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 546 毫秒
1.
Previous studies of the immune response in polytransfused Glanzmann thrombasthenia (GT) patients and in autoimmune thrombocytopenic purpura (AITP) have relied on serum analysis and have shown the frequent development of Abs directed against the alpha(IIb)beta(3) integrin. However, little is known about the molecular diversity of the humoral immune response to alpha(IIb)beta(3) due to the paucity of mAbs issuing from these pathologies. We have isolated human IgG anti-alpha(IIb)beta(3) binding fragments using combinatorial libraries of single-chain IgG created from the B cells of a GT and an AITP patient, both with serum Abs. Ab screening was performed using activated platelets or activated alpha(IIb)beta(3)-expressing Chinese hamster ovary cells. Sequencing of selected phage Abs showed that a broad selection of genes from virtually all V gene families had been used, indicating the diversity of the immune response. About one-half of the V(H) and V(L) segments of our IgG anti-alpha(IIb)beta(3) fragments displayed extensive hypermutations in the complementarity-determining region, supporting the idea that an Ag-driven immune response was occurring in both patients. The H chain complementarity-determining region 3 analysis of phage Abs revealed motifs other than the well-known RGD and KQAGDV integrin-binding sequences. To our knowledge, our study is the first to illustrate multiple human IgG anti-alpha(IIb)beta(3) reactivities and structural variations linked to the anti-platelet human immune response. Human alpha(IIb)beta(3) Abs preferentially directed against the activated form of the integrin were further characterized because platelet alpha(IIb)beta(3) inhibitors are potential therapeutic reagents for treating acute coronary syndromes. Currently available alpha(IIb)beta(3) antagonists do not specifically recognize the activated form of the integrin.  相似文献   

2.
Integrin cytoplasmic tails regulate integrin activation that is required for high affinity binding with ligands. The interaction of the integrin beta subunit tail with a cytoplasmic protein, talin, largely contributes to integrin activation. Here we report the cooperative interaction of the beta3 membrane-proximal and -distal residues in regulation of talin-mediated alpha IIb beta3 activation. Because a chimeric integrin, alpha IIb beta3/beta1, in which the beta3 tail was replaced with the beta1 tail was constitutively active, we searched for the residues responsible for integrin activation among the residues that differed between the beta3 and beta1 tails. Single amino acid substitutions of Ile-719 and Glu-749 in the beta3 membrane-proximal and -distal regions, respectively, with the corresponding beta1 residues or alanine rendered alphaIIbbeta3 constitutively active. The I719M/E749S double mutant had the same ligand binding activity as alpha IIb beta3/beta1. These beta3 mutations also induced alphaVbeta3 activation. Conversely, substitution of Met-719 or Ser-749 in the beta1 tail with the corresponding beta3 tail residue (M719I or S749E) inhibited alpha IIb beta3/beta1 activation, and the M719I/S749E double mutant inhibited ligand binding to a level comparable with that of the wild-type alpha IIb beta3. Knock down of talin by short hairpin RNA inhibited the I719M- and E749S-induced alpha IIb beta3 activation. These results suggest that the beta3 membrane-proximal and -distal residues cooperatively regulate talin-mediated alpha IIb beta3 activation.  相似文献   

3.
Ma YQ  Yang J  Pesho MM  Vinogradova O  Qin J  Plow EF 《Biochemistry》2006,45(21):6656-6662
The short cytoplasmic tails regulate activation of integrin adhesion receptors via clasping/unclasping of their membrane-proximal helices. Using integrin alpha(IIb)beta(3) as a model, we show that a previously reported activating mutation alpha(IIb)(R(995)D) that perturbs the electrostatic interface in the clasp only partially activates alpha(IIb)beta(3) and that extensive activation of the receptor is achieved by complete deletion of alpha(IIb) CT or triple mutations in alpha(IIb)(V(990)A/F(992)A/R(995)D) that disrupt both electrostatic and hydrophobic interfaces in the clasp. The results provide quantitative evidence for an equilibrium-based integrin activation process where shifting the equilibrium to the fully activated state requires total unclasping of the cytoplasmic tails. We further demonstrate that while the C-terminal region of the alpha(IIb) tail minimally influences alpha(IIb)beta(3) activation, the C-terminal region of the beta(3) tail is critically involved. A disease-causing mutation of S(752)P in this region, but not S(752)A, suppressed partial activation induced by R(995)D or the talin head domain but did not affect activation induced by alpha(IIb) truncation. NMR spectroscopy revealed that S(752)P but not the S(752)A mutation disrupted a C-terminal helix within the beta(3) tail, suggesting that the C-terminal helix may regulate the equilibrium-based clasping/unclasping process. Together, these data provide molecular insights into how distinct regions of the cytoplasmic tails differentially and cooperatively regulate integrin activation.  相似文献   

4.
Adhesive interactions of platelet integrin alpha(IIb)beta3 with fibrinogen and fibrin are central events in hemostasis and thrombosis. However, the mechanisms by which alpha(IIb)beta3 binds these ligands remain incompletely understood. We have recently demonstrated that alpha(IIb)beta3 binds the gamma365-383 sequence in the gammaC-domain of fibrin(ogen). This sequence contains neither the AGDV nor the RGD recognition motifs, known to bind alpha(IIb)beta3, suggesting the different specificity of the integrin. Here, using peptide arrays, mutant fibrinogens, and recombinant mutant gammaC-domains, we have examined the mechanism whereby alpha(IIb)beta3 binds gamma365-383. The alpha(IIb)beta3-binding activity was localized within gamma370-381, with two short sequences, gamma370ATWKTR375 and gamma376WYSMKK381, being able to independently bind the integrin. Furthermore, recognition of alpha(IIb)beta3 by gamma370-381 depended on four basic residues, Lys373, Arg375, Lys380, and Lys381. Simultaneous replacement of these amino acids and deletion of the gamma408AGDV411 sequence in the recombinant gammaC-domain resulted in the loss of alpha(IIb)beta3-mediated platelet adhesion. Confirming the critical roles of the identified residues, abnormal fibrinogen Kaiserslautern, in which gammaLys380 is replaced by Asn, demonstrated delayed clot retraction and impaired alpha(IIb)beta3 binding. Also, a mutant recombinant fibrinogen modeled after the naturally occurring variant Osaka V (gammaArg375 --> Gly) showed delayed clot retraction and reduced binding to purified alpha(IIb)beta3. These results identify the gamma370-381 sequence of fibrin(ogen) as the binding site for alpha(IIb)beta3 involved in platelet adhesion and clot retraction and define the new recognition specificity of this integrin.  相似文献   

5.
The functional regulation of integrins is a major determinant of cell adhesion, migration and tissue maintenance. The binding of cytoskeletal proteins to various sites of integrin cytoplasmic domains is a key mechanism of this functional regulation. Expression of recombinant integrin alpha(IIb)beta(3) and alpha(M)beta(2) lacking the GFFKR-region in CHO cells results in constitutively activated integrins. In contrast, CHO cells stably expressing either a GFFKR-deleted alpha(V(del))beta(3) or a FF to AA-substituted alpha(V(AA))beta(3) do not reveal a constitutively activated integrin. Adhesion to immobilized fibrinogen is strongly impaired in alpha(V(del))beta(3) or alpha(V(AA))beta(3)-expressing cells, whereas it is not impaired in alpha(IIb)beta(3) and alpha(M)beta(2), both lacking the GFFKR-region. In a parallel plate flow chamber assay, alpha(V)beta(3)-expressing cells adhere firmly to fibrinogen and spread even at shear rates of 15 to 20 dyn/cm(2), whereas alpha(V(del))beta(3) or alpha(V(AA))beta(3) cells are detached at 15 dyn/cm(2). Actin stress fiber formation and focal adhesion plaques containing alpha(V)beta(3) are observed in alpha(V)beta(3) cells but not in alpha(V(del))beta(3) or alpha(V(AA))beta(3)-expressing cells. As an additional manifestation of impaired outside-in signaling, phosphorylation of pp125(FAK) was reduced in these cells. In summary, we report that the GFFKR-region of the alpha(V)-cytoplasmic domain and in particular two phenylalanines are essential for integrin alpha(V)beta(3) function, especially for outside-in signaling. Our results suggest that the two beta(3)-integrins alpha(IIb)beta(3) and alpha(V)beta(3) are differentially regulated via their GFFKR-region.  相似文献   

6.
This study investigates three aspects of the adhesive interaction operating between platelet glycoprotein Ib/IX and integrin alpha(IIb)beta(3). These include the following: 1) examining the sufficiency of GPIb/IX and integrin alpha(IIb)beta(3) to mediate irreversible cell adhesion on immobilized von Willebrand factor (vWf) under flow; 2) the ability of the vWf-GPIb interaction to induce integrin alpha(IIb)beta(3) activation independent of endogenous platelet stimuli; and 3) the identification of key second messengers linking the vWf-GPIb/IX interaction to integrin alpha(IIb)beta(3) activation. By using Chinese hamster ovary cells transfected with GPIb/IX and integrin alpha(IIb)beta(3), we demonstrate that these receptors are both necessary and sufficient to mediate irreversible cell adhesion under flow, wherein GPIb/IX mediates cell tethering and rolling on immobilized vWf, and integrin alpha(IIb)beta(3) mediates cell arrest. Moreover, we demonstrate direct signaling between GPIb/IX and integrin alpha(IIb)beta(3). Studies on human platelets demonstrated that vWf binding to GPIb/IX is able to induce integrin alpha(IIb)beta(3) activation independent of endogenous platelet stimuli under both static and physiological flow conditions (150-1800 s(-)(1)). Analysis of the key second messengers linking the vWf-GPIb interaction to integrin alpha(IIb)beta(3) activation demonstrated that the first step in the activation process involves calcium release from internal stores, whereas transmembrane calcium influx is a secondary event potentiating integrin alpha(IIb)beta(3) activation.  相似文献   

7.
The integrin alpha(v)beta(3) is the major receptor mediating the attachment of osteoclasts to the extracellular matrix in bone and plays a critical role in bone resorption and bone remodeling. Most of the ligands interacting with the alpha(v)beta(3) receptor contain an Arg-Gly-Asp (RGD) motif. Recently, we have identified two small RGD peptides, containing a benzophenone moiety at either the carboxyl or amino terminus, that photo-cross-linked within the beta(3)[99-118] [Bitan, G., et al. (1999) Biochemistry 38, 3414-3420] or the beta(3)[167-171] [Bitan, G., et al. (2000) Biochemistry 39, 11014-11023] sequence, respectively, of the alpha(v)beta(3) receptor in a selective fashion. Here, we report the synthesis of a photoreactive analogue of echistatin (a 49-amino acid peptide), a potent RGD-containing antagonist of the alpha(v)beta(3) receptor both in vitro and in vivo. This bioactive analogue is substituted at position 45 with a p-benzoyl moiety (pBz(2)), located within the flexible C-terminal domain and removed 20 amino acid residues from the R(24)GD(26) triad. This C-terminal domain was reported to contribute to receptor binding affinity by acting as an auxiliary binding site. The radiolabeled (125)I-[Arg(35),Lys(45)(N(epsilon)-pBz(2))]-echistatin photo-cross-links effectively to a site within the beta(3)[209-220] sequence. Residues in this domain have been reported to be part of the metal ion-dependent adhesion site (MIDAS). Receptor fragments overlapping this domain were reported to bind to fibrinogen and block fibrinogen binding to alpha(IIb)beta(3), the platelet integrin receptor. Taken together, position 45 in echistatin, located within an auxiliary binding site in echistatin, cross-links to a site distinct from the two previously reported sites, beta(3)[99-118] and beta(3)[167-171], which cross-link to photophores flanking the RGD triad. These cross-linking data support the hypothesis that the ligand-bound conformation of the integrin beta(3) subunit differs from the known conformation of I domains.  相似文献   

8.
Previous evidence suggests that interactions between integrin cytoplasmic domains regulate integrin activation. We have constructed and validated recombinant structural mimics of the heterodimeric alpha(IIb)beta(3) cytoplasmic domain. The mimics elicited polyclonal antibodies that recognize a combinatorial epitope(s) formed in mixtures of the alpha(IIb) and beta(3) cytoplasmic domains but not present in either isolated tail. This epitope(s) is present within intact alpha(IIb)beta(3), indicating that interaction between the tails can occur in the native integrin. Furthermore, the combinatorial epitope(s) is also formed by introducing the activation-blocking beta(3)(Y747A) mutation into the beta(3) tail. A membrane-distal heptapeptide sequence in the alpha(IIb) tail ((997)RPPLEED) is responsible for this effect on beta(3). Membrane-permeant palmitoylated peptides, containing this alpha(IIb) sequence, specifically blocked alpha(IIb)beta(3) activation in platelets. Thus, this region of the alpha(IIb) tail causes the beta(3) tail to resemble that of beta(3)(Y747A) and suppresses activation of the integrin.  相似文献   

9.
We have reconstituted the platelet glycoprotein (GP) Ib-IX-mediated activation of the integrin alpha(IIb)beta(3) in a recombinant DNA expression model, and show that 14-3-3 is important in GPIb-IX signaling. CHO cells expressing alpha(IIb)beta(3) adhere poorly to vWF. Cells expressing GPIb-IX adhere to vWF in the presence of botrocetin but spread poorly. Cells coexpressing integrin alpha(IIb)beta(3) and GPIb-IX adhere and spread on vWF, which is inhibited by RGDS peptides and antibodies against alpha(IIb)beta(3). vWF binding to GPIb-IX also activates soluble fibrinogen binding to alpha(IIb)beta(3) indicating that GPIb-IX mediates a cellular signal leading to alpha(IIb)beta(3) activation. Deletion of the 14-3-3-binding site in GPIbalpha inhibited GPIb-IX-mediated fibrinogen binding to alpha(IIb)beta(3) and cell spreading on vWF. Thus, 14-3-3 binding to GPIb-IX is important in GPIb-IX signaling. Expression of a dominant negative 14-3-3 mutant inhibited cell spreading on vWF, suggesting an important role for 14-3-3. Deleting both the 14-3-3 and filamin-binding sites of GPIbalpha induced an endogenous integrin-dependent cell spreading on vWF without requiring alpha(IIb)beta(3), but inhibited vWF-induced fibrinogen binding to alpha(IIb)beta(3). Thus, while different activation mechanisms may be responsible for vWF interaction with different integrins, GPIb-IX-mediated activation of alpha(IIb)beta(3) requires 14-3-3 interaction with GPIbalpha.  相似文献   

10.
In platelets, alpha(IIb)beta(3) exists in a form that cannot bind adhesive proteins in the plasma; although it can interact with immobilized fibrinogen it cannot interact with immobilized von Willebrand factor in the vessel wall. Soluble agonists such as thrombin convert alpha(IIb)beta(3) to a form that recognizes soluble and immobilized ligands. Attempts to reconstitute alpha(IIb)beta(3) activation in a non-hematopoietic, nucleated cell system have been unsuccessful. In the present study, we have developed a transfected Chinese hamster ovary cell model in which alpha(IIb)beta(3) activation is induced by signaling across glycoprotein (GP) Ib-IX by its ligand, von Willebrand factor. GPIb-IX activates not only the transfected alpha(IIb)beta(3) but also endogenous alpha(v)beta(3). Activation of the pathways leading to integrin activation occurred even in cells transfected with GPIb-IX lacking the domain on GPIbalpha that binds 14-3-3 or that which binds actin-binding protein. These studies demonstrate that signals induced by interaction of GPIb-IX with von Willebrand factor lead to alpha(IIb)beta(3) activation and suggest that the signaling pathways by which GPIb-IX induces alpha(IIb)beta(3) activation are different to those used by thrombin. Elucidation of these differences may provide insights into therapeutic ways in which to inhibit integrin activation in selective clinical settings.  相似文献   

11.
We have investigated the calcium signaling relationship between the two major platelet adhesion receptors, glycoprotein Ib/V/IX (GPIb/V/IX) and integrin alpha(IIb)beta(3), involved in regulating platelet adhesion on von Willebrand factor (vWf) under flow. Our studies demonstrate that GPIb engagement of immobilized vWf elicits a transient calcium spike that may function to promote reversible arrest of translocating platelets. Subsequent integrin alpha(IIb)beta(3) engagement of vWf promotes sustained calcium oscillations that are essential for the maintenance of irreversible adhesion. GPIb-induced calcium spikes appear distinct from those initiated by integrin alpha(IIb)beta(3), in that the former are exclusively mediated through release of intracellular calcium stores via a signaling mechanism independent of PI 3-kinase. In contrast, integrin alpha(IIb)beta(3)-dependent calcium flux involves a PI 3-kinase-dependent signaling mechanism linked to intracellular calcium mobilization and subsequent transmembrane calcium influx. Studies employing the caged calcium chelator (o-nitrophenyl-EGTA) demonstrate that transient calcium spikes initiate a transient phase of platelet arrest that is converted to irreversible adhesion with the development of sustained oscillatory calcium flux. These studies demonstrate the existence of a dual step calcium signaling mechanism utilized by GPIb and integrin alpha(IIb)beta(3) that serves to regulate the dynamics of platelet adhesion under flow.  相似文献   

12.
By superimposing data obtained by photo-cross-linking RGD-containing ligands to the human alpha(V)beta(3) integrin onto the crystal structure of the ectopic domain of this receptor (Xiong et al. (2001) Science 294, 339-345), we have identified the binding site for the RGD triad within this integrin. We synthesized three novel analogues of the 49-amino acid disintegrin, echistatin: [Bpa(21),Leu(28)]-, [Bpa(23),Leu(28)]-, and [Bpa(28)]echistatin. Each contains a photoreactive p-benzoyl-phenylalanyl (Bpa) residue in close proximity to the RGD motif which spans positions 24-26; together, the photoreactive positions flank the RGD motif. The analogues bind with high affinity to the purified recombinant alpha(V)beta(3) integrin, but very poorly to the closely related human alpha(IIb)beta(3) platelet integrin. While echistatin analogues containing Bpa in either position 23 or 28 cross-link specifically and almost exclusively to the beta(3) subunit of alpha(V)beta(3), [Bpa(21),Leu(28)]echistatin cross-links to both the alpha(V) and the beta(3) subunits, with cross-linking to the former favored. [Bpa(23),Leu(28)]echistatin cross-links 10-30 times more effectively than the other two analogues. We identified beta(3)[109-118] as the domain that encompasses the contact site for [Bpa(28)]echistatin. This domain is included in beta(3)[99-118] (Bitan et al. (2000) Biochemistry 39, 11014-11023), a previously identified contact domain for a cyclic RGD-containing heptapeptide with a benzophenone moiety in a position that is similar to the placement of the benzophenone in [Bpa(28)]echistatin relative to the RGD triad. Recently, we identified beta(3)[209-220] as the contact site for an echistatin analogue with a photoreactive group in position 45, near the C-terminus of echistatin (Scheibler et al. (2001) Biochemistry 40, 15117-14126). Taken together, these results support the hypothesis that the very high binding affinity of echistatin to alpha(V)beta(3) results from two distinct epitopes in the ligand, a site including the RGD triad and an auxiliary epitope at the C-terminus of echistatin. Combining our results from photoaffinity cross-linking studies with data now available from the recently published crystal structure of the ectopic domain of alpha(V)beta(3), we characterize the binding site for the RGD motif in this receptor.  相似文献   

13.
Talin, consisting of a 47-kDa N-terminal head domain (residues 1-433) and a 190-kDa C-terminal rod domain (residues 434-2541), links integrins to the actin cytoskeleton. We previously reported that the binding stoichiometry of integrin alpha(IIb)beta(3):talin is approximately 2:1. More recently, an integrin binding site has been localized to the talin head domain. In the present study, we identified another integrin binding site at the C-terminal region of the talin rod domain. In a solid phase binding assay, RGD affinity-purified alpha(IIb)beta(3) bound in a dose-dependent manner to microtiter wells coated with the isolated 190-kDa proteolytic fragment of the talin rod domain. Additionally, alpha(IIb)beta(3) also bound to the talin rod domain captured by 8d4, an anti-talin monoclonal antibody. Polyclonal antibodies raised against a recombinant protein fragment corresponding to the entire talin rod domain (anti-talin-R) inhibited alpha(IIb)beta(3) binding to intact talin by approximately 50% but completely blocked alpha(IIb)beta(3) binding to the talin rod domain. To localize the integrin binding site, we examined alpha(IIb)beta(3) binding to recombinant polypeptide fragments corresponding to partial sequences of the talin rod domain. Whereas alpha(IIb)beta(3) bound effectively to talin-(1075-2541) and talin-(1984-2541), it failed to bind to talin-(434-1076) and talin-(434-1975). Furthermore, the binding of alpha(IIb)beta(3) to talin-(1984-2541) was inhibited by anti-talin-R. These results indicate that an integrin binding site is located within residues 1984-2541 of the talin rod domain. Thus, talin contains two integrin binding sites, one in the homologous FERM (band four-point-one, ezrin, radixin, moesin) domain and another near its C terminus. Because talin exists as an anti-parallel homodimer in focal adhesions, the two integrin binding sites in the adjacent talin molecules would be in close proximity with each other.  相似文献   

14.
We have investigated receptor function and epitope expression of recombinant alpha(IIb)beta(3) mutated at Cys(177) or Cys(273) in the I-like domain as well as Cys(598), located in the fourth repeat of the membrane-proximal cysteine-rich region and mutated in a Glanzmann's thrombasthenia type II patient. The beta(3) mutants beta(3)C177A, beta(3)C273A, and beta(3)C598Y exhibited a decreased electrophoretic mobility in SDS-polyacrylamide gel electrophoresis under nonreducing conditions, confirming the disruption of the respective disulfide loops. Despite reduced surface expression, the alpha(IIb)beta(3)C177A, alpha(IIb)beta(3)C273A, and alpha(IIb)beta(3)C598Y receptors mediated cell adhesion to immobilized fibrinogen and translocated into focal adhesion plaques. The beta(3)C598Y mutation, but not the beta(3)C177A or beta(3)C273A mutations, induced spontaneous binding of the ligand mimetic monoclonal antibody PAC-1, while the beta(3)C177A and beta(3)C273A mutants exhibited reduced complex stability in the absence of Ca(2+). Epitope mapping of function-blocking monoclonal antibodies (mAbs) allowed the identification of two distinct subgroups; mAbs A2A9, pl2-46, 10E5, and P256 did not interact with alpha(IIb)beta(3)C273A and bound only weakly to alpha(IIb)beta(3)C177A, while mAbs AP2, LM609 and 7E3 bound normally to mutant alpha(IIb)beta(3)C273A, but interacted only weakly with mutant alpha(IIb)beta(3)C177A. Furthermore, a cryptic epitope recognized by mAb 4D10G3 and not exposed on wild type alpha(IIb)beta(3) became accessible only on mutant alpha(IIb)beta(3)C177A and was mapped to the 60-kDa chymotrypsin fragment of beta(3). Finally, the ligand-induced binding site (LIBS) epitopes AP5, D3, LIBS1, and LIBS2 were spontaneously expressed on all three mutants independent of RGDS or dithiothreitol treatment. Our results provide evidence that disruption of a single cysteine disulfide bond in the cysteine-rich repeat domain, but not in the I-like domain, activates integrin alpha(IIb)beta(3). In contrast, disruption of each of the disulfide bonds in the two long insertions of the I-like domain predicted to be in close contact with the alpha subunit beta-propeller domain affect the stability of the alpha(IIb)beta(3) heterodimer and inhibit complex-specific mAb binding without affecting the RGD binding capacity of the metal ion-dependent adhesion site-like domain.  相似文献   

15.
Several human and rodent tumor cell lines were examined for the presence of integrin mRNAs by dot- and Northern-blot analysis. All tumor cells tested expressed mRNAs for alpha 5, alpha IIb, beta 1 and beta 3. The mRNA of beta 2 integrin was not detectable and that of alpha V integrin was found only in certain cells. Northern blotting was carried out in three selected tumor cell lines: Clone A, HEL and B16a. An apparent difference in the mRNA species coding for the alpha IIb beta 3 integrin, but not for alpha 5 and beta 1, was found. Our result suggests that alternative splicing of integrin genes may be one of the important mechanisms in regulating alpha IIb beta 3 expression and function in different tumor cells.  相似文献   

16.
The affinity of integrin alpha(IIb)beta(3) for fibrinogen is controlled by inside-out signals that are triggered by agonists like thrombin. Agonist treatment of platelets also activates Rap1b, a small GTPase known to promote integrin-dependent adhesion of other cells. Therefore, we investigated the role of Rap1b in alpha(IIb)beta(3) function by viral transduction of GFP-Rap1 chimeras into murine megakaryocytes, which exhibit inside-out signaling similar to platelets. Expression of constitutively active GFP-Rap1b (V12) had no effect on unstimulated megakaryocytes, but it greatly augmented fibrinogen binding to alpha(IIb)beta(3) induced by a PAR4 thrombin receptor agonist (p < 0.01). The Rap1b effect was cell-autonomous and was prevented by pre-treating cells with cytochalasin D or latrunculin A to inhibit actin polymerization. Rap1b-dependent fibrinogen binding to megakaryocytes was blocked by POW-2, a novel monovalent antibody Fab fragment specific for high affinity murine alpha(IIb)beta(3). In contrast to GFP-Rap1b (V12), expression of GFP-Rap1GAP, which deactivates endogenous Rap1, inhibited agonist-induced fibrinogen binding (p < 0.01), as did dominant-negative GFP-Rap1b (N17) (p < 0.05). None of these treatments affected surface expression of alpha(IIb)beta(3). These studies establish that Rap1b can augment agonist-induced ligand binding to alpha(IIb)beta(3) through effects on integrin affinity, possibly by modulating alpha(IIb)beta(3) interactions with the actin cytoskeleton.  相似文献   

17.
Synthesis and SARs of new integrin alpha(V)beta(3) antagonists based on an N-substituted dibenzazepinone scaffold are described. Variation of spacer and guanidine mimetic led to potent compounds exhibiting an IC(50) towards alpha(V)beta(3) in the nanomolar range, high selectivity versus integrin alpha(IIb)beta(3) and efficacy in functional cellular assays.  相似文献   

18.
Hemostasis and thrombosis (blood clotting) involve fibrinogen binding to integrin alpha(IIb)beta(3) on platelets, resulting in platelet aggregation. alpha(v)beta(3) binds fibrinogen via an Arg-Asp-Gly (RGD) motif in fibrinogen's alpha subunit. alpha(IIb)beta(3) also binds to fibrinogen; however, it does so via an unstructured RGD-lacking C-terminal region of the gamma subunit (gammaC peptide). These distinct modes of fibrinogen binding enable alpha(IIb)beta(3) and alpha(v)beta(3) to function cooperatively in hemostasis. In this study, crystal structures reveal the integrin alpha(IIb)beta(3)-gammaC peptide interface, and, for comparison, integrin alpha(IIb)beta(3) bound to a lamprey gammaC primordial RGD motif. Compared with RGD, the GAKQAGDV motif in gammaC adopts a different backbone configuration and binds over a more extended region. The integrin metal ion-dependent adhesion site (MIDAS) Mg(2+) ion binds the gammaC Asp side chain. The adjacent to MIDAS (ADMIDAS) Ca(2+) ion binds the gammaC C terminus, revealing a contribution for ADMIDAS in ligand binding. Structural data from this natively disordered gammaC peptide enhances our understanding of the involvement of gammaC peptide and integrin alpha(IIb)beta(3) in hemostasis and thrombosis.  相似文献   

19.
Fertilin beta (also known as ADAM2), a mammalian sperm protein that mediates gamete cell adhesion during fertilization, is a member of the ADAM protein family whose members have disintegrin domains with homology to integrin ligands found in snake venoms. Fertilin beta utilizes an ECD sequence within its disintegrin domain to interact with the egg plasma membrane; the Asp is especially critical. Based on what is known about different integrin subfamilies and their ligands, we sought to characterize fertilin beta binding sites on mouse eggs, focusing on integrin subfamilies that recognize short peptide sequences that include an Asp residue: the alpha(5)/alpha(8)/alpha(v)/alpha(IIb) or RGD-binding subfamily (alpha(5)beta(1), alpha(8)beta(1), alpha(V)beta(1), alpha(V)beta(3), alpha(V)beta(5), alpha(V)beta(6), alpha(V)beta(8), and alpha(IIb)beta(3)) and the alpha(4)/alpha(9) subfamily (alpha(4)beta(1), alpha(9)beta(1), and alpha(4)beta(7)). We tested peptide sequences known to perturb interactions mediated by these integrins in two different assays for fertilin beta binding. Peptides with the sequence MLDG, which perturb alpha(4)/alpha(9) integrin-mediated interactions, significantly inhibit fertilin beta binding to eggs, which suggests a role for a member of this integrin subfamily as a fertilin beta receptor. RGD peptides, which perturb alpha(5)/alpha(8)/alpha(v)/alpha(IIb) integrin-mediated interactions, have partial inhibitory activity. The anti-alpha(6) antibody GoH3 has little or no inhibitory activity. An antibody to the integrin-associated tetraspanin protein CD9 inhibits the binding of a multivalent presentation of fertilin beta (immobilized on beads) but not soluble fertilin beta, which we speculate has implications for the role of CD9 in the strengthening of fertilin beta-mediated cell adhesion but not in initial ligand binding.  相似文献   

20.
Concrete structural features of integrin alpha(IIb)beta(3) on the surface of platelets (at rest and after activation) have been obtained from epitope maps based on cross-competition among monoclonal antibodies directed against the alpha(IIb) subunit calf-2 domain and the beta(3) subunit betaA domain of alpha(IIb)beta(3). At rest, the observed intersubunit interface is formed by the sequence stretches beta(3)-(150-216), alpha(IIb) light chain-(1-92), and alpha(IIb) heavy chain-(826-856); and the alpha(IIb) interchain interface is formed by the two latter sequence stretches, disulfide-bonded between alpha(IIb) heavy chain Cys(826) and alpha(IIb) light chain Cys(9). These structural features agree with those observed in the alpha(IIb)beta(3) rudimentary connectivity map in solution and with the alpha(v)beta(3) V-shaped crystal structure (Xiong, J.-P., Zhang, R., Dunker, R., Scott, D. L., Joachimiak, A., Goodman, S. L., and Arnaout, M. A. (2001) Science 294, 339-345), but they disagree with the domain disposition suggested by the actual ultrastructural model. The epitope maps in platelets activated by ADP, thrombin receptor activation peptide, and arachidonic acid differ not only from those in platelets at rest, but also among themselves. The structural rearrangements observed confirm the presence in activated platelets of the crystallographically observed knee and argue against the switchblade mechanism proposed for activation (Beglova, N., Blacklow, S. C., Takagi, J., and Springer, T. A. (2002) Nat. Struct. Biol. 9, 282-287), demonstrate the existence of alpha(IIb)beta(3) agonist-specific activation states, explain the specificity for ligand binding and functional inhibition for some agonists, and predict the existence of agonist-specific final effectors and receptor activation mechanisms. The distinct non-reciprocal competition patterns observed at rest and after activation support the agonist-specific activation states and the existence of intrasubunit and intersubunit allosteric effects, previously proposed as the mechanism for alpha(IIb)beta(3) transmembrane activation.  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