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.
A plasminogen activator secreted from human kidney cells was highly purified by affinity chromatography on an anti-urokinase IgG-Sepharose column. The purified plasminogen activator was inactive and had a single-chain structure and a Mr of 50,000. It not only did not incorporate diisopropyl fluorophosphate, which reacts with active site serine residue in urokinase, but also did not bind to p-aminobenzamidine-immobilized CH-Sepharose, to which urokinase bind via its side-chain binding pocket present in active center. The plasminogen activator was converted to the active two-chain form with the same Mr by catalytic amounts of plasmin. Its potential enzymatic activity was quenched completely by anti-urokinase IgG, but not by anti-tissue plasminogen activator Ig. These results indicate that the plasminogen activator is an inactive proenzyme form of human urokinase. Therefore, the plasminogen activator was termed single-chain pro-urokinase. The cleavage of single-chain pro-urokinase by plasmin induced conformational change which followed the generation of reactive serine residue at active site, the increase enzyme activity and the reduction of its high affinity for fibrin. These findings suggest that conformational change occurs in both regions responsible for enzyme activity and affinity for fibrin upon activation of single-chain pro-urokinase.  相似文献   

2.
Activation of plasminogen by pro-urokinase. II. Kinetics   总被引:3,自引:0,他引:3  
The kinetics of the activation of plasminogen by recombinant pro-urokinase obtained by expression of human urokinase cDNA in Escherichia coli was studied. The conversion of pro-urokinase (U) and plasminogen (P) to urokinase (u) and plasmin (p) is represented by a sequence of three reactions which each obey Michaelis-Menten kinetics, i.e. (Formula: see text). In this model, pro-urokinase formally behaves as an enzyme in Reaction I and as a substrate in reaction II. The experimentally measured overall rates of formation of urokinase and plasmin are in good agreement with those calculated from the kinetic parameters and the initial concentrations of pro-urokinase and plasminogen, confirming the validity of the model. It appears that recombinant pro-urokinase is an equally potent activator of plasminogen (k2/Km = 0.05 microM-1 s-1), as in urokinase (k"2/K"m = 0.02 microM-1 s-1). This is due to the fact that the proenzyme, which is virtually inactive toward low Mr substrates for urokinase, forms an intermediate of the Michaelis-Menten type with plasminogen, with a much higher affinity than that of the active enzyme with its substrate. This is an exceptional phenomenon among the serine proteases.  相似文献   

3.
The relative fibrin-binding, fibrinolytic and fibrinogenolytic properties of single-chain pro-urokinase, an inactive proenzyme form of human urokinase purified from cultured human kidney cells, and urokinase were compared. The affinity of single-chain pro-urokinase for fibrin was much higher than that of urokinase. In Vitro thrombolytic studies showed that single-chain pro-urokinase is approximately three times more potent in fibrinolysis than urokinase and that it does not degrade fibrinogen in the plasma at a concentration, at which complete plasma clot lysis takes place; whereas, urokinase extensively degrades the fibrinogen in the plasma. These specific, potent thrombolytic properties of single-chain pro-urokinase seem to be due to its high affinity for fibrin and to its conversion from the inactive single-chain form to the active two-chain form on the thrombus by the catalytic amount of plasmin generated during coagulation. This single-chain pro-urokinase obtained from human kidney cells by tissue culture should prove advantageous than urokinase in thrombolytic therapy.  相似文献   

4.
Plasma kallikrein was found to be a good activator of pro-urokinase, the inactive zymogen form of urokinase. The complete activation of pro-urokinase by plasma kallikrein was obtained in 2 h with an enzyme/substrate weight ratio of 1/30. The rate of activation of pro-urokinase by plasma kallikrein was comparable to that catalyzed by plasmin and trypsin. The rate of activation of pro-urokinase by factor XIIa was approximately one-seventh of that by plasma kallikrein. The activation of the zymogen was due to the cleavage of a single internal peptide bond, resulting in the conversion of a single chain pro-urokinase (Mr = 55,000) into two-chain urokinase (Mr = 33,000 and 22,000), and these two chains were linked by a disulfide bond(s). These results indicate an important role of plasma kallikrein for the activation of pro-urokinase in the factor XII-dependent intrinsic pathway of fibrinolysis. Thrombin also converted pro-urokinase to a two-chain form that was not activatable by plasmin, plasma kallikrein, and factor XIIa. Thrombin specifically cleaved the Arg 156-Phe 157 bond which is located 2 residues prior to the activation site of Lys 158-Ile 159.  相似文献   

5.
alpha 2-antiplasmin (alpha 2-AP) exerts its inhibitory effect on fibrinolysis by rapidly inhibiting the plasmin evolved; in addition, it has been suggested that interference with the binding of plasminogen to fibrin, a function shared with histidine-rich glycoprotein (HRGP), may also be significant in inhibition of fibrinolysis. To elucidate if plasminogen binding by these two alpha 2-globulins may decrease the generation of plasmin by tissue-type plasminogen activator (t-PA) at the surface of fibrin, a system mimicking the fibrin/plasma interface was used. Attempts were made to differentiate the plasminogen binding from the plasmin inhibitory function of alpha 2-AP. The activation of human Glu-plasminogen (native plasminogen with NH2-terminal glutamic acid) by fibrin-bound t-PA was performed in a plasma environment using either normal plasma, alpha 2-AP- or HRGP-depleted plasmas supplemented with increasing amounts of the lacking protein, or in a reconstituted system with purified plasminogen and various concentrations of alpha 2-AP and HRGP. The activation of Glu-plasminogen in alpha 2-AP-depleted plasma containing a normal concentration of HRGP produced a time-dependent increase in the generation of plasmin. The addition of 1 microM-alpha 2-AP to this plasma prevented the formation of Lys-derivatives and produced a marked decrease (42%) in the number of plasminogen-binding sites. In contrast, the addition of 1.5 microM-HRGP to HRGP-depleted plasma containing a normal amount of alpha 2-AP produced only a modest (17%) decrease in the amount of plasmin(ogen) bound. Moreover, in a purified system the amount of plasminogen-binding sites and thereby of plasmin generated at the surface of fibrin in the presence of both alpha-2 globulins was similar to the amount generated in the presence of alpha 2-AP alone. These results indicate clearly that the formation of reversible complexes between plasminogen and alpha 2-AP does not interfere with the binding and activation of plasminogen at the fibrin surface. In contrast, the inhibition of plasmin by alpha 2-AP decreases importantly the number of plasminogen-binding sites (carboxyl-terminal lysines) and inhibits thereby the accelerated phase of fibrinolysis. It can be concluded that interference of the binding of plasminogen to fibrin by alpha 2-AP during plasminogen activation, does not play a significant role in inhibition of fibrinolysis, and that the plasminogen-binding effect of HRGP, if any, is obscured by the important inhibitory effect of alpha 2-AP.  相似文献   

6.
The fibrinolytic system comprises a proenzyme, plasminogen, which can be converted to the active enzyme, plasmin, which degrades fibrin. Plasminogen activation is mediated by plasminogen activators, which are classified as either tissue-type plasminogen activators (t-PA) or urokinase-type plasminogen activators (u-PA). Inhibition of the fibrinolytic system may occur at the level of the activators or at the level of generated plasmin. Plasmin has a low substrate specificity, and when circulating freely in the blood it degrades several proteins including fibrinogen, factor V, and factor VIII. Plasma does, however, contain a fast-acting plasmin inhibitor, alpha 2-antiplasmin, which inhibits free plasmin extremely rapidly but which reacts much slower with plasmin bound to fibrin. A "systemic fibrinolytic state" may, however, occur by extensive activation of plasminogen and depletion of alpha 2-antiplasmin. Clot-specific thrombolysis therefore requires plasminogen activation restricted to the vicinity of the fibrin. Two physiological plasminogen activators, t-PA and single-chain u-PA (scu-PA) induce clot-specific thrombolysis, via entirely different mechanisms, however. t-PA is relatively inactive in the absence of fibrin, but fibrin strikingly enhances the activation rate of plasminogen by t-PA. This is explained by an increased affinity of fibrin-bound t-PA for plasminogen and not by alteration of the catalytic rate constant of the enzyme. The high affinity of t-PA for plasminogen in the presence of fibrin thus allows efficient activation on the fibrin clot, while no significant plasminogen activation by t-PA occurs in plasma. scu-PA has a high affinity for plasminogen (Km = 0.3 microM) but a low catalytic rate constant (kcat = 0.02 sec-1). However, scu-PA does not activate plasminogen in plasma in the absence of a fibrin clot, owing to the presence of (a) competitive inhibitor(s). Fibrin-specific thrombolysis appears to be due to the fact that fibrin reverses the competitive inhibition. The thrombolytic efficacy and fibrin specificity of natural and recombinant t-PA has been demonstrated in animal models of pulmonary embolism, venous thrombosis, and coronary artery thrombosis. In all these studies intravenous infusion of t-PA at sufficiently high rates caused efficient thrombolysis in the absence of systemic fibrinolytic activation. The efficacy and relative fibrinogen-sparing effect of t-PA was recently confirmed in three multicenter clinical trials in patients with acute myocardial infarction.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
Previous results have shown that the autoantibody eluted from the glomeruli of rats with active Heymann nephritis contain a population of antibodies not only to the putative autoantigen of the disease, gp330, but alos to plasminogen. Since gp330 has been shown to serve as a receptor for plasminogen, we have analyzed the effects of autoantibody on plasminogen-binding to gp330 and activation of plasminogen to plasmin by urokinase. Autoantibody does not inhibit the binding of plasminogen to gp330. The change in the conformation of plasminogen when its lysine-binding sites are occupied or after conversion to plasmin results in a significant decrease in autoantibody-binding. The most significant effect of autoantibody on this system is the inhibition of plasminogen activation to plasmin by urokinase. The binding of autoantibody to plasminogen acts as a competitive inhibitor of the reaction by apparently blocking access of urokinase to plasminogen's activation site. These results indicate that autoantibody obtained from the immune deposits in the glomeruli of rats with active Heyman nephritis does not inhibit the binding of plasminogen to gp330 but does significantly alter the urokinase catalyzed activation of plasminogen to plasmin.  相似文献   

8.
The human T cell-associated serine proteinase-1 (HuTSP-1) is expressed by activated T lymphocytes and is exocytosed upon their interaction with target cells. Here, we report that HuTSP-1 is able to convert single-chain human pro-urokinase into the active two-chain enzyme. Time-dependent activation by HuTSP-1 of recombinant human pro-urokinase as well as natural pro-urokinase derived from human melanoma cells was demonstrated in a chromogenic assay specific for active urokinase type plasminogen activator and in immunoblotting experiments revealing the conversion of single-chain into two-chain urokinase. Control experiments excluded plasmin as the activating agent. These data suggest a novel pathway for plasmin generation during T cell-mediated processes such as immune responses and extravasation of immune cells.  相似文献   

9.
Enzyme kinetic plots relating the initial rate of activation of pro-urokinase to urokinase by plasmin, according to the concentration of substrate, were smooth downward curves and indicated that an apparent decrease in binding affinity occurred with increase in the concentration of pro-urokinase. Such nonlinear plots were obtained with plasmin 1 and also plasmin 2. Over sections of each curve it was possible to estimate apparent kinetic constants. At the uppermost concentrations of substrate tested, these were Km 2.9 microM and kcat 35.5 min-1 for plasmin 1, and at the lowermost concentrations, Km 9.5 nM and kcat 2.0 min-1. Linear plots were obtained when the single proteolytic cleavage was made by K5-plasmin or undegraded plasmin in the presence of 1.0 mM 6-aminohexanoic acid (6-AHa). Constants were estimated for catalysis of this reaction by K5 plasmin to be Km 6.0 microM and kcat 38 min-1 (r = 0.987). The catalytic efficiency of plasmin, at the lowermost concentrations of pro-urokinase tested, was therefore 33-fold higher than that of K5-plasmin. Plotting of data for the cleavage of pro-urokinase by plasmin 1 (in the absence of 6-AHa) according to the model of Hill, gave a slope of 0.5 at the lowermost concentrations of pro-urokinase increasing to 1.0 at higher concentrations (greater than 0.3 microM); such a profile is characteristic of negative cooperativity. The rates of formation of plasmin and urokinase in a mixture containing a low concentration of plasminogen and pro-urokinase were measured and compared to those predicted by a computer program designed to calculate theoretical rates using available kinetic data. The observed rates of generation of both plasmin and urokinase coincided to those predicted from the negative cooperativity model. The mechanism of the negative cooperativity may reside in a conformational change induced by binding of pro-urokinase to the kringle structure of plasmin. This property may be of significance in controlling the fibrinolytic properties of the urokinase-type plasminogen activator system.  相似文献   

10.
The kinetics of plasminogen activation catalysed by urokinase and tissue-type plasminogen activator were investigated. Kinetic measurements are performed by means of a specific chromogenic peptide substrate for plasmin, D-valyl-L-leucyl-L-lysine 4-nitroanilide. Two methods are proposed for the analysis of the resulting progress curve of nitroaniline formation in terms of zymogen-activation kinetics: a graphical transformation of the parabolic curve and transformation of the curve for nitroaniline production into a linear progress curve by the addition of a specific inhibitor of plasmin, bovine pancreatic trypsin inhibitor. The two methods give similar results, suggesting that the reaction between activator and plasminogen is a simple second-order reaction at least at plasminogen concentrations up to about 10 microM. The kinetics of both Glu1-plasminogen (residues 1-790) and Lys77-plasminogen (residues 77-790) activation were investigated. The results confirm previous observations showing that trans-4-(aminomethyl)cyclohexane-1-carboxylic acid at relatively low concentrations enhances the activation rate of Glu1-plasminogen but not that of Lys77-plasminogen. At higher concentrations both Glu1- and Lys77-plasminogen activation are inhibited. The concentration interval for the inhibition of urokinase-catalysed reactions is shown to be very different from that of the tissue-plasminogen activator system. Evidence is presented indicating that binding to the active site of urokinase (KD = 2.0 mM) is responsible for the inhibition of the urokinase system, binding to the active site of tissue-plasminogen activator is approx. 100-fold weaker, and inhibition of the tissue-plasminogen activator system, when monitored by plasmin activity, is mainly due to plasmin inhibition. Poly-D-lysine (Mr 160 000) causes a marked enhancement of plasminogen activation catalysed by tissue-plasminogen activator but not by urokinase. Bell-shaped curves of enhancement as a function of the logarithm of poly-D-lysine concentration are obtained for both Glu1- and Lys77-plasminogen activation, with a maximal effect at about 10 mg/litre. The enhancement of Glu1-plasminogen activation exerted by trans-4-(aminomethyl)cyclohexane-1-carboxylic acid is additive to that of poly-D-lysine, whereas poly-D-lysine-induced enhancement of Lys77-plasminogen activation is abolished by trans-4-(aminomethyl)cyclohexane-1-carboxylic acid. Analogies are drawn up between the effector functions of poly-D-lysine and fibrin on the catalytic activity of tissue-plasminogen activator.  相似文献   

11.
Single-chain urokinase-type plasminogen activator (scu-PA), a potential therapeutic reagent for thrombosis, is activated in plasma by plasmin. The activated enzyme is further digested by plasmin to generate low-molecular-weight urokinase (LMW-UK), which has no affinity for fibrin. To circumvent this dual effect of plasmin, we synthesized in Escherichia coli a variant of scu-PA, which is not converted to LMW-UK on treatment with plasmin. In another variant, the activation cleavage site was modified such that activation by plasmin was slowed down and that inactivation by thrombin was greatly diminished. The combination of these variants may be applicable as an effective thrombolytic reagent for clinical use.  相似文献   

12.
The influence of angiostatin K1-4.5--a fragment of the heavy chain of plasmin and a powerful inhibitor of angiogenesis--on kinetic parameters (k(Pg) and K(Pg)) of human Glu-plasminogen activation under the action of urokinase (uPA) not having affinity for fibrin and fibrin-specific tissue plasminogen activator (tPA) was investigated. Angiostatin does not affect the k(Pg) value, but increases the value K(Pg) urokinase plasminogen activation. A decrease in the k(Pg) value and an increase in the K(Pg) value were found for fibrin-stimulated plasminogen activation by tPA with increasing concentrations of angiostatin. The obtained results show that angiostatin is competitive inhibitor of the uPA activator activity, while it inhibits the activator activity of tPA by mixed type. Such an influence ofangiostatin on the kinetic constants ofthe urokinase plasminogen activation suggests that angiostatin dose dependent manner replaces plasminogen in the binary enzyme-substrate complex uPA-Pg. In case of fibrin-stimulated plasminogen activation by tPA, both zymogen and tPA are bound to fibrin with formation of the effective triple tPA-Pg-fibrin complex. Angiostatin replaces plasminogen both from the fibrin surface and from the enzyme-substrate tPA-Pg complex that leads to a decrease in k(Pg) and an increase in K(Pg) of plasminogen activation. Inhibition constants by angioststin (Ki) of plasminogen-activator activities of uPA and tPA determined by Dixon method were found to be 0.59 +/- 0.04 and 0.12 +/- 0.05 microM, respectively.  相似文献   

13.
The effect of tissue plasminogen activator (TPA) or urokinase on the specific binding of human Glu-plasminogen to fibrin I formed in plasma by clotting with Reptilase was studied using 125I-plasminogen and 131I-fibrinogen. In the absence of TPA, small amounts of plasminogen were bound to fibrin I. TPA induced binding of plasminogen to plasma fibrin I that was dependent upon the concentrations of TPA and plasminogen as well as upon the time of incubation. Plasminogen binding occurred in association with fibrin clot lysis and the formation in the clot supernatant of alpha 2-plasmin inhibitor-plasmin complexes. Urokinase also induced binding of plasminogen to plasma fibrin I that was concentration- and time-dependent. The molecular form of plasminogen bound to the fibrin I plasma clot was identified as Glu-plasminogen by dodecyl sulfate-polyacrylamide gel electrophoresis and by fast performance liquid chromatography. Further studies demonstrated that fibrin I formed from fibrinogen that had been progressively degraded by plasmin-bound Glu-plasminogen. The mole ratio of plasminogen bound increased with the time of plasmin digestion. Glu-plasminogen did not bind to fibrin I formed from fibrinogen progressively digested by human leukocyte elastase, thereby demonstrating the specificity of plasmin. These studies demonstrate that plasminogen activators regulate the binding of Glu-plasminogen to fibrin I by catalyzing plasmin-mediated modifications in the fibrin substrate.  相似文献   

14.
The effects of purified soluble fibrin and of fibrinogen fragments (fibrin mimic) on the activation of Lys-plasminogen (i.e. plasminogen residues 77-790) to plasmin by streptokinase.plasminogen activator complex and by tissue-type plasminogen activator were studied. Dissociation constants of both activators were estimated to lie in the range 90-160 nM (fibrin) and 16-60 nM (CNBr-cleavage fragments of fibrinogen). The kinetic mechanism for both types of activator comprised non-essential enzyme activation via a Rapid Equilibrium Ordered Bireactant sequence. In order to relate the fibrin affinity of plasminogen activators to their fibrinolytic potency, the rate of lysis of supported human plasma clots formed in the presence of unmodified or active-centre-acylated precursors of plasminogen activators was studied as a function of the concentration of enzyme derivative. The concentrations of unmodified enzyme giving 50% lysis/h in this assay were 0.9, 2.0 and 11.0 nM for tissue-type plasminogen activator, streptokinase.plasmin(ogen) and urokinase respectively. However, the potencies of active-centre-acylated derivatives of these enzymes suggested that acylated-tissue plasminogen activator and streptokinase.plasminogen complexes of comparable hydrolytic stability were of comparable potency. Both types of acyl-enzyme were significantly more potent than acyl-urokinases.  相似文献   

15.
Photoaffinity labeling of human plasmin using 4-azidobenzoylglycyl-L-lysine inhibits clot lysis activity, while the activity toward the active-site titrant, p-nitrophenyl-p'-guanidinobenzoate, or alpha-casein are maintained. Photoaffinity labeling of native Glu-plasminogen with the same reagent causes incorporation of approximately 1.5 mol label per mol plasminogen. This labeled plasminogen can be activated to plasmin by either urokinase or streptokinase. The resulting plasmin has full clot lysis activity and can be subsequently photoaffinity labeled with a loss of clot lysis activity. The rate of activation of labeled plasminogen by urokinase is increased relative to that of native plasminogen. epsilon-Aminocaproic acid blocks incorporation of photoaffinity label into both plasminogen and plasmin, indicating that the labeling is specific to the lysine-binding sites. The labels are located in the kringle 1+2+3 fragment in either photoaffinity-labeled plasminogen or plasmin. These results indicate that the specific lysine-binding site blocked in plasmin acts in concert with the active-site in binding and using fibrin as a substrate. This clot lysis regulating site is not available for labeling in plasminogen, but is exposed or changed upon activation to plasmin. The different lysine-binding sites labeled in plasminogen may regulate the conformation of the molecule as evidence by an enhanced rate of activation to plasmin.  相似文献   

16.
Plasminogen activation by single-chain urokinase-type plasminogen activator or pro-urokinase (pro-UK) is accompanied by the generation of two-chain urokinase (UK) by plasmin which provides a positive feedback. In the present study, the time course of the activation of Glu-plasminogen and of Lys-plasminogen (10 microM) by pro-UK (1.0 nM) was studied. In the presence of native plasminogen (Glu-plasminogen), three distinct phases with different rates of plasmin generation were observed. The initial phase was slow and corresponded to the intrinsic activity of pro-UK as reflected by the activity of a plasmin-resistant mutant (Lys158----Ala). This was followed by a second phase which had the most rapid rate. The third phase had a plasminogen activation rate which was significantly slower than the second and paralleled the rate of activation by UK (1.0 nM). The second phase coincided with the time at which there was only about 50% conversion of pro-UK to UK, whereas the final phase coincided with essentially complete conversion. In the presence of fibrin fragment E-2 (20 microM), previously shown to strongly promote plasminogen activation by pro-UK, the identical phenomenon was observed, but at one-tenth the concentration of pro-UK. The most rapid rate of plasmin generation again coincided with transitional (25-60%) pro-UK to UK conversion. To further examine this phenomenon, the rate of pro-UK to UK conversion was controlled by using kallikrein in the presence of a plasmin inhibitor. In this experiment, the activation of Glu-plasminogen bound to solid-phase fibrin was measured. A similar three-phase sequence was observed, the highest rate of plasmin generation coinciding with about 45% conversion of pro-UK to UK. A mechanism for this transitional state phenomenon was postulated based on the established significantly higher affinity of pro-UK than of UK for Glu-plasminogen. This exceptional property for a proenzyme may enable a transient activity to be generated during the transition from pro-UK to UK corresponding to the more favorable KM of pro-UK and the kcat of UK. This hypothesis was supported by the results from experiments in which Lys-plasminogen was substituted for the Glu form. No transitional state activity was observed, consistent with the relatively high KM of pro-UK against Lys-plasminogen.  相似文献   

17.
A procedure is presented for purifying a novel proteinase inhibitor in human plasma whose apparent unique biological property is to inhibit efficiently the lysis of fibrin clots induced by plasminogen activator. The final product is homogeneous as judged by disc gel electrophoresis, and immunoelectrophoresis. Its molecular weight estimated by sodium dodecyl sulfate gel electrophoresis or sedimentation equilibrium is 67,000 and 63,000, respectively. The inhibitor is a glycoprotein consisting polypeptide chain containing 11.7% carbohyrate. It migrates in the alpha2-globulin region in immunoelectrophoresis. The inhibitor is chemically and immunologically different from all the other known inhibitors in plasma. Inhibition of plasmin by the inhibitor is almost instantaneous even at 0 degrees, in contrast to the slow inhibition of urokinase (plasminogen activator in urine). Plasminogen activation by urokinase-induced clot lysis is inhibited by the inhibitor mainly through a mechanism of instantaneous inhibition of plasmin formed and not through the inhibition of urokinase. The inhibitor also inhibits trypsin. Consequently, it is suggested that this newly identified inhibitor is named alpha2-plasmin inhibitor or alpha2-proteinase inhibitor. A specific antibody directed against the inhibitor neutralizes virtually all inhibitory activity of plasma to activator-induced clot lysis. Immunochemical quantitation of the inhibitor was specific antiserum to the inhibitor and the purified inhibitor as a standard indicates that the concentration of the inhibitory in the serum of a healthy man is in or near the range of 5 to 7 mg/100 ml, which is the lowest concentration among the concentration of the proteinase inhibitors in plasma. The inhibitor and plasmin, trypsin, or urokinase form a complex which cannot be dissociated with denaturing and reducing agents. The formation of the enzyme-inhibitor complex occurs on a 1:1 molar basis and is associated with the cleavage of a unique peptide bone, which is most clearly demonstrated in the interaction of the inhibitor and beta-trypsin. In the complex formation between the inhibitor and plasmin, the inhibitor is cross-linked with the light chain which contains the active site of plasmin. It is suggested that, in a fashion analogous to complex formation between alpha1-antitrypsin and trypsin, the cross-links are formed between the active site serine of the enzyme and the newly formed COOH-terminal residue of the inhibitor, with cleavage of a peptide bond.  相似文献   

18.
Mechanisms of plasminogen activation by mammalian plasminogen activators   总被引:4,自引:0,他引:4  
H R Lijnen  D Collen 《Enzyme》1988,40(2-3):90-96
Plasminogen activators convert the proenzyme plasminogen to the active serine protease plasmin by hydrolysis of the Arg560-Val561 peptide bond. Physiological plasminogen activation is however regulated by several additional molecular interactions resulting in fibrin-specific clot lysis. Tissue-type plasminogen activator (t-PA) binds to fibrin and thereby acquires a high affinity for plasminogen, resulting in efficient plasmin generation at the fibrin surface. Single-chain urokinase-type plasminogen activator (scu-PA) activates plasminogen directly but with a catalytic efficiency which is about 20 times lower than that of urokinase. In plasma, however, it is inactive in the absence of fibrin. Chimeric plasminogen activators consisting of the NH2-terminal region of t-PA (containing the fibrin-binding domains) and the COOH-terminal region of scu-PA (containing the active site), combine the mechanisms of fibrin specificity of both plasminogen activators. Combination of t-PA and scu-PA infusion in animal models of thrombosis and in patients with coronary artery thrombosis results in a synergic effect on thrombolysis, allowing a reduction of the therapeutic dose and elimination of side effects on the hemostatic system.  相似文献   

19.
The reciprocal activation of plasminogen and prourokinase (pro-u-PA) is an important mechanism in the initiation and propagation of local fibrinolytic activity. We have found that a bacterial lipopeptide compound, surfactin C (3-20 microM), enhances the activation of pro-u-PA in the presence of plasminogen. This effect accompanied increased conversions of both pro-u-PA and plasminogen to their two-chain forms. Surfactin C also elevated the rate of plasminogen activation by two-chain urokinase (tcu-PA) while not affecting plasmin-catalyzed pro-u-PA activation and amidolytic activities of tcu-PA and plasmin. The intrinsic fluorescence of plasminogen was increased, and molecular elution time of plasminogen in size-exclusion chromatography was shortened in the presence of surfactin C. These results suggested that surfactin C induced a relaxation of plasminogen conformation, thus leading to enhancement of u-PA-catalyzed plasminogen activation, which in turn caused feedback pro-u-PA activation. Surfactin C was active in enhancing [125I]fibrin degradation both by pro-u-PA/plasminogen and tcu-PA/plasminogen systems. In a rat pulmonary embolism model, surfactin C (1 mg/kg, i.v.) elevated 125I plasma clot lysis when injected in combination with pro-u-PA. The present results provide first evidence that pharmacological relaxation of plasminogen conformation leads to enhanced fibrinolysis in vivo.  相似文献   

20.
Prourokinase-induced plasminogen activation is complex and involves three distinct reactions: (1) plasminogen activation by the intrinsic activity of prourokinase; (2) prourokinase activation by plasmin; (3) plasminogen activation by urokinase. To further understand some of the mechanisms involved, the effects of epsilon-aminocaproic acid (EACA), a lysine analogue, on these reactions were studied. At a low range of concentrations (10-50 microM), EACA significantly inhibited prourokinase-induced (Glu-/Lys-) plasminogen activation, prourokinase activation by Lys-plasmin, and (Glu-/Lys-) plasminogen activation by urokinase. However, no inhibition of plasminogen activation by Ala158-prourokinase (a plasmin-resistant mutant) occurred. Therefore, the overall inhibition of EACA on prourokinase-induced plasminogen activation was mainly due to inhibition of reactions 2 and 3, by blocking the high-affinity lysine binding interaction between plasmin and prourokinase, as well as between plasminogen and urokinase. These findings were consistent with kinetic studies which suggested that binding of kringle 1-4 of plasmin to the N-terminal region of prourokinase significantly promotes prourokinase activation, and that binding of kringle 1-4 of plasminogen to the C-terminal lysine158 of urokinase significantly promotes plasminogen activation. In conclusion, EACA was found to inhibit, rather than promote, prourokinase-induced plasminogen activation due to its blocking of the high-affinity lysine binding sites on plasmin(ogen).  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