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 31 毫秒
1.
125I-labeled polymeric fibrin hydrolyzed with plasmin, Val442-plasmin (miniplasmin, Lys530-plasmin (microplasmin) and trypsin has been studied for radioactivity of its separate electrophoretic bands. The reaction of hydrolysis was stopped at a moment of a two-fold decrease of the fibrin clot turbidity (t1/2) at the wave length 350 nm. For plasmin, miniplasmin, microplasmin and trypsin taken in the same caseinolytic activities t1/2 was 12.4, 40.0 164.1 and 76.8 min, respectively. Differences in composition of fibrin digests taken at t1/2, are demonstrated: the content of high-molecular components of digests decreases in the order of plasmin greater than miniplasmin greater than microplasmin greater than trypsin, thus showing differences in the processes of fibrin clot structure disruption by the enzymes.  相似文献   

2.
Fibrinolysis and fibrinogenolysis by Val442-plasmin   总被引:2,自引:0,他引:2  
Elastase cleavage of Lys77-plasmin results in the formation of Val442-plasmin. This result suggests that small, active plasmin fragments can be produced even under conditions of high plasminogen activator levels such as occur in vivo. We examined the effect of the generation of such fragments by studying the degradation of fibrinogen and fibrin by Val442-plasmin. Val442-plasmin lysis of fibrinogen yielded the same products as obtained with Lys77-plasmin, but at a slightly lower rate. Lysine inhibited fibrinogenolysis by both Lys77-plasmin and Val442-plasmin. The marked inhibition observed at concentrations higher than 10 mM lysine occurred to the same extent for both proteases. In addition, the products and rate of fibrinolysis were the same for both proteases. These results indicate that the lysine binding regions present in Lys77-plasmin but absent in Val442-plasmin do not determine the rate, reaction products, or lysine inhibition of fibrinolysis and fibrinogenolysis by plasmin.  相似文献   

3.
Antiplasminogen monoclonal antibody IV-1c (IV-1c) binds to Val 709-Gly 718 site of plasminogen (Pg) protease domain, which is far removed from the active site. Pg-IV-1c complex formation induces catalytic activity in proenzymes active site. Influence of IV-1c binding to plasmin (Pm) on Pm catalytic properties has not been investigated yet. Data on catalytic properties of Pm in equimolar Pm-IV-1c complex are presented. It was found that Pm and mini-Pm amidolytic and caseinolytic activity was twice as high as in Pm-IV-1c and mini-Pm-IV-1c complexes. 20 mM 6-AHA and 100 mM arginine did not influence this rise. The increase of amidolytic activity is connected with reduction of K(m) of S 2251 hydrolysis reaction for Pm and mini-Pm from 0.125 and 0.43 to 0.05 and 0.23 mM, correspondingly. Kcat remains almost the same. Fibrinolytic and fibrinogenolytic activity of Pm in Pm-IV-1c complex decreased to 20% of initial value alpha 2-Antiplasmin inhibited Pm activity in complex Pm-IV-1c by 80%. Pm-IV-1c complex did not activate free Pg, but activated equimolar Pg-IV-1c complex. Affinity of IV-1c to Pm and Pg was the same as C50 approximately 1.5 nM. Binding of Pm with IV-1c in a complex: a) leads to increase of Pm active site affinity to LMW substrates; b) causes steric hindrances for fibrin/fibrinogen access to Pm active site; c) proceeds with the same affinity for Pm and Pg, that indicates to invariable Val 709-Gly 718 site conformation after Pg transition in Pm.  相似文献   

4.
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   

5.
Thermodynamic and kinetic parameters for the binding of the bovine basic pancreatic trypsin inhibitor (BPTI, Kunitz inhibitor) to human Glu1-, Lys77-, Val442- and Val561-plasmin (EC 3.4.21.7) have been determined between pH 3.0 and 9.5, and from 5.0 to 45.0 degrees C. The inhibitor-binding properties to human Glu1-, Lys77-, Val442- and Val561-plasmin suggest a possible role of BPTI in modulating plasmin activity when the inhibitor is used therapeutically.  相似文献   

6.
A numerical model based on the convective-diffusive transport of reacting and adsorbing proteolytic enzymes within erodible fibrous biopolymers was used to predict lysis fronts moving across biogels such as fibrin or collagen. The fiber structure and the transport properties of solutes in fibrin (or collagen) were related to the local extent of dissolution within the dissolving structure. An accounting for solubilization of adsorbed species into solution from the eroding fiber phase provided for complete conservation of mass in reacting systems containing over 10 species. At conditions of fibrinolysis typical of clinical situations, the model accurately predicted the dynamic rate of lysis front movement for plasmin, urokinase, and tissue plasminogen activator (tPA)-mediated lysis of fibrin gels measured in vitro. However, under conditions of extremely fast fibrinolysis using high enzyme concentrations, fibrinolytic fronts moved very rapidly (>0.1 mm/mm)-faster than predicted for diffusionlimited reactions-at nearly constant velocity for over 2 h, indicating non-Fickian behavior. This was due to proteolysis-mediated retraction of dissolving fibrin fibers that resulted in fiber convection and front-sharpening within 3 mum of the reaction front, as observed by digitally enhanced microscopy. In comparing the model to fibrinolysis measurements using human lys(77)-plasmin, the average first order rate constant for non-crosslinked fibrin bond cleavage by fibrin-bound plasmin was calculated to be 5s(-1) assuming that 10 cleavages per fibrin monomer were required to solubilize each monomer. The model accurately predicted lysis front movement using pressure-driven permeation of plasmin or urokinase into fibrin as well as literature data obtained under well- mixed conditions for tPA-mediated fibrinolysis. This numerical formulation provides predictive capability for optimization of proteolytic systems which include thrombolytic therapy, wound healing, controlled drug release, and tissue engineering applications. (c) 1995 John Wiley & Sons, Inc.  相似文献   

7.
The first step in the activation of blood coagulation factor XIII (FXIII) is the proteolytic cleavage of the potentially active A subunit (FXIII-A) by thrombin at Arg37-Gly38. Both fibrin formation and FXIII-A Val34Leu polymorphism influence the rate of proteolytic activation of purified factor XIII, however their relative importance and interaction in determining the time of onset and the rate of FXIII activation in whole plasma have not yet been explored. In the present study it was shown that in plasma, fibrin formation preceded the truncation of FXIII-A by thrombin, the activation process took place exclusively on the surface of newly formed fibrin and activated FXIII remained associated with the fibrin clot. The time of fibrin formation closely correlated with the time of FXIII activation, while there was no significant relationship between the time of FXIII activation and FXIII-A Val34Leu genotype. However, in the case of Leu34 variant the lag phase between fibrin formation and FXIII-A truncation was significantly shorter than in the case of Val34 variant. The results suggest that in whole plasma the onset of FXIII activation is determined by fibrin formation, while the rate of activation is modulated by Val34Leu polymorphism.  相似文献   

8.
Plasminogen activation catalysed by tissue-type plasminogen activator (t-PA) has been examined in the course of concomitant fibrin formation and degradation. Plasmin generation has been measured by the spectrophotometric method of Petersen et al. (Biochem. J. 225 (1985) 149-158), modified so as to allow for light scattering caused by polymerized fibrin. Glu1-, Lys77- and Val442-plasminogen are activated in the presence of fibrinogen, des A- and des AB-fibrin and the rate of plasmin formation is found to be greatly enhanced by both des A- and des AB-fibrin polymer. Plasmin formation from Glu1- and Lys77-plasminogen yields a sigmoidal curve, whereas a linear increase is obtained with Val442-plasminogen. The rate of plasmin formation from Glu1- and Lys77-plasminogen declines in parallel with decreasing turbidity of the fibrin polymer effector. In order to study the effect of polymerization, this has been inhibited by the synthetic polymerization site analogue Gly-Pro-Arg-Pro, by fibrinogen fragment D1 or by prior methylene blue-dependent photooxidation of the fibrinogen used. Inhibition of polymerization by Gly-Pro-Arg-Pro reduces plasmin generation to the low rate observed in the presence of fibrinogen. Antipolymerization with fragment D1 or photooxidation has the same effect on Glu1-plasminogen activation, but only partially reduces and delays the stimulatory effect on Lys77- and Val442-plasminogen activation. The results suggest that protofibril formation (and probably also gelation) of fibrin following fibrinopeptide release is essential to its stimulatory effect. The gradual increase and subsequent decline in the rate of plasmin formation from Glu1- or Lys77-plasminogen during fibrinolysis may be explained by sequential exposure, modification and destruction of different t-PA and plasminogen binding sites in fibrin polymer.  相似文献   

9.
Binding of streptokinase (SK) to plasminogen (Pg) induces conformational activation of the zymogen and initiates its proteolytic conversion to plasmin (Pm). The mechanism of coupling between conformational activation and Pm formation was investigated in kinetic studies. Parabolic time courses of Pg activation by SK monitored by chromogenic substrate hydrolysis had initial rates (v(1)) representing conformational activation and subsequent rates of activity increase (v(2)) corresponding to the rate of Pm generation determined by a specific discontinuous assay. The v(2) dependence on SK concentration for [Lys]Pg showed a maximum rate at a Pg to SK ratio of approximately 2:1, with inhibition at high SK concentrations. [Glu]Pg and [Lys]Pg activation showed similar kinetic behavior but much slower activation of [Glu]Pg, due to an approximately 12-fold lower affinity for SK and an approximately 20-fold lower k(cat)/K(m). Blocking lysine-binding sites on Pg inhibited SK.Pg* cleavage of [Lys]Pg to a rate comparable with that of [Glu]Pg, whereas [Glu]Pg activation was not significantly affected. The results support a kinetic mechanism in which SK activates Pg conformationally by rapid equilibrium formation of the SK.Pg* complex, followed by intermolecular cleavage of Pg to Pm by SK.Pg* and subsequent cleavage of Pg by SK.Pm. A unified model of SK-induced Pg activation suggests that generation of initial Pm by SK.Pg* acts as a self-limiting triggering mechanism to initiate production of one SK equivalent of SK.Pm, which then converts the remaining free Pg to Pm.  相似文献   

10.
The effect of heparin on the proteolytic and fibrinolytic activities of plasmin and plasminogen was studied. Heparin at a concentration of 6.3.10(-6) M did not change the caseinolytic activity of plasmin and plasminogen stimulated by streptokinase but suppressed their fibrinolytic activity. At concentrations from 2.10(-8) to 0.5.10(-6) M heparin increased, whereas at 1.10(-6)-4.10(-6) M reduced the time of desAAfibrin clot half-lysis by plasmin. Within the concentration range of 2.10(-8) to 4.10(-6) M heparin did not change the time of the clot half-lysis by glu-plasminogen and slightly decreased the time of fibrin clot half-lysis by lys-plasminogen in the presence of the tissue activator. It was supposed that heparin inhibits the fibrinolytic effect of plasmin by way of formation of complexes with plasmin and reduction of plasmin specificity to the solid phase substrate, i. e., polymeric fibrin.  相似文献   

11.
Wang S  Reed GL  Hedstrom L 《Biochemistry》1999,38(16):5232-5240
Plasminogen (Plgn) is usually activated by proteolytic cleavage of Arg561-Val562. The new N-terminal amino group of Val562 forms a salt bridge with Asp740, creating the active protease plasmin (Pm). However, streptokinase (SK) binds to Plgn, generating an active protease in a poorly understood, nonproteolytic process. We hypothesized that the N-terminus of SK, Ile1, substitutes for the N-terminal Val562 of Pm, forming an analogous salt bridge with Asp740. SK initially forms an inactive complex with Plgn, which subsequently rearranges to create an active complex; this rearrangement is rate limiting at 4 degrees C. SK.Plgn efficiently hydrolyzes amide substrates at 4 degrees C, although DeltaIle1-SK. Plgn has no amidolytic activity. DeltaIle1-SK prevents formation of wild-type SK.Plgn. These results indicate that DeltaIle1-SK forms the initial inactive complex with plasminogen, but cannot form the active complex. However, when the experiment is performed at 37 degrees C, amidolytic activity is observed when DeltaIle1-SK is added to plasminogen. SDS-PAGE analysis demonstrates that the amidolytic activity results from the formation of DeltaIle1-SK.Pm. To further demonstrate that the activity of DeltaIle1-SK requires the conversion of Plgn to Pm, we characterized the reaction of SK with a mutant microplasminogen, Arg561Ala-microPlgn, that cannot be converted to microplasmin. Amidolytic activity is observed when Arg561Ala-microPlgn is incubated with wild-type SK at 37 degrees C; however, no amidolytic activity is observed in the presence of DeltaIle1-SK. These observations demonstrate that the amidolytic activity of DeltaIle1-SK at 37 degrees C requires the conversion of Plgn to Pm. Our findings indicate that Ile1 of SK is required for the nonproteolytic activation of Plgn by SK and are consistent with the hypothesis that Ile1 of SK substitutes for Val562 of Pm.  相似文献   

12.
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   

13.
Active-site-inhibited plasmin was prepared by inhibition with d-valyl-l-phenylalanyl-l-lysylchloromethane or by bovine pancreatic trypsin inhibitor (Kunitz inhibitor). Active-site-inhibited Glu-plasmin binds far more strongly to fibrin than Glu-plasminogen [native human plasminogen with N-terminal glutamic acid (residues 1–790)]. This binding is decreased by α2-plasmin inhibitor and tranexamic acid, and is, in the latter case, related to saturation of a strong lysine-binding site. In contrast, α2-plasmin inhibitor and tranexamic acid have only weak effects on the binding of Glu-plasminogen to fibrin. This demonstrates that its strong lysine-binding site is of minor importance to its binding to fibrin. Active-site-inhibited Lys-plasmin and Lys-plasminogen (Glu-plasminogen lacking the N-terminal residues Glu1–Lys76, Glu1–Arg67 or Glu1–Lys77)display binding to fibrin similar to that of active-site inhibited Glu-plasmin. In addition, α2-plasmin inhibitor or tranexamic acid similarly decrease their binding to fibrin. Glu-plasminogen and active-site-inhibited Glu-plasmin have the same gross conformation, and conversion into their respective Lys- forms produces a similar marked change in conformation [Violand, Sodetz & Castellino (1975) Arch. Biochem. Biophys. 170, 300–305]. Our results indicate that this change is not essential to the degree of binding to fibrin or to the effect of α2-plasmin inhibitor and tranexamic acid on this binding. The conversion of miniplasminogen (Glu-plasminogen lacking the N-terminal residues Glu1–Val441) into active-site-inhibited miniplasmin makes no difference to the degree of binding to fibrin, which is similarly decreased by the addition of tranexamic acid and unaffected by α2-plasmin inhibitor. Active-site-inhibited Glu-plasmin, Lys-plasmin and miniplasmin have lower fibrin-binding values in a plasma system than in a purified system. Results with miniplasmin(ogen) indicate that plasma proteins other than α2-plasmin inhibitor and histidine-rich glycoprotein decrease the binding of plasmin(ogen) to fibrin.  相似文献   

14.
The predominant high molecular weight products of plasmic digestion of human crosslinked fibrin Fragments DD, E and (DD)E complex were purified by column gel filtration in a non-dissociating buffer or by ion-exchange chromatography on DEAE-cellulose. The structure of the degradation products was studied by proteolytic degradation, polyacrylamide gel electrophoresis immunodiffusion and sucrose density gradient centrifugation. Unaltered derivatives were very resistant to proteolytic degradation by plasmin. In the the presence of 10 mM EDTA the (DD)E complex did not dissociate, but similar to Fragment DD, became susceptible to plasmic degradation forming Fragment D derivatives. The (DD)E complex dissociated in 3 M urea at pH 5.5, had an altered conformation as evidenced by its aggregability and by its increased susceptibility to degradation by plasmin resulting in the formation of Fragment d. The gammagamma chain remnants of Fragment DD were attacked first, followed by cleavage of the beta chain remnants. It is concluded that plasmin resistance is a function of the intact structure and it is not directly dependent on the presence of the crosslink bonds or calcium ions.  相似文献   

15.
Interaction of human plasmin with human alpha 2-macroglobulin   总被引:2,自引:0,他引:2  
The steady-state kinetic parameters of plasmin and the alpha 2-macroglobulin (alpha 2M)-plasmin complex toward the chromogenic substrate Val-Leu-Lys-p-nitroanilide (S-2251), in the presence and absence of plasmin competitive inhibitors, have been determined. At pH 7.4 and 22 degrees C, the Km values for plasmin and alpha 2M-plasmin for S-2251 were 0.13 +/- 0.02 mM and 0.3 +/- 0.03 mM. The kcat of this reaction, when catalyzed by alpha 2M-plasmin, was 6.0 +/- 0.5 s-1, a value significantly decreased from the kcat of 11.0 +/- 1.0 s-1, determined when free plasmin was the enzyme. KI values for benzamidine of 0.50 +/- 0.05 mM and 0.23 +/- 0.02 mM were obtained for S-2251 hydrolysis, as catalyzed by alpha 2M-plasmin and plasmin, respectively. When leupeptin was the competitive inhibitor, KI values of 5.0 +/- 0.65 microM and 1.0 +/- 0.1 microM were obtained when alpha 2M-plasmin and plasmin, respectively, were the enzymes employed for catalysis of S-2251 hydrolysis. The comparative rates of reaction of the peptide inhibitor Trasylol (Kunitz basic pancreatic inhibitor) with plasmin and alpha 2M-plasmin were also determined. A concentration of Trasylol of at least 3 orders of magnitude greater for alpha 2M-plasmin than for free plasmin was required to observe inhibition rates on comparable time scales.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Ternary complex formation of tissue plasminogen activator (TPA) and plasminogen (Plg) with thrombospondin (TSP) or histidine-rich glycoprotein (HRGP) has been demonstrated using an enzyme-linked immunosorbent assay, an affinity bead assay, and a rocket immunoelectrophoresis assay. The formation of these complexes was specific, concentration dependent, saturable, lysine binding site-dependent, and inhibitable by fluid phase plasminogen. Apparent Kd values were approximately 12-36 nM for the interaction of TPA with TSP-Plg complexes and 15-31 nM with HRGP-Plg complexes. At saturation the relative molar stoichiometry of Plg:TPA was 3:1 within the TSP-containing complexes and 1:1 within HRGP-containing complexes. The activation of Plg to plasmin by TPA on TSP- and HRGP-coated surfaces was studied using a synthetic fluorometric plasmin substrate (D-Val-Leu-Lys-7-amino-4-trifluoromethyl coumarin). Kinetic analysis demonstrated a marked increase in the affinity of TPA for plasminogen in the presence of surface-associated TSP or HRGP. Compared to fluid phase activation or activation on fibronectin- or Factor VIII-related antigen-coated surfaces there was a 35-fold increase in efficiency of plasmin generation. A substantial amount (up to 71%) of the plasmin formed remained surface-associated and was found to be protected from inhibition by alpha 2-plasmin inhibitor. Greater than 200-fold increase in inhibitor concentration was required to effect 50% inhibition. Complex formation of locally released tissue plasminogen activator with Plg immobilized on TSP or HRGP surfaces may thus play an important role in effecting proteolytic events in nonfibrin-containing microenvironments.  相似文献   

17.
The alpha(2)-plasmin inhibitor (A2PI) is a main physiological regulator of the trypsin-like serine proteinase plasmin. It is composed of an N-terminal 15 amino acid fibrin cross-linking polypeptide, a 382-residue serpin domain, and a flexible C-terminal segment. The latter, peptide Asn(398)-Lys(452), and its Lys452Ala mutant were expressed as recombinant proteins in Escherichia coli (r-A2PIC and r-A2PICmut, respectively). CD and NMR analyses indicate that r-A2PIC is flexible, loosely folded, and with low content of regular secondary structure. Functional characterization via intrinsic fluorescence ligand titrations shows that r-A2PIC interacts with the isolated plasminogen kringle 1 (r-K1) (K(a) approximately 69.9 mM(-)(1)), K4 (K(a) approximately 45.7 mM(-)(1)), K5 (K(a) approximately 4.3 mM(-)(1)), and r-K2 (K(a) approximately 3.2 mM(-)(1)), all of which are known to exhibit lysine-binding capability. The affinities of these kringles for r-A2PIC are consistently larger than those reported for the ligand N(alpha)-acetyllysine, a mimic of a C-terminal Lys residue. The r-A2PICmut, with a C-terminal Ala residue, also interacts with r-K1 and K4, although with approximately 5-fold lesser affinities relative to r-A2PIC, demonstrating that while Lys(452) plays a major role in the binding, internal residues in r-A2PIC tether the kringles. (1)H NMR spectroscopy shows that key aromatic residues within the K4 lysine-binding site (LBS), namely, Trp(25), Trp(62), Phe(64), Trp(72), and Tyr(74), selectively respond to the addition of r-A2PIC and r-A2PICmut, indicating that these interactions proceed via the kringles' canonical LBS. We conclude that r-A2PIC docks to kringles primarily through lysine side chains and that Lys(452) most definitely enhances the binding. This suggests that multiple Lys residues within A2PI could contribute, perhaps in a zipper-like fashion, to its binding to the in-tandem, multikringle array that configures the plasmin heavy chain.  相似文献   

18.
The progressive inhibition of plasmin by pancreatic trypsin inhibitor and by alpha 2-plasmin inhibitor in the presence of D-valyl-L-leucyl-L-lysine 4-nitroanilide was investigated. The kinetics with plasmin were compared with those with miniplasmin. The kinetic properties of two functionally different forms of alpha 2-plasmin inhibitor described by Clemmensen [(1979) in The Physiological Inhibitors of Coagulation and Fibrinolysis (Collen. D., Wiman, B & Verstraete, M., eds.), pp 131-136, Elsevier, Amsterdam] were characterized. The two forms differ in their plasminogen-binding capability, and this difference can account for a difference in secondary site interaction suggested from the kinetics. The binding of inhibitor to miniplasmin is a simple pseudo-first-order reaction with both pancreatic trypsin inhibitor and the two alpha 2-plasmin inhibitor forms. Such simple kinetics are also observed for the reaction between plasmin and the non-plasminogen-binding form of alpha 2-plasmin inhibitor. More complicated kinetics are obtained for the reaction between plasmin and the alpha 2-plasmin inhibitor form that binds to plasminogen. With both forms of the alpha 2-plasmin inhibitor, a complex stable to acetic acid/urea and gel electrophoresis is present and fully developed 15 s after initiation of the reaction with plasmin.  相似文献   

19.
We previously showed that coagulation factor Xa (FXa) enhances activation of the fibrinolysis zymogen plasminogen to plasmin by tissue plasminogen activator (tPA). Implying that proteolytic modulation occurs in situ, intact FXa (FXaα) must be sequentially cleaved by plasmin or autoproteolysis, producing FXaβ and Xa33/13, which acquire necessary plasminogen binding sites. The implicit function of Xa33/13 in plasmin generation has not been demonstrated, nor has FXaα/β or Xa33/13 been studied in clot lysis experiments. We now report that purified Xa33/13 increases tPA-dependent plasmin generation by at least 10-fold. Western blots confirmed that in situ conversion of FXaα/β to Xa33/13 correlated to enhanced plasmin generation. Chemical modification of the FXaα active site resulted in the proteolytic generation of a product distinct from Xa33/13 and inhibited the enhancement of plasminogen activation. Identical modification of Xa33/13 had no effect on tPA cofactor function. Due to its overwhelming concentration in the clot, fibrin is the accepted tPA cofactor. Nevertheless, at the functional level of tPA that circulates in plasma, FXaα/β or Xa33/13 greatly reduced purified fibrin lysis times by as much as 7-fold. This effect was attenuated at high levels of tPA, suggesting a role when intrinsic plasmin generation is relatively low. FXaα/β or Xa33/13 did not alter the apparent size of fibrin degradation products, but accelerated the initial cleavage of fibrin to fragment X, which is known to optimize the tPA cofactor activity of fibrin. Thus, coagulation FXaα undergoes proteolytic modulation to enhance fibrinolysis, possibly by priming the tPA cofactor function of fibrin.  相似文献   

20.
It was found that fragments E (Mr = 45 000), DH (Mr = 95 000) and DL (Mr = 82 000) decrease the rate of plasmin hydrolysis of fibrin that is not cross-linked with factor XIII; the most effective inhibitor is fragment DL. The Kd values for the interactions of fragments E, DH and DL with plasmin are equal to 0.15, 0.4 and 0.04 microM, respectively.  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