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 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   

2.
We have examined the antithrombin effects of various phosphate-containing polyanions (including linear polyphosphates, polynucleotides and the phosphoserine glycoprotein, phosvitin) on the glycosaminoglycan-binding plasma proteinase inhibitors, antithrombin III (ATIII) and heparin cofactor II (HCII). These phosphate-containing polyanions accelerate the HCII-thrombin reaction, as much as 1600-fold in the case of phosvitin. The HCII-thrombin reaction with both phosvitin and polynucleotides appears to follow the ternary complex mechanism. The HCII-thrombin complex is rapidly formed in the presence of these phosphate polyanions (each at 10 micrograms/ml) when 125I-labeled thrombin is incubated with human plasma (ex vivo). None of these phosphate polyanions accelerate the ATIII-thrombin reaction. Our results suggest that the antithrombotic effect of these phosphate-containing polyanions is mediated by HCII activation and not by ATIII.  相似文献   

3.
The interactions of two proteinase inhibitors, heparin cofactor II and antithrombin, with thrombin are potentiated by heparin. Using two methods, we have studied the potentiating effects of a series of heparin (poly)saccharides with high affinity for antithrombin and mean Mr ranging from approx. 1700 to 18,800. First, catalytic amounts of heparin (poly)saccharide were added to purified systems containing thrombin and either heparin cofactor II or antithrombin. Residual thrombin activity was determined with a chromogenic substrate. It was found that only the higher-Mr polysaccharides (Mr greater than 8000) efficiently catalysed thrombin inhibition by heparin cofactor II, there being a progressive catalytic effect with increasing Mr of the polysaccharide. Weak accelerating effects were noted with low-Mr saccharides (Mr less than 8000). This contrasted with the well-characterized interaction of heparin with antithrombin and thrombin, where heparin oligosaccharides of Mr less than 5400 had absolutely no ability to accelerate the reaction, while (poly)saccharides of Mr exceeding 5400 showed rapidly increasing catalytic activity with increasing Mr. Secondly, these and other heparin preparations were added in a wide concentration range to plasma with which 125I-labelled thrombin was then incubated for 30 s. Inhibited thrombin was determined from the distribution of labelled thrombin amongst inhibitor-thrombin complexes, predominantly antithrombin-thrombin and heparin cofactor II-thrombin complexes. In this situation, where the inhibitors competed for thrombin and for the (poly)saccharides, it was found that, provided the latter were of high affinity for antithrombin and exceeded a Mr of 5400, thrombin inhibition in plasma was mediated largely through antithrombin. Polysaccharides of Mr exceeding 8000 that were of low affinity for antithrombin accelerated thrombin inhibition in plasma through their interaction with heparin cofactor II. High concentrations of saccharides of Mr 1700-5400 exhibited a size-dependent acceleration of thrombin inhibition, not through their interaction with antithrombin, but through their interaction with heparin cofactor II.  相似文献   

4.
The heparin cofactor II (HCII)/thrombin inhibition reaction is enhanced by various carboxylate polyanions. In the presence of polyaspartic acid, the HCII/thrombin reaction is accelerated more than 1000-fold with the second-order rate constant increasing from 3.2 x 10(4) M-1 min-1 (in the absence of polyAsp) to 3.6 x 10(7) M-1 min-1 as the polyAsp concentration is increased from 1 to 250 micrograms/ml. This accelerating effect was observed for HCII/thrombin, though to varying degrees, with other carboxylate polyanions. In contrast to HCII, the rate of antithrombin III inhibition of thrombin was decreased in the presence of polyAsp. The HCII/thrombin complex is rapidly formed in the presence of 10 micrograms/ml polyAsp when 125I-labeled-thrombin is incubated with plasma. It is possible that at physiological sites rich in carboxylate polyanions, thrombin may be preferentially inhibited by HCII.  相似文献   

5.
The anticoagulant properties of mast cell product, chondroitin sulphate E   总被引:2,自引:0,他引:2  
The anticoagulant potency in vitro of chondroitin sulphate E has been found to be similar to that of the heparinoids. In purified systems chondroitin sulphate E was shown to be principally an activator of heparin cofactor II. Maximum acceleration of heparin cofactor II:thrombin interaction was 185-fold (9.3 X 10(7) M-1 min-1), antithrombin III:thrombin interaction was 11-fold (4.16 X 10(6) M-1 min-1) and antithrombin III:factor Xa was 146-fold (3.86 X 10(6) M-1 min-1). Chondroitin sulphate E was observed to prolong the thrombin clotting time of fibrinogen in the absence of antithrombin III and heparin cofactor II. The effect appeared to be related to interference in thrombin:fibrinogen interaction rather than in fibrin monomer polymerization.  相似文献   

6.
Heparin cofactor II (Mr = 65,600) was purified 1800-fold from human plasma to further characterize the structural and functional properties of the protein as they compare to antithrombin III (Mr = 56,600). Heparin cofactor II and antithrombin III are functionally similar in that both proteins have been shown to inhibit thrombin at accelerated rates in the presence of heparin. There was little evidence for structural homology between heparin cofactor II and antithrombin III when high performance liquid chromatography-tryptic peptide maps and NH2-terminal sequences were compared. A partially degraded form of heparin cofactor II was also obtained in which a significant portion (Mr = 8,000) of the NH2 terminus was missing. The rates of thrombin inhibition (+/- heparin) by native and partially degraded-heparin cofactor II were not significantly different, suggesting that the NH2-terminal region of the protein is not essential either for heparin binding or for thrombin inhibition. A significant degree of similarity was found in the COOH-terminal regions of the proteins when the primary structures of the reactive site peptides, i.e. the peptides which are COOH-terminal to the reactive site peptide bonds cleaved by thrombin, were compared. Of the 36 residues identified, 19 residues in the reactive site peptide sequence of heparin cofactor II could be aligned with residues in the reactive site peptide from antithrombin III. While the similarities in primary structure suggest that heparin cofactor II may be an additional member of the superfamily of proteins consisting of antithrombin III, alpha 1-antitrypsin, alpha 1-antichymotrypsin and ovalbumin, the differences in structure could account for differences in protease specificity and reactivity toward thrombin. In particular, a disulfide bond which links the COOH-terminal (reactive site) region of antithrombin III to the remainder of the molecule and is important for the heparin-induced conformational change in the protein and high affinity binding of heparin does not appear to exist in heparin cofactor II. This observation provides an initial indication that while the reported kinetic mechanisms of action of heparin in accelerating the heparin cofactor II/thrombin and antithrombin III/thrombin reactions are similar, the mechanisms and effects of heparin binding to the two inhibitors may be different.  相似文献   

7.
Fucoidan, a sulfated fucopolysaccharide of marine algae is able to inhibit the proliferation of arterial smooth muscle cells half maximally at a concentration of 80 to 100 micrograms/ml culture medium. In comparable concentrations heparin was significantly less active than the fucopolysaccharide. Sulfation of fucoidan was found to be essential for expression of antiproliferative activity. The inhibitory effect of fucoidan is a time-dependent event with highest effectiveness during the first 6 h. Fucoidan does not influence the overall rate of synthesis of cell proteins and glycoconjugates, but led to substantial alterations in the synthesis and secretion of fibronectin and thrombospondin. Immunoprecipitation and quantitation revealed that the incorporation of [35S]methionine into fibronectin is reduced whereas thrombospondin synthesis was increased. The effect on fibronectin was not shared by heparin. Desulfation of the fucopolysaccharide abolished the observed modulation. Binding experiments with [125I]fucoidan indicate a saturable binding and a maximum of 2.8 x 10(6) bound molecules per cell. Fucoidan binding sites can be only partly displaced by heparin. The results suggest that both heparin and the structurally unrelated sulfated fucopolysaccharide act as an antiproliferative agent but differ in their modulation of cell metabolism.  相似文献   

8.
The minimum concentrations of heparin, dermatan sulfate, hirudin, and D-Phe-Pro-ArgCH2Cl required to delay the onset of prothrombin activation in contact-activated plasma also prolong the lag phases associated with both factor X and factor V activation. Heparin and dermatan sulfate prolong the lag phases associated with the activation of the three proteins by catalyzing the inhibition of endogenously generated thrombin. Thrombin usually activates factor V and factor VIII during coagulation. The smallest fragment of heparin able to catalyze thrombin inhibition by antithrombin III is an octadecasaccharide with high affinity for antithrombin III. In contrast, a dermatan sulfate hexasaccharide with high affinity for heparin cofactor II can catalyze thrombin inhibition by heparin cofactor II. A highly sulfated bis(lactobionic acid amide), LW10082 (Mr 2288), which catalyzes thrombin inhibition by heparin cofactor II and has both antithrombotic and anticoagulant activities, has been synthesized. In this study, we determined how the minimum concentration of LW10082 required to delay the onset of intrinsic prothrombin activation achieved this effect. We demonstrate that, like heparin and dermatan sulfate, LW10082 delays the onset of intrinsic prothrombin activation by prolonging the lag phase associated with both factor X and factor V activation. In addition, LW10082 is approximately 25% as effective as heparin and 10 times as effective as dermatan sulfate in its ability to delay the onset of prothrombin activation. The strong anticoagulant action of LW10082 is consistent with previous reports which show that the degree of sulfation is an important parameter for the catalytic effectiveness of sulfated polysaccharides on thrombin inhibition.  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.
Heparin cofactor II and antithrombin III are plasma proteins functionally similar in their ability to inhibit thrombin at accelerated rates in the presence of heparin. To further characterize the structural and functional properties of human heparin cofactor II as compared to antithrombin III, we studied the possible significance of arginyl and tryptophanyl residues and the changes in protein structure and activity during guanidinium chloride (GdmCl) denaturation. Both antithrombin and heparin cofactor activities of heparin cofactor II are inactivated by the arginine-specific reagent, 2,3-butanedione. Saturation kinetics are observed during modification and suggest formation of a reversible protease inhibitor-butanedione complex. Quantitation of arginyl residues following butanedione modification shows a loss of about four residues for total inactivation, one of which is essential for antithrombin activity. Arginine-modified heparin cofactor II did not bind to heparin-agarose and implies a role for the other modified arginyl residues during heparin cofactor activity. N-Bromosuccinimide oxidation (20 mol of reagent/mol of protein) of heparin cofactor II results in modification of approximately two tryptophanyl residues with no concomitant loss of heparin cofactor activity. Moreover, there is no enhancement of intrinsic protein fluorescence during heparin binding to the native inhibitor. Circular dichroism measurements show that the structural transition of heparin cofactor II during denaturation is distinctly biphasic, yielding midpoints at 0.6 and 2.6 M GdmCl. Functional protease inhibitory activities are affected to the same extent following denaturation-renaturation at various GdmCl concentrations. The results indicate that arginyl residues are critical for both antithrombin and heparin binding activities. In contrast, tryptophanyl residues are apparently not essential for heparin-dependent interactions. The results also suggest that heparin cofactor II contains two structural domains which unfold at different GdmCl concentrations.  相似文献   

11.
AimsFucoidan, a sulfated polysaccharide extracted from brown seaweed (F. vesiculosus) is recognized as an effective anticoagulant but its anti-lipidemic potency has not been well defined. We investigated the effect of fucoidan on lipoprotein lipase (LPL) secretion by human adipocytes.Main methodsLPL mRNA and protein expressions were measured using semi-quantitative RT-PCR, ELISA and immunohistochemistry in cultured adipocytes with or without fucoidan treatment. LPL enzyme activity was determined by a fluorometric assay.Key findingsIn cultured adipocytes, fucoidan induced LPL secretion in a dose- and time-dependent manner. An initial increase in LPL was maintained at a significant level but much slower than that in heparin-treated cells. Fucoidan also dose-dependently induced a cofactor of LPL, the apolipoprotein C-II (ApoC-II) secretion. In fucoidan-treated cells, LPL mRNA was time-dependently increased and LPL protein expression was also inceased. Treatment with both heparin and fucoidan showed no further increase in media LPL activity compared to heparin alone. In the conditioned medium from fucoidan-treated cells followed for 4 h, LPL activity decayed exponentially with half-life of about 180 min. In addition, the extracellular LPL mass in cycloheximide (a protein synthesis inhibitor) and fucoidan-treated cells did not change markedly, but LPL shifted significantly from active to inactive form.SignificanceThese results suggest that fucoidan acts like heparin by releasing LPL in addition to increasing the intracellular transport and decreasing the degradation of LPL in the medium. Furthermore, LPL and ApoC-II secretion induced by fucoidan may be involved in regulating plasma triglyceride lowering clearance.  相似文献   

12.
The selectin-mediated rolling of leukocytes along the endothelial cells is a prerequisite step followed by firm adhesion and extravasation into the inflamed tissue. This initial contact can be suppressed by sulphated polysaccharides. We have studied the effect of sulphated polysaccharides on the ultimate polymorphonuclear leukocyte (PMN) recruitment and plasma leakage in rabbit skin in response to intradermal injection of various inflammatory mediators. PMN infiltration evoked by various PMN chemoattractants (FMLP, C5a desArg, LTB(4) and IL-8) was significantly inhibited after intravenous injection of dextran sulphate (25 mg/kg), heparin (2 x 90 mg/kg) or fucoidan (1 mg/kg). PMN-dependent plasma leakage was equally well reduced by the different sulphated polymers. Vascular permeability induced by histamine or thrombin acting via a PMN-independent mechanism was not reduced. Fucoidan was the only polysaccharide able to suppress IL-1-induced PMN infiltration for 60-70%. Local administration of dextran sulphate had no effect on PMN-dependent plasma leakage. Differential inhibition of PMN recruitment was determined after injection of dextran sulphate or fucoidan depending on the type of insult. Therefore, these results suggest that different adhesion pathways are utilized during PMN recruitment in vivo in response to chemoattractants and IL-1.  相似文献   

13.
14.
Structure and activity of a unique heparin-derived hexasaccharide   总被引:2,自引:0,他引:2  
A hexasaccharide representing a major sequence in porcine mucosal heparin has been enzymatically prepared from heparin. Its structure was determined by an integrated approach using chemical, enzymatic, and spectroscopic methods. Two-dimensional 1H homonuclear COSY, C-H correlation NMR, and selective irradiation were used to assign many of the NMR resonances. In addition, new techniques including sulfate determination by ion chromatography and Fourier transform IR and californium plasma desorption mass spectroscopy have been applied, resulting in an unambiguous structural assignment of delta IdoAp2S(1----4)-alpha-D-GlcNp2S6S(1----4)-alpha-L-IdoAp++ +(1----4)-alpha-D-GlcNA cp6S-(1----4)-beta-D-GlcAp(1----4)-alpha-D-GlcNp2S3S6S (where delta IdoA represents 4-deoxy-alpha-L-threo-hex-4-enopyranosyluronic acid, p represents pyranose, and GlcA and IdoA represent glucuronic and iduronic acid). This hexasaccharide contains a portion of the antithrombin III-binding site and has a Kd of 4 X 10(-5) M. Unlike other small heparin oligosaccharides, which are specific for coagulation factor Xa, it inhibits both factors IIa and Xa equally through antithrombin III. This hexasaccharide may have the unique capacity to act primarily through heparin cofactor II to inhibit thrombin (factor IIa) and shows over half of heparin's heparin cofactor II-mediated anti-factor IIa activity. These studies suggest the occurrence of contiguous binding sites on heparin for Xa, antithrombin III, and heparin cofactor II.  相似文献   

15.
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   

16.
Plasminogen activator inhibitor 1 (PAI-1), a member of the serpin superfamily of proteins, has been demonstrated previously to interact functionally with the glycosaminoglycan heparin (Ehrlich, H.J., Keijer, J., Preissner, K. T., Klein Gebbink, R., and Pannekoek, H. (1991) Biochemistry 30, 1021-1028). Heparin specifically enhances the rate of association between PAI-1 and thrombin about 2 orders of magnitude, whereas no effect is detected with other serine proteases (e.g. factor Xa). For the heparin-dependent serpins antithrombin III and heparin cofactor II, basic amino acid residues in and around the helix D subdomain were proposed to be involved in the binding of glycosaminoglycans. Here we employed site-directed mutagenesis of full-length PAI-1 cDNA to identify the amino acid residues that mediate heparin binding. To that end, 15 single-point mutants of PAI-1, each having individual arginyl, lysyl, or histidyl residues replaced by a neutral (alanyl) residue ("ala-scan"), and one double mutant were constructed, expressed in Escherichia coli, and purified to apparent homogeneity. The purified biologically active proteins were subjected to the following analyses: (i) heparin-dependent inhibition of thrombin; (ii) heparin-dependent formation of sodium dodecyl sulfate-stable complexes with thrombin; and (iii) binding to and elution from heparin-Sepharose. Based on the data presented, we propose that the amino acid residues Lys65, Lys69, Arg76, Lys80, and Lys88 constitute major determinants for heparin binding of PAI-1. These residues are located in and around the helix D domain and are conserved in the other heparin-dependent thrombin inhibitors, antithrombin III and heparin cofactor II.  相似文献   

17.
F Lian  L He  N S Colwell  P Lollar  D M Tollefsen 《Biochemistry》2001,40(29):8508-8513
A monoclonal IgG isolated from a patient with multiple myeloma has been shown to bind to exosite II of thrombin, prolong both the thrombin time and the activated partial thromboplastin time (aPTT) when added to normal plasma, and alter the kinetics of hydrolysis of synthetic peptide substrates. Although the IgG does not affect cleavage of fibrinogen by thrombin, it increases the rate of inhibition of thrombin by purified antithrombin approximately 3-fold. Experiments with plasma immunodepleted of antithrombin or heparin cofactor II confirm that prolongation of the thrombin time requires antithrombin. By contrast, prolongation of the aPTT requires neither antithrombin nor heparin cofactor II. The IgG delays clotting of plasma initiated by purified factor IXa but has much less of an effect on clotting initiated by factor Xa. In a purified system, the IgG decreases the rate of activation of factor VIII by thrombin. These studies indicate that binding of a monoclonal IgG to exosite II prolongs the thrombin time indirectly by accelerating the thrombin-antithrombin reaction and may prolong the aPTT by interfering with activation of factor VIII, thereby diminishing the catalytic activity of the factor IXa/VIIIa complex.  相似文献   

18.
The interference of S protein with the heparin-catalyzed inhibition of thrombin by antithrombin III was studied in a purified system and in plasma. The effect of S protein to counteract heparin activity was documented by kinetic analysis of the initial phase of the inhibition reaction. Addition of S protein induced a concentration-dependent reduction of the inhibition rate, reflected in a decrease of the apparent pseudo-first-order rate constant by a factor of 5-8 in the presence of a twofold molar excess of S protein over antithrombin III. A non-competitive interaction of S protein with the thrombin--antithrombin-III--heparin inhibition reaction with Ki = 0.6 microM was found. While the association constant of thrombin--antithrombin III in the presence of 0.05 U/ml heparin amounted to 2.5 X 10(8) M-1, an approximately 200-fold decrease of this value was observed in the presence of S protein. The fast formation of the covalent complex between thrombin and antithrombin III in the presence of heparin was impaired as a result of the presence of S protein, as was shown by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. In the absence of heparin the inhibition of thrombin by antithrombin III alone was not influenced by S protein. The heparin-counteracting activity of S protein was found to be mainly expressed in the range of 0.01-0.1 U/ml heparin, thereby shifting the point of 50% inhibition of thrombin from 0.003 U/ml to 0.1 U/ml heparin with a second-order rate constant of k2 = 1.4 X 10(6) M-1. A direct interaction of S protein with heparin was demonstrated by crossed immunoelectrophoresis with purified proteins as well as in plasma and serum. The analysis of plasma and serum by crossed immunoelectrophoresis against rabbit anti-(human S protein) serum revealed an additional cathodal peak in the serum sample, resulting from the interaction of S protein with serum components. These findings not only indicate a direct interaction of S protein with heparin in the onset of the inhibition of thrombin by antithrombin-III--heparin, but also a contribution of S protein during enzyme-inhibitor complex formation.  相似文献   

19.
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   

20.
Heparin and dermatan sulfate increase the rate of inhibition of thrombin by heparin cofactor II (HCII) approximately 1000-fold by providing a catalytic template to which both the inhibitor and the proteinase bind. A variant form of HCII that binds heparin but not dermatan sulfate has been described recently in two heterozygous individuals (Andersson, T.R., Larsen, M.L., and Abildgaard, U. (1987) Thromb. Res. 47, 243-248). We have now purified the variant HCII (designated HCIIOslo) from the plasma of ne of these individuals. HCIIOslo or normal HCII (11 nM) was incubated with thrombin (9 nM) for 1 min in the presence of heparin or dermatan sulfate. Fifty percent inhibition of thrombin occurred at 26 micrograms/ml dermatan sulfate with normal HCII and greater than 1600 micrograms/ml dermatan sulfate with HCIIOslo. In contrast, inhibition of thrombin occurred at a similar concentration of heparin (1.0-1.5 micrograms/ml) with both inhibitors. To identify the mutation in HCIIOslo, DNA fragments encoding the N-terminal 220 amino acid residues of HCII were amplified from leukocyte DNA by the Taq DNA polymerase chain reaction and both alleles were cloned. A point mutation (G----A) resulting in substitution of His for Arg-189 was found in one allele. The same mutation was constructed in the cDNA of native HCII by oligonucleotide-directed mutagenesis and expressed in Escherichia coli. The recombinant HCIIHis-189 reacted with thrombin in the presence of heparin but not dermatan sulfate, confirming that this mutation is responsible for the functional abnormality in HCIIOslo.  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