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.
At the site of vascular injury, von Willebrand factor (VWF) mediates platelet adhesion to subendothelial connective tissue through binding to the N-terminal domain of the alpha chain of platelet glycoprotein Ib (GPIbalpha). To elucidate the molecular mechanisms of the binding, we have employed charged-to-alanine scanning mutagenesis of the soluble fragment containing the N-terminal 287 amino acids of GPIbalpha. Sixty-two charged amino acids were changed singly or in small clusters, and 38 mutant constructs were expressed in the supernatant of 293T cells. Each mutant was assayed for binding to several monoclonal antibodies for human GPIbalpha and for ristocetin-induced and botrocetin-induced binding of 125I-labeled human VWF. Mutations at Glu128, Glu172, and Asp175 specifically decreased both ristocetin- and botrocetin-induced VWF binding, suggesting that these sites are important for VWF binding of platelet GPIb. Monoclonal antibody 6D1 inhibited ristocetin- and botrocetin-induced VWF binding, and a mutation at Glu125 specifically reduced the binding to 6D1. In contrast, antibody HPL7 had no effect for VWF binding, and mutant E121A reduced the HPL7 binding. Mutations at His12 and Glu14 decreased the ristocetin-induced VWF binding with normal botrocetin-induced binding. Crystallographic modeling of the VWF-GPIbalpha complex indicated that Glu128 and Asp175 form VWF binding sites; the binding of 6D1 to Glu125 interrupts the VWF binding of Glu128, but HPL7 binding to Glu121 has no effect on VWF binding. Moreover, His12 and Glu14 contact with Glu613 and Arg571 of VWF A1 domain, whose mutations had shown similar phenotype. These findings indicated the novel binding sites required for VWF binding of human GPIbalpha.  相似文献   

2.
Zhang J  Ma Z  Dong N  Liu F  Su J  Zhao Y  Shen F  Wang A  Ruan C 《PloS one》2011,6(7):e22157
The size of von Willebrand factor (VWF), controlled by ADAMTS13-dependent proteolysis, is associated with its hemostatic activity. Many factors regulate ADAMTS13-dependent VWF proteolysis through their interaction with VWF. These include coagulation factor VIII, platelet glycoprotein 1bα, and heparin sulfate, which accelerate the cleavage of VWF. Conversely, thrombospondin-1 decreases the rate of VWF proteolysis by ADAMTS13 by competing with ADAMTS13 for the A3 domain of VWF. To investigate whether murine monoclonal antibodies (mAbs) against human VWF affect the susceptibility of VWF to proteolysis by ADAMTS13 in vitro, eight mAbs to different domains of human VWF were used to evaluate the effects on VWF cleavage by ADAMTS13 under fluid shear stress and static/denaturing conditions. Additionally, the epitope of anti-VWF mAb (SZ34) was mapped using recombinant proteins in combination with enzyme-linked immunosorbent assay and Western blot analysis. The results indicate that mAb SZ34 inhibited proteolytic cleavage of VWF by ADAMTS13 in a concentration-dependent manner under fluid shear stress, but not under static/denaturing conditions. The binding epitope of SZ34 mAb is located between A1555 and G1595 in the central A2 domain of VWF. These data show that an anti-VWF mAb against the VWF-A2 domain (A1555-G1595) reduces the proteolytic cleavage of VWF by ADAMTS13 under shear stress, suggesting the role of this region in interaction with ADAMTS13.  相似文献   

3.
We investigated the crucial hemostatic interaction between von Willebrand factor (VWF) and platelet glycoprotein (GP) Ibalpha. Recombinant VWF A1 domain (residues Glu(497)-Pro(705) of VWF) bound stoichiometrically to a GPIbalpha-calmodulin fusion protein (residues His(1)-Val(289) of GPIbalpha; GPIbalpha-CaM) immobilized on W-7-agarose with a K(d) of 3.3 microM. The variant VWF A1(R545A) bound to GPIbalpha-CaM 20-fold more tightly, mainly because the association rate constant k(on) increased from 1,100 to 8,800 M(-1) s(-1). The GPIbalpha mutations G233V and M239V cause platelet-type pseudo-von Willebrand disease, and VWF A1 bound to GPIbalpha(G233V)-CaM and GPIbalpha(M239V)-CaM with a K(d) of 1.0 and 0.63 microM, respectively. The increased affinity of VWF A1 for GPIbalpha(M239V)-CaM was explained by an increase in k(on) to 4,500 M(-1) s(-1). GPIbalpha-CaM bound with similar affinity to recombinant VWF A1, to multimeric plasma VWF, and to a fragment of dispase-digested plasma VWF (residues Leu(480)/Val(481)-Gly(718)). VWF A1 and A1(R545A) bound to platelets with affinities and rate constants similar to those for binding to GPIbalpha-CaM, and botrocetin had the expected positively cooperative effect on the binding of VWF A1 to GPIbalpha-CaM. Therefore, allosteric regulation by botrocetin of VWF A1 binding to GPIbalpha, and the increased binding affinity caused by mutations in VWF or GPIbalpha, are reproduced by isolated structural domains. The substantial increase in k(on) caused by mutations in either A1 or GPIbalpha suggests that productive interaction requires rate-limiting conformational changes in both binding sites. The exceptionally slow k(on) and k(off) provide important new constraints on models for rapid platelet tethering at high wall shear rates.  相似文献   

4.
Leukocyte-platelet interaction is important in mediating leukocyte adhesion to a thrombus and leukocyte recruitment to a site of vascular injury. This interaction is mediated at least in part by the beta2-integrin Mac-1 (CD11b/CD18) and its counter-receptor on platelets, glycoprotein Ibalpha (GPIbalpha). High molecular weight kininogen (HK) was previously shown to interact with both GPIbalpha and Mac-1 through its domains 3 and 5, respectively. In this study we investigated the ability of HK to interfere with the leukocyte-platelet interaction. In a purified system, HK binding to GPIbalpha was inhibited by HK domain 3 and the monoclonal antibody (mAb) SZ2, directed against the epitope 269-282 of GPIbalpha, whereas mAb AP1, directed to the region 201-268 of GPIbalpha had no effect. In contrast, mAb AP1 inhibited the Mac-1-GPIbalpha interaction. Binding of GPIbalpha to Mac-1 was enhanced 2-fold by HK. This effect of HK was abrogated in the presence of HK domains 3 or 5 or peptides from the 475-497 region of the carboxyl terminus of domain 5 as well as in the presence of mAb SZ2 but not mAb AP1. Whereas no difference in the affinity of the Mac-1-GPIbalpha interaction was observed in the absence or presence of HK, maximal binding of GPIbalpha to Mac-1 doubled in the presence of HK. Moreover, HK/HKa increased the Mac-1-dependent adhesion of myelomonocytic U937 cells and K562 cells transfected with Mac-1 to immobilized GPIbalpha or to GPIbalpha-transfected Chinese hamster ovary cells. Finally, Mac-1-dependent adhesion of neutrophils to surface-adherent platelets was enhanced by HK. Thus, HK can bridge leukocytes with platelets by interacting via its domain 3 with GPIbalpha and via its domain 5 with Mac-1 thereby augmenting the Mac-1-GPIbalpha interaction. These distinct molecular interactions of HK with leukocytes and platelets contribute to the regulation of the adhesive behavior of vascular cells and provide novel molecular targets for reducing atherothrombotic pathologies.  相似文献   

5.
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   

6.
Bitiscetin, a C-type lectin-like heterodimeric snake venom protein purified from Bitis arietans, binds to human von Willebrand factor (VWF) and induces the platelet membrane glycoprotein (GP) Ib-dependent platelet agglutination in vitro similar to botrocetin. In contrast with botrocetin which binds to the A1 domain of VWF, the A3 domain, a major collagen-binding site of VWF, was proposed to be a bitiscetin-binding site. In the competitive binding assay, neither bitiscetin nor botrocetin had an inhibitory effect on the VWF binding to the immobilized type III collagen on a plastic plate. The anti-VWF monoclonal antibody NMC-4, which inhibits VWF-induced platelet aggregation by binding to alpha4 helix of the A1 domain, also inhibited bitiscetin binding to the VWF. Binding of VWF to the immobilized bitiscetin was competitively inhibited by a high concentration of botrocetin. A panel of recombinant VWF, in which alanine-scanning mutagenesis was introduced to the charged amino acid residues in the A1 domain, showed that the bitiscetin-binding activity was reduced in mutations at Arg632, Lys660, Glu666, and Lys673 of the A1 domain. Those substituted at Arg629, Arg636, and Lys667, which decreased the botrocetin binding, showed no effect on the bitiscetin binding. These results indicate that bitiscetin binds to a distinct site in the A1 domain of VWF spanning over alpha4a, alpha5 helices and the loop between alpha5 and beta6 but close to the botrocetin- and NMC-4-binding sites. Monoclonal antibodies recognizing the alpha-subunit of bitiscetin specifically inhibited bitiscetin-induced platelet agglutination without affecting the binding between VWF and bitiscetin, suggesting that the alpha-subunit of bitiscetin is located on VWF closer to the GPIb-binding site than the beta-subunit is. Bitiscetin and botrocetin might modulate VWF by binding to the homologous region of the A1 domain to induce a conformational change leading to an increased accessibility to platelet GPIb.  相似文献   

7.
The monoclonal antibody 6B4 has a potent antithrombotic effect in nonhuman primates by binding to the flexible loop, also known as the beta-switch region (amino acids 230-242), of glycoprotein Ibalpha (GPIbalpha). This interaction blocks, in high shear stress conditions, the specific interaction between GPIbalpha and von Willebrand factor suppressing platelet deposition to the damaged vessel wall, a key event in the pathogenesis of arterial thrombosis. To understand the interactions between this antibody and its antigen at the amino acid level, we here report the identification of the paratope and epitope in 6B4 and GPIbalpha, respectively, by using computer modeling and site-directed mutagenesis. The docking programs ZDOCK (rigid body docking) and HADDOCK (flexible docking) were used to model the interaction of 6B4 with GPIbalpha and to delineate the respective paratope and epitope. 6B4 and GPIbalpha mutants were constructed and assayed for their capacity to bind GPIbalpha and 6B4, respectively. From these data, it is found that the paratope of 6B4 is mainly formed by five residues: Tyr(27D), Lys(27E), Asp(28), and Glu(93) located in light chain CDR1 and -3, respectively, and Tyr(100C) of the heavy chain CDR3. These residues form a valley, where the GPIbalpha flexible loop can bind via residues Asp(235) and Lys(237). The experimental results were finally used to build a more accurate docking model. Taken together, this information provides guidelines for the design of new derivatized lead compounds with antithrombotic properties.  相似文献   

8.
The extracellular calcium-binding domain (positions 138-286) of the matrix protein BM-40 possesses a binding epitope of moderate affinity for several collagen types. This epitope was predicted to reside in helix alphaA and to be partially masked by helix alphaC. Here we show that deletion of helix alphaC produces a 10-fold increase in collagen affinity similar to that seen after proteolytic cleavage of this helix. The predicted removal of the steric constraint was clearly demonstrated by the crystal structure of the mutant at 2.8 A resolution. This constitutively activated mutant was used to map the collagen-binding site following alanine mutagenesis at 13 positions. Five residues were crucial for binding, R149 and N156 in helix alphaA, and L242, M245 and E246 in a loop region connecting the two EF hands of BM-40. These residues are spatially close and form a flat ring of 15 A diameter which matches the diameter of a triple-helical collagen domain. The mutations showed similar effects on binding to collagens I and IV, indicating nearly identical binding sites on both collagens. Selected mutations in the non-activated mutant DeltaI also reduced collagen binding, consistent with the same location of the epitope but in a more cryptic form in intact BM-40.  相似文献   

9.
Site-directed mutagenesis was performed in the protease-sensitive region, between the lipoyl and catalytic domains and in the catalytic domain, of the dihydrolipoyl transacetylase component (E2p) of the pyruvate dehydrogenase complex from Azotobacter vinelandii. The interaction of the mutated enzymes with the peripheral components pyruvate dehydrogenase (E1p) and lipoamide dehydrogenase (E3) was studied by gel filtration experiments, analytical ultracentrifugation and reconstitution of the pyruvate dehydrogenase complex. Upon binding of peripheral components, the 24-subunit core of A. vinelandii wild-type E2p dissociates into tetramers. Four E1p or E3 dimers can bind to a tetramer. Binding is mutually exclusive, resulting in an active complex containing one E3 and three E1p dimers. Large deletions of the protease-sensitive region of E2p resulted in a total loss of the E1p and E3 binding. A small deletion (delta P361-R362) or the point mutation K367Q in the protease-sensitive region did not influence E3 binding, but affected E1p binding strongly, although with excess E1p almost complete reconstitution was reached. For E2p with the point mutation R416D in the N-terminal region of the catalytic domain only 16% overall activity could be measured in reconstituted complexes. This is due to a very weak E1p/E2p interaction, whereas the E3 binding was not affected. The point mutation R416D did not influence the catalytic activity of E2p, although a function for this residue in the formation of the active site was predicted from amino acid similarities with chloramphenicol acetyltransferase type III from Escherichia coli. Deletion of the complete Ala + Pro-rich sequence between the protease-sensitive region and the catalytic domain did not affect the enzymological properties of E2p, nor the affinity for E1p or E3. A further deletion of 20 N-terminal residues from the catalytic domain destroyed the E2p activity. From gel filtration experiments it was concluded that the quaternary structure was unaffected, as was E3 binding. E1p binding was lost and, in contrast to the wild-type enzyme, no dissociation of the core upon addition of E3 was observed. This mutant enzyme possesses, like E. coli E2p, six E3 binding sites and clearly shows that interaction of E3 or E1p with the E1p sites and dissociation are linked processes. It is concluded that the binding site for E3 is located on the N-terminal part of the protease-sensitive region. In contrast, the binding site for E1p consists of two regions, one located on the protease-sensitive region and one of the catalytic domain. These regions are separated by a flexible sequence of about 20 amino acids.  相似文献   

10.
In vitro platelet glycoprotein Ib (GPIb) binding of the human von Willebrand factor (VWF) increases markedly by exogenous modulators such as ristocetin or botrocetin, and the binding does not occur in normal circulation. GPIb binding sites have been assigned in the VWF A1 domain, which consists of a disulfide loop Cys1272(509)-Cys1458(695) where amino acid residues are numbered from the starting methionine as +1. The previous numbering from the N-terminal Ser of the mature processed VWF is indicated in parentheses. In contrast, several gain-of-function mutations have been found in two regions comprised of the disulfide loop and its N- and C-terminal flanking regions. In this study, Cys1222(459)-Tyr1271(508), Gln1238(475)-Tyr1271(508), Glu1260(497)-Tyr1271(508), and Asp1459(696)-Asp1472(709) were sequentially deleted of full-length multimeric recombinant VWF. Deletions at either side resulted in normal GPIb binding, indicating that the flanking regions are not GPIb binding sites. However, the addition of a mutation at Arg1308(545) on each deletion mutant resulted in spontaneous GPIb binding without requiring modulators, suggesting that both regions are important for the inhibition of GPIb binding. Spontaneous binding was completely inhibited by monoclonal antibodies that recognize the GPIb binding sites. Interestingly, mutant proteins with N-terminal but not C-terminal deletions lost binding to monoclonal antibodies B328, B710, and 23C7, which selectively inhibit ristocetin-induced GPI binding. Their epitopes were found at His1268(505) or Asp1269(506). The crystallographic structure of the A1 domain suggests that GPIb binding is influenced by the molecular interface between the two regions and that the antibody binding to the interface inhibits binding.  相似文献   

11.
At sites of vascular injury, von Willebrand factor (VWF) mediates platelet adhesion through binding to platelet glycoprotein Ib (GPIb). Previous studies identified clusters of charged residues within VWF domain A1 that were involved in binding GPIb or botrocetin. The contribution of 28 specific residues within these clusters was analyzed by mutating single amino acids to alanine. Binding to a panel of six conformation-dependent monoclonal antibodies was decreased by mutations at Asp(514), Asp(520), Arg(552), and Arg(611) (numbered from the N-terminal Ser of the mature processed VWF), suggesting that these residues are necessary for domain A1 folding. Binding of (125)I-botrocetin was decreased by mutations at Arg(629), Arg(632), Arg(636), and Lys(667). Ristocetin-induced and botrocetin-induced binding to GPIb both were decreased by mutations at Lys(599), Arg(629), and Arg(632); among this group the K599A mutant was unique because (125)I-botrocetin binding was normal, suggesting that Lys(599) interacts directly with GPIb. Ristocetin and botrocetin actions on VWF were dissociated readily by mutagenesis. Ristocetin-induced binding to GPIb was reduced selectively by substitutions at positions Lys(534), Arg(571), Lys(572), Glu(596), Glu(613), Arg(616), Glu(626), and Lys(642), whereas botrocetin-induced binding to GPIb was decreased selectively by mutations at Arg(636) and Lys(667). The binding of monoclonal antibody B724 involved Lys(660) and Arg(663), and this antibody inhibits (125)I-botrocetin binding to VWF. The crystal structure of the A1 domain suggests that the botrocetin-binding site overlaps the monoclonal antibody B724 epitope on helix 5 and spans helices 4 and 5. The binding of botrocetin also activates the nearby VWF-binding site for GPIb that involves Lys(599) on helix 3.  相似文献   

12.
In this study we examine for the first time the roles of the various domains of human RNase H1 by site-directed mutagenesis. The carboxyl terminus of human RNase H1 is highly conserved with Escherichia coli RNase H1 and contains the amino acid residues of the putative catalytic site and basic substrate-binding domain of the E. coli RNase enzyme. The amino terminus of human RNase H1 contains a structure consistent with a double-strand RNA (dsRNA) binding motif that is separated from the conserved E. coli RNase H1 region by a 62-amino acid sequence. These studies showed that although the conserved amino acid residues of the putative catalytic site and basic substrate-binding domain are required for RNase H activity, deletion of either the catalytic site or the basic substrate-binding domain did not ablate binding to the heteroduplex substrate. Deletion of the region between the dsRNA-binding domain and the conserved E. coli RNase H1 domain resulted in a significant loss in the RNase H activity. Furthermore, the binding affinity of this deletion mutant for the heteroduplex substrate was approximately 2-fold tighter than the wild-type enzyme suggesting that this central 62-amino acid region does not contribute to the binding affinity of the enzyme for the substrate. The dsRNA-binding domain was not required for RNase H activity, as the dsRNA-deletion mutants exhibited catalytic rates approximately 2-fold faster than the rate observed for wild-type enzyme. Comparison of the dissociation constant of human RNase H1 and the dsRNA-deletion mutant for the heteroduplex substrate indicates that the deletion of this region resulted in a 5-fold loss in binding affinity. Finally, comparison of the cleavage patterns exhibited by the mutant proteins with the cleavage pattern for the wild-type enzyme indicates that the dsRNA-binding domain is responsible for the observed strong positional preference for cleavage exhibited by human RNase H1.  相似文献   

13.
To explore the molecular basis of von Willebrand factor (VWF) clearance, an experimental model employing VWF-deficient mice was developed. Biodistribution was examined by the injection of radiolabeled VWF, which was primarily directed to the liver with minor amounts in other organs. Disappearance of VWF from plasma was characterized by a rapid initial phase (t((1/2))alpha = 13 min) and a slow secondary phase (t((1/2))beta = 3 h), with a mean residence time (MRT) of 2.8 h. A similar clearance was observed for VWF consisting of only high or low molecular weight multimers, indicating that, in our experimental model, clearance is independent of multimeric distribution. This allowed us to compare the survival of full-length VWF to truncated variants. Deletion of both the amino-terminal D'-D3 and carboxyl-terminal D4-CK domains resulted in a fragment with a similar clearance to wild-type VWF. Deletion of only the D'-D3 region was associated with an almost 2-fold lower recovery and increased clearance (MRT = 1.6 h), whereas deletion of only the D4-CK region resulted in a significantly reduced clearance (MRT = 4.5 h, p < 0.02). These results point to a role of the D'-D3 region in preventing clearance of VWF. Furthermore, replacement of D3 domain residue Arg-1205 by His resulted in a markedly increased clearance (MRT = 0.3 h; p = 0.004). Therefore, this mutation seems to abrogate the protective effect of the D'-D3 region. In vitro analysis of this mutant also revealed a 2-fold reduced affinity for VWF propeptide at low pH, showing that mutation of Arg-1205 results not only in an increased clearance rate but is also associated with an impaired pH-dependent interaction with VWF propeptide.  相似文献   

14.
Vitronectin and plasminogen activator inhibitor-1 (PAI-1) are proteins that interact in the circulatory system and pericellular region to regulate fibrinolysis, cell adhesion, and migration. The interactions between the two proteins have been attributed primarily to binding of the somatomedin B (SMB) domain, which comprises the N-terminal 44 residues of vitronectin, to the flexible joint region of PAI-1, including residues Arg-103, Met-112, and Gln-125 of PAI-1. A strategy for deletion mutagenesis that removes the SMB domain demonstrates that this mutant form of vitronectin retains PAI-1 binding (Schar, C. R., Blouse, G. E., Minor, K. M., and Peterson, C. B. (2008) J. Biol. Chem. 283, 10297-10309). In the current study, the complementary binding site on PAI-1 was mapped by testing for the ability of a battery of PAI-1 mutants to bind to the engineered vitronectin lacking the SMB domain. This approach identified a second, separate site for interaction between vitronectin and PAI-1. The binding of PAI-1 to this site was defined by a set of mutations in PAI-1 distinct from the mutations that disrupt binding to the SMB domain. Using the mutations in PAI-1 to map the second site suggested interactions between alpha-helices D and E in PAI-1 and a site in vitronectin outside of the SMB domain. The affinity of this second interaction exhibited a K(D) value approximately 100-fold higher than that of the PAI-1-somatomedin B interaction. In contrast to the PAI-1-somatomedin B binding, the second interaction had almost the same affinity for active and latent PAI-1. We hypothesize that, together, the two sites form an extended binding area that may promote assembly of higher order vitronectin-PAI-1 complexes.  相似文献   

15.
Botrocetin is a snake venom protein that enhances the affinity of the A1 domain of plasma von Willebrand factor (vWF) for the platelet receptor glycoprotein Ibalpha (GPIbalpha), an event that contributes to bleeding and host death. Here we describe a kinetic and crystallographic analysis of this interaction that reveals a novel mechanism of affinity enhancement. Using high-temporal-resolution microscopy, we show that botrocetin decreases the GPIbalpha off-rate two-fold in both human and mouse complexes without affecting the on-rate. The key to this behavior is that, upon binding of GPIbalpha to vWF-A1, botrocetin prebound to vWF-A1 makes no contacts initially with GPIbalpha, but subsequently slides around the A1 surface to form a new interface. This two-step mechanism and flexible coupling may prevent adverse alterations in on-rate of GPIbalpha for vWF-A1, and permit adaptation to structural differences in GPIbalpha and vWF in several prey species.  相似文献   

16.
Goodpasture (GP) disease is an autoimmune disorder in which autoantibodies against the alpha3(IV) chain of type IV collagen bind to the glomerular and alveolar basement membranes, causing progressive glomerulonephritis and pulmonary hemorrhage. Two major conformational epitope regions have been identified on the noncollagenous domain of type IV collagen (NC1 domain) of the alpha3(IV) chain as residues 17-31 (E(A)) and 127-141 (E(B)) (Netzer, K.-O. et al. (1999) J. Biol. Chem. 274, 11267-11274). To determine whether these regions are two distinct epitopes or form a single epitope, three GP sera were fractionated by affinity chromatography on immobilized NC1 chimeras containing the E(A) and/or the E(B) region. Four subpopulations of GP antibodies with distinct epitope specificity for the alpha3(IV)NC1 domain were thus separated and characterized. They were designated GP(A), GP(B), GP(AB), and GP(X), to reflect their reactivity with E(A) only, E(B) only, both regions, and neither, respectively. Hence, regions E(A) and E(B) encompass critical amino acids that constitute three distinct epitopes for GP(A), GP(B), and GP(AB) antibodies, respectively, whereas the epitope for GP(X) antibodies is located in a different unknown region. The GP(A) antibodies were consistently immunodominant, accounting for 60-65% of the total immunoreactivity to alpha3(IV)NC1; thus, they probably play a major role in pathogenesis. Regions E(A) and E(B) are held in close proximity because they jointly form the epitope for Mab3, a monoclonal antibody that competes for binding with GP autoantibodies. All GP epitopes are sequestered in the hexamer configuration of the NC1 domain found in tissues and are inaccessible for antibody binding unless dissociation of the hexamer occurs, suggesting a possible mechanism for etiology of GP disease. GP antibodies have the capacity to extract alpha3(IV)NC1 monomers, but not dimers, from native human glomerular basement membrane hexamers, a property that may be of fundamental importance for the pathogenesis of the disease.  相似文献   

17.
Physiological concentrations of NaCl inhibit the hydrolysis of von Willebrand factor (VWF) by ADAMTS-13. This effect is because of the specific binding of chloride ions to VWF. Urea-induced unfolding was measured in the presence of NaCl, CH3COONa, and NaClO4 at pH 8.0, 25 degrees C, for multimeric VWF, the recombinant A1-A2-A3 VWF domains, and the A1 domain. Chloride stabilizes the folded conformation of the A1-A2-A3 and A1 domains more efficiently than acetate but less strongly than perchlorate. Spectroscopic evidence showed that chloride binds to both the A1 and A1-A2 domain but not to the isolated A2 domain. Binding of Cl- to both wild type (WT) and the natural mutant p.R1306W A1-A2-A3 domains of VWF has a large heat capacity change equal to -1 and -0.4 kcal mol(-1) K(-1) for WT and p.R1306W A1-A2-A3 domains, respectively. This result implies that a burial of a vast apolar surface area is caused by conformational transitions linked to chloride binding. At any temperature, chloride affinity was higher for WT than for the mutant p.R1306W form. Chloride ions inhibit hydrolysis by ADAMTS-13 of the A1-A2-A3 and A1-A2 domains in the presence of either urea or high shear stress, whereas this effect was either absent or negligible in experiments using A2 and A2-A3 domains. These findings show that the A1 domain contains the binding site of chloride ions that control allosterically the proteolysis by ADAMTS-13 of the Tyr1605-Met1606 bond in the A2 domain and that the R1306W mutation of type 2B VWD quenches the binding of chloride ion to the A1 domain.  相似文献   

18.
Von Willebrand factor (VWF) plays a central role in hemostasis. Triggered by shear-stress, it adheres to platelets at sites of vascular injury. Inactivation of VWF has been associated to the shielding of its adhesion sites and proteolytic cleavage. However, the molecular nature of this shielding and its coupling to cleavage under shear-forces in flowing blood remain unknown. In this study, we describe, to our knowledge, a new force-sensory mechanism for VWF-platelet binding, which addresses these questions, based on a combination of molecular dynamics (MD) simulations, atomic force microscopy (AFM), and microfluidic experiments. Our MD simulations demonstrate that the VWF A2 domain targets a specific region at the VWF A1 domain, corresponding to the binding site of the platelet glycoprotein Ibα (GPIbα) receptor, thereby causing its blockage. This implies autoinhibition of the VWF for the binding of platelets mediated by the A1-A2 protein-protein interaction. During force-probe MD simulations, a stretching force dissociated the A1A2 complex, thereby unblocking the GPIbα binding site. Dissociation was found to be coupled to the unfolding of the A2 domain, with dissociation predominantly occurring before exposure of the cleavage site in A2, an observation that is supported by our AFM experiments. This suggests that the A2 domain prevents platelet binding in a force-dependent manner, ensuring that VWF initiates hemostasis before inactivation by proteolytic cleavage. Microfluidic experiments with an A2-deletion VWF mutant resulted in increased platelet binding, corroborating the key autoinhibitory role of the A2 domain within VWF multimers. Overall, autoinhibition of VWF mediated by force-dependent interdomain interactions offers the molecular basis for the shear-sensitive growth of VWF-platelet aggregates, and might be similarly involved in shear-induced VWF self-aggregation and other force-sensing functions in hemostasis.  相似文献   

19.
The interaction between platelet glycoprotein (GP) Ib-IX-V complex and von Willebrand factor (vWF) is the first step of the hemostatic response to vessel injury. In platelet-type von Willebrand disease, two mutations, G233V and M239V, have been described within the Cys209-Cys248 disulfide loop of GPIbalpha that compromise hemostasis by increasing the affinity for vWF. We have earlier shown that converting other residues in this region to valine alters the affinity of GPIbalpha for vWF, with mutations K237V and Q232V, respectively, showing the greatest increase and decrease in affinity. Here, we investigated further the effect of these two mutations on the kinetics of the GPIbalpha interaction with the vWF-A1 domain under dynamic flow conditions. We measured the cellular on- and off-rate constants of Chinese hamster ovary cells expressing GPIb-IX complexes containing wild-type or mutant GPIbalpha interacting with vWF-A1-coated surfaces at different shear stresses. We found that the gain-of-function mutant, K237V, rolled very slowly and continuously on vWF-A1 surface while the loss-of-function mutant, Q232V, showed fast, saltatory movement compared to the wild-type (WT). The off-rate constants, calculated based on the analysis of lifetimes of transient tethers formed on surfaces coated with limiting densities of vWF-A1, revealed that the Q232V and K237V dissociated 1.25-fold faster and 2.2-fold slower than the WT. The cellular on-rate constant of WT, measured in terms of tethering frequency, was threefold more and threefold less than Q232V and K237V, respectively. Thus, the gain- and loss-of-function mutations in GPIbalpha affect both the association and dissociation kinetics of the GPIbalpha-vWF-A1 bond. These findings are in contrast to the functionally similar selectin bonds where some of the mutations have been reported to affect only the dissociation rate.  相似文献   

20.
The neural cell adhesion molecule NCAM is capable of mediating cell-cell adhesion via homophilic interactions. In this study, three strategies have been combined to identify regions of NCAM that participate directly in NCAM-NCAM binding: analysis of domain deletion mutations, mapping of epitopes of monoclonal antibodies, and use of synthetic peptides to inhibit NCAM activity. Studies on L cells transfected with NCAM mutant cDNAs using cell aggregation and NCAM-covasphere binding assays indicate that the third immunoglobulin-like domain is involved in homophilic binding. The epitopes of four monoclonal antibodies that have been previously shown to affect cell-cell adhesion mediated by NCAM were also mapped to domain 3. Overlapping hexapeptides were synthesized on plastic pins and assayed for binding with these monoclonal antibodies. One of them (PP) reacted specifically with the sequence KYSFNY. Synthetic oligopeptides containing the PP epitope were potent and specific inhibitors of NCAM binding activity. A substratum containing immobilized peptide conjugates also exhibited adhesiveness for neural retinal cells. Cell attachment was specifically inhibited by peptides that contained the PP-epitope and by anti-NCAM univalent antibodies. The shortest active peptide has the sequence KYSFNYDGSE, suggesting that this site is directly involved in NCAM homophilic interaction.  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