed nor influenced the anticoagulant capacity of dextran sulfate toward the native alpha-thrombin. In accordance with our data thrombin consists of two rather strong interacting domains. It was shown further that its anion-binding exosite 1 may play a significant role in the interaction of the enzyme with dextran sulfate.  相似文献   

13.
Prothrombin (Pro) activation by factor Xa generates the thrombin catalytic site and exosites I and II. The role of fragment 1 (F1) in the pathway of exosite I expression during Pro activation was characterized in equilibrium binding studies using hirudin(54-65) labeled with 6-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)hexanoate ([NBD]Hir(54-65)(SO3-)) or 5-(carboxy)fluorescein ([5F]Hir(54-65)(SO3-)). [NBD]Hir(54-65)(SO3-) distinguished exosite I environments on Pro, prethrombin 1 (Pre 1), and prethrombin 2 (Pre 2) but bound with the same affinities as [5F]Hir(54-65)(SO3-). Conversion of Pro to Pre 1 caused a 7-fold increase in affinity for the peptides. Conversely, fragment 1.2 (F1.2) decreased the affinity of Pre 2 for [5F]Hir(54-65)(SO3-) by 3-fold. This was correlated with a 16-fold increased affinity of F1.2 for Pre 2 in comparison to thrombin, demonstrating an enhancing effect of F1 on F1.2 binding. The active intermediate, meizothrombin, demonstrated a 50- to 220-fold increase in exosite affinity. Free thrombin and thrombin.F1.2 complex bound [5F]Hir(54-65)(SO3-) with indistinguishable affinity, indicating that the effect of F1 on peptide binding was eliminated upon expression of catalytic activity and exosite I. The results demonstrate a new zymogen-specific role for F1 in modulating the affinity of ligands for exosite I. This may reflect a direct interaction between the F1 and Pre 2 domains in Pro that is lost upon folding of the zymogen activation domain. The effect of F1 on (pro)exosite I and the role of (pro)exosite I in factor Va-dependent substrate recognition suggest that the Pro activation pathway may be regulated by (pro)exosite I interactions with factor Va.  相似文献   

14.
Heparin cofactor II (HCII) is a plasma serine protease inhibitor whose ability to inhibit alpha-thrombin is accelerated by a variety of sulfated polysaccharides in addition to heparin and dermatan sulfate. Previous investigations have indicated that calcium spirulan (Ca-SP), a novel sulfated polysaccharide, enhanced the rate of inhibition of alpha-thrombin by HCII. In this study, we investigated the mechanism of the activation of HCII by Ca-SP. Interestingly, in the presence of Ca-SP, an N-terminal deletion mutant of HCII (rHCII-Delta74) inhibited alpha-thrombin, as native recombinant HCII (native rHCII) did. The second-order rate constant for the inhibition of alpha-thrombin by rHCII-Delta74 was 2.0 x 10(8) M(-1) min(-1) in the presence of 50 microgram/ml Ca-SP and 10, 000-fold higher than in the absence of Ca-SP. The rates of native rHCII and rHCII-Delta74 for the inhibition of gamma-thrombin were increased only 80- and 120-fold, respectively. Our results suggested that the anion-binding exosite I of alpha-thrombin was essential for the rapid inhibition reaction by HCII in the presence of Ca-SP and that the N-terminal acidic domain of HCII was not required. Therefore, we proposed a mechanism by which HCII was activated allosterically by Ca-SP and could interact with the anion-binding exosite I of thrombin not through the N-terminal acidic domain of HCII. The Arg(103) --> Leu mutant bound to Ca-SP-Toyopearl with normal affinity and inhibited alpha-thrombin in a manner similar to native rHCII. These results indicate that Arg(103) in HCII molecule is not critical for the interaction with Ca-SP.  相似文献   

15.
The inhibition of thrombin by antithrombin III (AT III) and heparin has been studied in pure systems to determine the kinetics of inhibition during human prothrombin activation. The present study shows that prothrombinase-catalyzed prothrombin activation resulted in the generation of thrombin and meizothrombin(des F1). In the absence of heparin the second-order rate constants of the inactivation of both thrombin and meizothrombin(des F1) formed in the reaction mixture appeared to be identical, k = 3.7 X 10(5) M-1 min-1. The rate constant of inhibition of purified thrombin was 6.5 X 10(5) M-1 min-1. In the presence of heparin the decay of the amidolytic activity was biexponential and could be modeled by a four-parameter equation to determine the pseudo first-order rate constants of inhibition as well as the composition of the reaction with respect to the levels of thrombin and meizothrombin(des F1). The ratio of thrombin over meizothrombin(des F1) varied with the initial prothrombin concentration. Heparin catalyzed the AT III inhibition of thrombin but not meizothrombin(des F1) formed during the prothrombin activation. Thrombin, generated by (Xa-Va-phospholipid-Ca2+) was inhibited by AT III/heparin more slowly than purified thrombin, and the saturation kinetics of the inhibition with respect to AT III differed from those found with purified thrombin.  相似文献   

16.
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   

17.
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   

18.
Thrombin acts on many protein substrates during the hemostatic process. Its specificity for these substrates is modulated through interactions at regions remote from the active site of the thrombin molecule, designated exosites. Exosite interactions can be with the substrate, cofactors such as thrombomodulin, or fragments from prothrombin. The relative activity of alpha-thrombin for fibrinogen is 10 times greater than that for protein C. However, the relative activity of meizothrombin for protein C is 14 times greater than that for fibrinogen. Modulation of thrombin specificity is linked to its Na(+)-binding site and residues in autolytic loop-2 that interact with the Na(+)-binding site. Recombinant prothrombins that yield recombinant meizothrombin (rMT) and rMT des-fragment 1 (rMT(desF1)) enable comparisons of the effects of mutations at the Na(+)-binding residue (Asp(554)) and deletion of loop-2 (Glu(466)-Thr(469)) on the relative activity of meizothrombin for several substrates. Hydrolysis of t-butoxycarbonyl-VPR-p-nitroanilide by alpha-thrombin, recombinant alpha-thrombin, or rMT(desF1) was almost identical, but that by rMT was only 40% of that by alpha-thrombin. Clotting of fibrinogen by rMT and rMT(desF1) was 12-16% of that by alpha-thrombin, as already known. Strikingly, however, although meizothrombins modified by substitution of Asp(554) with either Ala or Leu or by deletion of loop-2 had 6-8 and <1%, respectively, of the clotting activity of alpha-thrombin, the activity of these meizothrombins for protein C was increased to >10 times that of alpha-thrombin. It is proposed that interactions within thrombin that involve autolytic loop-2 and the Na(+)-binding site primarily enhance thrombin action on fibrinogen, but impair thrombin action on protein C.  相似文献   

19.
Vitronectin (VN) has been implicated as a major matrix-associated regulator component of plasminogen activation by serving as a potent stabilizing cofactor of plasminogen activator inhibitor-1 (PAI-1). The direct binding of heparin, plasminogen as well as PAI-1 in its latent and active form to immobilized VN was studied in the absence or presence of competitors. Monoclonal antibodies against the carboxyl-terminal portion of VN inhibited both PAI-1 and plasminogen binding, whereas heparin, heparan sulfate with a high degree of sulfation, or dextran sulfate interfered with PAI-1 binding (KD = 20 nM) only. Utilizing synthetic peptides encompassing overlapping sequences of the heparin-binding domain of VN, adjacent heparin and PAI-1-binding sites were localized within the sequence 348-370 of VN. Although a number of other serine protease inhibitors which do not form binary complexes with VN contain a reactive-site Ser at their P1'-position, a reactive-site P1' mutant of PAI-1 (Met----Ser) showed comparable if not increased binding to VN. Binding of Lys-plasminogen and active-site-blocked plasmin was at least 10-fold higher in affinity (KD = 85-100 nM) compared to Glu-plasminogen (KD approximately 1 microM) and could be inhibited by lysine analogs but not by glycosaminoglycans or PAI-1, indicating that heteropolar plasmin(ogen) binding of VN occurs to an adjacent segment upstream to the heparin and PAI-1-binding sites. This contention was further supported in binding studies with plasmin-modified VN which lost both heparin and PAI-1 binding but exhibited 2-3-fold higher capacity to bind plasminogen. The essential plasmin(ogen)-binding site was mapped by ligand blot analysis to the carboxyl-terminal portion of proteolytically trimmed VN (M(r) = 61,000). Moreover, treatment of the extracellular matrix of human umbilical vein endothelial cells with plasmin resulted in partial degradation of matrix-associated VN and concomitant release of PAI-1, but increased the ability of the matrix by about 2-fold to bind plasminogen. These results are indicative of differential interactions of VN with components of the plasminogen activation system, whereby plasmin itself may provoke the switch of VN from an anti-fibrinolytic into a pro-fibrinolytic cofactor. This process reflects a novel role for the adhesive protein and its degradation product(s) in the possible feedback regulation of localized plasmin formation at extracellular sites.  相似文献   

20.
Inactivation of thrombin (T) by the serpins heparin cofactor II (HCII) and antithrombin (AT) is accelerated by a heparin template between the serpin and thrombin exosite II. Unlike AT, HCII also uses an allosteric interaction of its NH2-terminal segment with exosite I. Sucrose octasulfate (SOS) accelerated thrombin inactivation by HCII but not AT by 2000-fold. SOS bound to two sites on thrombin, with dissociation constants (KD) of 10 ± 4 μm and 400 ± 300 μm that were not kinetically resolvable, as evidenced by single hyperbolic SOS concentration dependences of the inactivation rate (kobs). SOS bound HCII with KD 1.45 ± 0.30 mm, and this binding was tightened in the T·SOS·HCII complex, characterized by Kcomplex of ∼0.20 μm. Inactivation data were incompatible with a model solely depending on HCII·SOS but fit an equilibrium linkage model employing T·SOS binding in the pathway to higher order complex formation. Hirudin-(54–65)(SO3) caused a hyperbolic decrease of the inactivation rates, suggesting partial competitive binding of hirudin-(54–65)(SO3) and HCII to exosite I. Meizothrombin(des-fragment 1), binding SOS with KD = 1600 ± 300 μm, and thrombin were inactivated at comparable rates, and an exosite II aptamer had no effect on the inactivation, suggesting limited exosite II involvement. SOS accelerated inactivation of meizothrombin 1000-fold, reflecting the contribution of direct exosite I interaction with HCII. Thrombin generation in plasma was suppressed by SOS, both in HCII-dependent and -independent processes. The ex vivo HCII-dependent process may utilize the proposed model and suggests a potential for oversulfated disaccharides in controlling HCII-regulated thrombin generation.  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