首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The activation of Lys-plasminogen to plasmin by streptokinase was promoted by soluble fibrin such that Km was decreased and Vmax. increased. Enhancement was also observed when Glu-plasminogen was the substrate and was shared by the preformed streptokinase-plasminogen activator complex, indicating that the stimulation was not exerted primarily on the rate of active site formation.  相似文献   

2.
A comparative kinetic analysis of the enzymatic activities of one-chain and two-chain tissue-type plasminogen activator (t-PA) demonstrates that two-chain t-PA catalyzes the hydrolysis of the peptide substrate D-Val-Leu-Arg-pNA about 4-fold more effectively than one-chain t-PA. The difference is accounted for almost entirely by a corresponding difference is the kcat values of the enzymes, whereas the Km values are similar. The amidolytic activity of two-chain t-PA is not enhanced by intact or partially plasmin-degraded fibrin. In contrast, the activity of one-chain t-PA is stimulated up to 3.7-fold by intact fibrin and up to 4.7-fold by plasmin-degraded fibrin (fibrin X-fragment). The stimulatory effects are realized via increases in the kcat values. It appears thus that in the presence of fibrin the intrinsically inferior catalytic properties of one-chain t-PA become similar to the properties of two-chain t-PA. The dependency of the activity of one-chain t-PA on the concentration of fibrin monomer is consistent with a single association site of both proteins and an association constant of Kass = 6.25 x 10(6) l/mol. Stimulation of one-chain t-PA by plasmin-degraded fibrin is more complex and appears to involve two different binding sites with association constants of Kass = 0.67 x 10(9) l/mol and Kass = 3.85 x 10(6) l/mol, respectively. The stimulatory effects of fibrin and partially plasmin-degraded fibrin on one-chain t-PA are suppressed by epsilon-aminocaproic acid and by a monoclonal antibody directed against the lysine binding site of t-PA. The latter findings support the notion that fibrin activation of one-chain t-PA is mediated by the lysine binding site on kringel domains of the enzyme.  相似文献   

3.
The influence of antiplasmin on the interaction between fibrin and plasminogen was studied in plasma and in a purified system. The amount of plasminogen bound to fibrin was quantitated using trace amounts of 125I-labeled Glu-plasminogen (plasminogen with NH2-terminal glutamic acid) or 125I-labeled Lys-plasminogen (NH2-terminal lysine).When whole plasma was clotted, 5.2% of Glu-plasminogen was associated with the fibrin clot. In plasma clotted in the presence of 20 mM 6-amino-hexanoic acid only 1.4% of the plasminogen was bound to fibrin, indicating that about 4% of the plasma plasminogen specifically binds to fibrin. With Lys-plasminogen these values were approximately twice as high.When antiplasmin-depleted plasma was used, only slightly higher amounts of both types of plasminogen were associated with the fibrin. The adsorbed plasminogen was not significantly eluted with plasma or with purified antiplasmin at physiological concentrations.These findings indicate that antiplasmin does not play a significant role in the inhibition of the binding of plasminogen to fibrin or the dissociation of the plasminogen · fibrin complex.These observations in conjunction with previous findings on the kinetics of the plasmin-antiplasmin reaction suggest that the lysine-binding site of plasminogen, which is responsible both for its interaction with fibrin and its interaction with antiplasmin, plays an important role in the very fast neutralization of plasmin formed in circulating blood and serves to attach plasminogen to fibrin and thereby sequestrate plasmin formed in loco from circulating antiplasmin.  相似文献   

4.
Fibrin (Fn) enhances plasminogen (Pg) activation by tissue-type plasminogen activator (tPA) by serving as a template onto which Pg and tPA assemble. To explore the contribution of the Pg/Fn interaction to Fn cofactor activity, Pg variants were generated and their affinities for Fn were determined using surface plasmon resonance (SPR). Glu-Pg, Lys-Pg (des(1-77)), and Mini-Pg (lacking kringles 1-4) bound Fn with K(d) values of 3.1, 0.21, and 24.5 μm, respectively, whereas Micro-Pg (lacking all kringles) did not bind. The kinetics of activation of the Pg variants by tPA were then examined in the absence or presence of Fn. Whereas Fn had no effect on Micro-Pg activation, the catalytic efficiencies of Glu-Pg, Lys-Pg, and Mini-Pg activation in the presence of Fn were 300- to 600-fold higher than in its absence. The retention of Fn cofactor activity with Mini-Pg, which has low affinity for Fn, suggests that Mini-Pg binds the tPA-Fn complex more tightly than tPA alone. To explore this possibility, SPR was used to examine the interaction of Mini-Pg with Fn in the absence or presence of tPA. There was 50% more Mini-Pg binding to Fn in the presence of tPA than in its absence, suggesting that formation of the tPA-Fn complex exposes a cryptic site that binds Mini-Pg. Thus, our data (a) indicate that high affinity binding of Pg to Fn is not essential for Fn cofactor activity, and (b) suggest that kringle 5 localizes and stabilizes Pg within the tPA-Fn complex and contributes to its efficient activation.  相似文献   

5.
Pretreatment of native plasminogen with plasmin or activators resulted in a pronounced increase in the binding of plasminogen to fibrin. The pretreated plasminogen was considered to be identical to the proteolytically degraded proenzyme with NH2-terminal lysine, valine or methionine, which is formed as an intermediate stage during activation of plasminogen. Bound plasminogen could be extracted by 6-aminohexanoic acid indicating a reversible binding between plasminogen and fibrin. Adsorption of pretreated plasminogen decreased when increasing concentrations of 6-aminohexanoic acid or trans-4-aminomethylcyclohexane-1-carboxylic acid (t-AMCHA) were present during fibrin formation. The concentration of amino acid producing a decrease in the binding of pretreated plasminogen to 0.5 of the amount bound in the absence of amino acid was 8.0-10(-5) M with 6-aminohexanoic acid and 1.7.10-5 M with t-AMCHA. The decrease in binding is most likely related to an effect of the amino acids on plasminogen, since agarose gel electrophoresis of pretreated plasminogen in the presence of 6-aminohexanoic acid or t-AMCHA showed a cathodic shift in mobility at the same range of concentrations of amino acid, which produced the decrease in binding of plasminogen to fibrin. Evidence is provided that the decrease in binding of proteolytically degraded plasminogen may result in an inhibition of fibrinolysis caused by activators.  相似文献   

6.
Interaction of plasminogen and fibrin in plasminogen activation   总被引:2,自引:0,他引:2  
Glu1-, Lys77-, miniplasminogens, kringle 1-3, kringle 1-5A, and kringle 1-5R were able to bind with fibrin, while microplasminogen and kringle 4 did not bind significantly. Kringle 1-5A, but not kringle 1-3, effectively inhibited the binding of Glu1-, Lys77-, and miniplasminogens with fibrin. Miniplasminogen also inhibited the binding of Glu1-plasminogen with fibrin. The binding of kringle 1-3 with fibrin was blocked by mini- or Glu1-plasminogen. It is therefore evident that there are two fibrin-binding domains in plasminogen and that the one in kringle 5 is of higher affinity than that in kringle 1-3. CNBr cleavage products of fibrinogen effectively enhanced the activation of Glu1-, Lys77-, or miniplasminogens, but not microplasminogen, by tissue-type plasminogen activator. Kringle 1-5, but not kringle 1-3, dose-dependently inhibited the enhancement by fibrinogen degradation products of Glu1-plasminogen activation by the activator. Lysine and epsilon-aminocaproic acid could inhibit the binding of plasminogens and plasminogen derivatives with fibrin and block the enhancement effect of fibrinogen degradation products on plasminogen activation. The data clearly illustrate that the binding of plasminogen with fibrin, mainly determined by kringle 5, is essential for effective activation by tissue-type plasminogen activator. However, the presence of kringle 1-4 in the plasminogen molecule is required for the full enhancing effect since the kcat/Km of miniplasminogen activation in the presence of fibrinogen degradation products was 8.2 microM-1 min-1 which is significantly less than 52.0 microM-1 min-1 of Glu1-plasminogen.  相似文献   

7.
8.
Out of 29 disulfide bonds in human fibrinogen, 7 were cleaved during limited reduction under nondenaturing conditions in calcium-free buffer: 2 A alpha 442Cys-A alpha 472Cys and 2 gamma 326Cys-gamma 339Cys intrachain disulfide bonds in the carboxy-terminal ends of the A alpha- and gamma-chains and the symmetrical disulfide bonds at gamma 8Cys, gamma 9Cys, and A alpha 28Cys. We studied the loss of thrombin clottability that followed limited reduction and the increase in the susceptibility of the fibrinogen A alpha 19-A alpha 20 bond to hydrolysis by thrombin. Using differential scanning calorimetry, we show that the extent of unfolding and denaturation of specific domains following limited reduction is small. Heat absorption peaks corresponding to the melting of the major regions of compact structure give high calorimetric enthalpies, as in untreated nonreduced fibrinogen, indicating that substantial regions of native structure are still present in partially reduced fibrinogen. Thrombin releases fibrinopeptide A at an identical rate as in nonreduced fibrinogen while fibrinopeptide B release is slower. Sedimentation velocity studies show that thrombin treatment leads to complex formation; however, gelation does not occur. Amino-terminal analysis indicates that the second thrombin cleavage in the A alpha-chain at A alpha 19-A alpha 20 takes place only after fibrinopeptide A release. Thus, the loss of clottability appears to result from perturbation of carboxy-terminal polymerization sites, probably a consequence of gamma 326Cys-gamma 339Cys intrachain disulfide bond cleavage. The thrombin-treated partially reduced fibrinogen remains soluble in buffered saline and fully expresses at least one epitope, B beta 15-21, unique to fibrin. Furthermore, this nonclottable form accelerates the tissue plasminogen activator dependent conversion of plasminogen to plasmin.  相似文献   

9.
The kinetic of plasmin, Va1442-plasmin, Lys530-plasmin inhibition reaction by alpha 2-antiplasmin as well as interaction of the inhibitor with different derivatives of the plasminogen and its fragments were studied. It was shown that plasmin, mini- and micro-plasmin activity decreased by 97, 88 and 85%, respectively, for equimolar ratio 1:1 of the inhibitor. The value of the inhibition reached its maximum in 1-2, 5-10 and 10-15 min, respectively. The constants of the complex formation rate were 1.4 x 10(6); 1.7 x 10(5) and 6.2 x 10(4) M-1s-1 for the plasmin, mini- and micro-plasmin with alpha 2-antiplasmin, respectively. Both 10(-2) M 6-aminohexanoic acid and 10(-1) M arginine reduced the complex formation rate between plasmin, mini-plasmin and alpha 2-antiplasmin to the value of the rate reaction between micro-plasmin and inhibitor. alpha 2-Antiplasmin bound with all investigated derivatives and fragments of plasminogen. The amount of inhibitor decreased in the series: plasmin, kringle 1-3, kringle 4, mini-plasminogen, micro-plasminogen. The kringle 1-4 and kringle 5 were determined to control the rate of reaction between enzyme and inhibitor, being not necessary for the inhibition. The comparison of the inhibitor interaction with DPP-plasmin, mini-plasminogen and micro-plasminogen displayed the possibility of the additional region existence in catalytic domain. This region participated in the complex with alpha 2-antiplasmin formation. It is supposed that the multisite interaction between plasmin and alpha 2-antiplasmin provides for the specificity and efficiency the inhibitor action.  相似文献   

10.
On the interaction of F-actin with fibrin   总被引:5,自引:0,他引:5  
  相似文献   

11.
The conformational properties of xanthans with partially hydrolyzed side chains were in vestigated by optical rotation, CD, and differential scanning calorimetry (DSC). All variants displayed the well-known temperature-driven, cooperative order–disorder transition, and both optical rotation and DSC showed that the transition temperature was essentially independent of the content of terminal β-mannose. It was found that up to 80% of the changes in the specific optical rotation accompanying the transition reflects conformational changes linked to the terminal β-mannose in the side chains. Modification of the sidechains also affected the CD when xanthan was in the ordered state, but in this case the data suggest that the glucuronic acid is the major component determining the magnitude of the CD signal. DSC measurements showed that the transition enthalpy (ΔHcal) increased linearly with the fraction of β-mannose, again indicating that a significant part (up to 80%) of ΔHcal reflects conformational changes in the side chains. The conformational transition of the xanthan variants generally showed a higher degree of cooperativity (sharper transition) than unmodified, pyruvated xanthan. Calculation of the cooperativity parameter σ by means of the Zimm–Bragg theory (OR data) or from the ratio between ΔHcal and the van't Hoff enthalpy (ΔHvH) using DSC data showed a correlation between σ and the content of β-mannose, but the two methods gave different results when the content of β-mannose approached 100%. The ionic strength dependence of the transition temperature, expressed as d (log I)/d(T?1m), was nearly identical for intact xanthan and a sample containing only 6% of the terminal β-mannose. Application of the Manning polyelectrolyte theory does not readily account for the observed ΔHcal values, neither does it provide new information on the nature of the ordered and disordered conformations in xanthan. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
The binding of tissue-type plasminogen activator (t-PA) to fibrin is mediated both by its finger domain and by its kringle-2 domain. In this report, we investigate the relative affinities of these domains for lysine. Human recombinant t-PA deletion-mutant proteins were prepared and their ability to bind to lysine-Sepharose was investigated. Mutants containing the kringle-2 domain bound to lysine-Sepharose, whereas mutants lacking this domain but containing the finger domain, the epidermal growth factor domain or the kringle-1 domain did not bind to lysine-Sepharose. Mutant proteins containing the kringle-2 domain could be specifically eluted from lysine-Sepharose with epsilon-amino caproic acid. This lysine derivative also abolished fibrin binding by the kringle-2 domain but had no effect on the fibrin-binding property of the finger domain. Thus, a lysine-binding site is involved in the interaction of the kringle-2 domain with fibrin but not in the interaction of the finger domain with fibrin. The implications of the nature of these two distinct interactions of t-PA with fibrin on plasminogen activation by t-PA will be discussed.  相似文献   

13.
14.
The reaction between plasminogen activators and plasminogen activator inhibitor-1 is characterized by an initial rapid formation of an inactive reversible complex. The second-order association rate constant (k1) of complex formation of recombinant two-chain tissue-type plasminogen activator (rt-PA) or recombinant two-chain urokinase-type plasminogen activator (rtcu-PA) by recombinant plasminogen activator inhibitor-1 (rPAI-1) is 2.9 +/- 0.4 x 10(7) M-1 s-1 (mean +/- S.D., n = 30) and 2.0 +/- 0.6 x 10(7) M-1 s-1 (n = 12), respectively. Different molecular forms of tissue- or urokinase-type plasminogen activator which do not form covalent complexes with rPAI-1, including rt-PA-Ala478 (rt-PA with the active-site Ser478 mutagenized to Ala) and anhydro-urokinase (rtcu-PA with the active-site Ser356 converted to dehydroalanine) reduced k1 in a concentration-dependent manner, compatible with 1:1 stoichiometric complex formation between rPAI-1 and these ligands. The apparent dissociation constant (KD) of the complex between rPAI-1 and rt-PA-Ala478, determined as the concentration of rt-PA-Ala478 which reduced k1 to 50% of its control value, was 3-5 nM. Corresponding concentrations of active-site-blocked two-chain rt-PA were 150-250-fold higher. The concentration of anhydro-urokinase which reduced k1 to 50% was 4-6 nM, whereas that of active-site-blocked rtcu-PA was 100-250-fold higher. Recombinant single-chain urokinase-type plasminogen activator had an apparent KD of about 2 microM. These results suggest that inhibition of rt-PA or rtcu-PA by rPAI-1 proceeds via a reversible high affinity interaction which does not require a functional active site but which is markedly reduced following inactivation of the enzymes with active-site titrants.  相似文献   

15.
Immobilization of proteins on partially hydrolyzed agarose beads   总被引:2,自引:0,他引:2  
Treatment of agarose beads with mild acid (0.2 M HCl, 55 degrees C, several hours) hydrolyzes some of the glycosidic bonds between D-galactosyl residues and 3,6-anhydro-L-galactosyl residues, and thus produces aldehydo-groups useful for immobilization of amino compounds by reductive amination with NaCNBH3. More than 20 mg (0.3 mumol) of bovine serum albumin could be coupled per gram of partially hydrolyzed agarose beads. Arthrobacter neuraminidase immobilized by this method was useful for desialylation of sialyl glycoconjugates, and was found not to leach from the gel and to be much more thermostable than the free enzyme.  相似文献   

16.
Thrombin mediates the balance between coagulant and fibrinolytic forces and has numerous cellular effects. This intricate balance is maintained by biochemical mechanisms that regulate thrombin activity. Disruption of this balance could lead to bleeding or thrombosis. Once thrombin is generated, two major mechanisms regulate its activity. By binding fibrin, thrombin's activity is localized to the thrombus, a process that limits its systemic procoagulant effects. Thrombin that escapes into the circulation is efficiently inactivated by plasma inhibitors, such as antithrombin, or is sequestered by thrombomodulin on the endothelium. Although thrombin's interaction with fibrin limits its systemic effects, fibrin-bound thrombin resists inactivation and can produce a local procoagulant stimulus that triggers thrombus growth. Direct thrombin inhibitors were developed, at least in part, to target fibrin-bound thrombin. These agents are finding their niche for the prevention and treatment of venous and arterial thrombosis. The mechanisms by which thrombin binds fibrin are reviewed in this paper. As well, the potential pathological consequences of thrombin's interaction with fibrin are discussed.  相似文献   

17.
18.
Summary 1. Cereal straw was partially hydrolyzed with dilute sulphuric or hydrochloric acid at elevated temperatures, yielding about 20% of matter assimilable byTorulopsis utilis (calculated on oven-dry straw), which consisted chiefly of xylose, with small amounts of glucose and acetic acid.2. Experiments on a laboratory scale with non-fermentable substrates like ethanol and acetic acid showed that also in these cases an ample aeration is an essential condition for a good yield.3. Acetic acid is harmful to the yeast, even in small concentrations, when the pH is lower than 5.0.4. Small quantities of glucose or acetic acid stimulate the conversion of xylose byTorulopsis utilis.5. With an adequate aeration and an initial pH of 5.5–6.0 (maintained at this level until the acetic acid has disappeared and then lowered to 4.0–4.5) satisfactory conversion rates and yields could be obtained on straw extracts as mentioned sub 1.6. The results of the laboratory experiments mentioned sub 5 could be reproduced in a semi-technical installation with a capacity of 200 l straw extract.7. Good results were obtained in this installation with aVogelbusch aeration device, while foaming could be adequately controlled by means of a rotating wire screen.  相似文献   

19.
D L Higgins  G A Vehar 《Biochemistry》1987,26(24):7786-7791
Tissue-type plasminogen activator (t-PA) plays a central role in fibrinolysis in vivo. Although it is known to bind to fibrin, the dissociation constant (Kd) and number of moles bound per mole of fibrin monomer (n) have never been measured directly. In this study, the binding of both the one-chain form and the two-chain form of recombinant, human t-PA to fibrin was measured. Although more one-chain t-PA than two-chain t-PA is bound to fibrin, the Kd's and n's were within experimental error of each other. Significantly more t-PA is bound to clots made from fibrinogen which has been digested with plasmin than to clots made from intact fibrinogen. The additional binding was shown to be due to the formation of new set(s) of binding site(s) with dissociation constants that are 2-4 orders of magnitude tighter than the binding site present on clots made from intact fibrinogen. epsilon-Aminocaproic acid was capable of competing for the loose binding site present on both intact and degraded fibrin but had little effect on the binding of t-PA to the new site(s) formed by plasmin digestion. This increase in binding caused by plasmin-mediated proteolysis of fibrin suggests a possible mechanism for a positive regulation capable of accelerating fibrinolysis.  相似文献   

20.
A recombinant chimeric plasminogen activator (GHRP-scu-PA-32K), consisting of the tetrapeptide Gly-His-Arg-Pro fused to the N-terminus of the low-molecular single-chain urokinase-type plasminogen activator (Leu144-Leu411), was produced by expression in CHO cells. The stable expression cell line was selected for large-scale expression. The product was purified by antibody-Sepharose affinity chromatography with a recovery of 67%. The apparent molecular weight of purified GHRP-scu-PA-32K was 33 kDa according to SDS-PAGE. Its specific activity was 150000 IU/mg protein according to fibrin plate determination. The conversion of single-chain to two-chain molecules mediated by plasmin was comparable for GHRP-scu-PA-32K (K(m)=4.9 microM, k(2)=0.35 s(-1)) and scu-PA-32K. The activation of plasminogen by GHRP-scu-PA-32K (K(m)=1.02 microM, k(2)=0.0028 s(-1)) was also similar to that of scu-PA-32K. The fibrin binding of GHRP-scu-PA-32K was 2.5 times higher than that of scu-PA-32K at a fibrin concentration of 3.2 mg/ml. In contrast to scu-PA-32K in vitro 125I-fibrin-labeled plasma clot lysis, GHRP-scu-PA had a higher thrombolytic potency, whereas it depleted less fibrinogen in plasma. These results show that GHRP-scu-PA-32K as expected is a potential thrombolytic agent.  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