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.
The mechanism whereby fragments of streptokinase (SK) derived from its N terminus (e.g., SK1-59 or SK1-63) enhance the low plasminogen (PG)-activating ability of other fragments, namely SK64-386, SK60-414, SK60-387, and SK60-333 (reported previously), has been investigated using a synthetic peptide approach. The addition of either natural SK1-59, or chemically synthesized SK16-59, at saturation (about 500-fold molar excess) generated amidolytic and PG activation capabilities in equimolar mixtures of human plasminogen (HPG) and its complementary fragment (either SK60-414 or SK56-414, prepared by expression of truncated SK gene fragments in Escherichia coli) that were approximately 1.2- and 2.5-fold, respectively, of that generated by equimolar mixtures of native SK and HPG. Although in the absence of SK1-59 equimolar mixtures of SK56-414 and HPG could generate almost 80% of amidolytic activity, albeit slowly, less than 2% level of PG activation could be observed under the same conditions, indicating that the contribution of the N-terminal region lay mainly in imparting in SK56-414 an enhanced ability for PG activation. The ability of various synthetic peptides derived from the amino-terminal region (SK16-51, SK16-45, SK37-59, SK1-36, SK16-36, and SK37-51) to (1) complement equimolar mixtures of SK56-414 and HPG for the generation of amidolytic and PG activation functions, (2) inhibit the potentiation of SK56-414 and HPG by SK16-59, and (3) directly inhibit PG activation by the 1:1 SK-HPG activator complex was tested. Apart from SK16-59, SK16-51, and 16-45, the ability to rapidly generate amidolytic potential in HPG in the presence of SK56-414 survived even in the smaller SK-peptides, viz., SK37-59 and SK37-51. However, this ability was abolished upon specifically mutating the sequence -LTSRP-, present at position 42-46 in native SK. Although SK16-51 retained virtually complete ability for potentiation of PG activation in comparison to SK16-59 or SK1-59, this ability was reduced by approximately fourfold in the case of SK16-45, and completely abolished upon further truncation of the C-terminal residues to SK16-36 or SK1-36. Remarkably, however, these peptides not only displayed ability to bind PG, but also showed strong inhibition of PG activation by the native activator complex in the micromolar range of concentration; the observed inhibition, however, could be competitively relieved by increasing the concentration of substrate PG in the reaction, suggesting that this region in SK contains a site directed specifically toward interaction with substrate PG. This conclusion was substantiated by the observation that the potentiation of PG activating ability was found to be considerably reduced in a peptide (SK25-59) in which the sequence corresponding to this putative locus (residues 16-36) was truncated at the middle. On the other hand, fragments SK37-51 and SK37-59 did not show any inhibition of the PG activation by native activator complex. Taken together, these findings strongly support a model of SK action wherein the HPG binding site resident in the region 37-51 helps in anchoring the N-terminal domain to the strong intermolecular complex formed between HPG and the region 60-414. In contrast, the site located between residues 16 and 36 is qualitatively more similar to the previously reported PG interacting site (SK254-273) present in the core region of SK, in being involved in the relatively low-affinity enzyme-substrate interactions of the activator complex with PG during the catalytic cycle.  相似文献   

2.
The role of Asn-linked oligosaccharide in the functional properties of both human tissue-type plasminogen activator (t-PA) and a genetic variant of t-PA was studied. Nonglycosylated and glycosylated wild-type t-PA were produced in mammalian cells which express recombinant t-PA. These proteins were compared in fibrin binding and 125I-labeled fibrin clot lysis assays, using purified components. The nonglycosylated form showed higher fibrin binding, as well as higher fibrinolytic potency than the glycosylated form. Subsequently, prevention of glycosylation of a t-PA variant which lacked the finger and epidermal growth factor domains (delta FE), was carried out in an attempt to enhance its fibrinolytic activity. Glycosylation was prevented by changing Asn to Gln; at Asn-117 to produce delta FE1X t-PA, and at Asn-117, -184, and -448 to produce delta FE3X t-PA. All variants were similar to wild-type t-PA in their catalytic dependence on fibrinogen fragments, fibrinolytic activity in fibrin autography analysis, and plasminogen activator activity. In a clot lysis assay, using citrated human plasma, the fibrinolytic potency of the variants were comparable to that of wild-type t-PA at activator concentrations of 17-51 nM (approximately 1-3 micrograms/ml). At 0.5-5.1 nM (approximately 0.03-0.3 micrograms/ml), however, the variant proteins had lower fibrinolytic potency than wild-type t-PA. Fifty percent lysis in 1.5 h for wild-type, delta FE, delta FE1X, and delta FE3X t-PA, required 2.5, 10, 7.5, and 5.5 nM t-PA, respectively. The fibrinogenolytic activity in human plasma was measured for wild-type, delta FE, delta FE1X, and delta FE3X t-PA, and showed significant fibrinogen depletion after 3 h of incubation at 51 nM, decreasing to 11, 11, 50, and 72% of basal levels, respectively. These data indicate that partial or total nonglycosylated t-PA variants have a higher fibrinolytic versus fibrinogenolytic ratio than their fully glycosylated counterparts.  相似文献   

3.
Streptokinase (SK) binds to plasminogen (Pg) to form a complex that converts substrate Pg to plasmin. Residues 1-59 of SK regulate its capacity to induce an active site in bound Pg by a nonproteolytic mechanism and to activate substrate Pg in a fibrin-independent manner. We analyzed 24 SK mutants to better define the functional properties of SK-(1-59). Mutations within the alphabeta1 strand (residues 17-26) of SK completely prevented nonproteolytic active site induction in bound Pg and rendered SK incapable of protecting plasmin from inhibition by alpha2-antiplasmin. However, when fibrin-bound, the activities of alphabeta1 strand mutants were similar to that of wild-type (WT) SK and resistant to alpha2-antiplasmin. Mutation of Ile1 of SK also prevented nonproteolytic active site induction in bound Pg. However, unlike alphabeta1 strand mutants, the functional defect of Ile1 mutants was not relieved by fibrin, and complexes of Ile1 mutants and plasmin were resistant to alpha2-antiplasmin. Plasmin enhanced the activities of alphabeta1 strand and Ile1 mutants, suggesting that SK-plasmin complexes activated mutant SK.Pg complexes by hydrolyzing the Pg Arg561-Val562 bond. Mutational analysis of Glu39 of SK suggested that a salt bridge between Glu39 and Arg719 of Pg is important, but not essential, for nonproteolytic active site induction in Pg. Deleting residues 1-59 rendered SK dependent on plasmin and fibrin to generate plasminogen activator (PA) activity. However, the PA activity of SK-(60-414) in the presence of fibrin was markedly reduced compared with WT SK. Despite its reduced PA activity, the fibrinolytic potency of SK-(60-414) was greater than that of WT SK at higher (but not lower) SK concentrations due to its capacity to deplete plasma Pg. These studies define mechanisms by which the SK alpha domain regulates rapid active site induction in bound Pg, contributes to the resistance of the SK-plasmin complex to alpha2-antiplasmin, and controls fibrin-independent Pg activation.  相似文献   

4.
Recombinant plasmids were constructed by fusing the gene fragments encoding the full-length (1-191aa) and the truncated (1-40aa and l-69aa) HCV core proteins (HCc) respectively to the core gene of HBV at the position of amino acid 144 and expressed inE. coli. The products were analyzed by ELISA, Western blotting as well as the immunization of the mice. The results showed that those fusion proteins (B144C191, B144C69, B144C40) possessed the dual antigenicity and immunogenicity of both hepatitis B virus core antigen (HBcAg) and hepatitis C virus core protein (HCc). Analysis by electron microscopy and CsCI density gradient ultra-centrifugation revealed that similar to the HBcAg itself, all fusion proteins were able to form particles. Comparison of the antigenicity and immunogenicity of those fusion proteins showed that the length of HCc gene fused to HBcAg had no much effect on the antigenicity and immunogenicity of HBcAg, however, B144C69 and B144C40 induced higher titres antibodies against HCc than B144C191. Using those fusion proteins, ELISA for screening of antibodies against both HBV and HCV in human sera was also established.  相似文献   

5.
Recombinant plasmids were constructed by fusing the gene fragments encoding the full-length (1-191aa) and the truncated (1-40aa and 1-69aa) HCV core proteins (HCc) respectively to the core gene of HBV at the position of amino acid 144 and expressed in E. coli. The products were analyzed by ELISA, Western blotting as well as the immunization of the mice. The results showed that those fusion proteins (B144C191, B144C69, B144C40) possessed the dual antigenicity and immunogenicity of both hepatitis B virus core antigen (HBcAg) and hepatitis C virus core protein (HCc). Analysis by electron microscopy and CsCl density gradient ultra-centrifugation revealed that similar to the HBcAg itself, all fusion proteins were able to form particles. Comparison of the antigenicity and immunogenicity of those fusion proteins showed that the length of HCc gene fused to HBeAg had no much effect on the antigenicity and immunogenicity of HBcAg, however, B144C69 and B144C40 induced higher titres antibodies against HCc than B14d  相似文献   

6.
Plasminogen activators (PAs) are proteases that convert plasminogen to plasmin. Plasmin, in turn, is a protease that can lyse a fibrin clot and, therefore, PAs have a primary role in fibrinolysis. Two PAs, urokinase (UK) and streptokinase (SK), have been available for therapeutic use for years. Unfortunately, both can cause systemic fibrinogenolysis and other side effects which have limited their use. Interest has focused on a different enzyme, tissue plasminogen activator (t-PA), which will cause specific clot lysis without systemic problems. The gene for t-PA has been cloned and many biotechnology firms are preparing to produce t-PA for therapeutic use. The properties and potential for therapy of t-PA are reviewed and compared to new forms of other activators, such as pro-urokinase. How the interactions of PAs and inhibitors may affect the use of PAs is also discussed.  相似文献   

7.
Neutrophil extracellular traps are networks of DNA and associated proteins produced by nucleosome release from activated neutrophils in response to infection stimuli and have recently been identified as key mediators between innate immunity, inflammation, and hemostasis. The interaction of DNA and histones with a number of hemostatic factors has been shown to promote clotting and is associated with increased thrombosis, but little is known about the effects of DNA and histones on the regulation of fibrin stability and fibrinolysis. Here we demonstrate that the addition of histone-DNA complexes to fibrin results in thicker fibers (increase in median diameter from 84 to 123 nm according to scanning electron microscopy data) accompanied by improved stability and rigidity (the critical shear stress causing loss of fibrin viscosity increases from 150 to 376 Pa whereas the storage modulus of the gel increases from 62 to 82 pascals according to oscillation rheometric data). The effects of DNA and histones alone are subtle and suggest that histones affect clot structure whereas DNA changes the way clots are lysed. The combination of histones + DNA significantly prolongs clot lysis. Isothermal titration and confocal microscopy studies suggest that histones and DNA bind large fibrin degradation products with 191 and 136 nm dissociation constants, respectively, interactions that inhibit clot lysis. Heparin, which is known to interfere with the formation of neutrophil extracellular traps, appears to prolong lysis time at a concentration favoring ternary histone-DNA-heparin complex formation, and DNase effectively promotes clot lysis in combination with tissue plasminogen activator.  相似文献   

8.
A low Mr form (Mr 32,000) of single-chain urokinase-type plasminogen activator (scu-PA) was isolated from conditioned culture medium of a human lung adenocarcinoma cell line, CALU-3 (ATCC, HTB-55). The purified material (scu-PA-32k) consists of a single polypeptide chain and is immunologically similar to Mr 33,000 urokinase. Its NH2-terminal sequence is identical to that beginning at Leu-144 of Mr 54,000 urokinase. Whereas low Mr urokinase is derived from mature Mr 54,000 scu-PA by limited hydrolysis by plasmin first of the Lys-158-Ile-159 peptide bond and then of the Lys-136-Lys-137, scu-PA-32k is generated by specific hydrolysis of the Glu-143-Leu-144 peptide bond by an unidentified protease. scu-PA-32k resembles its Mr 54,000 scu-PA counterpart by its very low activity on chromogenic substrates for urokinase, by plasminogen-dependent fibrinolytic activity on fibrin plates, and by the lack of specific binding to fibrin. It activates plasminogen directly with high affinity, Km = 0.9 microM, but low turnover number, kcat = 0.0028 s-1. It is converted to fully active two-chain urokinase by plasmin with Km = 12 microM and kcat = 0.3 s-1. Like Mr 54,000 scu-PA, it causes significant lysis of a 125I-labeled fibrin clot in human plasma with relatively less fibrinogen breakdown as compared to urokinase. scu-PA-32k, which also has conserved fibrin specificity, represents a molecular variant which may be more suitable for large scale production as a fibrin-specific thrombolytic agent by recombinant DNA technology.  相似文献   

9.
Factor VII Activating Protease (FSAP) is a plasma protease affecting both coagulation and fibrinolysis. Although a role in hemostasis is still unclear, the identification of additional physiologic substrates will help to elucidate its role in this context. FSAP has been reported to cleave fibrinogen, but the functional consequences of this are not known. We have therefore undertaken this study to determine the implications of this cleavage for fibrin-clot formation and its lysis. Treatment of human fibrinogen with FSAP released an N-terminal peptide from the Bβ chain (Bβ1-53) and subsequently the fibrinopeptide B; within the Aα chain a partial truncation of the αC-region by multiple cleavages was seen. The truncated fibrinogen showed a delayed thrombin-catalyzed polymerization and formed fibrin clots of reduced turbidity, indicative of thinner fibrin fibers. Confocal laser scanning and scanning electron microscopy of these clots revealed a less coarse fibrin network with thinner fibers and a smaller pore size. A lower pore size was also seen in permeability studies. Unexpectedly, FSAP-treated fibrinogen or plasma exhibited a significantly faster tPA-driven lysis, which correlated exclusively with cleavage of fibrinogen and not with activation of plasminogen activators. Similar observations were also made in plasma after activation of endogenous zymogen FSAP, but not in plasma of carrier of the rare Marburg I single nucleotide polymorphism. In conclusion, altering fibrin clot properties by fibrinogenolysis is a novel function of FSAP in the vasculature, which facilitates clot lysis and may in vivo contribute to reduced fibrin deposition during thrombosis.  相似文献   

10.
Zhai P  Wakeham N  Loy JA  Zhang XC 《Biochemistry》2003,42(1):114-120
The bacterial protein streptokinase (SK) activates human plasminogen (Pg) into the fibrinolytic protease plasmin (Pm). Roughly 40 residues from the SK C-terminal domain are mobile in the crystal structure of SK complexed with the catalytic domain of Pm, and the functions of this C-tail remain elusive. To better define its roles in Pg activation, we constructed and characterized three C-terminal truncation mutants containing SK residues 1-378, 1-386, and 1-401, respectively. They exhibit gradually reduced amidolytic activity and Pg-activator activity, as well as marginally decreased binding affinity toward Pg, as more of the C-terminus is deleted. As compared with full-length SK, the shortest construct, SK(1-378), exhibits an 80% decrease in amidolytic activity (k(cat)/K(M)), an 80% decrease in Pg-activator activity, and a 30% increase in the dissociation constant toward the Pg catalytic domain. The C-terminal truncation mutations did not attenuate the resistance of the SK-Pm complex to alpha(2)-antiplasmin. Attempts at using a purified C-tail peptide to rescue the activity loss of the truncation mutants failed, suggesting that the integrity of the SK C-terminal peptide is important for the full function of SK.  相似文献   

11.
Cleavage of Arg(561)-Val(562) in plasminogen (Pg) generates plasmin (Pm) through a classical activation mechanism triggered by an insertion of the new amino terminus into a binding pocket in the Pg catalytic domain. Streptokinase (SK) circumvents this process and activates Pg through a unique nonproteolytic mechanism postulated to be initiated by the intrusion of Ile(1) of SK in place of Val(562). This hypothesis was evaluated in equilibrium binding and kinetic studies of Pg activation with an SK mutant lacking Ile(1) (SK(2--414)). SK(2--414) retained the affinity of native SK for fluorescein-labeled [Lys]Pg and [Lys]Pm but induced no detectable conformational activation of Pg. The activity of SK(2--414) was partially restored by the peptides SK(1--2), SK(1--5), SK(1--10), and SK(1--15), whereas Pg(562--569) peptides were much less effective. Active site-specific fluorescence labeling demonstrated directly that the active catalytic site was formed on the Pg zymogen by the combination of SK(1--10) and SK(2--414), whereas sequence-scrambled SK(1-10) was inactive. The characterization of SK(1--10) containing single Ala substitutions demonstrated the sequence specificity of the interaction. SK(1--10) did not restore activity to the further truncated mutant SK(55-414), which was correlated with the loss of binding affinity of SK(55--414) for labeled [Lys]Pm but not for [Lys]Pg. The studies support a mechanism for conformational activation in which the insertion of Ile(1) of SK into the Pg amino-terminal binding cleft occurs through sequence-specific interactions of the first 10 SK residues. This event and the preferentially higher affinity of SK(2--414) for the activated proteinase domain of Pm are thought to function cooperatively to trigger the conformational change and stabilize the active zymogen conformation.  相似文献   

12.
THE human plasma protein, fibrinogen, is a disulphide bonded1 dimer2, each unit containing an Aα, Bβ and 8 chain*, interconnected by disulphide bridges3. Thrombin (E.C.3.4.4.-13) releases fibrinopeptides A and B from the Aα and Bβ chains respectively4 to form fibrin monomer (α2β2γ2) ? which polymerizes to form fibrin polymer or clotted fibrin. This polymer, following factor XIII (plasma transglutaminase, fibrin stabilizing factor) mediated crosslinking among the α chains and among the γ chains5, is one of the major and initiating constituents of a thrombus. Fibrinolytic activators, for example, streptokinase (SK) and urokinase (UK), are of thrombolytic value as they convert the thrombus plasminogen to plasmin (E.C.3.4.4.14) which by fibrinolytic action dissolves the thrombus. Whereas the interaction of fibrinogen and plasmin has been well studied6–9, little is known concerning the mechanism of plasmin mediated fibrin clot lysis. I report here on the mechanism of non-cross-linked fibrin clot solubilization in near physiological conditions.  相似文献   

13.
On the mechanism of fibrin-specific plasminogen activation by staphylokinase   总被引:10,自引:0,他引:10  
The mechanism of plasminogen activation by recombinant staphylokinase was studied both in the absence and in the presence of fibrin, in purified systems, and in human plasma. Staphylokinase, like streptokinase, forms a stoichiometric complex with plasminogen that activates plasminogen following Michaelis-Menten kinetics with Km = 7.0 microM and k2 = 1.5 s-1. In purified systems, alpha 2-antiplasmin inhibits the plasminogen-staphylokinase complex with k1(app) = 2.7 +/- 0.30 x 10(6) M-1 s-1 (mean +/- S.D., n = 12), but not the plasminogen-streptokinase complex. Addition of 6-aminohexanoic acid induces a concentration-dependent reduction of k1(app) to 2.0 +/- 0.17 x 10(4) M-1 s-1 (mean +/- S.D., n = 5) at concentrations greater than or equal to 30 mM, with a 50% reduction at a 6-aminohexanoic acid concentration of 60 microM. Staphylokinase does not bind to fibrin, and fibrin stimulates the initial rate of plasminogen activation by staphylokinase only 4-fold. Staphylokinase induces a dose-dependent lysis of a 0.12-ml 125I-fibrin-labeled human plasma clot submersed in 0.5 ml of citrated human plasma; 50% lysis in 2 h is obtained with 17 nM staphylokinase and is associated with only 5% plasma fibrinogen degradation. Corresponding values for streptokinase are 68 nM and more than 90% fibrinogen degradation. In the absence of a fibrin clot, 50% fibrinogen degradation in human plasma in 2 h requires 790 nM staphylokinase, but only 4.4 nM streptokinase. These results suggest the following mechanism for relatively fibrin-specific clot lysis with staphylokinase in a plasma milieu. In plasma in the absence of fibrin, the plasminogen-staphylokinase complex is rapidly neutralized by alpha 2-antiplasmin, thus preventing systemic plasminogen activation. In the presence of fibrin, the lysine-binding sites of the plasminogen-staphylokinase complex are occupied and inhibition by alpha 2-antiplasmin is retarded, thus allowing preferential plasminogen activation at the fibrin surface.  相似文献   

14.
In a previous report we showed that plasmin-dependent lysis of a fibrin polymer, produced from purified components, was totally blocked if annexin II heterotetramer (AIIt) was present during fibrin polymer formation. Here, we show that AIIt inhibits fibrin clot lysis by stimulation of plasmin autodegradation, which results in a loss of plasmin activity. Furthermore, the C-terminal lysine residues of its p11 subunit play an essential role in the inhibition of fibrin clot lysis by AIIt. We also found that AIIt binds to fibrin with a K(d) of 436 nm and a stoichiometry of about 0.28 mol of AIIt/mol of fibrin monomer. The binding of AIIt to fibrin was not dependent on the C-terminal lysines of the p11 subunit. Furthermore, in the presence of plasminogen, the binding of AIIt to fibrin was increased to about 1.3 mol of AIIt/mol of fibrin monomer, suggesting that AIIt and plasminogen do not compete for identical sites on fibrin. Immunohistochemical identification of p36 and p11 subunits of AIIt in a pathological clot provides important evidence for its role as a physiological fibrinolytic regulator. These results suggest that AIIt may play a key role in the regulation of plasmin activity on the fibrin clot surface.  相似文献   

15.
Streptokinase purified from Streptococcus equinus VIT_VB2 isolated from bovine milk sample was immobilized in various solid supports namely entrapment in agarose gel, calcium alginate beads and gelatin gel by cross-linking with formaldehyde. Immobilization of streptokinase in calcium alginate beads showed maximum efficiency (81.8?±?1.06%) when compared with entrapment with agarose gel (55.6?±?2.17%) and cross-linked gelatin formaldehyde gel (71.0?±?1.54%). The purified SK activity was expressed maximum in calcium alginate (1%) and gelatin gel (0.25%) with 1292.68?±?1.33 and 1121.9?±?1.2?U?mL?1, respectively. Similarly, SK entrapped in gelatin gel and calcium alginate showed maximum in vitro blood clot lysis activity with 77.67?±?2.64% and 76.16?±?2.72%, respectively. The immobilized SK in gelatin gel showed complete clot lysis within 15?min; hence, this application of the study could be used in the treatment of superficial thrombophlebitis, phlebitis, and venous thrombosis. These beads were used for three repeated cycles to check the conversion of substrates into their products, and we concluded that SK can be immobilized in the suitable matrices. Therefore, this helps in the drug-delivery strategies in highly efficient way, moreover, economically competent process in the pharmaceutics.  相似文献   

16.
《The Journal of cell biology》1994,126(4):1089-1097
The extracellular domain of the myelin P0 protein is believed to engage in adhesive interactions and thus hold the myelin membrane compact. We have previously shown that P0 can behave as a homophilic adhesion molecule through interactions of its extracellular domains (Filbin, M. T., F. S. Walsh, B. D. Trapp, J. A. Pizzey, and G. I. Tennekoon. 1990. Nature (Lond.) 344:871-872). To determine if the cytoplasmic domain of P0 must be intact for the extracellular domains to adhere, we compared the adhesive capabilities of P0 proteins truncated at the COOH-terminal to the full-length P0 protein. P0 cDNAs lacking nucleotides coding for the last 52 or 59 amino acids were transfected into CHO cells, and surface expression of the truncated proteins was assessed by immunofluorescence, surface labeling followed by immunoprecipitation, and an ELISA. Cell lines were chosen that expressed at least equivalent amounts of the truncated P0 proteins at the surface as did a cell line expressing the full-length P0. The adhesive properties of these three cell lines were compared. It was found that when a suspension of single cells was allowed to aggregate for a period of 60 min, only the cells expressing the full-length P0 had formed large aggregates, while the cells expressing the truncated P0 molecules were still mostly single cells indistinguishable from the control cells. Furthermore, 25-30% of the full-length P0 was insoluble in NP40, indicative of an interaction with the cytoskeleton, whereas only 5-10% of P0 lacking 52 amino acids and none of P0 lacking 59 amino acids were insoluble. These results suggest that for the extracellular domain of P0 to behave as a homophilic adhesion molecule, its cytoplasmic domain must be intact, and most probably, it is interacting with the cytoskeleton.  相似文献   

17.
Clotting and fibrinolysis are initiated simultaneously in vivo, and fibrinolysis usually occurs without any individualized lysis front (intrinsic fibrinolysis). We have developed a novel model to assess whether morphological changes resulting from intrinsic fibrinolysis are similar to those previously reported at the lysis front using externally applied lytic agents. Fibrin assembly and fibrinolysis were followed in real-time by confocal microscopy using gold-labeled fibrinogen molecules. An increase in fiber absorbance (30%, p < 0.01) and a decrease in fiber diameter (60%, p < 0.01) due to the ongoing accumulation and packing of fibrin molecules were the most significant detectable features occurring during fibrin assembly. Similar features with a similar magnitude were observed during fibrin dissolution, but in the reverse order and with a 3-fold increase in duration. Then, lysing fibers were progressively transected laterally, and thinner fibers were cleaved at a 2.5-fold faster rate than thicker fibers (p < 0.001). Frayed lysing fibers were seen to interact progressively with adjoining fibers (agglomeration), leading to a 76 and 88% increase in the network pore diameter (p < 0.05) and fiber diameter (p < 0.01), respectively. At the maximum decrease in fiber absorbance (46%, p < 0.05), the network suddenly collapsed with the release of large fragments that gradually vanished. Morphological changes of fibrin that occur during intrinsic fibrinolysis are similar as those observed next to the lysis front, although they are not restricted spatially to the clot/surrounding milieu interface but are observed through the entire clot.  相似文献   

18.
Streptokinase (SK) activates human fibrinolysis by inducing non-proteolytic activation of the serine proteinase zymogen, plasminogen (Pg), in the SK.Pg* catalytic complex. SK.Pg* proteolytically activates Pg to plasmin (Pm). SK-induced Pg activation is enhanced by lysine-binding site (LBS) interactions with kringles on Pg and Pm, as evidenced by inhibition of the reactions by the lysine analogue, 6-aminohexanoic acid. Equilibrium binding analysis and [Lys]Pg activation kinetics with wild-type SK, carboxypeptidase B-treated SK, and a COOH-terminal Lys414 deletion mutant (SKDeltaK414) demonstrated a critical role for Lys414 in the enhancement of [Lys]Pg and [Lys]Pm binding and conformational [Lys]Pg activation. The LBS-independent affinity of SK for [Glu]Pg was unaffected by deletion of Lys414. By contrast, removal of SK Lys414 caused 19- and 14-fold decreases in SK affinity for [Lys]Pg and [Lys]Pm binding in the catalytic mode, respectively. In kinetic studies of the coupled conformational and proteolytic activation of [Lys]Pg, SKDeltaK414 exhibited a corresponding 17-fold affinity decrease for formation of the SKDeltaK414.[Lys]Pg* complex. SKDeltaK414 binding to [Lys]Pg and [Lys]Pm and conformational [Lys]Pg activation were LBS-independent, whereas [Lys]Pg substrate binding and proteolytic [Lys]Pm generation remained LBS-dependent. We conclude that binding of SK Lys414 to [Lys]Pg and [Lys]Pm kringles enhances SK.[Lys]Pg* and SK.[Lys]Pm catalytic complex formation. This interaction is distinct structurally and functionally from LBS-dependent Pg substrate recognition by these complexes.  相似文献   

19.
Kringles K1-3, K4 and K5 are studied for their effect on tissue plasminogen activator-induced fibrin clot lysis in the presence of Glu- and Lys-plasminogen. It is established that kringles K4 and K5 inhibit fibrinolysis of Glu-plasminogen, and K1-3--that of Lys-plasminogen. The role of plasminogen molecule kringles in the plasminogen interaction with fibrin polymer is discussed.  相似文献   

20.
Although several recent studies employing various truncated fragments of streptokinase (SK) have demonstrated that the high-affinity interactions of this protein with human plasminogen (HPG) to form activator complex (SK-HPG) are located in the central region of SK, the exact location and nature of such HPG interacting site(s) is still unclear. In order to locate the "core" HPG binding ability in SK, we focused on the primary structure of a tryptic fragment of SK derived from the central region (SK143-293) that could bind as well as activate HPG, albeit at reduced levels in comparison to the activity of the native, full-length protein. Because this fragment was refractory to further controlled proteolysis, we took recourse to a synthetic peptide approach wherein the HPG interacting properties of 16 overlapping 20-mer peptides derived from this region of SK were examined systematically. Only four peptides from this set, viz., SK234-253, SK254-273, SK274-293, and SK263-282, together representing the contiguous sequence SK234-293, displayed HPG binding ability. This was established by a specific HPG-binding ELISA as well as by dot blot assay using 125I-labeled HPG. These results showed that the minimal sequence with HPG binding function resided between residues 234 and 293. None of the synthetic SK peptides was found to activate HPG, either individually or in combination, but, in competition experiments where each of the peptides was added prior to complex formation between SK and HPG, three of the HPG binding peptides (SK234-253, SK254-273, and SK274-293) inhibited strongly the generation of a functional activator complex by SK and HPG. This indicated that residues 234-293 in SK participate directly in intermolecular contact formation with HPG during the formation of the 1:1 SK-HPG complex. Two of the three peptides (SK234-253 and SK274-293), apart from interfering in SK-HPG complex formation, also showed inhibition of the amidolytic activity of free HPN by increasing the K(m) by approximately fivefold. A similar increase in K(m) for amidolysis by HPN as a result of complexation with SK has been interpreted previously to arise from the steric hinderance at or near the active site due to the binding of SK in this region. Thus, our results suggest that SK234-253 and SK274-293 also, like SK, bound close to the active site of HPN, an event that was reflected in the observed alteration in its substrate accessibility. By contrast, whereas the intervening peptide (SK254-273) could not inhibit amidolysis by free HPN, it showed a marked inhibition of the activation of "substrate" PG (human or bovine plasminogen) by activator complex, indicating that this particular region is intimately involved in interaction of the SK-HPG activator complex with substrate plasminogen during the catalytic cycle. This finding provides a rational explanation for one of the most intriguing aspects of SK action, i.e., the ability of the SK-HPG complex to catalyze selectively the activation of substrate molecules of PG to PN, whereas free HPN alone cannot do so. Taken together, the results presented in this paper strongly support a model of SK action in which the segment 234-293 of SK, by virtue of the epitopes present in residues 234-253 and 274-293, binds close to the active center of HPN (or, a cryptic active site, in the case of HPG) during the intermolecular association of the two proteins to form the equimolar activator complex; the segment SK254-273 present in the center of the core region then imparts an ability to the activator complex to interact selectively with substrate PG molecules during each PG activation cycle.  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