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.
Inhibition of human factor IXa by human antithrombin.   总被引:2,自引:0,他引:2  
A procedure is presented for the purification of Factor IX from human plasma. The final product is homogeneous as judged by disc gel electrophoresis and sodium dodecyl sulfate gel electrophoresis. Furthermore, it is completely free of other coagulation component activities. Factor IX is converted to its enzymatically active form by the addition of small quantities of Factor IXa in the presence of calcium ions. This activated species is added to purified antithrombin-heparin cofactor and the interaction is studied in the presence and absence of heparin. Antithrombin-heparin cofactor is found to be a progressive, time-dependent inhibitor of Factor IXa and neutralizes approximately 57% of this enzyme's proteolytic activity within 30 min. The addition of heparin dramatically accelerates the rate of this interaction with virtually complete inhibition of Factor IXa occurring within 15 s. Sodium dodecyl sulfate gel electrophoresis of reduced and nonreduced proteins indicates that antithrombin-heparin cofactor functions as a potent inhibitor of Factor IXa by forming an undissociable complex with the enzyme which is stable in the presence of denaturing or reducing agents (or both). This complex represents a 1:1 stoichiometric combination of enzyme and inhibitor. Heparin increases the rate of formation of this complex without affecting its dissociability or stoichiometry.  相似文献   

2.
Human antithrombin-heparin cofactor is a naturally occurring plasma inhibitor of serine proteases generated during activation of the coagulation and fibrinolytic systems. We have demonstrated that purified preparations of this inhibitor also neutralize the esterolytic activity of human kallikrein as well as the ability of the enzyme to release kinins. When an excess of inhibitor is present, the inactivation process follows pseudo-first-order kinetics. Furthermore, the addition of heparin to mixtures of kallikrein and antithrombin-heparin cofactor results in a doubling of the rate and extent of enzyme neutralization. Disc gel analysis of incubation mixtures of kallikrein and excess antithrombin-heparin cofactor, with and without heparin, revealed that the enzyme band had vanished in conjunction with the emergence of two new electrophoretic species. These two new components probably represent stoichiometric complexes of kallikrein and antithrombin-heparin cofactor since a twofold increase in the concentration of enzyme doubled the concentration of these new molecular species. In plasmas which contain adequate levels of other protease inhibitors, antithrombin-heparin cofactor does not appear to be a quantitatively important inactivator of kallikrein. This is suggested by our finding that the addition of heparin at concentrations as high as 50 units/ml did not increase the inhibitory capacity of normal plasma directed against this enzyme. However, plasma from patients with hereditary angioedema had little neutralizing activity directed against kallikrein and revealed a marked increase in this inhibitory capacity when therapeutic concentrations of heparin were added. Our observation suggests that this acidic mucopolysaccharide may prove useful in controlling acute attacks of this disorder.  相似文献   

3.
The separation of active and inactive forms of heparin.   总被引:38,自引:0,他引:38  
Heparin has been fractionated into two distinct forms. The isolation of these species was accomplished by sucrose density gradient centrifugation of heparin mixed with antithrombin-heparin cofactor. Approximately 13 of this mucopolysaccharide was bound to antithrombin-heparin cofactor and had potent anticoagulant activity. This component was clearly separated from the remaining 23 of the heparin which could not form a stable complex with antithrombin-heparin cofactor and had minimal anticoagulant activity.  相似文献   

4.
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   

5.
The purpose of this study was to compare three heparin-binding plasma proteinase inhibitors in order to identify common and unique features of heparin binding and heparin-enhanced proteinase inhibition. Experiments with antithrombin, heparin cofactor, and protein C inhibitor were performed under identical conditions in order to facilitate comparisons. Synthetic peptides corresponding to the putative heparin binding regions of antithrombin, heparin cofactor, and protein C inhibitor bound to heparin directly and interfered in heparin-enhanced proteinase inhibition assays. All three inhibitors obeyed a ternary complex mechanism for heparin-enhanced thrombin inhibition, and the optimum heparin concentration was related to the apparent heparin affinity of the inhibitor. The maximum inhibition rate and rate enhancement due to heparin appeared to be unique properties of each inhibitor. In assays with heparin oligosaccharides of known size, only the antithrombin-thrombin reaction exhibited a sharp threshold for rate enhancement at 14-16 saccharide units. Acceleration of antithrombin inhibition of factor Xa, heparin cofactor inhibition of thrombin, and protein C inhibitor inhibition of thrombin, activated protein C, and factor Xa did not require a minimum saccharide size. The differences in heparin size dependence and rate enhancement of proteinase inhibition by these inhibitors might reflect differences in the importance of the ternary complex mechanism and other mechanisms, alterations in inhibitor reactivity, and orientation effects in heparin-enhanced proteinase inhibition.  相似文献   

6.
The plasma clearance of 125I-labeled human heparin cofactor II and its complex with thrombin was studied in mice to determine whether a specific mechanism exists for the catabolism of the inhibitor-proteinase complex. Initial studies demonstrated that murine plasma contains a heparin cofactor II-like inhibitor as shown by the presence of a dermatan sulfate-sensitive thrombin inhibitor. Human heparin cofactor II cleared from the circulation of mice with an apparent half-life of 80 min while heparin cofactor II-thrombin complexes cleared with an apparent half-life of only 10 min. The specificity of the clearance mechanism was investigated by clearance competition studies involving coinjection of excess unlabeled heparin cofactor II-alpha-thrombin, antithrombin III-alpha-thrombin, or alpha 1-proteinase inhibitor-elastase, and by tissue distribution studies. The results demonstrated that the clearance of 125I-labeled heparin cofactor II-alpha-thrombin is a receptor-mediated process, and that the same hepatocyte receptor system recognizes complexes containing heparin cofactor II, antithrombin III, and alpha 1-proteinase inhibitor.  相似文献   

7.
Although fibrin-bound thrombin is resistant to inactivation by heparin.antithrombin and heparin.heparin cofactor II complexes, indirect studies in plasma systems suggest that the dermatan sulfate.heparin cofactor II complex can inhibit fibrin-bound thrombin. Herein we demonstrate that fibrin monomer produces a 240-fold decrease in the heparin-catalyzed rate of thrombin inhibition by heparin cofactor II but reduces the dermatan sulfate-catalyzed rate only 3-fold. The protection of fibrin-bound thrombin from inhibition by heparin.heparin cofactor II reflects heparin-mediated bridging of thrombin to fibrin that results in the formation of a ternary heparin.thrombin.fibrin complex. This complex, formed as a result of three binary interactions (thrombin.fibrin, thrombin.heparin, and heparin.fibrin), limits accessibility of heparin-catalyzed inhibitors to thrombin and induces conformational changes at the active site of the enzyme. In contrast, dermatan sulfate binds to thrombin but does not bind to fibrin. Although a ternary dermatan sulfate. thrombin.fibrin complex forms, without dermatan sulfate-mediated bridging of thrombin to fibrin, only two binary interactions exist (thrombin.fibrin and thrombin. dermatan sulfate). Consequently, thrombin remains susceptible to inactivation by heparin cofactor II. This study explains why fibrin-bound thrombin is susceptible to inactivation by heparin cofactor II in the presence of dermatan sulfate but not heparin.  相似文献   

8.
Bovine plasma protein C inhibitor was purified; it was then characterized in comparison with human protein C inhibitor. The specific inhibitory activity of the purified inhibitor for bovine activated protein C was 8,500 times that of the inhibitor in plasma. The purified inhibitor showed a single band with Mr 56,000 by SDS-PAGE at pH 7.0, and two bands at pH 8.8, a major one with Mr 56,000 and a minor one with Mr 105,000, under both unreduced and reduced conditions. The pI range of the inhibitor was between 4.4 and 6.1. The Mr of the inhibitor was reduced by treatment with neuraminidase, O-glycanase, and also with glycopeptidase-A, suggesting that the inhibitor has both Asn-linked and Ser/Thr-linked carbohydrate chains. Twenty-seven of the NH2-terminal 49 amino acid residues of the bovine inhibitor, which lacks the first 4 residues from the NH2-terminal amino acid sequence of human inhibitor, were identical to those of the human inhibitor. The bovine inhibitor inhibited bovine and human activated protein C, human thrombin, Factor Xa, Factor XIa, and plasma kallikrein with Ki = 1.0, 5.2, 2.6, 3.0, 1.3 X 10(-8) M, and 4.5 X 10(-9) M, respectively. The inhibitory rates for activated protein C and thrombin were accelerated significantly in the presence of heparin or negatively charged dextran sulfate. However, the acceleration by heparin or dextran sulfate for the inhibition of Factor Xa, Factor XIa, and plasma kallikrein was not significant. The bovine inhibitor did not inhibit human Factor XIIa or plasmin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The effects of autolysis on the structure of chicken calpain II.   总被引:2,自引:2,他引:0       下载免费PDF全文
Heparin catalyses the inhibition of two key enzymes of blood coagulation, namely Factor Xa and thrombin, by enhancing the antiproteinase activities of plasma antithrombin III and heparin cofactor II. In addition, heparin can directly inhibit the activation of Factor X and prothrombin. The contributions of each of these effects to the anticoagulant activity of heparin have not been delineated. We therefore performed experiments to assess how each of these effects of heparin contributes to its anticoagulant activity by comparing the effects of heparin, pentosan polysulphate and D-Phe-Pro-Arg-CH2Cl on the intrinsic pathway of coagulation. Unlike heparin, pentosan polysulphate catalyses only the inhibition of thrombin by plasma. D-Phe-Pro-Arg-CH2Cl is rapid enough an inhibitor of thrombin so that when added to plasma no complexes of thrombin with its inhibitors are formed, whether or not the plasma also contains heparin. Heparin (0.66 microgram/ml) and pentosan polysulphate (6.6 micrograms/ml) completely inhibited the intrinsic-pathway activation of 125I-prothrombin to 125I-prothrombin fragment 1 + 2 and 125I-thrombin. On the addition of thrombin, a good Factor V activator, to the plasma before each sulphated polysaccharide, the inhibition of prothrombin activation was demonstrable only in the presence of higher concentrations of the sulphated polysaccharide. D-Phe-Pro-Arg-CH2Cl also completely inhibited the intrinsic-pathway activation of prothrombin in normal plasma. The inhibitory effect of D-Phe-Pro-Arg-CH2Cl was reversed if thrombin was added to the plasma before D-Phe-Pro-Arg-CH2Cl. The inhibition of the activation of prothrombin by the three agents was also abolished with longer times with re-added Ca2+. Reversal of the inhibitory effects of heparin and pentosan polysulphate was associated with the accelerated formation of 125I-thrombin-antithrombin III and 125I-thrombin-heparin cofactor complexes respectively. These results suggest that the anticoagulant effects of heparin and pentosan polysulphate are mediated primarily by their ability to inhibit the thrombin-dependent activation of Factor V, thereby inhibiting the formation of prothrombinase complex, the physiological activator of prothrombin.  相似文献   

10.
Protein C inhibitor is a plasma protein whose ability to inhibit activated protein C, thrombin, and other enzymes is stimulated by heparin. These studies were undertaken to further understand how heparin binds to protein C inhibitor and how it accelerates proteinase inhibition. The region of protein C inhibitor from residues 264-283 was identified as the heparin-binding site. This differs from the putative heparin-binding site in the related proteins antithrombin and heparin cofactor. The glycosaminoglycan specificity of protein C inhibitor was relatively broad, including heparin and heparan sulfate, but not dermatan sulfate. Non-sulfated and non-carboxylated polyanions also enhanced proteinase inhibition by protein C inhibitor. Heparin accelerated inhibition of alpha-thrombin, gamma T-thrombin, activated protein C, factor Xa, urokinase, and chymotrypsin, but not plasma kallikrein. The ability of glycosaminoglycans to accelerate proteinase inhibition appeared to depend on the formation of a ternary complex of inhibitor, proteinase, and glycosaminoglycan. The optimum heparin concentration for maximal rate stimulation varied from 10 to 100 micrograms/ml and was related to the apparent affinity of the proteinase for heparin. There was no obvious relationship between heparin affinity and maximum inhibition rate or degree of rate enhancement. The affinity of the resultant protein C inhibitor-proteinase complex was also not related to inhibition rate enhancement, and the results showed that decreased heparin affinity of the complex is not an important part of the catalytic mechanism of heparin. The importance of protein C inhibitor as a regulator of the protein C system may depend on the relatively large increase in heparin-enhanced inhibition rate for activated protein C compared to other proteinases.  相似文献   

11.
Protein C inhibitor (PCI) is a heparin-binding serine proteinase inhibitor (serpin) which is thought to be a physiological regulator of activated protein C (APC). The residues F353-R354-S355 (P2-P1-P1′) constitute part of the reactive site loop of PCI with the R-S peptide bond being cleaved by the proteinase. Changing the reactive site P1 and P2 residues to those of either proteinase nexin-1, α1-proteinase inhibitor or heparin cofactor II resulted in a decrease in inhibitory activity towards thrombin and APC. Changing the P2 residue F353 → P generated a rPCI which was a better thrombin inhibitor, but was 10-fold less active with APC. While these results support the concept that the P1 and P2 residues are important in the specificity of PCI, they suggest that the reactive site residues are not the only determinant of serpin specificity. Kinetic analysis of the rPCI variants was consistent with PCI operating by a mechanism similar to that proposed for other serpins. In this model an intermediary complex forms between inhibitor and proteinase that can proceed to either cleavage of the inhibitor as substrate or formation of an inactive complex.  相似文献   

12.
Rezaie AR  Olson ST 《Biochemistry》2000,39(39):12083-12090
Heparin catalyzes the inhibition of factor Xa by antithrombin mainly through an allosteric activation of the serpin inhibitor, but an alternative heparin bridging mechanism has been suggested to enhance the catalysis in the presence of physiologic calcium levels due to calcium interactions with the Gla domain exposing a heparin binding exosite in factor Xa. To provide direct evidence for this bridging mechanism, we studied the heparin-catalyzed reaction of antithrombin with factor Xa, Gla-domainless factor Xa (GDFXa), and a heparin binding exosite mutant of GDFXa in the absence and presence of calcium using rapid kinetic methods. The pseudo-first-order rate constant for factor Xa inhibition by antithrombin complexed with a long-chain approximately 70-saccharide heparin showed a saturable dependence on inhibitor concentration in the presence but not in the absence of 2.5 mM Ca(2+), indicating the formation of an intermediate heparin-serpin-proteinase encounter complex with a dissociation constant of approximately 90 nM prior to formation of the stable serpin-proteinase complex with a rate constant of approximately 20 s(-1). Similar saturation kinetics were observed for the inhibition of GDFXa by the antithrombin-heparin complex, except that Ca(2+) was not required for the effect. By contrast, no Ca(2+)-dependent saturation of the inhibition rate constant was detectable over the same range of inhibitor concentrations for reactions of either a short-chain approximately 26-saccharide high-affinity heparin-antithrombin complex with factor Xa or the long-chain heparin-antithrombin complex with the heparin binding exosite mutant, GDFXa R240A. These findings suggest that binding of full-length heparin chains to an exosite of factor Xa in the presence of Ca(2+) produces a chain-length-dependent lowering of the dissociation constant for assembly of the intermediate heparin-antithrombin-factor Xa encounter complex, resulting in a several 100-fold rate enhancement by a heparin bridging mechanism.  相似文献   

13.
Thrombin possesses two positively charged surface domains, termed exosites, that orient substrates and inhibitors for reaction with the enzyme. Because the exosites also allosterically modulate thrombin's activity, we set out to determine whether the structure or function of the exosites changes when thrombin forms complexes with antithrombin, heparin cofactor II, or alpha(1)-antitrypsin (M358R), serpins that utilize both, one, or neither of the exosites, respectively. Using a hirudin-derived peptide to probe the integrity of exosite 1, no binding was detected when thrombin was complexed with heparin cofactor II or alpha(1)-antitrypsin (M358R), and the peptide exhibited a 55-fold lower affinity for the thrombin-antithrombin complex than for thrombin. Bound peptide or HD-1, an exosite 1-binding DNA aptamer, was displaced from thrombin by each of the three serpins. Thrombin binding to fibrin also was abrogated when the enzyme was complexed with serpins. These data reveal that, regardless of the initial mode of interaction, the function of exosite 1 is lost when thrombin is complexed by serpins. In contrast, the integrity of exosite 2 is largely retained when thrombin is complexed by serpins, because interaction with heparin or an exosite 2-directed DNA aptamer was only modestly altered. The disorganization of exosite 1 that occurs when thrombin is complexed by serpins is consistent with results of protease sensitivity studies and crystallographic analysis of a homologous enzyme-serpin complex.  相似文献   

14.
Reactions between near equimolar amounts of antithrombin and Factors IXa or Xa resulted in the formation of a free proteolytically modified, two-chain form of the inhibitor, in addition to the inactive antithrombin-protease complexes. The modified inhibitor produced by either enzyme was electrophoretically identical with that formed in the reaction with thrombin. As in the latter reaction, the formation of the modified antithrombin by Factor Xa was increased in the presence of heparin, while only small amounts were produced by Factor IXa both in the absence and presence of the polysaccharide. NH2-terminal sequence analyses of the isolated modified inhibitor formed by Factor Xa showed that a single Arg-Ser bond in the COOH-terminal end of the inhibitor had been cleaved. This cleavage site is identical with that identified in free thrombin-modified antithrombin. The purified antithrombin-Factor IXa and antithrombin-Factor Xa complexes were dissociated by ammonia or hydroxylamine into free enzyme and a modified two-chain form of the inhibitor. Electrophoresis studies and NH2-terminal sequence analyses showed that the modified antithrombin obtained from either complex was identical with that produced in free form by the two enzymes and also with the modified inhibitor that is released from the antithrombin-thrombin complex. The fact that identical results were obtained for the reactions between antithrombin and three enzymes with different specificities strongly suggests that the observed Arg-Ser cleavage site is the active site of antithrombin.  相似文献   

15.
Preparations of low molecular weight porcine heparin with an average specific anticoagulant activity of 94 units/mg were fractionated into "active" and "relatively inactive" forms of the mucopolysaccharide of approximately 6000 daltons each. The active fraction was further subdivided into various species with descending but significant affinities for the protease inhibitor as well as decreasing but substantial anticoagulatn potencies. "Highly active" heparin (approximately 8% of the low molecular weight pool) possesses a specific anticoagulant activity of 350 +/- 10 units/mg. The relatively inactive fraction (67% of the low molecular weight pool) exhibits a specific anticoagulant activity of 4 +/- 1 units/mg. The binding of highly active heparin to antithrombin is accurately described by a single-site binding model with a KHep-ATDISS of approximately 1 X 10(-7) M. Variations in this binding parameter secondary to changes in environmental variables indicate that charge-charge interactions as well as an increase in entropy are critical to the formation of the highly active heparin-antithrombin complex. The interaction of relatively inactive heparin with the protease inhibitor is characterized by an apparent KHep-ATDISS of 1 X 10(-4) M. In large measure, this is due to small amounts of residual active mucopolysaccharide (0.5%). The ability of the highly active heparin to accelerate the thrombin-antithrombin interaction was also examined. We were able to demonstrate that the mucopolysaccharide acts as a catalyst in this process and is able to initiate multiple rounds of enzyme-inhibitor complex formation. The rate of enzyme neutralization is increased to a maximum of 2300-fold as the concentration of heparin is raised until the inhibitor is saturated with mucopolysaccharide. Further increases in heparin concentration result in a reduction in the speed of enzyme neutralization. This appears to be due to the formation of thrombin-heparin complexes. A mathematical model is given which provides a relationship between the initial velocity of enzyme neutralization and reactant concentrations.  相似文献   

16.
Acidic and non-acidic forms of rabbit thrombomodulin were studied with regard to their effects on the inhibition of thrombin by antithrombin in the presence of exogenous heparin. The non acidic form was obtained by proteolytic cleavage of a polyanionic component (presumably a sulfated polysaccharide) from the parent acidic form of thrombomodulin, and purified by ion-exchange chromatography. It was previously found that the acidic form of thrombomodulin increases the rate of thrombin inactivation by antithrombin. The present study showed that thrombin bound to acidic thrombomodulin was inactivated at a lower rate by antithrombin in the presence of exogenous heparin than was free thrombin or thrombin bound to the non-acidic form of thrombomodulin. The data suggest that the acidic component of thrombomodulin is primarily responsible for the retardation of thrombin-antithrombin complex formation in the presence of exogenous heparin. It is proposed that the polyanionic component of thrombomodulin blocks a site on thrombin required for heparin binding, thus rendering the antithrombin-heparin complex ineffective.  相似文献   

17.
Factor XII deficiency has been postulated to be a risk factor for thrombosis suggesting that factor XII is an antithrombotic protein. The biochemical mechanism leading to this clinical observation is unknown. We have previously reported high molecular weight kininogen (HK) inhibition of thrombin-induced platelet aggregation by binding to the platelet glycoprotein (GP) Ib-IX-V complex. Although factor XII will bind to the intact platelet through GP Ibalpha (glycocalicin) without activation, we now report that factor XIIa (0. 37 microm), but not factor XII zymogen, is required for the inhibition of thrombin-induced platelet aggregation. Factor XIIa had no significant effect on SFLLRN-induced platelet aggregation. Moreover, an antibody to the thrombin site on protease-activated receptor-1 failed to block factor XII binding to platelets. Inhibition of thrombin-induced platelet aggregation was demonstrated with factor XIIa but not with factor XII zymogen or factor XIIf, indicating that the conformational exposure of the heavy chain following proteolytic activation is required for inhibition. However, inactivation of the catalytic activity of factor XIIa did not affect the inhibition of thrombin-induced platelet aggregation. Factor XII showed displacement of biotin-labeled HK (30 nm) binding to gel-filtered platelets and, at concentrations of 50 nm, was able to block 50% of the HK binding, suggesting involvement of the GP Ib complex. Antibodies to GP Ib and GP IX, which inhibited HK binding to platelets, did not block factor XII binding. However, using a biosensor, which monitors protein-protein interactions, both HK and factor XII bind to GP Ibalpha. Factor XII may serve to regulate thrombin binding to the GP Ib receptor by co-localizing with HK, to control the extent of platelet aggregation in vivo.  相似文献   

18.
The interaction of high-molecular-weight (HMW) kininogen, Factor XII and prekallikrein with sulfatide was studied by fluorescence polarization. Fluorescein-conjugated derivatives of HMW kininogen, Factor XII and prekallikrein were prepared by reacting the purified bovine factors with fluorescein isothiocyanate (FITC). The apparent dissociation constant (Kd) for the binding of FITC-labeled HMW kininogen (F-HMW kininogen) with sulfatide was calculated to be 3.2 (+/- 0.3) X 10(-8) M. This binding was partially inhibited by three kininogen derivatives, fragment 1 X 2, kinin-free protein and fragment 1 X 2-light chain, but not by kinin and fragment 1 X 2-free protein. In the presence of Factor XII, the binding of F-HMW kininogen with sulfatide was strongly inhibited, suggesting that the zymogen and the protein cofactor compete for the same or a closely related binding site on the sulfatide surface. In contrast, the binding of FITC-labeled Factor XII (F-Factor XII) with sulfatide was weakly inhibited by HMW kininogen but not by prekallikrein. The Kd value for binding of F-Factor XII with sulfatide was calculated to be 2.0 (+/- 0.3) X 10(-8) M. F-Prekallikrein did not interact with sulfatide. Moreover, the fluorescence polarization value of F-HMW kininogen decreased in the presence of prekallikrein, leveling off at a one-to-one molar ratio of prekallikrein to F-HMW kininogen. The Kd value for binding of F-HMW kininogen-light chain (F-light chain) with prekallikrein was calculated to be 3.8 (+/- 0.6) X 10(-8) M and the stoichiometry was estimated as 1 to 1.2 on a molar basis from the Scatchard plot.  相似文献   

19.
Inhibition of bovine factor IXa and factor Xabeta by antithrombin III.   总被引:10,自引:0,他引:10  
Factor IXa and factor Xabeta are serine proteases which participate in the middle phase of blood coagulation. These two enzymes are inhibited by antithrombin III by the formation of an enzyme-inhibitor complex containing 1 mol of enzyme and 1 mol of antithrombin III. The complex was readily demonstrated by sodium dodecyl sulfate polyacrylamide gel electrophoresis and loss of coagulant or esterase activity at increasing concentrations of inhibitor. The inactivation of factor IXa by antithrombin III was relatively slow, but the reaction was greatly accelerated by the addition of heparin.  相似文献   

20.

Background

Factor VIII is the cofactor for Factor X activation by Factor IXa. Activated Factor X, Factor Xa, in turn activates prothrombin in a sequence that leads to fibrin clot formation at the site of vascular injury. Although the biochemistry of the cascade has been well studied, the molecular mechanism underlying the cofactor role of Factor VIII is not understood.

Methods

We screened a bacterial peptide display library with Factor IXa and Factor X co-immobilized on tosylactivated Dynabeads which were then used as platelet surrogates. Validation of peptide selection procedure and comparison of Factor VIII-like cofactor activity of oxidoreductases was performed using COATEST assays. Determination of Factor VIII as a folding catalyst with potential disulphide isomerase activity was determined using the RNase A renaturation assay.

Results

We set out to identify the cofactor requirements of the Factor IXa/Factor X procoagulant complex by random peptide display, and isolated a peptide with the active-site sequence, CGPC, of thioredoxin. This peptide was able to activate Factor X in a Factor IXa-dependent manner. Redox catalysts or oxidoreductases with homologous active-site vicinal cysteines such as PDI and DsbA also mimicked Factor VIII in their requirement of Factor IXa in Factor X activation. However, the cofactor activity of these peptides was up to a 1000-fold lower than that of Factor VIII and they were therefore unable to catalyse blood coagulation. Factor X activation by PDI and by Factor VIII was abolished by oxidation in an isolated system, which implies a possible role for thiol–disulphide exchange in the activity of the tenase complex. Using scrambled RNase A as a surrogate substrate, we also found that Factor VIII could renature this enzyme.

Conclusion

Our findings suggest that Factor VIII may be a specialized folding catalyst with disulphide isomerase activity. We suggest that it is this activity that may underlie its cofactor function in Factor X activation, and that this function is interchangeable with classical oxidoreductases.

General significance

The possible involvement of thiol–disulphide interchange as a mechanism underlying Factor VIII cofactor activity may provide some insight into the biochemistry of the intrinsic tenase complex.  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