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 human platelets by alpha-thrombin is mediated at least in part by cleavage of protease-activated G-protein-coupled receptors, PAR-1 and PAR-4. Platelet glycoprotein Ibalpha also has a high affinity binding site for alpha-thrombin, and this interaction contributes to platelet activation through a still unknown mechanism. In the present study the hypothesis that GpIbalpha may contribute to platelet activation by modulating the hydrolysis of PAR-1 on the platelet membrane was investigated. Gel-filtered platelets from normal individuals were stimulated by alpha-thrombin, and the kinetics of PAR-1 hydrolysis by enzyme was followed with flow cytometry using an anti-PAR-1 monoclonal antibody (SPAN 12) that recognizes only intact PAR-1 molecules. This strategy allowed measurement of the apparent k(cat)/K(m) value for thrombin hydrolysis of PAR-1 on intact platelets, which was equal to 1.5 +/- 0.1 x 10(7) m(-1) sec(-1). The hydrolysis rate of PAR-1 by thrombin was measured under conditions in which thrombin binding to GpIb was inhibited by different strategies, with the following results. 1) Elimination of GpIbalpha on platelet membranes by mocarhagin treatment reduced the k(cat)/K(m) value by about 6-fold. 2) A monoclonal anti-GpIb antibody reduced the apparent k(cat)/K(m) value by about 5-fold. 3) An oligonucleotide DNA aptamer, HD22, which binds to the thrombin heparin-binding site (HBS) and inhibits thrombin interaction with GpIbalpha, reduced the apparent k(cat)/K(m) value by about 5-fold. 4) Displacement of alpha-thrombin from the binding site on GpIb using PPACK-thrombin reduced the apparent k(cat)/K(m) value by about 5-fold, and 5) mutation at the HBS of thrombin (R98A) caused a 5-fold reduction of the apparent k(cat)/K(m) value of PAR-1 hydrolysis. Altogether these results show that thrombin interaction with GpIb enhances the specificity of thrombin cleavage of PAR-1 on intact platelets, suggesting that GpIb may function as a "cofactor" for PAR-1 activation by thrombin.  相似文献   

2.
Binding of ADP to platelets enhances the binding of fibrinogen to Gp IIb-IIIa, the specific platelet receptor for adhesive proteins. The linkage between ADP and fibrinogen binding is indirect since ADP does not bind to the same receptor as fibrinogen. We have recently proposed that a third component, once affected by ADP binding, induces a conformational transition of the fibrinogen receptor from the low to the high affinity state, which is responsible for platelet aggregation [De Cristofaro, R., Landolfi, R., Castagnola, M., De Candia, E., Di Cera, E., & Wyman, J. (1988) Proc. Natl. Acad. Sci. USA 85, 8473-8476]. In the present study we provide evidence that this component should be identified with the platelet Na+/H+ antiport. Inhibition of the antiport by pharmacological agents such as amiloride, or else by decreasing extracellular Na+, results in a marked decrease of fibrinogen binding to platelets.  相似文献   

3.
The activation of human platelets by alpha-thrombin is mediated in part by cleavage of the protease-activated receptor (PAR) 1 and 4 and by the glycoprotein (Ib)alpha, (Gp(Ib)alpha), which binds with high affinity to alpha-thrombin. Recent studies have shown that the thrombin domain referred to as heparin binding site (HBS) is involved in the interaction with the platelet Gp(Ib)alpha. The HBS is rich in basic amino acids. To identify the key amino acid residues involved in the binding to Gp(Ib)alpha, we have performed alanine scanning mutagenesis of the basic HBS R93, R97, R101, R233, K236, K240, R233/K236/Q239, as well as of the neutral Q239 residues, located in different regions of the domain. For comparison, mutation at R67 within the fibrinogen recognition site (FRS) of thrombin was performed as well. Solid-phase binding experiments showed that the Kd of thrombin-GpIb interaction was reduced 22-fold for R93A, 8-fold for R97A, 13-fold for R101A, 29-fold for R233A, 21-fold for K236A, 5-fold for K240A, and 31-fold for the triple mutant R233A/K236A/Q239A, while the Q239A and R67A forms did not show any significant affinity change. The platelet activating capacity of these mutants was evaluated as well. Using gel-filtered platelets, the EC50 value of thrombin-induced aggregation was from 5- to 13-fold higher in the HBS mutants than in the WT form, and was linearly and positively correlated with the corresponding Kd values pertaining to thrombin binding to GpIb. Measurements of PAR-1 hydrolysis on the platelet membrane showed that the HBS mutants R233A, R101A, R93A, K236A, and R233/K236/Q239 forms had a reduction of the apparent kcat/Km value. These results are a consequence of a defective binding to GpIb, which is known to optimize the interaction with PAR-1 in situ. A confirm of this hypothesis came from the demonstration that the kcat/Km value pertaining to the hydrolysis by the HBS-mutated thrombins of the synthetic PAR-1 38-60 peptide in solution was similar to that one obtained with the WT form. In conclusion, these experiments provide a structural and functional mapping of the thrombin HBS subregions involved in the binding to the platelet Gp(Ib)alpha and in the cell activation.  相似文献   

4.
Activation of platelets by the serine protease thrombin is a critical event in haemostasis. This process involves the binding of thrombin to glycoprotein Ibα (GpIbα) and cleavage of protease-activated receptors (PARs). The N-terminal extracellular domain of GpIbα contains an acidic peptide stretch that has been identified as the main thrombin binding site, and both anion binding exosites of thrombin have been implicated in GpIbα binding, but it remains unclear how they are involved. This issue is of critical importance for the mechanism of platelet activation by thrombin. If both exosites bind to GpIbα, thrombin could potentially act as a platelet adhesion molecule or receptor dimerisation trigger. Alternatively, if only a single site is involved, GpIbα may serve as a cofactor for PAR-1 activation by thrombin. To determine the involvement of thrombin's two exosites in GpIbα binding, we employed the complementary methods of mutational analysis, binding studies, X-ray crystallography and NMR spectroscopy. Our results indicate that the peptide corresponding to the C-terminal portion of GpIbα and the entire extracellular domain bind exclusively to thrombin's exosite II. The interaction of thrombin with GpIbα thus serves to recruit thrombin activity to the platelet surface while leaving exosite I free for PAR-1 recognition.  相似文献   

5.
The expression of the elongated fibrinogen gamma chain, termed gamma', derives from alternative splicing of mRNA and causes an insertion sequence of 20 amino acids. This insertion domain interacts with the anion-binding exosite (ABE)-II of thrombin. This study investigated whether and how gamma' chain binding to ABE-II affects thrombin interaction with its platelet receptors, i.e. glycoprotein Ibalpha (GpIbalpha), protease-activated receptor (PAR) 1, and PAR4. Both synthetic gamma' peptide and fibrinogen fragment D*, containing the elongated gamma' chain, inhibited thrombin-induced platelet aggregation up to 70%, with IC(50) values of 42+/-3.5 and 0.47+/-0.03 microm, respectively. Solid-phase binding and spectrofluorimetric assays showed that both fragment D* and the synthetic gamma' peptide specifically bind to thrombin ABE-II and competitively inhibit the thrombin binding to GpIbalpha with a mean K(i) approximately 0.5 and approximately 35 microm, respectively. Both these gamma' chain-containing ligands allosterically inhibited thrombin cleavage of a synthetic PAR1 peptide, of native PAR1 molecules on intact platelets, and of the synthetic chromogenic peptide D-Phe-pipecolyl-Arg-p-nitroanilide. PAR4 cleavage was unaffected. In summary, fibrinogen gamma' chain binds with high affinity to thrombin and inhibits with combined mechanisms the platelet response to thrombin. Thus, its variations in vivo may affect the hemostatic balance in arterial circulation.  相似文献   

6.
The present study has shown that calcium inhibits the heparin-catalyzed antithrombin III/thrombin reaction. The initial rate of thrombin (4.0 nM) inhibition by antithrombin III (200 nM) in the presence of heparin (2.5 ng/ml) decreased from 3.6 nM/min (in the absence of calcium) to 0.12 nM/min in the presence of 10 mM calcium. In the absence of heparin, the initial rate of thrombin inhibition by antithrombin III was not affected by calcium. The heparin-catalyzed antithrombin III/thrombin reaction is described by the general rate equation for a random-order, bireactant, enzyme-catalyzed reaction (M. J. Griffith (1982) J. Biol. Chem. 257, 13899-13902). As such, the reaction is saturable with respect to both thrombin and antithrombin III. The apparent kinetic parameters for the heparin-catalyzed antithrombin III/thrombin reaction were determined in the presence and absence of calcium. The apparent heparin/antithrombin III dissociation constant values were not measurably different in the presence of 0, 1.0, and 3.0 mM calcium. The apparent heparin/thrombin dissociation constant value increased from 7.0 nM, in the absence of calcium, to 10 and 30 nM in the presence of 1.0 and 3.0 mM calcium, respectively. The maximum reaction velocity, at saturation with respect to both proteins, was not affected by calcium. It is concluded that calcium binds to functional groups within the heparin molecule which are required for thrombin binding.  相似文献   

7.
Heparin binding to rabbit histidine-rich glycoprotein (HRG) was studied in a purified system, allowing for determination of a heparin dissociation constant of approximately 5.5 X 10(-8) M for the interaction with HRG at pH 7.0. The strong interaction between heparin and HRG was demonstrated to be competitive with the binding of both antithrombin and thrombin to the heparin chain. HRG was further tested as a modulator of the anticoagulant activity of heparin by comparing rates of the heparin-catalyzed reaction between antithrombin and thrombin in the presence and absence of added HRG. The heparin-antithrombin-thrombin reaction was modeled using the formalism of a two-substrate enzyme-catalyzed reaction with heparin as the enzyme and HRG analyzed as an enzyme inhibitor. HRG was shown to compete with both antithrombin and thrombin for binding to heparin by this kinetic analysis. Thus, both the kinetic and heparin-binding data indicate that the mechanism by which HRG modulates heparin anticoagulant activity involves competition for heparin with both the inhibitor and the protease. Inhibition by HRG of the heparin-catalyzed reaction was found to be highly dependent on pH, with a sharp increase in inhibition from about 15% to greater than 90% observed as pH was lowered from 7.4 to 7.0. Since little change in the rate of the heparin-catalyzed inhibition of thrombin by antithrombin occurs in this pH region, the dramatic change in HRG inhibition seen upon pH titration may reflect increasing ionic interaction between heparin and HRG due to the protonation of histidine residues which occurs in this pH region.  相似文献   

8.
A R Rezaie 《Biochemistry》1999,38(44):14592-14599
The cofactors heparin, vitronectin (VN), and thrombomodulin (TM) modulate the reactivity of alpha-thrombin with plasminogen activator inhibitor (PAI-1). While heparin and VN accelerate the reaction by approximately 2 orders of magnitude, TM protects alpha-thrombin from rapid inactivation by PAI-1 in the presence of VN. To understand how these cofactors function, we studied the kinetics of PAI-1 inactivation of alpha-thrombin, the exosite 1 variant gamma-thrombin, the exosite 2 mutant R93,97,101A thrombin, and recombinant meizothrombin in both the absence and presence of these cofactors. Heparin and VN accelerated the second-order association rate constant [k(2) = (7.9 +/- 0.5) x 10(2) M(-)(1) s(-)(1)] of alpha-thrombin with PAI-1 approximately 200- and approximately 240-fold, respectively. The k(2) value for gamma-thrombin [(7.9 +/- 0.7) x 10(1) M(-)(1) s(-)(1)] was impaired 10-fold, but was enhanced by heparin and VN approximately 280- and approximately 75-fold, respectively. Similar to inactivation of gamma-thrombin, PAI-1 inactivation of alpha-thrombin in complex with the epidermal growth factor-like domains 4-6 of TM (TM4-6) was impaired approximately 10-fold. The exosite 2 mutant R93,97,101A thrombin, which was previously shown not to bind heparin, and meizothrombin, in which exosite 2 is masked, reacted with PAI-1 at similar rates in both the absence and presence of heparin [k(2) = (1.3-1.5) x 10(3) M(-)(1) s(-)(1) for R93,97,101A thrombin and k(2) = (3.6-5.1) x 10(2) M(-)(1) s(-)(1) for meizothrombin]. Unlike heparin, however, VN enhanced the k(2) of R93,97,101A thrombin and meizothrombin inactivation approximately 80- and approximately 30-fold, respectively. Continuous kinetic analysis as well as competition kinetic studies in the presence of S195A thrombin suggested that the accelerating effect of VN or heparin occurs primarily by lowering the dissociation constant (K(d)) for formation of a noncovalent, Michaelis-type complex. Analysis of these results suggest that (1) heparin binds to exosite 2 of alpha-thrombin to accelerate the reaction by a template mechanism, (2) VN accelerates PAI-1 inactivation of alpha-thrombin by lowering the K(d) for initial complex formation by an unknown mechanism that does not require binding to either exosite 1 or exosite 2 of alpha-thrombin, (3) alpha-thrombin may have a binding site for PAI-1 within or near exosite 1, and (4) TM occupancy of exosite 1 partially accounts for the protection of thrombin from rapid inactivation by PAI-1 in the presence of vitronectin.  相似文献   

9.
Factor XI (FXI) binds specifically and reversibly to high affinity sites on the surface of stimulated platelets (Kd app of approximately 10 nm; Bmax of approximately 1,500 sites/platelet) utilizing residues exposed on the Apple 3 domain in the presence of high molecular weight kininogen and Zn2+ or prothrombin and Ca2+. Because the FXI receptor in the platelet membrane is contained within the glycoprotein Ibalpha subunit of the glycoprotein Ib-IX-V complex (Baglia, F. A., Badellino, K. O., Li, C. Q., Lopez, J. A., and Walsh, P. N. (2002) J. Biol. Chem. 277, 1662-1668), we utilized mocarhagin, a cobra venom metalloproteinase, to generate a fragment (His1-Glu282) of glycoprotein Ibalpha that contains the leucine-rich repeats of the NH2-terminal globular domain and excludes the macroglycopeptide portion of glycocalicin, the soluble extracytoplasmic portion of glycoprotein Ibalpha. This fragment was able to compete with FXI for binding to activated platelets (Ki of 3.125 +/- 0.25 nm) with a potency similar to that of intact glycocalicin (Ki of 3.72 +/- 0.30 nm). However, a synthetic glycoprotein Ibalpha peptide, Asp269-Asp287, containing a thrombin binding site had no effect on the binding of FXI to activated platelets. Moreover, the binding of 125I-labeled thrombin to glycocalicin was unaffected by the presence of FXI at concentrations up to 10(-5) m. The von Willebrand factor A1 domain, which binds the leucine-rich repeats, inhibited the binding of FXI to activated platelets. Thus, we examined the effect of synthetic peptides of each of the seven leucine-rich repeats on the binding of 125I-FXI to activated platelets. All leucine-rich repeat (LRR) peptides derived from glycoprotein Ibalpha were able to inhibit FXI binding to activated platelets in the following order of decreasing potency: LRR7, LRR1, LRR4, LRR5, LRR6, LRR3, and LRR2. However, the leucine-rich repeat synthetic peptides derived from glycoprotein Ibbeta and Toll protein had no effect. We conclude that FXI binds to glycoprotein Ibalpha at sites comprising the leucine-rich repeat sequences within the NH2-terminal globular domain that are separate and distinct from the thrombin-binding site.  相似文献   

10.
Calcium spirulan (Ca-SP), a novel sulfated polysaccharide, increases the rate of thrombin inhibition by heparin cofactor II (HCII) more than 1000-fold through a mechanism not requiring the amino-terminal acidic domain of HCII. Activation of HCII by Ca-SP was molecular-weight dependent. Furthermore, HD22, an aptamer that binds exosite II of thrombin, produced a concentration-dependent, 15-fold reduction at 5 microM in the rate of thrombin inhibition by HCII with Ca-SP, suggesting that Ca-SP interacts with exosite II of thrombin. Mutations of Lys173 to Leu (K173L) and Arg189 to Leu (R189L) in the HCII molecule resulted in large decreases in the rate of thrombin inhibition mediated by Ca-SP and in the NaCl concentration needed for elution from Ca-SP-Toyopearl. Mutations of Lys173 to Arg (K173R) and Arg189 to Lys (R189K) showed inhibition of thrombin similar to wild-type rHCII (wt-rHCII). These results indicate that Ca-SP binds to the positive charges of Lys173 and Arg189 on the HCII molecule. In the thrombin inhibitory process by HCII, Ca-SP appears to play as a template by binding to both thrombin and HCII.  相似文献   

11.
The active site of the prothrombin activation intermediate meizothrombin(desF1) was probed using several fluorosulfonylphenyl spin labels specific for the active serine hydroxyl of serine proteases. The mobilities of the thrombin species inhibited with the nitroxide spin labelsm-IV [4-(2,2,6,6-tetramethyl-piperidine-1-oxyl)-m-(fluorosulfonyl)benzamide] andm-V [3-(2,2,5,5-tetramethyl-pyrrolidine-1-oxyl)-m-(fluorosulfonyl)benzamide], which are sensitive to differences betweenα- andγ-thrombin, were quite similar to that ofα-thrombin. That is, no major conformational differences between meizothrombin(desF1) andα-thrombin were observed in this region of the extended active site. On the other hand,p-IV [4-(2,2,6,6-tetramethyl piperidine-1-oxyl)-p-(fluorosulfonyl)benzamide],p-V [3-(2,2,5,5-tetramethylpyrrolidine-1-oxyl)-p-(fluorosulfonyl)benzamide], andm-VII [N-[m-(fluorosulfonyl)phenyl]-4-N-(2,2,6,6-tetramethyl-piperidine-1-oxyl)urea], which probe an apolar binding region of bovine thrombin, exhibited large differences in mobility betweenα-thrombin and meizothrombin(desF1). The conformational consequences of indole binding to spin-labeled thrombin species demonstrated that both species also possess an indole-binding site. However, the nitroxide mobility changes upon indole binding to the spin-labeled protein derivative were somewhat different between the two thrombin species under study. In addition, the effects of the benzamidine binding were quite similar for each labeled protein. Thus is appears that, while both species posses a fully functional active site, the region in meizothrombin(desF1) probed by spin labelsp-IV,p-V, andm-VII, which corresponds to the apolar binding region, differs in conformation fromα-thrombin.  相似文献   

12.
The glycoprotein (GP) Ib-IX complex is a platelet surface receptor that binds thrombin as one of its ligands, although the biological significance of thrombin interaction remains unclear. In this study we have used several approaches to investigate the GPIb alpha-thrombin interaction in more detail and to study its effect on the thrombin-induced elaboration of fibrin. We found that both glycocalicin and the amino-terminal fragment of GPIb alpha reduced the release of fibrinopeptide A from fibrinogen by about 50% by a noncompetitive allosteric mechanism. Similarly, GPIb alpha caused in thrombin an allosteric reduction in the rate of turnover of the small peptide substrate d-Phe-Pro-Arg-pNA. The K(d) for the glycocalicin-thrombin interaction was 1 microm at physiological ionic strength but was highly salt-dependent, decreasing to 0.19 microm at 100 mm NaCl (Gamma(salt) = -4.2). The salt dependence was characteristic of other thrombin ligands that bind to exosite II of this enzyme, and we confirmed this as the GPIb alpha-binding site on thrombin by using thrombin mutants and by competition binding studies. R68E or R70E mutations in exosite I of thrombin had little effect on its interaction with GPIb alpha. Both the allosteric inhibition of fibrinogen turnover caused by GPIb alpha binding to these mutants, and the K(d) values for their interactions with GPIb alpha were similar to those of wild-type thrombin. In contrast, R89E and K248E mutations in exosite II of thrombin markedly increased the K(d) values for the interactions of these thrombin mutants with GPIb alpha by 10- and 25-fold, respectively. Finally, we demonstrated that low molecular weight heparin (which binds to thrombin exosite II) but not hirugen (residues 54-65 of hirudin, which binds to exosite I of thrombin) inhibited thrombin binding to GPIb alpha. These data demonstrate that GPIb alpha binds to thrombin exosite II and in so doing causes a conformational change in the active site of thrombin by an allosteric mechanism that alters the accessibility of both its natural substrate, fibrinogen, and the small peptidyl substrate d-Phe-Pro-Arg-pNA.  相似文献   

13.
Although fibrin-bound thrombin is resistant to inactivation by heparin.antithrombin and heparin.heparin cofactor II complexes, indirect studies in plasma systems suggest that the dermatan sulfate.heparin cofactor II complex can inhibit fibrin-bound thrombin. Herein we demonstrate that fibrin monomer produces a 240-fold decrease in the heparin-catalyzed rate of thrombin inhibition by heparin cofactor II but reduces the dermatan sulfate-catalyzed rate only 3-fold. The protection of fibrin-bound thrombin from inhibition by heparin.heparin cofactor II reflects heparin-mediated bridging of thrombin to fibrin that results in the formation of a ternary heparin.thrombin.fibrin complex. This complex, formed as a result of three binary interactions (thrombin.fibrin, thrombin.heparin, and heparin.fibrin), limits accessibility of heparin-catalyzed inhibitors to thrombin and induces conformational changes at the active site of the enzyme. In contrast, dermatan sulfate binds to thrombin but does not bind to fibrin. Although a ternary dermatan sulfate. thrombin.fibrin complex forms, without dermatan sulfate-mediated bridging of thrombin to fibrin, only two binary interactions exist (thrombin.fibrin and thrombin. dermatan sulfate). Consequently, thrombin remains susceptible to inactivation by heparin cofactor II. This study explains why fibrin-bound thrombin is susceptible to inactivation by heparin cofactor II in the presence of dermatan sulfate but not heparin.  相似文献   

14.
The endothelial cell surface membrane protein thrombomodulin binds thrombin with high affinity and acts as both a cofactor for protein C activation and an inhibitor of fibrinogen hydrolysis. We have previously shown that bovine thrombomodulin is a competitive inhibitor of fibrinogen binding to thrombin but has no effect on thrombin activity toward tripeptide substrates or antithrombin III. Hence, thrombomodulin and fibrinogen may share macromolecular specificity sites on thrombin which are distinct from the active site. In this investigation, we have studied the interaction of thrombin-thrombomodulin with fibrinogen and various thrombin derivatives. We show that fibrinogen is a competitive inhibitor of thrombomodulin binding to thrombin, with a Kis = 10 microM. Thrombin derivatives (bovine (pyridoxal phosphate)4-thrombin and human thrombin Quick I), which bind fibrinogen with much reduced affinity, are shown to also interact with thrombomodulin with greatly reduced affinity. These results are consistent with the hypothesis that thrombomodulin and fibrinogen share macromolecular specificity sites on thrombin.  相似文献   

15.
We investigated the regulation of arachidonic acid liberation catalyzed by group-IV cytosolic phospholipase A2 (cPLA2) in human platelets upon stimulation with thrombin through interaction with protease-activated receptor-1 (PAR-1) or glycoprotein Ib. Leupeptin, a protease inhibitor, completely inhibited thrombin-induced arachidonic acid liberation and Ca2+ mobilization, with inhibition of its protease activity. However, preincubation with thrombin in the presence of leupeptin potentiated Ca2+ ionophore-induced arachidonic acid liberation. The preincubation did not affect the intracellular Ca2+ level or cPLA2 activity in response to ionomycin. Human leukocyte elastase, which cleaves glycoprotein Ib, did not inhibit the enhancement of arachidonic acid liberation by thrombin in the presence of leupeptin. However, the effect of thrombin with leupeptin was abolished by a peptide corresponding to residues 54-65 of hirudin (hirudin peptide), which impairs the binding of thrombin to PAR-1. Furthermore, Phe-Pro-Arg chloromethyl ketone (PPACK)-thrombin, which binds to platelets but has no protease activity, also enhanced Ca2+ ionophore-induced arachidonic acid liberation. In contrast, trypsin with leupeptin did not mimic the effect of thrombin with leupeptin, and furthermore trypsin-induced arachidonic acid liberation was insensitive to hirudin peptide. On the basis of the present results, we suggest that thrombin may accelerate cPLA2-catalyzed arachidonic acid liberation through non-proteolytic action toward PAR-1 but not toward glycoprotein Ib in co-operation with the proteolytic action leading to Ca2+ mobilization.  相似文献   

16.
Trumbo TA  Maurer MC 《Biochemistry》2002,41(8):2859-2868
In blood coagulation, thrombin helps to activate factor XIII by cleaving the activation peptide at the R37-G38 peptide bond. The more easily activated factor XIII V34L has been correlated with protection from myocardial infarction. V34L and V29F factor XIII mutant peptides were designed to further characterize substrate binding to thrombin. HPLC kinetic studies have been carried out on thrombin hydrolysis of FXIII activation peptide (28-41), FXIII (28-41) V34L, FXIII (28-41) V29F, and FXIII (28-41) V29F V34L. The V34L mutations lead to improvements in both K(m) and k(cat) whereas the V29F mutation primarily affects K(m). Interactions of the peptides with thrombin have been monitored by 1D proton line broadening NMR and 2D transferred NOESY studies. The results were compared with previously published X-ray crystal structures of thrombin-bound fibrinogen Aalpha (7-16), thrombin receptor PAR1 (38-60), and factor XIII (28-37). In solution, the (34)VVPR(37) and (34)LVPR(37) segments of the factor XIII activation peptide serve as the major anchor points onto thrombin. The N-terminal segments are proposed to interact transiently with the enzyme surface. Long-range NOEs from FXIII V29 or F29 toward (34)V/LVPR(37) have not been observed by NMR studies. Overall, the kinetic and NMR results suggest that the factor XIII activation peptide binds to thrombin in a manner more similar to the thrombin receptor PAR1 than to fibrinogen Aalpha. The V29 and V34 positions affect, in different ways, the ability of thrombin to effectively hydrolyze the activation peptide. Mutations at these sites may prove useful in controlling factor XIII activation.  相似文献   

17.
Glycoprotein (GP) Ib-IX-V binds von Willebrand factor (VWF), initiating thrombosis at high shear stress. The VWF-A1 domain binds the N-terminal domain of GPIbalpha (His1-Glu282); this region contains seven leucine-rich repeats (LRR) plus N- and C-terminal flanking sequences and an anionic sequence containing three sulfated tyrosines. Our previous analysis of canine/human and human/canine chimeras of GPIbalpha expressed on Chinese hamster ovary (CHO) cells demonstrated that LRR2-4 (Leu60-Glu128) were crucial for GPIbalpha-dependent adhesion to VWF. Paradoxically, co-crystal structures of the GPIbalpha N-terminal domain and GPIbalpha-binding VWF-A1 under static conditions revealed that the LRR2-4 sequence made minimal contact with VWF-A1. To resolve the specific functional role of LRR2-4, we compared wild-type human GPIbalpha with human GPIbalpha containing a homology domain swap of canine for human sequence within Leu60-Glu128 and a reverse swap (canine GPIbalpha with human Leu60-Glu128) for the ability to support adhesion to VWF under flow. Binding of conformation-specific anti-GPIbalpha antibodies and VWF binding in the presence of botrocetin (which does not discriminate between species) confirmed equivalent expression of wild-type and mutant receptors in a functional form competent to bind ligand. Compared with CHO cells expressing wild-type GPIbalpha, cells expressing GPIbalpha, where human Leu60-Glu128 sequence was replaced by canine sequence, supported adhesion to VWF at low shear rates but became increasingly ineffective as shear increased from 50 to 2000 s(-1). Together, these data demonstrate that LRR2-4, encompassing a pronounced negative charge patch on human GPIbalpha, is essential for GPIbalpha.VWF-dependent adhesion as hydrodynamic shear increases.  相似文献   

18.
Assembly of ternary thrombin-heparin-fibrin complexes, formed when fibrin binds to exosite 1 on thrombin and fibrin-bound heparin binds to exosite 2, produces a 58- and 247-fold reduction in the heparin-catalyzed rate of thrombin inhibition by antithrombin and heparin cofactor II, respectively. The greater reduction for heparin cofactor II reflects its requirement for access to exosite 1 during the inhibitory process. Protection from inhibition by antithrombin and heparin cofactor II requires ligation of both exosites 1 and 2 because minimal protection is seen when exosite 1 variants (gamma-thrombin and thrombin Quick 1) or an exosite 2 variant (Arg93 --> Ala, Arg97 --> Ala, and Arg101 --> Ala thrombin) is substituted for thrombin. Likewise, the rate of thrombin inhibition by the heparin-independent inhibitor, alpha1-antitrypsin Met358 --> Arg, is decreased less than 2-fold in the presence of soluble fibrin and heparin. In contrast, thrombin is protected from inhibition by a covalent antithrombin-heparin complex, suggesting that access of heparin to exosite 2 of thrombin is hampered when ternary complex formation occurs. These results reveal the importance of exosites 1 and 2 of thrombin in assembly of the ternary complex and the subsequent protection of thrombin from inhibition by heparin-catalyzed inhibitors.  相似文献   

19.
    
The active site of the prothrombin activation intermediate meizothrombin(desF1) was probed using several fluorosulfonylphenyl spin labels specific for the active serine hydroxyl of serine proteases. The mobilities of the thrombin species inhibited with the nitroxide spin labelsm-IV [4-(2,2,6,6-tetramethyl-piperidine-1-oxyl)-m-(fluorosulfonyl)benzamide] andm-V [3-(2,2,5,5-tetramethyl-pyrrolidine-1-oxyl)-m-(fluorosulfonyl)benzamide], which are sensitive to differences between- and-thrombin, were quite similar to that of-thrombin. That is, no major conformational differences between meizothrombin(desF1) and-thrombin were observed in this region of the extended active site. On the other hand,p-IV [4-(2,2,6,6-tetramethyl piperidine-1-oxyl)-p-(fluorosulfonyl)benzamide],p-V [3-(2,2,5,5-tetramethylpyrrolidine-1-oxyl)-p-(fluorosulfonyl)benzamide], andm-VII [N-[m-(fluorosulfonyl)phenyl]-4-N-(2,2,6,6-tetramethyl-piperidine-1-oxyl)urea], which probe an apolar binding region of bovine thrombin, exhibited large differences in mobility between-thrombin and meizothrombin(desF1). The conformational consequences of indole binding to spin-labeled thrombin species demonstrated that both species also possess an indole-binding site. However, the nitroxide mobility changes upon indole binding to the spin-labeled protein derivative were somewhat different between the two thrombin species under study. In addition, the effects of the benzamidine binding were quite similar for each labeled protein. Thus is appears that, while both species posses a fully functional active site, the region in meizothrombin(desF1) probed by spin labelsp-IV,p-V, andm-VII, which corresponds to the apolar binding region, differs in conformation from-thrombin.  相似文献   

20.
An experimental strategy based on solution viscosity perturbation allowed us to study the energetics of amide substrates,p-aminobenzamidine (p-ABZ) and proflavin binding to the catalytic site of two proteolyzed forms of α-thrombin, i.e. ζ- and γT-thrombin. These thrombin derivatives are cleaved at the Leu144-Gly150 loop and at the fibrinogen recognition exosite (FRS), respectively. A phenomenological analysis of thermodynamic data showed that the amide substrates andp-ABZ interactions with ζ-thrombin were respectively, associated with a chemical compensation (i.e. the linear relationship between entropy and enthalpy of binding) and a hydrophobic phenomenon (i.e. a change in the standard heat capacity). The latter was slightly lower than that previously observed for a α-thrombin (0.78±0.25versus1.01±0.17 kcal/mol K). Both phenomenon were absent in γT-thrombin. The interaction of a α-, ζ- and γT-thrombin with macromolecular substrates that “bridge-bind” to both the catalytic site (CS) and fibrinogen recognition exosite (FRS), such as fibrinogen and the cleavable platelet receptor (CPR), was also evaluated. These interactions ere studied by following fibrinopeptide A (FpA) release and by measuring intraplatelet Ca2+changes induced by thrombin-CPR interaction. It was found that the free energy of activation (RTlnkcat/Km) for both fibrinogen and CPR hydrolysis followed the same hierarchy, i.e. α>ζ>γ. Moreover, the values of ΔCpfot α-, δ- and γT- thrombin interaction withp-ABZ were found to be linearly correlated to the free energy of activation for both fibrinogen and CPR cleavage. In conclusion, these data demonstrate that: (1) the Leu144-Gly150 loop and the FRS are both involved in the conformational transition linked to the binding ofp-aminobenzamidine to the thrombin active site; (2) the extent of thrombin's capacity to undergo conformational transitions in α-, ζ and γTforms is positively correlated to the free energy of activation for hydrolysis of macromolecular substrates interacting with both the catalytic domain and the FRS.f2f3  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