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 ability of normal and malignant blood-borne cells to extravasate correlates with the activity of an endo-beta-D-glucuronidase (heparanase) which degrades heparan sulfate (HS) in the subendothelial extracellular matrix (ECM). The association of malignancy with different types of coagulopathies prompted us to study the effect of thrombin (EC 3.4.21.5), a serine protease elaborated during activation of the clotting cascade, on the ability of heparanase to degrade the ECM-HS. The circulating zymogen form of thrombin, prothrombin, was converted to proteolytically active thrombin during incubation with ECM. Thrombin generation by the ECM was time and dose dependent, reaching maximal conversion by 6 h incubation at 3 U/ml of prothrombin. Heparanase-mediated release of low Mr HS cleavage products from sulfate-labeled ECM was stimulated four- to sixfold in the presence of alpha-thrombin, but there was no effect on degradation of soluble HS. Similar results were obtained with heparanase preparations derived from mouse lymphoma and human hepatoma cell lines and from human placenta. Incubation of ECM with alpha-thrombin alone resulted in release of nearly intact high-Mr labeled proteoglycans. Thrombin stimulation of heparanase action was dose and time dependent, reaching a maximal value at 24 h incubation with 1 microM alpha-thrombin. The effect of modified thrombin preparations correlated with their proteolytic activity. Catalytically blocked preparations of thrombin (e.g., DIP-alpha-thrombin, MeSO2-alpha-thrombin) failed to facilitate heparanase action, while catalytically modified preparations (e.g., gamma-thrombin, NO2-alpha-thrombin) exerted only a slight enhancement. Antithrombin III (ATIII) and hirudin both inhibited thrombin-stimulated heparanase degradation of ECM-bound HS. Heparanase action was also facilitated by ECM-immobilized thrombin to an extent which was similar to that induced by soluble thrombin. This result implies that thrombin sequestered by the subendothelial ECM and protected from interaction with its natural inhibitor ATIII (Bar-Shavit et al., 1989, J. Clin. Invest. 84, 1096-1104) may participate locally in cellular invasion during tumor metastasis, inflammation, and autoimmunity.  相似文献   

2.
1. Crude globulin from more than 1,000 liters of citrated bovine plasma has been used in developing a procedure for moderately large scale separation of clotting factors. Fraction A, prothrombin, kinase, and thrombin fractions were prepared. Fraction A contained both kinase and accessory thromboplastin, the latter predominating when fraction A was diluted. 2. When prothrombin was activated by kinase, the rate of thrombin production was enhanced by the addition of platelets, or brain lipid, or dilute fraction A. These accessory thromboplastins caused this acceleration only when calcium chloride was added. Even with calcium, they were not effective unless kinase was present. 3. In contrast, the action of kinase was not entirely dependent on either ionic calcium or accessory thromboplastin. The concentrated kinase fraction activated prothrombin in the presence of excess oxalate. Although kinase often contaminates highly purified thrombins, it is probably distinct from thrombin. The ratio of kinase to thrombin was 100 times as great in the kinase fraction as in the thrombin fraction. 4. The kinase fraction, diluted 45,000-fold, to protein-nitrogen concentrations as low as 0.02 microgram per ml., accelerated the conversion of crude prokinase in three-stage tests. 5. The findings are consistent with the following concept of the basic enzymatic mechanism: See PDF for Structure It is now added that calcium and accessory thromboplastin exert their effects by impinging on the basic mechanism, in a chemically secondary or indirect manner.  相似文献   

3.
Sustained generation of alpha-thrombin and its breakdown forms at sites of thromboses has focused attention on the roles thrombin may play in vascular responses to thrombosis and injury. We have previously shown that alpha-thrombin stimulates many growth signals in cultured rat aortic smooth muscle cells (VSMC). To characterize thrombin growth mechanisms, we studied the effects on cultured VSMC of gamma-thrombin (catalytically active with obstructed anion-binding site required for clotting activity) and D-phenylalanyl-L-prolyl-L-arginine chloromethylketone-alpha-thrombin (catalytically inactive with intact anion-binding exosite) on cultured VSMC. Either derivative alone failed to increase growth, but in combination at 130 nM each, they caused a 75 +/- 5% increase in protein synthesis, similar to that observed with alpha-thrombin. This increase in protein synthesis was related to activation of protein kinase C (PKC) and Na+/H+ exchange, because only in combination could the derivatives increase phosphorylation of a 76,000-dalton PKC substrate and alkalinize the cells. Activation of PKC was correlated with a synergistic effect of the derivatives on diacylglycerol formation at 2 min (maximum, 55 +/- 1% combined increase vs. 24 +/- 9% and 4 +/- 4% individual increases with gamma- and D-phenylalanyl-L-prolyl-L-arginine chloromethylketone-alpha-thrombin alone, respectively, p less than 0.05). The derivatives stimulated PKC without increasing inositol trisphosphate, intracellular Ca2+, or expression of the protooncogene, c-fos. Thus, thrombin stimulation of Na+/H+ exchange, diacylglycerol formation, and growth of VSMC can be distinguished from thrombin mobilization of [Ca2+]i and induction of c-fos mRNA. These data indicate the presence of more than one mechanism for thrombin-mediated signaling events in cultured VSMC. Our results also suggest that various thrombin forms retained in clots may have significant effects on VSMC growth and function.  相似文献   

4.
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   

5.
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   

6.
Human prothrombin was acetylated to produce a modified prothrombin that upon activation by platelet-bound prothrombinase generates a form of thrombin that does not activate platelets but retains its amidolytic activity on a chromogenic peptide substrate. If normal prothrombin is used in such an assay, the thrombin that is generated activates the platelets in a feedback manner, accelerating the rate of thrombin generation and thereby preventing accurate measurement of the initial platelet procoagulant activity. Acetylation of prothrombin was carried out over a range of concentrations of sulfo-N-succinimidyl acetate (SNSA). Acetylation by 3 mM SNSA at room temperature for 30 min at pH 8.2 in the absence of metal ions produced a modified prothrombin that has <0.1% clotting activity (by specific prothrombin clotting assay), but it is activated by factor Xa (in the presence of either activated platelets or factor Va + anionic phospholipid) to produce thrombin activity that is measurable with a chromogenic substrate. Because the feedback action on the platelets is blocked, thrombin generation is linear, allowing quantitative measurement of the initial platelet activation state.  相似文献   

7.
Human neutrophil cathepsin G or bovine chymotrypsin proteolytically cleaved human alpha-thrombin at the B-chain Trp148-Thr149 bond generating a new form, zeta-thrombin. While incubation of alpha-thrombin with cathepsin G at pH 7.4 and 37 degrees C resulted in a partial loss of fibrinogen clotting activity, 86 +/- 13% of the clotting activity and 99 +/- 16% of the active sites titratable with p-nitrophenyl p-guanidinobenzoate were retained upon controlled passage of alpha-thrombin through chymotrypsin-Sepharose 4B at pH 6.2 or 7.4 and 24 degrees C (n = 15). Kinetic parameters for H-D-hexahydrotyrosyl-Ala-Arg p-nitroanilide were Km = 1.52 +/- 0.60 vs 1.32 +/- 0.18 microM and kcat = 51.9 +/- 2.9 vs 35.8 +/- 6.4 s-1 with alpha-thrombin vs chymotrypsin-prepared zeta-thrombin (n = 4 vs 3), respectively (I = 0.15 M, pH 7.4, and 24 degrees C). Some 95% of the clotting activity was lost when zeta-thrombin was passed through trypsin-Sepharose 4B under conditions for converting alpha- to nonclotting beta- and subsequently gamma-thrombin. The resulting gamma-like thrombins eluted bimodally with 260 and 310 mM NaCl when applied to Amberlite CG-50 resin [cross-linked poly(methylacrylic acid)] developed with a linear salt gradient in 50 mM Tris at pH 7.4 and 24 degrees C. These elution peaks correspond to 240, 330, and 350 mM NaCl for gamma-, alpha-, and zeta-thrombin, respectfully, implying that the anion-binding exosite is partially destroyed in gamma-like thrombins but is intact in zeta-thrombin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The effect of 95- (HRgpA) and 50-kDa gingipain R (RgpB), arginine-specific cysteine proteinases from periodontopathogenic bacterium Porphyromonas gingivalis on human prothrombin activation was investigated. Each enzyme released thrombin from prothrombin in a dose- and time-dependent manner with the former enzyme, containing adhesion domains, being 17-fold more efficient than the single chain RgpB. A close correlation between the generation of fibrinogen clotting activity and amidolytic activity indicated that alpha-thrombin was produced by gingipains R, and this was confirmed by SDS-polyacrylamide gel electrophoresis, thrombin active site labeling, and amino-terminal sequence analysis of prothrombin digestion fragments. Significantly, the catalytic efficiency of HRgpA to generate thrombin (k(cat)/K(m) = 1.2 x 10(6) m(-)1 s(-)1) was 100-fold higher than that of RgpB (k(cat)/K(m) = 1.2 x 10(4) m(-)1 s(-)1). The superior prothrombinase activity of HRgpA over RgpB correlates with the fact that only the former enzyme was able to clot plasma, and kinetic data indicate that prothrombin activation can occur in vivo. At P. gingivalis-infected periodontitis sites HRgpA may be involved in the direct production of thrombin and, therefore, in the generation of prostaglandins and interleukin-1, both have been found to be associated with the development and progression of the disease. Furthermore, by taking into account that the P. gingivalis bacterium has been immunolocalized in carotid atherosclerotic plaques at thrombus formation sites (Chiu, B. (1999) Am. Heart J. 138, S534-S536), our results indicate that bacterial proteinases may potentially participate in the pathogenesis of cardiovascular disease associated with periodontitis.  相似文献   

9.
Meizothrombin and thrombin formation were quantitated during factor Xa-catalyzed activation of human prothrombin in reaction systems containing purified proteins and in plasma. In the purified system considerable amounts of meizothrombin accumulated when prothrombin was activated by factor Xa (with or without accessory components) under initial steady state conditions. The ratio of the rates of meizothrombin and thrombin formation was not influenced by variation of the pH, temperature, or ionic strength of the reaction medium. When 2 microM prothrombin was activated by the complete prothrombinase complex (factor Xa, factor Va, Ca2+, and phospholipid) 80-90% of the initially formed reaction product was meizothrombin. Lowering the prothrombin concentration from 2 to 0.03 microM caused a gradual decrease in the ratio of meizothrombin/thrombin formation from 5 to 0.6. When the phosphatidylserine content of the phospholipid vesicles was varied between 20 and 1 mol % and prothrombin activation was analyzed at 2 microM prothrombin the relative amount of meizothrombin formed decreased from 85 to 55%. With platelets, cephalin, or thromboplastin as procoagulant lipid, thrombin was the major reaction product and only 30-40% of the activation product was meizothrombin. We also analyzed complete time courses of prothrombin activation both with purified proteins and in plasma. In reaction systems with purified proteins substantial amounts of meizothrombin accumulated under a wide variety of experimental conditions. However, little or no meizothrombin was detected in plasma in which coagulation was initiated via the extrinsic pathway with thromboplastin or via the intrinsic pathway with kaolin plus phospholipid (cephalin, platelets, or phosphatidylserine-containing vesicles). Thus, thrombin was the only active prothrombin activation product that accumulated during ex vivo coagulation experiments in plasma.  相似文献   

10.
By means of DEAE-Sephadex A-50 column chromatography, Trimeresurus gramineus venom was separated into 12 fractions. Fraction 8 had marked anticoagulant action in the tests of whole blood clotting time, calcium clotting time and plasma prothrombin time. Fraction 8 was rechromatographed on Sephadex G-100, then on DEAE-Sephadex A-50 again, and finally on Sephadex G-100, and a single peak was obtained. The patterns of microzone and disc electrophoresis also showed a single band. A single symmetrical boundary with 1.70 Svedberg units was obtained by ultracentrifugation. The estimated molecular weight was 19 500. The isoelectric point was pH 4.5. Chemical analysis showed that the anticoagulant principle was a glycoprotein and that it was thermolabile. The anticoagulant activity of this purified principle was 3.5 times higher than that of the crude venom. Fraction 5 potentiated its anticoagulant activity to 10 times higher than that of the crude venom. This principle did not possess caseinolytic, tosyl-L-arginine methyl ester esterase, phospholipase A, phosphodiesterase, alkaline phosphomonoesterase, fibrinolytic, hemorrhagic or local irritating activities. The purified anticoagulant principle did not destroy fibrinogen, induce fibrinolysis, inactivate thrombin nor interfere with the interaction between thrombin and fibrinogen. However, a marked inhibition of prothrombin activation was caused by the anticoagulant principle. The inhibition of prothrombin activation was not due to the destruction of prothrombin or its activation factors, but due to an interference in the interaction between prothrombin and its activation factors because of the reversible binding of these factors with the anticoagulant principle of the venom.  相似文献   

11.
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   

12.
Protein C activation by alpha-thrombin on the surface of endothelial cells depends on an essential membrane-glycoprotein cofactor, thrombomodulin. In the present study we have monitored the activity of thrombin-thrombomodulin complexes on human saphenous-vein endothelial cells (HSVEC) or on the endothelial cell line EA.hy 926. Cell monolayers were exposed for 5 min to 8.5 nM human alpha-thrombin and then washed to remove unbound thrombin. The cells were then incubated at 37 degrees C for 5-180 min. At the end of the respective incubation periods, purified human protein C (120 nM) was added in order to assay the activity of the thrombin-thrombomodulin complexes present on the cell surface. HSVEC pre-exposed to thrombin retained their full capacity to promote protein C activation up to 90 min after free thrombin was removed. This capacity then decreased slowly to reach 56% of control value after 180 min of incubation. Original activity was 3.8 +/- 0.9 pmol of activated protein C formed/min per ml per 10(6) cells (mean +/- S.E.M., n = 5). The capacity of protein C activation of EA.hy 926 cells remained constant for 120 min after free thrombin was removed, then decreased to 76% of control after 180 min. Original activity was 2.0 +/- 0.4 pmol of activated protein C formed/min per ml per 10(6) cells (mean +/- S.E.M., n = 3). Similar results were obtained with cells fixed with 3% paraformaldehyde. However, during the 5-180 min incubation period, non-fixed cells of both types were capable of significantly internalizing fluorescent acetylated low-density lipoprotein. In the experimental protocol used here, an eventual inhibition of thrombin internalization by protein C can be excluded, as protein C is only added at the end of the incubation period. We conclude that there is no evidence of rapid internalization of thrombin-thrombomodulin complexes on HSVEC or the EA.hy 926 cell line, as assessed by the ability of membrane-bound thrombin to activate protein C.  相似文献   

13.
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   

14.
Prothrombin contains two kringle domains that are removed during activation to the blood clotting enzyme alpha-thrombin. By analogy with other kringle-containing proteins the prothrombin kringles may play a role in the protein-protein interactions necessary for prothrombin activation. Four monoclonal antibodies to prothrombin kringle 2 have been produced against human prothrombin, and a fifth monoclonal antibody was produced against a synthetic peptide consisting of amino acid residues 216-231 of kringle 2. Each antibody was tested for its ability to block prothrombin activation by factor Xa. In the presence of phosphatidylcholine/phosphatidylserine vesicles and factor Va, two of the antibodies, alpha HII-3 and alpha HII-4, inhibited prothrombin activation at a 90 and 50% level, respectively. Two other monoclonal antibodies (alpha HII-6 and alpha HII-7) and the antipeptide antibody (alpha HII-5) had no effect on prothrombin activation. When factor Xa was the catalyst alone, antibody alpha HII-3 lost the ability to inhibit prothrombin activation whereas antibody alpha HII-4 again partially inhibited the reaction. When human platelets were the reaction surface, the patterns of inhibition by the anti-fragment 2 antibodies were identical to that observed with phospholipid vesicles. These data suggest a role for prothrombin fragment 2 in activation, possibly by mediating the interaction of substrate prothrombin with factor Xa or factor Va on the phospholipid surface.  相似文献   

15.
In order to investigate ligand binding sites in alpha-thrombin that interact with nonpolymerized fibrin, fibrinogen was conjugated (with CNBr) to Sepharose 4B and converted to the nonpolymerized fibrin resin with alpha-thrombin. Human alpha-thrombin was bound to the resin at 22 degrees C and eluted with a linear NaCl gradient [50-300 mM in 50 mM tris(hydroxymethyl)aminomethane hydrochloride, pH 7.6] with midpeak elution occurring at an ionic strength that corresponds to 170 +/- 5 mM NaCl. Among various ligands examined, ATP and its analogues caused alpha-thrombin to elute with 125 mM or less salt. Apparent dissociation constants were estimated by the dependence of elution volume on ligand concentration. The most potent ligands for desorption from the column were anionic (e.g., adenine nucleotides), which also inhibit thrombin esterolytic/amidolytic and clotting activity [Conery, B. G., & Berliner, L. J. (1983) Biochemistry 22, 369-375]. The desorption series was at 10 mM concentrations: ATP = ADP greater than pyrophosphate greater than citrate greater than oxalate greater than PO4(3-). Contrastingly, serotonin and related apolar compounds did not cause dissociation of alpha-thrombin from the fibrin resin, even though several of these substances inhibit fibrinogen clotting and esterolytic/amidolytic activities of the enzyme. These data imply that independent sites for apolar and anionic binding in alpha-thrombin are required for converting fibrinogen into clottable fibrin and that alpha-thrombin-fibrin binding involves an anionic site.  相似文献   

16.
The binding of 125I-labeled prothrombin fragment I. prethrombin I and alpha-thrombin to native and papain-treated tissue thromboplastin in the presence of CaCl2 of EDTA was studied. The experimental curves plotted in the Scatchard coordinates testify to the presence in thromboplastin of two types of fragment I binding sites: those with a high (Kd = 7.6 x 10(-6) M) and moderate (Kd = 1.3 x 10(-8) M) binding affinity. The parameters of fragment I binding and their changes reproduced, for the most part, the mode of prothrombin binding observed in previous studies. The experimental results provide indirect evidence in favour of a hydrophobic role of Ca(2+)-dependent binding of prothrombin fragment I to thromboplastin. The binding of prethrombin I was nonspecific and Ca(2+)-independent, whereas alpha-thrombin showed a relatively high level of nonspecific electrostatic binding which was competitively inhibited by Ca2+. Thromboplastin proteins interacted (both directly and in a Ca(2+)-independent fashion) with all the prothrombin derivatives under study.  相似文献   

17.
Cyanate reacts with unchanged amino groups of various proteins in a specific irreversible carbamylation reaction. The effect of this molecule on the clotting process and the effects of carbamylation on the clotting proteins and platelet functions were investigated in vitro. An immediate effect on the clotting proteins, not related to pH, was seen in the screening tests prothrombin time, partial thromboplastin time and thrombin time at the highest concentration (100 mM), to a lesser degree at the lower concentration (10 mM). These changes reflected decreases of 19 and 36% respectively in Factor V and X activity, an inhibition of 63-75% of Factors VII, IX, X and XI activity, and 80% inhibition of thrombin activity. The inhibitory changes of carbamylation increased with time. No changes were seen in the activity of Factors I and VIII. Platelet function studies revealed no inhibition of Factor III release; aggregation was abnormal only at high concentrations with epinephrine and collagen induction and partially reversible by resuspension in normal plasma.  相似文献   

18.
Hirugen, a synthetic dodecapeptide corresponding to the carboxyl-terminal amino acids 53-64 of hirudin, binds within a deep groove in thrombin that contains a cationic region referred to as the anion-binding exosite. This region is important in many of the binary interactions of thrombin with macromolecular substrates and cofactors. Fluorescein-labeled hirugen was used to probe which steps in the prothrombin activation process generate this anion-binding exosite. Two activation cleavage sites exist in bovine prothrombin. Cleavage at Arg274-Thr275 releases the activation fragments to generate the thrombin precursor, prethrombin 2. Cleavage of prothrombin within a disulfide loop at Arg323-Ile324 leads to formation of meizothrombin with no loss of peptide material but with formation of amidolytic activity. Cleavage of the same bond in prethrombin 2 generates thrombin. Hirugen, labeled at the amino terminus with fluorescein isothiocyanate, does not bind to prothrombin but does bind to thrombin (Kd = 9.6 +/- 1.2 x 10(-8) M), prethrombin 2 (Kd = 1.3 +/- 0.1 x 10(-7) M), thrombin-fragment-2 complex (Kd = 1.1 +/- 0.2 x 10(-6) M), and meizothrombin (Kd = 1.6 +/- 0.5 x 10(-8) M). Prothrombin fragment-2 and hirugen both bind independently to thrombin. A ternary complex can form with hirugen and fragment-2 and either thrombin or prethrombin 2, suggesting that fragment-2 and hirugen bind to discrete sites. Hirugen also alters the active site conformation of thrombin as detected by modulation of synthetic substrate hydrolytic activity. These studies suggest that conformational changes, rather than alleviating steric hindrance, are responsible for the formation of the hirugen-binding site during prothrombin activation. Furthermore, this conformational change can be effected by the cleavage of either of the two bonds required for activation of prothrombin.  相似文献   

19.
H Hofmann  C Bon 《Biochemistry》1987,26(3):772-780
In this paper, we show that the procoagulant action of Bothrops atrox venom is due in part to a protein component that activates prothrombin. The venom prothrombin activator was purified by ion-exchange chromatography and gel filtration. It was separated from a protease by affinity chromatography in a p-aminobenzamidine-CH-Sepharose column. It is a protein of about Mr 70,000, consisting of a single polypeptide chain. We have studied the kinetics of activation of prothrombin under different experimental conditions. The prothrombin activator from B. atrox venom is insensitive to reagents of serine and thiol proteases but is inactivated by ion chelators and by various divalent ions. These results suggest that it is a metalloenzyme. The prothrombin activator from B. atrox venom is inactive on the chromogenic substrates S-2337 and S-2238, and it is selective for prothrombin since it does not act on other blood coagulation factors such as fibrinogen and factor X. We have also studied the pattern of peptide cleavages produced in the human prothrombin molecule during the activation by the activator from B. atrox venom and compared it to that obtained with ecarin, a prothrombin activator from Echis carinatus venom. In the presence of thrombin inhibitors, e.g., hirudin, we found that the activators from B. atrox venom and ecarin act in a similar, or identical, manner by producing a thrombin intermediate, meizothrombin. In the absence of thrombin inhibitors, several peptides are generated, and alpha-thrombin is produced as a consequence of meizothrombin action.  相似文献   

20.
In addition to the 7-, 5- and 2-carboxyglutamyl varieties of dicoumarol-induced prothrombins (Malhotra, O.P. (1979) Thromb. Res. 15, 427-463), we have isolated two more atypical prothrombins, one containing 1.1 +/- 0.1 gamma-carboxyglutamic acid, '1-carboxyglutamyl prothrombin,' and the other less than 0.2, '0-carboxyglutamyl prothrombin.' Both variants showed a single component by analytical polyacrylamide-gel electrophoresis in the absence and in the presence of sodium dodecyl sulfate and contained antigenic activity indistinguishable from that of normal prothrombin. The pI of both proteins as assessed by electrofocusing was 4.835 +/- 0.015, compared with 4.58 for 10- and 7-, 4.75 for 5- and 4.81 for 2-carboxyglutamyl materials. By the two-stage prothrombin assay procedure, the 1- and 0-carboxyglutamyl variants generated thrombin, respectively 19 and 13% of normal prothrombin, and their activation times ranged from 4 to 7 h, compared with 7 min for normal. Kinetic studies, utilizing the one-stage coagulation assay, showed that both Km and tmin (minimal clotting time) increase proportionally with the decrease of gamma-carboxyglutamyl residues (from 10 to 7, 5, 2, 1 and 0 gamma-carboxyglutamic acids). Each of the five (partially) acarboxy prothrombins owe their clotting activity to gamma-carboxyglutamyl residues and not to the presence of some normal prothrombin molecules.  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