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 0 毫秒
1.
An investigation was made to determine whether it is possible to attract tissue plasminogen activator (tPA) to the site of a thrombus by means of an antibody with affinites for both tPA and fibrin. A bispecific antibody conjugate was constructed by cross-linking two monoclonal antibodies: one specific for tPA, the other specific for fibrin. The bispecific antibody enhanced fibrinolysis by capturing tPA at the site of a fibrin deposit. In an in vitro quantitative fibrinolysis assay, the relative fibrinolytic potency of tPA bound to the bispecific antibody was 13 times greater than that of tPA and 200 times greater than that of urokinase. When fibrin was treated with the bispecific antibody before being mixed with tPA, the relative fibrinolytic potency of tPA was enhanced 14-fold. This capture also occurred when the concentration of tPA present in the assay was less than the concentration of tPA present in normal human plasma. In a human plasma clot assay, samples containing both the bispecific antibody and tPA exhibited significantly more lysis than did samples containing tPA alone. In spite of the increased clot lysis effected by the presence of bispecific antibody, there was no significant increase in fibrinogen or alpha 2-antiplasmin degradation at equal tPA concentrations. The ability of the bispecific antibody to concentrate exogenous tPA in vivo was then examined in the rabbit jugular vein model. Systemic infusion of a small amount of tPA (10,000 units) produced no significant increment in thrombolysis over the level of spontaneous lysis (14 +/- 8%). However, the simultaneous infusion of 10,000 units of tPA and 2 mg of bispecific antibody resulted in 42 +/- 14% (p less than 0.01) lysis. These results suggest that a molecule capable of binding both fibrin and tPA with high affinity could enhance thrombolysis in the circulation by capturing endogenous tPA.  相似文献   

2.
Type 1 plasminogen activator inhibitor binds to fibrin via vitronectin   总被引:2,自引:0,他引:2  
Type 1 plasminogen activator inhibitor (PAI-1), the primary inhibitor of tissue-type plasminogen activator (t-PA), circulates as a complex with the abundant plasma glycoprotein, vitronectin. This interaction stabilizes the inhibitor in its active conformation In this report, the effects of vitronectin on the interactions of PAI-1 with fibrin clots were studied. Confocal microscopic imaging of platelet-poor plasma clots reveals that essentially all fibrin-associated PAI-1 colocalizes with fibrin-bound vitronectin. Moreover, formation of platelet-poor plasma clots in the presence of polyclonal antibodies specific for vitronectin attenuated the inhibitory effects of PAI-1 on t-PA-mediated fibrinolysis. Addition of vitronectin during clot formation markedly potentiates PAI-1-mediated inhibition of lysis of (125)I-labeled fibrin clots by t-PA. This effect is dependent on direct binding interactions of vitronectin with fibrin. There is no significant effect of fibrin-associated vitronectin on fibrinolysis in the absence of PAI-1. The binding of PAI-1 to fibrin clots formed in the absence of vitronectin was characterized by a low affinity (K(d) approximately 3.5 micrometer) and rapid loss of PAI-1 inhibitory activity over time. In contrast, a high affinity and stabilization of PAI-1 activity characterized the cooperative binding of PAI-1 to fibrin formed in the presence of vitronectin. These findings indicate that plasma PAI-1.vitronectin complexes can be localized to the surface of fibrin clots; by this localization, they may modulate fibrinolysis and clot reorganization.  相似文献   

3.
Plasminogen activator inhibitor-1 (PAI-1) accumulates within thrombi and forming whole blood clots. To explore this phenomenon at the molecular level, PAI-1 binding to fibrin was examined. The experiments were performed by adding 125I-PAI-1, which retains its complete tissue-type plasminogen (t-PA) inhibitory activity, to fibrin matrices formed in 2-cm2 tissue culture wells. Guanidine HCl-activated PAI-1 binding was reversible and was inhibited in the presence of excess, unlabeled PAI-1. Activated 125I-PAI-1 recognized 2 sites on fibrin: a very small number of high affinity sites (Kd less than 1 nM) and principally a large number of low affinity sites with an approximate Kd of 3.8 microM. Latent PAI-1 bound to fibrin at a site indistinguishable from the lower affinity site recognized by activated PAI-1. Fibrin, pretreated with activated PAI-1, was protected from t-PA-mediated plasmin degradation in a PAI-1 dose-responsive manner (IC50 = 12.3 nM). Clot protection correlated with partial occupancy of the low affinity PAI-1 binding site on fibrin and was due to the formation of sodium dodecyl sulfate-stable, PAI-1.t-PA complexes. Latent PAI-1 (27 nM) did not protect the fibrin from dissolution. The localization of PAI-1 to a thrombus by virtue of its fibrin binding potential could result in significant protection of the thrombus from the degradative effects of the fibrinolytic system.  相似文献   

4.
The dissolution of blood clots by plasmin is normally initiated in vivo by the activation of plasminogen to plasmin through the activity of tissue plasminogen activator (t-PA). The rate of plasminogen activation can be stimulated several orders of magnitude by the presence of fibrin-related proteins. Here we describe the kinetic analysis of both recombinant human t-PA (wild-type) and a t-PA variant produced by site-directed mutagenesis in which the original sequence from amino acids 296 to 299, KHRR, has been altered to AAAA. This tetra-alanine variant form of t-PA, K296A/H297A/R298A/R299A t-PA, we refer to as "KHRR" t-PA here. The plasminogen activating kinetics of wild-type t-PA (Activase alteplase) showed a catalytic efficiency which changed over 100-fold dependent on the stimulator in the assay. The lowest rate was in the absence of a stimulator. The following stimulators showed increasing ability to accelerate the catalytic efficiency of the reaction: fibrinogen, fragments of fibrinogen obtained by digestion with plasmin, fibrin, and slightly degraded fibrin. This increase in efficiency was driven primarily by decreases in the Michaelis constant (KM) of the reaction, whereas the catalytic rate constant (kcat) of the reaction did not change significantly. The "KHRR" variant of t-PA displayed novel kinetics with all stimulators tested. In the absence of a stimulator or with the poorer stimulators (fibrinogen and fibrinogen fragments), the KM values of the reaction with Activase alteplase and "KHRR" t-PA were similar. The kcat however, was lower with "KHRR" t-PA than with wild-type t-PA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
In this study, we identified lysine residues in the fibrinogen Aalpha chain that serve as substrates during transglutaminase (TG)-mediated cross-linking of plasminogen activator inhibitor 2 (PAI-2). Comparisons were made with alpha(2)-antiplasmin (alpha(2)-AP), which is known to cross-link to lysine 303 of the Aalpha chain. A 30-residue peptide containing Lys-303 specifically competed with fibrinogen for cross-linking to alpha(2)-AP but not for cross-linking to PAI-2. Further evidence that PAI-2 did not cross-link via Lys-303 was the cross-linking of PAI-2 to I-9 and des-alphaC fibrinogens, which lack 100 and 390 amino acids from the C terminus of the Aalpha chain, respectively. PAI-2 or alpha(2)-AP was cross-linked to fibrinogen and digested with trypsin or endopeptidase Glu-C, and the resulting peptides analyzed by mass spectrometry. Peptides detected were consistent with tissue TG (tTG)-mediated cross-linking of PAI-2 to lysines 148, 176, 183, 457 and factor XIIIa-mediated cross-linking of PAI-2 to lysines 148, 230, and 413 in the Aalpha chain. alpha(2)-AP was cross-linked only to lysine 303. Cross-linking of PAI-2 to fibrinogen did not compete with alpha(2)-AP, and the two proteins utilized different lysines in the Aalpha chain. Therefore, PAI-2 and alpha(2)-AP can cross-link simultaneously to the alpha polymers of a fibrin clot and promote resistance to lysis.  相似文献   

6.
A complex of d-dimer noncovalently associated with fragment E ((DD)E), a degradation product of cross-linked fibrin that binds tissue plasminogen activator (t-PA) and plasminogen (Pg) with affinities similar to those of fibrin, compromises the fibrin specificity of t-PA by stimulating systemic Pg activation. In this study, we examined the effect of thrombin-activable fibrinolysis inhibitor (TAFI), a latent carboxypeptidase B (CPB)-like enzyme, on the stimulatory activity of (DD)E. Incubation of (DD)E with activated TAFI (TAFIa) or CPB (a) produces a 96% reduction in the capacity of (DD)E to stimulate t-PA-mediated activation of Glu- or Lys-Pg by reducing k(cat) and increasing K(m) for the reaction; (b) induces the release of 8 mol of lysine/mol of (DD)E, although most of the stimulatory activity is lost after release of only 4 mol of lysine/mol (DD)E; and (c) reduces the affinity of (DD)E for Glu-Pg, Lys-Pg, and t-PA by 2-, 4-, and 160-fold, respectively. Because TAFIa- or CPB-exposed (DD)E produces little stimulation of Glu-Pg activation by t-PA, (DD)E is not degraded into fragment E and d-dimer, the latter of which has been reported to impair fibrin polymerization. These data suggest a novel role for TAFIa. By attenuating systemic Pg activation by (DD)E, TAFIa renders t-PA more fibrin-specific.  相似文献   

7.
TNK-tissue plasminogen activator (TNK-t-PA), a bioengineered variant of tissue-type plasminogen activator (t-PA), has a longer half-life than t-PA because the glycosylation site at amino acid 117 (N117Q, abbreviated N) has been shifted to amino acid 103 (T103N, abbreviated T) and is resistant to inactivation by plasminogen activator inhibitor 1 because of a tetra-alanine substitution in the protease domain (K296A/H297A/R298A/R299A, abbreviated K). TNK-t-PA is more fibrin-specific than t-PA for reasons that are poorly understood. Previously, we demonstrated that the fibrin specificity of t-PA is compromised because t-PA binds to (DD)E, the major degradation product of cross-linked fibrin, with an affinity similar to that for fibrin. To investigate the enhanced fibrin specificity of TNK-t-PA, we compared the kinetics of plasminogen activation for t-PA, TNK-, T-, K-, TK-, and NK-t-PA in the presence of fibrin, (DD)E or fibrinogen. Although the activators have similar catalytic efficiencies in the presence of fibrin, the catalytic efficiency of TNK-t-PA is 15-fold lower than that for t-PA in the presence of (DD)E or fibrinogen. The T and K mutations combine to produce this reduction via distinct mechanisms because T-containing variants have a higher K(M), whereas K-containing variants have a lower k(cat) than t-PA. These results are supported by data indicating that T-containing variants bind (DD)E and fibrinogen with lower affinities than t-PA, whereas the K and N mutations have no effect on binding. Reduced efficiency of plasminogen activation in the presence of (DD)E and fibrinogen but equivalent efficiency in the presence of fibrin explain why TNK-t-PA is more fibrin-specific than t-PA.  相似文献   

8.
Elevated levels of lipoprotein(a) [Lp(a)] are associated with an increased risk of atherothrombotic disease, but the mechanism(s) by which Lp(a) potentiates atherogenesis is unknown. The extensive homology of apolipoprotein(a) [apo(a)] to plasminogen has led us and others to postulate that Lp(a) may impair fibrinolysis. We have previously shown that Lp(a) inhibits fibrin stimulation of plasminogen activation by tissue-type plasminogen activator (t-PA); however, we and other investigators have been unable to demonstrate direct inhibition of t-PA by Lp(a) in solution. We now report that t-PA binds reversibly and saturably to surface-bound Lp(a) and to low-density lipoprotein (LDL) and that as a result of this binding activation of plasminogen by t-PA is inhibited. The catalytic efficiency (kcat/Km) of t-PA when bound to polystyrene surface-bound fibrinogen increased 2.9-fold compared to t-PA bound to control wells. When bound to surface-bound Lp(a), however, the catalytic efficiency of t-PA was reduced 9.5-fold compared to t-PA bound to control wells; likewise, by binding to surface-bound LDL, the catalytic efficiency of t-PA was reduced 16-fold compared to the control. Studies with defined monoclonal antibodies suggest that major determinants of t-PA binding are its active site, the LDL receptor binding domain of apolipoprotein B-100 (apoB-100), and apo(a). These data suggest a unique mechanism by which Lp(a) and LDL incorporated in an atheroma can inhibit endogenous fibrinolysis and thereby contribute to the genesis of atherothrombotic disease.  相似文献   

9.
This study examined the effect of the preovulatory gonadotropin surge on the temporal and spatial regulation of tissue plasminogen activator (tPA), urokinase plasminogen activator (uPA), and uPA receptor (uPAR) mRNA expression and tPA, uPA, and plasmin activity in bovine preovulatory follicles and new corpora lutea collected at approximately 0, 6, 12, 18, 24, and 48 h after a GnRH-induced gonadotropin surge. Messenger RNAs for tPA, uPA, and uPAR were increased in a temporally specific fashion within 24 h of the gonadotropin surge. Localization of tPA mRNA was primarily to the granulosal layer, whereas both uPA and uPAR mRNAs were detected in both the granulosal and thecal layers and adjacent ovarian stroma. Activity for tPA was increased in follicular fluid and the preovulatory follicle apex and base within 12 h after the gonadotropin surge. The increase in tPA activity in the follicle base was transient, whereas the increased activity in the apex was maintained through the 24 h time point. Activity for uPA increased in the follicle apex and base within 12 h of the gonadotropin surge and remained elevated. Plasmin activity in follicular fluid also increased within 12 h after the preovulatory gonadotropin surge and was greatest at 24 h. Our results indicate that mRNA expression and enzyme activity for both tPA and uPA are increased in a temporally and spatially specific manner in bovine preovulatory follicles after exposure to a gonadotropin surge. Increased plasminogen activator and plasmin activity may be a contributing factor in the mechanisms of follicular rupture in cattle.  相似文献   

10.
BackgroundMechanisms of fibrin-specificity of tissue plasminogen activator (tPA) and recombinant staphylokinase (STA) are different, therefore we studied in vitro the possibility of the synergy of their combined thrombolytic action.MethodsThrombolytic effects of tPA, STA and their combinations were measured by lysis rate of human plasma clot and side effects were evaluated by decreasing in fibrinogen, plasminogen and α2-antiplasmin levels in the surrounding plasma at 37 °C in vitro.ResultsSTA and tPA induced dose- and time-dependent clot lysis: 50% lysis in 2 h was obtained with 30 nM tPA and 75 nM STA, respectively. At these concentrations, tPA produced greater degradation of plasma fibrinogen than STA. According to a mathematical analysis of dose–response curves by the isobole method, combinations of tPA and STA caused a considerable synergistic thrombolytic effect. The simultaneous and sequential combinations of tPA (< 4 nM) and STA (< 35 nM) induced a significant fibrin-specific synergistic thrombolysis, which was more pronounced in 2 h at simultaneous combinations than at sequential addition of STA after 30 min of tPA action. Simultaneous combination of 2.5 nM tPA and 15 nM STA showed a maximal 3-fold increase in thrombolytic effect compared to the expected total effect of the individual agents. Sequential combinations caused a lower depletion of plasma proteins compared to simultaneous combinations.ConclusionsThe simultaneous and sequential combinations of tPA and STA possessed synergistic fibrin-specific thrombolytic action on clot lysis in vitro.General significanceThe results show that combined thrombolysis may be more effective and safer than thrombolysis with each activator alone.  相似文献   

11.
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   

12.
13.
Vitronectin immobilized onto polystyrene microtiter wells was demonstrated to specifically bind plasminogen in a concentration-dependent manner, yielding an estimated KD = 0.4 microM. Heparin only moderately interfered with the vitronectin-plasminogen interaction, whereas high concentrations of 6-amino-hexanoic acid inhibited binding. Utilizing a ligand-blotting procedure in which plasminogen was reacted with proteolytic fragments of vitronectin, transblotted onto nitrocellulose, the plasminogen-binding site of vitronectin was localized to the heparin-binding domain of the adhesive protein. Moreover, vitronectin was found to inhibit in a dose-dependent fashion the fibrin(ogen)-induced activation of plasminogen by tissue plasminogen activator. These results provide the first evidence for a novel vitronectin-mediated control of plasminogen activation potentially relevant for directional clot-lysis and plasmin-dependent proteolysis in extracellular matrices.  相似文献   

14.
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   

15.
Placental microvillous membranes exhibited saturable binding of urokinase-type plasminogen activator with plateau achieved by 30 min at 4 degrees C and 10 min at 37 degrees C. The binding was essentially irreversible. The capacity was about 8 pmol urokinase per mg membrane protein. Half-maximal displacement of 125I-labelled urokinase was achieved with about 1.0 nM unlabelled urokinase when using 75 micrograms membrane protein/ml. 125I-labelled urokinase did not bind when treated with diisopropylfluorophosphate to block the catalytic activity. Single-chain urokinase (prourokinase), devoid of catalytic activity, did not bind. Catalytically active tissue-type plasminogen activator did compete with 125I-labelled urokinase for binding although less efficiently than urokinase. Binding activity remained in the 100,000 x g pellet after treatment of the membranes with 3 M KCl, alkaline stripping at pH 12 or extraction by the detergent Triton X-100. The binding was essentially blocked by antibodies against plasminogen activator inhibitor-type-2 (PAI-2). Sodium dodecyl sulfate polyacrylamide gel electrophoresis of solubilized membranes with bound 125I-labelled urokinase showed that the urokinase-PAI-2 complexes largely migrated in fractions corresponding to a very large Mr although no clearly defined peaks were observed. It is suggested that PAI-2 occurs in a form anchored to syncytiotrophoblast microvilli, possibly to the cytoskeleton.  相似文献   

16.
1. Serum-free conditioned medium from L-cells or L-cells treated with the tumor-promotor phorbol myristate acetate (PMA) was analyzed for plasminogen activator (PA) and plasminogen activator inhibitor (PAI) activity. Conditioned medium from control or PMA-treated cells did not contain detectable PA activity when assayed by SDS-PAGE and zymography. 2. Conditioned medium from PMA-treated cells, but not control cells, contained a PAI of Mr = 40,000 da when assayed by reverse zymography. 3. The L-cell PAI formed SDS-stable complexes with purified human (homo sapiens) urokinase and tissue plasminogen activator, as well as, mouse (Mus musculus) urinary PA. 4. These results indicate that biochemical and immunological differences between human and mouse urokinase and human urokinase and human tissue plasminogen activator do not influence the interaction of the L-cell PAI with these enzymes.  相似文献   

17.
Maspin is a member of the serpin family with a reactive center loop that is incompatible with proteinase inhibition by the serpin conformational change mechanism. Despite this there are reports that maspin might regulate uPA-dependent processes in vivo. Using exogenous and endogenous fluorescence, we demonstrate here that maspin can bind uPA and tPA in both single-chain and double-chain forms, with K(d) values between 300 and 600 nM. Binding is at an exosite on maspin close to, but outside of, the reactive center loop and is therefore insensitive to mutation of Arg(340) within the reactive center loop. The binding site on tPA does not involve the proteinase active site, with the result that maspin can bind to S195A tPA that is already complexed to plasminogen activator inhibitor-1. The ability of maspin to bind these proteinases without involvement of the reactive center loop leaves the latter free to engage in additional, as yet unidentified, maspin-protein interactions that may serve to regulate the properties of the exosite-bound proteinase. This may help to reconcile apparently conflicting studies that demonstrate the importance of the reactive center loop in certain maspin functions, despite the inability of maspin to directly inhibit tPA or uPA catalytic activity in in vitro assays through engagement between its reactive center loop and the active site of the proteinase.  相似文献   

18.
The interaction between type 1 plasminogen activator inhibitor (PAI-1) and fragments of vitronectin (Vn) was investigated. The PAI-1-binding domain was not destroyed when Vn was cleaved by treatment with either acid or CNBr. Acid-cleaved Vn was fractionated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and analyzed by PAI-1 ligand binding. The smallest fragment (Mr 40,000) that retained PAI-1 binding function was sequenced and shown to contain the NH2 terminus of the molecule. Further cleavage of this fragment by treatment with CNBr generated a Mr 35,000 fragment (Pro52-Asp239) that did not interact with PAI-1, and a Mr 6,000 NH2-terminal fragment (Asp1-Met51) that spanned the somatomedin B domain and contained the RGD (cell binding) sequence. The purified Mr 6,000 fragment competed with immobilized Vn for PAI-1 binding, and formed complexes with activated PAI-1. These complexes could be immunoprecipitated by antibodies to PAI-1. Synthetic peptides containing the RGD sequence had no effect on the binding of this fragment to PAI-1. These results suggest that the cell-binding and PAI-1 binding sequences of Vn occupy distinct regions in the NH2-terminal somatomedin B domain of the molecule.  相似文献   

19.
Solid phase synthesis of the hexapeptide, GPRVVE, which represents the amino terminal six amino acids of the alpha-chain of human fibrin, yielded a product that contained a modified glutamic acid. The nature of the modification was established as the Friedel-Crafts acylation product of the peptide and anisole, the latter reagent employed in the HF deblocking step. The anisoylated peptide selectively enhanced the activation rate of native [Glu1]plasminogen by recombinant tissue plasminogen activator, accelerated clot lysis, and retarded the polymerization of nascent fibrin.  相似文献   

20.
V Gurewich 《Enzyme》1988,40(2-3):97-108
Single chain urokinase (SC-UK) is a precursor of 55 kd two-chain UK (TC-UK). Treatment with catalytic proportions of plasmin or kallikrein converts SC-UK to TC-UK as a consequence of cleavage of its Lys158-Ile159 peptide bond. This plasmin-mediated activation of SC-UK induces a positive feedback secondary reaction and complicates measurement of its activity against its natural substrate, Glu-plasminogen. The fibrin-selective effect of pro-UK-induced clot lysis is not related to fibrin binding. Rather, a conformational change in Glu-plasminogen, conferred when it binds to certain carboxy-terminal lysine residues on fibrin, has been implicated in this mechanism. This is complementary to t-PA. Fibrin-bound t-PA was found to exclusively activate plasminogen bound to certain internal lysine residues. Their complementariness is believed to explain their synergism in fibrinolysis.  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