首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The molecular and genetic basis of a compound heterozygote for dys- and hypoprothrombinemia was analyzed. Abnormal nucleotide sequences of the human prothrombin gene were screened by PCR-single-strand conformation polymorphism (PCR-SSCP) with endonuclease digestion and mutated primer-mediated PCR-RFLP. A single nucleotide substitution responsible for dysprothrombinemia of prothrombin Tokushima was detected, as were three polymorphisms. The mutation for hypoprothrombinemia was detected by PCR-single-strand conformation polymorphism (PCR-SSCP) with endonuclease digestion in exon 6, near MboII-RFLP and NcoI-RFLP. Sequencing of PCR-amplified genomic DNA revealed a single base insertion of thymine (T) at position 4177. The resulting frameshift mutation caused both an altered amino acid sequence from codon 114 and a premature termination codon (i.e., TGA) at codon 174 in exon 7. Because exon 7 encodes the kringle 2 domain preceding the thrombin sequence, this frameshift leads to the null prothrombin phenotype. The inheritance of the hypoprothrombinemia gene from the father to the proband was proved by PCR-SSCP with endonuclease digestion and mutated primer-mediated PCR-RFLP.  相似文献   

2.
R A Henriksen  K G Mann 《Biochemistry》1988,27(26):9160-9165
A congenitally dysfunctional form of prothrombin, prothrombin Quick, was isolated from the plasma of an individual with less than 2% of normal prothrombin activity. Following activation of prothrombin Quick, two dysfunctional thrombins, thrombin Quick I and thrombin Quick II, were isolated. Functional characterization of thrombin Quick I indicated an increase in KM and a decrease in kcat, relative to thrombin, for release of fibrinopeptide A. Comparison of kcat/KM for thrombin Quick I to the value obtained for thrombin yielded a relative catalytic efficiency of 0.012 for thrombin Quick I [Henriksen, R. A., & Owen, W. G. (1987) J. Biol. Chem. 262, 4664-4669]. Lysyl endopeptidase digestor of reduced and S-carboxymethylated thrombin and thrombin Quick I has resulted in the identification of an altered peptide in this dysthrombin. Edman degradation of the isolated peptide has shown that the altered residue in this protein is Arg-382 which is replaced by Cys. This could result from a point mutation in the Arg codon, CGC, to yield TGC. Together, these results indicate that Arg-382 is a critical residue in determining the specificity of thrombin toward fibrinogen. Similar relative activities for thrombin Quick I in stimulating platelet aggregation, in the release of prostacyclin from human umbilical vein endothelium, and in the release of fibrinopeptide A suggest that these activities of thrombin share the same specificity determinants.  相似文献   

3.
Binding of prothrombin, prethrombin 1, prethrombin 2 and thrombin to fibrinogen-Sepharose was studied. Thrombin and prethrombin 2 bound to fibrinogen-Sepharose, while prethrombin 1 and prothrombin did not. Bound thrombin and prethrombin 2 were recovered from the column by eluting with 0.1 M NaCl/0.05 M Tris-HCl buffer (pH 7.4). The affinity of thrombin and prethrombin 2 to fibrinogen-Sepharose depended on ionic strength and reached a maximum at 50 mm concentration. Prethrombin 2 interacts with fibrinogen as well as thrombin; and prothrombin fragment 1.2 is not important in the formation of this complex. Thus, prethrombin 2, which is a precursor of thrombin without measurable enzymatic activity and which lacks the single cleavage at Arg-322-Ile-323 present in thrombin, has the same or very similar structural conformation as thrombin and has the same macromolecular substrate recognition site. These results confirm the earlier results that active center is not necessary in fibrinogen-thrombin interaction.  相似文献   

4.
Structural studies on a hereditary abnormal prothrombin, prothrombin Salakta, have been performed to identify the difference responsible for its reduced fibrinogen clotting activity and its reduced esterase activity. Amino acid composition and sequence analyses of a peptide isolated from a lysylendopeptidase digest of the abnormal thrombin indicated that Glu-466 had been replaced by Ala. This amino acid substitution can result from a single nucleotide change in the codon for Glu-466 (GAG----GCG). The model building and the molecular dynamics simulation of thrombin Salakta suggest that the Glu-466----Ala substitution would change the proper conformation around the substrate binding site containing Trp-468, which is a unique surface loop on the thrombin molecule. This is the experimental and theoretical evidence supporting the role of the surface loop containing Trp-468 for the proper conformation of the substrate binding site.  相似文献   

5.
The zymogen prothrombin is composed of fragment 1 containing a Gla domain and kringle-1, fragment 2 containing kringle-2, and a protease domain containing A and B chains. The prothrombinase complex assembled on the surface of platelets converts prothrombin to thrombin by cleaving at Arg-271 and Arg-320. The three-dimensional architecture of prothrombin and the molecular basis of its activation remain elusive. Here we report the first x-ray crystal structure of prothrombin as a Gla-domainless construct carrying an Ala replacement of the catalytic Ser-525. Prothrombin features a conformation 80 Å long, with fragment 1 positioned at a 36° angle relative to the main axis of fragment 2 coaxial to the protease domain. High flexibility of the linker connecting the two kringles suggests multiple arrangements for kringle-1 relative to the rest of the prothrombin molecule. Luminescence resonance energy transfer measurements detect two distinct conformations of prothrombin in solution, in a 3:2 ratio, with the distance between the two kringles either fully extended (54 ± 2 Å) or partially collapsed (≤34 Å) as seen in the crystal structure. A molecular mechanism of prothrombin activation emerges from the structure. Of the two sites of cleavage, Arg-271 is located in a disordered region connecting kringle-2 to the A chain, but Arg-320 is well defined within the activation domain and is not accessible to proteolysis in solution. Burial of Arg-320 prevents prothrombin autoactivation and directs prothrombinase to cleave at Arg-271 first. Reversal of the local electrostatic potential then redirects prothrombinase toward Arg-320, leading to thrombin generation via the prethrombin-2 intermediate.  相似文献   

6.
Diethyl pyrocarbonate (ethoxyformic anhydride) was used to modify histidyl residues in prothrombin. Diethyl pyrocarbonate inactivated the potential fibrinogen-clotting activity of prothrombin with a second-order rate constant of 70 M-1 min-1 at pH 6.0 and 25 degrees C. The difference spectrum of the modified protein had a maximum absorption at 240 nm which is characteristic of N-carbethoxyhistidine. The pH dependence for inactivation suggested the participation of a residue with a pKa of 6.2. Addition of hydroxylamine to ethoxyformylated prothrombin reversed the loss of fibrinogen-clotting activity. No structural differences were detected between the native and modified proteins using fluorescence emission and high-performance size-exclusion chromatography. The tyrosine and tryptophan content was not altered, but approximately 1-2 amino groups were modified. Statistical analysis of residual enzyme activity and extent of modification indicates that among 7 histidyl residues modified per molecule, there is 1 essential histidine (not in the active site) involved in the potential fibrinogen-clotting activity of prothrombin. To further examine its properties, the modified prothrombin was activated to thrombin using Echis carinatus venom protease. There was no difference in the catalytic activity of thrombin obtained from either native or ethoxyformylated prothrombin, as measured by H-D-Phe-pipecolyl-Arg-p-nitroanilide (D-Phe-Pip-Arg-NA) hydrolysis. However, thrombin produced from the modified protein showed a loss of fibrinogen-clotting activity but had a comparable apparent Ki value (about 20 microM) to thrombin from native prothrombin when fibrinogen was used as a competitive inhibitor during D-Phe-Pip-Arg-NA hydrolysis. The similarity in Ki values indicated that thrombin derived from diethyl pyrocarbonate-modified prothrombin does not have an altered fibrinogen-binding site. Although the histidyl residue involved during inactivation has not been identified, the results suggest that a histidyl residue in the thrombin portion of prothrombin is essential for interaction with fibrinogen.  相似文献   

7.
1. A method is described for the preparation and titration of prothrombin and thrombin. 2. Confirming the views of Morawitz, Howell (1916–17, 1925), and Bordet, thrombin cannot be regarded as an artificial by-product of coagulation (Wooldridge, Nolf (both quoted from Morawitz)). Calcium, a platelet factor, and a plasma factor (prothrombin) interact to form thrombin, and this then acts upon fibrinogen to form fibrin. The amount and rate of thrombin formation in the first reaction are independent of the presence or absence of fibrinogen. After a variable latent period, thrombin suddenly appears in large quantities, coincident with or immediately preceding the deposition of fibrin if fibrinogen is present. 3. The amount of thrombin formed in a mixture of prothrombin, Ca and platelets is independent of the platelet or Ca concentration, and depends primarily upon the amount of prothrombin used. The platelets (or cephalin) enormously accelerate the transformation of prothrombin to thrombin, and this acceleration seems to be their physiological rôle in the coagulation process. 4. Contrary to previous reports, platelets have not been demonstrated to contain significant quantities of prothrombin. 5. The available data do not allow any definite decision as to whether the platelet factor actually combines with prothrombin to form thrombin, or merely catalyzes the transformation. The very slow formation of thrombin in the complete absence of platelets may be due to dissolved traces of platelet material released during the physical manipulation of the plasma (centrifuging, Berkefeld filtration). 6. There was no evidence for a species-specific activity of platelets in the transformation of prothrombin to thrombin.  相似文献   

8.
Factor XIII is activated by thrombin, and this reaction is enhanced by the presence of fibrin(ogen). Using a substrate-based screening assay for factor XIII activity complemented by kinetic analysis of activation peptide cleavage, we show by using thrombin mutants of surface-exposed residues that Arg-178, Arg-180, Asp-183, Glu-229, Arg-233, and Trp-50 of thrombin are necessary for direct activation of factor XIII. These residues define a low specificity site known to be important also for both protein C activation and for inhibition of thrombin by antithrombin. The enhancing effect of fibrinogen occurs as a consequence of its conversion to fibrin and subsequent polymerization. Surface residues of thrombin further involved in high specificity fibrin-enhanced factor XIII activation were identified as His-66, Tyr-71, and Asn-74. These residues represent a distinct interaction site on thrombin (within exosite I) also employed by thrombomodulin in its cofactor-enhanced activation of protein C. In competition experiments, thrombomodulin inhibited fibrin-enhanced factor XIII activation. Based upon these and prior published results, we propose that the polymerization process forms a fibrin cofactor that acts to approximate thrombin and factor XIII bound to separate and complementary domains of fibrinogen. This enables enhanced factor XIII activation to be localized around the fibrin clot. We also conclude that proximity to and competition for cofactor interaction sites primarily directs the fate of thrombin.  相似文献   

9.
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   

10.
Albumin Tokushima is a Japanese genetic variant of human serum albumin. Two homozygous and 6 heterozygous subjects with this variant were found in a family. Albumin Tokushima was purified from sera of the homozygous subjects. Its amino acid composition and amino-terminal sequence were determined and compared with those of a normal serum albumin. Albumin Tokushima with the amino-terminal sequence of Arg-Gly-Val-Phe-His-Arg-Asp-Ala-His-Lys-Ser-Glu-Val-Ala-His-Arg-Phe-Lys- Asp- Leu-Gly-Glu-Glu-Asn-Phe was found to be the same abnormal proalbumin as proalbumin Lille (Abdo, Y. et al. (1981) FEBS Lett. 131, 286-288). The isoelectric points of albumin Tokushima were pH 4.70 and 4.90 as compared with pH 5.05 and 5.25 of a normal serum albumin. Albumin Tokushima was converted to normal serum albumin by purified cathepsin B in vitro. Albumin Tokushima can bind Ni2+ at 4 degrees C but binds little at 37 degrees C.  相似文献   

11.
Bothrojaracin is a potent and specific alpha-thrombin inhibitor (Kd approximately 0.6 nM) isolated from Bothrops jararaca venom. It binds to both of thrombin's anion-binding exosites (1 and 2), thus inhibiting the ability of the enzyme to act upon several natural macromolecular substrates, such as fibrinogen, platelet receptor, protein C, and factor V. Additionally, bothrojaracin interacts with prothrombin (Kd approximately 30 nM), as previously determined by a solid-phase assay. However, there is no information concerning the effect of this interaction on prothrombin activation and whether the binding of bothrojaracin can occur in plasma. Here, we show that bothrojaracin specifically interacts with prothrombin in human plasma. It is an effective anticoagulant after activation of the intrinsic pathway of blood coagulation, and analysis of prothrombin conversion in plasma shows that bothrojaracin strongly reduces alpha-thrombin formation. To determine whether this effect is due exclusively to inhibition of feedback reactions involving the thrombin-induced activation of factors V and VIII, we analyzed the effect of bothrojaracin on the activation of purified prothrombin by Oxyuranus scutellatus venom. As with plasma, bothrojaracin greatly inhibited thrombin formation, suggesting a direct interference in the prothrombin activation by the enzyme found in this venom (scuterin, a prothrombin activator described as a factor Xa/factor Va-like complex). Altogether, we suggest that bothrojaracin exerts its anticoagulant effect in plasma by two distinct mechanisms: (1) it binds generated thrombin and inhibits exosite 1 dependent activities such as fibrinogen clotting and factor V activation, and (2) it interacts with prothrombin and decreases its proteolytic activation. Thus, bothrojaracin may be useful in the search for thrombin inhibitors that bind both the zymogen and the active enzyme.  相似文献   

12.
The consequences of the combined effects of fibrin II monomer (FnIIm) and heparin (H) on the hydrolysis of peptidyl p-nitroanilide substrates by thrombin (IIa), the cleavage of prothrombin by thrombin and the thrombin-catalyzed release of fibrinopeptides from fibrinogen have been studied at pH 7.4 and I 0.15. The effects of fibrin II monomer and heparin on chromogenic substrate hydrolysis can be described by a hyperbolic mixed inhibition model in which substrate can interact with four possible enzyme species (IIa, IIa.H, IIa.FnIIm, and IIa.FnIIm.H) that arise as a result of random formation of a ternary complex among thrombin, fibrin II monomer, and heparin (Hogg, P. J. and Jackson, C. M. (1990) J. Biol. Chem. 265, 241-247). The formation of the ternary IIa.FnIIm.H complex results in an increase in the Km values of 7.03 +/- 1.17-fold (1.37-9.65 microM) and 1.94 +/- 0.60-fold (38.1-73.9 microM) for H-D-Ile-Pro-Arg-pNA and Cbz-Gly-Pro-Arg-pNA hydrolysis, respectively, and a decrease in the kc values of 0.45 +/- 0.08-fold (49.5-22.3 s-1) and 0.52 +/- 0.05-fold (93.1-48.4 s-1). Fibrin II monomer and heparin in combination also decrease the efficiency (kc/Km) with which thrombin cleaves prothrombin to produce Fragment 1 and Prethrombin 1 by 2.3-fold from 607 +/- 30 to 264 +/- 13 M-1 s-1. In contrast to the effects of fibrin II monomer and heparin on thrombin hydrolysis of chromogenic substrates, its proteolysis of prothrombin and its inactivation by antithrombin III (Hogg, P. J., and Jackson, C. M. (1989) Proc. Natl. Acad. Sci. U. S. A. 86, 3619-3623), these components have no discernible influence on the ability of thrombin to cleave fibrinogen. These observations indicate that the substrate specificity of thrombin is altered when it is bound in a complex with fibrin II monomer and heparin and suggest that the catalytic efficiency of thrombin for its physiological substrates will be affected differentially by these interactions. Such ternary complex formation involving thrombin, fibrin II monomer, and heparin may provide a mechanism for selectively regulating thrombin action.  相似文献   

13.
In the presence of a procoagulant fraction (Echis carinatus procoagulant) isolated from the venom of the saw-scaled viper Echis carinatus sochureki, purified human prothrombin (P1) is completely converted to thrombin. The first step is the removal of an NH2-terminal peptide (F1) representing approximately one-third of the prothrombin molecule. The remaining peptide (P2) is then cleaved by the action of E.c. procoagulant to yield a two-chain, disulfide-bridged protein (P'2) which has the same molecular weight as P2. P'2 has enzymic (thrombin) activity, as evidence by incorporation of radiolabeled diisopropylphosphate into its heavy chain (TB), hydrolysis of p-toluenesulfonylarginine methyl ester, and clotting of fibrinogen. Relative to thrombin, its esterolytic activity greatly exceeds its clot-promoting activity. Examination of the polypeptide chains obtained by reducing P'2 has shown that its larger chain (TB) is indistinguishable from the heavy chain of thrombin. Its other chain (F2TA) consists of the light chain (TA) of thrombin bound by peptide linkage to the protion of the prothrombin molecule which had been adjacent to F1. Removal of this portion (F2) is catalyzed by thrombin (and, evidently, by P'2), but not by the E.c. procoagulant. When F2 is removed from P'2, the remaining two-chian protein is indistinguishable from thrombin by any of the criteria applied--molecular weight, subunit chain composition, or enzymic activity. Polyacrylamide gel electrophoresis was carried out in sodium dodecyl sulfate before and after disulfide reduction of samples generated in the presence and in the absence of diisopropylphosphorofluoridate, which inhibits thrombin but not the E.c. procoagulant. Such experiments showed that thrombin (and probably P'2), as well as E.c. procoagulant, catalyzes the release of F1. Furthermore, thrombin brings about the cleavage of F1 to yield a two-chain, disulfidebridged protein (F'1). These observations, particularly those made in the course of characterizine P'2, have led to the conclusion that cleavage of the peptide bond connecting the TA and TB portions of the prothrombin molecule (or its derivatives) produces a serine active center and, hence, a molecule possessing thrombin activity. This cleavage is catalyzed by the E.c. procoagulant but not by thrombon itself.  相似文献   

14.
The effect of desialylation of fibrinogen on its conversion to fibrin has been investigated with particular reference to the kinetics of clot formation and structure. Also examined was the role of sialic acid in fibrinogen (factor I) poor in factor XIII (fibrinstabilizing factor) and factor I containing F XIII. The removal of more than 90% of the sialic acid of fibrinogen does not alter the thrombin clotting time, the clot solubility in monochloroacetic acid, the extent of cross-linking in the fibrin polymer, or the firmness and elasticity of the evolved clot. The data indicate that the sialic acid residues of fibrinogen do not contribute significantly to its conversion to fibrin by thrombin.  相似文献   

15.
The conversion of prothrombin into thrombin by Factor Xa requires the cleavage of two peptide bonds in prothrombin. Dependent on the order of cleavage, prethrombin 2 or meizothrombin occurs as intermediate. Since prethrombin 2 has as yet been the only observed intermediate, prothrombin activation is generally considered to proceed via prethrombin 2. In this paper we present new methods that allow differentiation between meizothrombin and thrombin formed during the initial phase of prothrombin activation. These methods, which make use of the different reactivities of meizothrombin and thrombin toward fibrinogen and antithrombin III plus heparin, enabled us to show the generation of considerable amounts of meizothrombin during Factor Xa-catalyzed prothrombin activation. Both meizothrombin and thrombin incorporated the active site-directed fluorescent chloromethyl ketone 5-dimethylaminonaphthalene-1-sulfonyl-Glu-Gly-Arg-CH2Cl. Gel electrophoretic analysis of chloromethyl ketone-treated aliquots of prothrombin activation mixtures confirmed meizothrombin formation. These observations demonstrate that prothrombin may also be converted into thrombin via meizothrombin.  相似文献   

16.
The tetradecapeptide Ac-D-F-L-A-E-G-G-G-V-R-G-P-R-V-OMe, which mimics residues 7f-20f of the A alpha-chain of human fibrinogen, has been co-crystallized with bovine thrombin from ammonium sulfate solutions in space group P2(1) with unit cell dimensions of a = 83.0 A, b = 89.4 A, c = 99.3 A, and beta = 106.6 degrees. Three crystallographically independent complexes were located in the asymmetric unit by molecular replacement using the native bovine thrombin structure as a model. The standard crystallographic R-factor is 0.167 at 2.3-A resolution. Excellent electron density could be traced for the decapeptide, beginning with Asp-7f and ending with Arg-16f in the active site of thrombin; the remaining 4 residues, which have been cleaved from the tetradecapeptide at the Arg-16f/Gly-17f bond, are not seen. Residues 7f-11f at the NH2 terminus of the peptide form a single turn of alpha-helix that is connected by Gly-12f, which has a positive phi angle, to an extended chain containing residues 13f-16f. The major specific interactions between the peptide and thrombin are 1) a hydrophobic cage formed by residues Tyr-60A, Trp-60D, Leu-99, Ile-174, Trp-215, Leu-9f, Gly-13f, and Val-15f that surrounds Phe-8f; 2) a hydrogen bond linking Phe-8f NH to Lys-97 O;3) a salt link between Glu-11f and Arg-173; 4) two antiparallel beta-sheet hydrogen bonds between Gly-14f and Gly-216; and 5) the insertion of Arg-16f into the specificity pocket. Binding of the peptide is accompanied by a considerable shift in two of the loops near the active site relative to human D-phenyl-L-prolyl-L-arginyl chloromethyl ketone (PPACK)-thrombin.  相似文献   

17.
The activating effect of Na(+) on thrombin is allosteric and depends on the conformational transition from a low activity Na(+)-free (slow) form to a high activity Na(+)-bound (fast) form. The structures of these active forms have been solved. Recent structures of thrombin obtained in the absence of Na(+) have also documented inactive conformations that presumably exist in equilibrium with the active slow form. The validity of these inactive slow form structures, however, is called into question by the presence of packing interactions involving the Na(+) site and the active site regions. Here, we report a 1.87A resolution structure of thrombin in the absence of inhibitors and salts with a single molecule in the asymmetric unit and devoid of significant packing interactions in regions involved in the allosteric slow --> fast transition. The structure shows an unprecedented self-inhibited conformation where Trp-215 and Arg-221a relocate >10A to occlude the active site and the primary specificity pocket, and the guanidinium group of Arg-187 penetrates the protein core to fill the empty Na(+)-binding site. The extreme mobility of Trp-215 was investigated further with the W215P mutation. Remarkably, the mutation significantly compromises cleavage of the anticoagulant protein C but has no effect on the hydrolysis of fibrinogen and PAR1. These findings demonstrate that thrombin may assume an inactive conformation in the absence of Na(+) and that its procoagulant and anticoagulant activities are closely linked to the mobility of residue 215.  相似文献   

18.
Previous results indicate extensive similarity of the active site regions of thrombin (EC 3.4.21.5) and Thrombin Quick, a congenital dysthrombin. A binding defect of Thrombin Quick toward fibrinogen is indicated by an increased KI when fibrinogen is present as a competitive inhibitor in the hydrolysis of tosyl-Gly-Pro-Arg-p-nitroanilide. In the present study, Thrombin Quick I is shown to have an activity of 1.3 and 34%, respectively, toward fibrinogen and prothrombin. Like the activity observed in prothrombin hydrolysis, Thrombin Quick I was 30% as effective as thrombin in stimulating release of thromboxane from platelets. Thrombin Quick was 1.7 and 2.4%, as effective as thrombin in stimulating platelet aggregation and prostacyclin production, respectively. Based on the activity of Thrombin Quick I in the reactions investigated, it is concluded that 1) the three cellular responses studied are initiated by proteolytic action of thrombin, 2) thrombin stimulation of aggregation and thromboxane release from platelets occurs via two different receptors, 3) the thrombin cellular interaction resulting in platelet aggregation and prostacyclin release must involve the thrombin active site as well as a secondary binding site required for optimal interaction with fibrinogen, and 4) the release of thromboxane from platelets does not involve the interaction of thrombin at the extrinsic binding site.  相似文献   

19.
The effect of bovine thrombomodulin on the specificity of bovine thrombin   总被引:8,自引:0,他引:8  
Bovine lung thrombomodulin is purified and used to investigate the basis of the change in substrate specificity of bovine thrombin when bound to thrombomodulin. Bovine thrombomodulin is a single polypeptide having an apparent molecular weight of 84,000 and associates with thrombin with high affinity and rapid equilibrium, to act as a potent cofactor for protein C activation and antagonist of reactions of thrombin with fibrinogen, heparin cofactor 2, and hirudin. Bovine thrombomodulin inhibits the clotting activity of thrombin with Kd less than 2.5 nM. Kinetic analysis of the effect of bovine thrombomodulin on fibrinopeptide A hydrolysis by thrombin indicates competitive inhibition with Kis = 0.5 nM. The active site of thrombin is little perturbed by thrombomodulin, as tosyl-Gly-Pro-Arg-p-nitroanilide hydrolysis and inhibition by antithrombin III are unaffected. Insensitivity of the reaction with antithrombin III is likewise observed with thrombin bound to thrombomodulin on intact endothelium. Antithrombin III-heparin, human heparin cofactor 2, and hirudin inhibit thrombin-thrombomodulin more slowly than thrombin. These effects may arise from a decrease in Ki of the inhibitors for thrombin-thrombomodulin or from changes in the active site not detected by tosyl-Gly-Pro-Arg-p-nitroanilide or antithrombin III. Bovine prothrombin fragment 2 inhibits thrombin clotting activity (Kd less than 7.5 microM) and acts as a competitive inhibitor of protein C activation (Kis = 2.1 microM). The data are consistent with a mechanism whereby thrombomodulin alters thrombin specificity by either binding to or allosterically altering a site on thrombin distinct from the catalytic center required for binding or steric accommodation of fibrinogen, prothrombin fragment 2, heparin cofactor 2, and hirudin.  相似文献   

20.
Thrombin bound to thrombomodulin activates thrombin-activable fibrinolysis inhibitor (TAFI) and protein C much more efficiently than thrombin alone. Although thrombomodulin has been proposed to alter the thrombin active site, the recently determined structure of the thrombin-thrombomodulin complex does not support this proposal. In this study, the contribution of amino acids near the activation site of TAFI toward thrombomodulin dependence was determined, utilizing four variants of TAFI with specific substitutions in the P6-P'3 region surrounding the Arg-92 cleavage site. Two point mutants had either the Ser-90 or Asp-87 of TAFI replaced with Ala, a third mutant had the thrombin activation site of the fibrinogen Bbeta-chain substituted into positions 91-95 of TAFI, and a fourth mutant had the thrombin activation site of protein C substituted into positions 90-95 of TAFI. Each of these mutants was expressed, purified, and characterized with respect to activation kinetics and functional properties of the enzyme. Even though fibrinogen is poorly cleaved by thrombin-thrombomodulin, the fibrinogen activation site does not significantly alter the thrombomodulin dependence of TAFI activation. The TAFI variant with the protein C activation sequence is only slowly activated by thrombin-thrombomodulin, and not at all by free thrombin. Mutating Asp-87 to Ala increases the catalytic efficiency of activation 3-fold both in the presence and absence of thrombomodulin, whereas mutating Ser-90 to Ala effects only minor kinetic differences compared with wild type TAFI. The thermal stabilities and antifibrinolytic properties of the enzymes were not substantially altered by any of the mutations that allowed for efficient activation of the enzyme. We conclude that residues in the P6-P'3 region of TAFI do not determine the thrombomodulin dependence of activation, which lends support to the argument that the role of thrombomodulin is to optimally orient thrombin and its substrate, rather than to allosterically alter the specificity of the thrombin active site.  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