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 62 毫秒
1.
Structural data of integrin αIIbβ3 have been interpreted as supporting a model in which: 1) the receptor exists primarily in a “bent,” low affinity conformation on unactivated platelets and 2) activation induces an extended, high affinity conformation prior to, or following, ligand binding. Previous studies found that “clasping” the αIIb head domain to the β3 tail decreased fibrinogen binding. To study the role of αIIb extension about the genu, we introduced a disulfide “clamp” between the αIIb thigh and calf-1 domains. Clamped αIIbβ3 had markedly reduced ability to bind the large soluble ligands fibrinogen and PAC-1 when activated with monoclonal antibody (mAb) PT25-2 but not when activated by Mn2+ or by coexpressing the clamped αIIb with a β3 subunit containing the activating mutation N339S. The clamp had little effect on the binding of the snake venom kistrin (Mr 7,500) or αIIbβ3-mediated adhesion to immobilized fibrinogen, but it did diminish the enhanced binding of mAb AP5 in the presence of kistrin. Collectively, our studies support a role for αIIb extension about the genu in the binding of ligands of 340,000 and 900,000 Mr with mAb-induced activation but indicate that it is not an absolute requirement. Our data are consistent with αIIb extension resulting in increased access to the ligand-binding site and/or facilitating the conformational change(s) in β3 that affect the intrinsic affinity of the binding pocket for ligand.  相似文献   

2.
Integrins are the major cell adhesion molecules responsible for cell attachment to the extracellular matrix. The strength of integrin-mediated adhesion is controlled by the affinity of individual integrins (integrin activation) as well as by the number of integrins involved in such adhesion. The positive correlation between integrin activation and integrin clustering had been suggested previously, but several trials to induce integrin clustering by dimerization of the transmembrane domain or tail region of integrin α subunits failed to demonstrate any change in integrin activation. Here, using platelet integrin αIIbβ3 as a model system, we showed that there is intermolecular lateral interaction between integrins through the transmembrane domains, and this interaction can enhance the affinity state of integrins. In addition, when integrin clustering was induced through heteromeric lateral interactions using bimolecular fluorescence complementation, we could observe a significant increase in the number of active integrin molecules. Because the possibility of intermolecular interaction would be increased by a higher local concentration of integrins, we propose that integrin clustering can shift the equilibrium in favor of integrin activation.  相似文献   

3.
4.
Integrin-linked kinase (ILK) is an important signaling regulator that assembles into the heteroternary complex with adaptor proteins PINCH and parvin (termed the IPP complex). We recently reported that ILK is important for integrin activation in a Chinese hamster ovary (CHO) cell system. We previously established parental CHO cells expressing a constitutively active chimeric integrin (αIIbα6Bβ3) and mutant CHO cells expressing inactive αIIbα6Bβ3 due to ILK deficiency. In this study, we further investigated the underlying mechanisms for ILK-dependent integrin activation. ILK-deficient mutant cells had trace levels of PINCH and α-parvin, and transfection of ILK cDNA into the mutant cells increased not only ILK but also PINCH and α-parvin, resulting in the restoration of αIIbα6Bβ3 activation. In the parental cells expressing active αIIbα6Bβ3, ILK, PINCH, and α-parvin were co-immunoprecipitated, indicating the formation of the IPP complex. Moreover, short interfering RNA (siRNA) experiments targeting PINCH-1 or both α- and β-parvin mRNA in the parent cells impaired the αIIbα6Bβ3 activation as well as the expression of the other components of the IPP complex. In addition, ILK mutants possessing defects in either PINCH or parvin binding failed to restore αIIbα6Bβ3 activation in the mutant cells. Kindlin-2 siRNA in the parental cells impaired αIIbα6Bβ3 activation without disturbing the expression of ILK. For CHO cells stably expressing wild-type αIIbβ3 that is an inactive form, overexpression of a talin head domain (THD) induced αIIbβ3 activation and the THD-induced αIIbβ3 activation was impaired by ILK siRNA through a significant reduction in the expression of the IPP complex. In contrast, overexpression of all IPP components in the αIIbβ3-expressing CHO cells further augmented THD-induced αIIbβ3 activation, whereas they did not induce αIIbβ3 activation without THD. These data suggest that the IPP complex rather than ILK plays an important role and supports integrin activation probably through stabilization of the active conformation.  相似文献   

5.
Agonist-stimulated platelet activation triggers conformational changes of integrin αIIbβ3, allowing fibrinogen binding and platelet aggregation. We have previously shown that an octapeptide, p1YMESRADR8, corresponding to amino acids 313–320 of the β-ribbon extending from the β-propeller domain of αIIb, acts as a potent inhibitor of platelet aggregation. Here we have performed in silico modelling analysis of the interaction of this peptide with αIIbβ3 in its bent and closed (not swing-out) conformation and show that the peptide is able to act as a substitute for the β-ribbon by forming a clasp restraining the β3 hybrid and βI domains in a closed conformation. The involvement of species-specific residues of the β3 hybrid domain (E356 and K384) and the β1 domain (E297) as well as an intrapeptide bond (pE315-pR317) were confirmed as important for this interaction by mutagenesis studies of αIIbβ3 expressed in CHO cells and native or substituted peptide inhibitory studies on platelet functions. Furthermore, NMR data corroborate the above results. Our findings provide insight into the important functional role of the αIIb β-ribbon in preventing integrin αIIbβ3 head piece opening, and highlight a potential new therapeutic approach to prevent integrin ligand binding.  相似文献   

6.
Integrins are bidirectional, allosteric transmembrane receptors that play a central role in hemostasis and arterial thrombosis. Using cryo-electron microscopy, multireference single-particle reconstruction methods, and statistics-based computational fitting approaches, we determined three-dimensional structures of human integrin αIIbβ3 embedded in a lipid bilayer (nanodiscs) while bound to domains of the cytosolic regulator talin and to extracellular ligands. We also determined the conformations of integrin in solution by itself to localize the membrane and the talin-binding site. To our knowledge, our data provide unprecedented three-dimensional information about the conformational states of intact, full-length integrin within membrane bilayers under near-physiological conditions and in the presence of cytosolic activators and extracellular ligands. We show that αIIbβ3 integrins exist in a conformational equilibrium clustered around four main states. These conformations range from a compact bent nodule to two partially extended intermediate conformers and finally to a fully upright state. In the presence of nanodiscs and the two ligands, the equilibrium is significantly shifted toward the upright conformation. In this conformation, the receptor extends ∼20 nm upward from the membrane. There are no observable contacts between the two subunits other than those in the headpiece near the ligand-binding pocket, and the α- and β-subunits are well separated with their cytoplasmic tails ∼8 nm apart. Our results indicate that extension of the ectodomain is possible without separating the legs or extending the hybrid domain, and that the ligand-binding pocket is not occluded by the membrane in any conformations of the equilibrium. Further, they suggest that integrin activation may be influenced by equilibrium shifts.  相似文献   

7.
《Biophysical journal》2023,122(3):533-543
The platelet integrin αIIbβ3 undergoes long-range conformational transitions associated with its functional conversion from inactive (low-affinity) to active (high-affinity) during hemostasis. Although new conformations that are intermediate between the well-characterized bent and extended states have been identified, their molecular dynamic properties and functions in the assembly of adhesions remain largely unexplored. In this study, we evaluated the properties of intermediate conformations of integrin αIIbβ3 and characterized their effects on the assembly of adhesions by combining all-atom simulations, principal component analysis, and mesoscale modeling. Our results show that in the low-affinity, bent conformation, the integrin ectodomain tends to pivot around the legs; in intermediate conformations, the headpiece becomes partially extended, away from the lower legs. In the fully open, active state, αIIbβ3 is flexible, and the motions between headpiece and lower legs are accompanied by fluctuations of the transmembrane helices. At the mesoscale, bent integrins form only unstable adhesions, but intermediate or open conformations stabilize the adhesions. These studies reveal a mechanism by which small variations in ligand binding affinity and enhancement of the ligand-bound lifetime in the presence of actin retrograde flow stabilize αIIbβ3 integrin adhesions.  相似文献   

8.
Under physiological conditions, circulating platelets are discoid in shape.1 On these platelets, the fibrinogen receptor (integrin αIIbβ3) is in a low-affinity state, unable to bind soluble fibrinogen (Fg). Activation by agonists such as ADP and thrombin leads to a change in the conformation of the integrin αIIbβ3 through a process known as inside-out signaling. This enables the integrin to bind soluble Fg, which initiates a cascade of events referred to as outside-in signaling.2 Outside-in signaling control processes, such as platelet spreading and clot retraction, by regulating small G-proteins such as RhoA, Rac and cdc42.Key words: platelets, integrin αIIbβ3, Galpha13, RhoA, clot retraction, thrombin, fibrinogenThe majority of the physiological platelet agonists (except collagen) induce inside-out signaling by binding to specific G-protein-coupled receptors (GPCRs). A G-protein plays a crucial role in translating the signal from GPCR to downstream effector molecules, ultimately leading to affinity modulation of integrin αIIbβ3. Platelets express nine Gα subunits; namely Gq, Gi1, Gi2, Gi3, Gz, G12, G13, Gs and G16. Previous studies have shown that a small G-protein, RhoA, is activated by the G12/13 family and plays a crucial role in calcium-independent platelet shape change.3 However, RhoA is also activated by αIIbβ3 and inhibits platelet spreading to trigger clot retraction.4 Recently, in a series of elegant experiments, Gong et al. have described the dynamic regulation of RhoA through a signaling crosstalk between Gα13 and αIIbβ3.5By generating mice in which the platelets were depleted of Gα13 using siRNA technology, Gong et al. investigated the role of Gα13-mediated signaling on platelet spreading on immobilized Fg.5 The confocal images very clearly showed that, in the absence of Gα13, platelets spread poorly on Fg, which was rescued by pretreatment with the Rho-kinase inhibitor Y27632, confirming previous findings that RhoA activated downstream of integrin αIIbβ3 inhibits platelet spreading. Interestingly, Gα13-depleted platelets failed to activate c-Src but accelerated RhoA activation. From these observations, the authors infer that Gα13 is important for integrin-mediated c-Src activation and RhoA inhibition, leading to increased cell spreading.5Since Gα13 regulates integrin-mediated cell spreading and c-Src activation, Gong et al. examined the interaction of Gα13 with αIIbβ3 using co-immunoprecipitation and GST pull-down assays.5 They found that the GTP-bound form of Gα13 shows enhanced interaction with the integrin β3 subunit. This interaction is required for the activation of c-Src and the inhibition of RhoA. However, they found that the inhibition of RhoA is transient. RhoA activation is suppressed for the first 15 min of platelet spreading, after which RhoA is activated. This initial suppression is rescued by blocking Gα13 and β3 cytoplasmic domain (β3-CD) interaction. Furthermore, they observed that RhoA activation parallels clot retraction.5 These findings indicate that Gα13 is a key regulator of platelet spreading and clot retraction phenomena.According to Gong et al., thrombin-induced inside-out signaling through GPCR leads to GTP loading of Gα13 (Fig. 1A). This GTP-bound Gα13 interacts with integrin β3-CD of ligand-bound integrin, thus facilitating c-Src activation, which leads to platelet spreading. Blockade of the interaction between Gα13 and β3-CD or cleavage of β3-CD by calpain results in clot retraction (Fig. 1B).Open in a separate windowFigure 1Schematic representation of the dynamic regulation of RhoA by Gα13 during platelet activation. (A) Activation of platelets by thrombin receptors coupled to Gα13 leads to the activation of RhoA, leading to platelet shape change. (B) The change in the conformation of integrin to a high-affinity form results in fibrinogen binding to αIIbβ3. Active Gα13 binds to the cytoplasmic domain of β3 leading to the activation of c-Src, resulting in platelet spreading. The rise in intracellular calcium activates calpain, which cleaves the β3 cytoplasmic domain, releasing c-Src, which, resulting in the activation of RhoA, leads to cell retraction. *Denotes GTP-bound active form of G-proteins.Perhaps the most significant and novel finding of the study is the identification of integrin αIIbβ3 as an effector of Gα13. The study also convincingly shows that Gα13 bound to integrin regulates RhoA via c-Src. Furthermore, achieving 80% knockdown of Gα13 in an in vivo setting using siRNA represents a technological advancement. Since Gα13 binds to integrin β3-CD in a 1:1 stoichiometry, it appears that only a small population of integrin is regulated by Gα13, as there are far less Gα13 molecules in a single platelet than the number of αIIbβ3 molecules. This will require further investigation. Gong et al. also finds that an appreciable amount of Gα13 is associated with β3 in resting platelets, which requires some explanation.5 It is also not clear if Gα13 remains bound to β3-CD or dissociates from the integrin during clot retraction.Overall, this is a paradigm-shifting study that establishes the importance of the dynamic regulation of RhoA by Gα13 in order to achieve efficient platelet spreading and clot retraction.  相似文献   

9.
Summary The localization of glycoprotein (GP) IIb/IIIa (integrin IIb3) in both resting and thrombin-activated platelets was studied immunocytochemically. By the pre-embedding method where only the GP IIb/IIIa molecules on the surface of platelets were immunostained, the distribution of protein A-colloidal gold label was randomly distributed along the surface membrane of resting platelets at a density of 18.0±2.7 gold particles/m of membrane. At 15 s after stimulation by 0.1 U/ml of thrombin in an unstirred platelet suspension, the spheroid-shaped platelets with pseudopodia still had normal numbers of -granules, and the density of gold particles was 19.7±3.6 particles/m. At 5 min, the -granules were no longer present because of the release reaction, and the density of gold particles significantly increased (27.0±3.7 particles/m; p<0.01). In immunostained ultra-thin frozen sections, the gold particles were detected not only on the surface membrane, including the open canalicular system (OCS), but also on the -granule membranes of resting platelets. At 30 s after thrombin stimulation the -granules fused with the OCS, resulting in the formation of a swollen OCS, which still had gold particles on its membrane. At 5 min, the gold particles were detected on the membrane of the swollen OCS located near the surface membrane, while very few gold particles were present on the membrane of the OCS in the central part of the platelets. These results demonstrate that -granule membrane GPIIb/IIIa translocates to the surface membrane through the membrane of the OCS. Also the translocation of -granule membrane GPIIb/IIIa gives rise to an actual increase in GPIIb/IIIa on the surface membrane during the release reaction induced by thrombin.  相似文献   

10.
Cationic membrane-proximal amino acids determine the topology of membrane proteins by interacting with anionic lipids that are restricted to the intracellular membrane leaflet. This mechanism implies that anionic lipids interfere with electrostatic interactions of membrane proteins. The integrin αIIbβ3 transmembrane (TM) complex is stabilized by a membrane-proximal αIIb(Arg995)-β3(Asp723) interaction; here, we examine the influence of anionic lipids on this complex. Anionic lipids compete for αIIb(Arg995) contacts with β3(Asp723) but paradoxically do not diminish the contribution of αIIb(Arg995)-β3(Asp723) to TM complex stability. Overall, anionic lipids in annular positions stabilize the αIIbβ3 TM complex by up to 0.50 ± 0.02 kcal/mol relative to zwitterionic lipids in a headgroup structure-dependent manner. Comparatively, integrin receptor activation requires TM complex destabilization of 1.5 ± 0.2 kcal/mol, revealing a sizeable influence of lipid composition on TM complex stability. We implicate changes in lipid headgroup accessibility to small molecules (physical membrane characteristics) and specific but dynamic protein-lipid contacts in this TM helix-helix stabilization. Thus, anionic lipids in ubiquitous annular positions can benefit the stability of membrane proteins while leaving membrane-proximal electrostatic interactions intact.  相似文献   

11.
A truncated form of the Agouti‐related protein (AgRP), a member of the cystine‐knot family, has shown promise as a scaffold for engineering novel peptides with new molecular recognition properties. In this study, we replaced a constrained six amino acid loop in AgRP with a nine amino acid loop containing an Arg–Gly–Asp integrin recognition motif, and randomized the neighboring residues to create a library of ~20 million AgRP variants. We displayed the AgRP mutants as fusions on the surface of yeast and used high‐throughput fluorescence‐activated cell sorting (FACS) to isolate peptides that bound specifically to the platelet integrin αIIbβ3, a clinically important target for the prevention and treatment of thrombosis. These AgRP peptides had equilibrium dissociation (KD) constants for αIIbβ3 integrin ranging from 60 to 90 nM, and did not bind to αvβ3, αvβ5, or α5β1 integrins. Using an alternate library screening strategy, we identified AgRP peptides that bound to both αIIbβ3 and αvβ3 integrins with KD values ranging from 40 to 70 nM and 20 to 30 nM, respectively, and did not bind to αvβ5 or α5β1 integrins. Unique consensus sequences were identified within both series of AgRP peptides suggesting alternative molecular recognition events that dictate different integrin binding specificities. In addition, the engineered AgRP peptides prevented platelet aggregation as well as or slightly better than the FDA‐approved cyclic peptide eptifibatide. Collectively, these data demonstrate that cystine‐knot peptides can be generated with high affinity and specificity to closely‐related integrins, and provide insights into molecular interactions between small, structured peptide ligands and their receptors. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

12.
Although vinculin is used frequently as a marker for integrin-mediated focal adhesion complexes, how it regulates the activation of integrin is mostly unknown. In this study, we examined whether vinculin would activate integrin in Chinese hamster ovary (CHO) cells expressing human integrin αIIbβ3. Silencing of vinculin by lentiviral transduction with a short hairpin RNA sequence affected the binding of PAC-1 (an antibody recognizing activated human αIIbβ3) to a constitutively active form of αIIbβ3 (α6Bβ3) expressed on CHO cells, while its inhibitory effects were much weaker than those of talin-1. Overexpression of an active form of vinculin without intramolecular interactions, but not the full length one, induced PAC-1 binding to native αIIbβ3 expressed on CHO cells in a manner dependent on talin-1. On the other hand, silencing of talin-1, but not vinculin, failed to induce cell spreading of α6Bβ3-CHO cells on fibrinogen, even in the presence of PT 25-2, a monoclonal antibody that activates αIIbβ3. Thus, an active form of vinculin could induce αIIbβ3 inside-out signaling through the actions of talin-1, while vinculin was dispensable for outside-in signaling.  相似文献   

13.
Talins and kindlins bind to the integrin β3 cytoplasmic tail and both are required for effective activation of integrin αIIbβ3 and resulting high-affinity ligand binding in platelets. However, binding of the talin head domain alone to β3 is sufficient to activate purified integrin αIIbβ3 in vitro. Since talin is localized to the cytoplasm of unstimulated platelets, its re-localization to the plasma membrane and to the integrin is required for activation. Here we explored the mechanism whereby kindlins function as integrin co-activators. To test whether kindlins regulate talin recruitment to plasma membranes and to αIIbβ3, full-length talin and kindlin recruitment to β3 was studied using a reconstructed CHO cell model system that recapitulates agonist-induced αIIbβ3 activation. Over-expression of kindlin-2, the endogenous kindlin isoform in CHO cells, promoted PAR1-mediated and talin-dependent ligand binding. In contrast, shRNA knockdown of kindlin-2 inhibited ligand binding. However, depletion of kindlin-2 by shRNA did not affect talin recruitment to the plasma membrane, as assessed by sub-cellular fractionation, and neither over-expression of kindlins nor depletion of kindlin-2 affected talin interaction with αIIbβ3 in living cells, as monitored by bimolecular fluorescence complementation. Furthermore, talin failed to promote kindlin-2 association with αIIbβ3 in CHO cells. In addition, purified talin and kindlin-3, the kindlin isoform expressed in platelets, failed to promote each other's binding to the β3 cytoplasmic tail in vitro. Thus, kindlins do not promote initial talin recruitment to αIIbβ3, suggesting that they co-activate integrin through a mechanism independent of recruitment.  相似文献   

14.
Pregnancy-specific glycoproteins (PSGs) are immunoglobulin superfamily members encoded by multigene families in rodents and primates. In human pregnancy, PSGs are secreted by the syncytiotrophoblast, a fetal tissue, and reach a concentration of up to 400 ug/ml in the maternal bloodstream at term. Human and mouse PSGs induce release of anti-inflammatory cytokines such as IL-10 and TGFβ1 from monocytes, macrophages, and other cell types, suggesting an immunoregulatory function. RGD tri-peptide motifs in the majority of human PSGs suggest that they may function like snake venom disintegrins, which bind integrins and inhibit interactions with ligands. We noted that human PSG1 has a KGD, rather than an RGD motif. The presence of a KGD in barbourin, a platelet integrin αIIbβ3 antagonist found in snake venom, suggested that PSG1 may be a selective αIIbβ3 ligand. Here we show that human PSG1 binds αIIbβ3 and inhibits the platelet – fibrinogen interaction. Unexpectedly, however, the KGD is not critical as multiple PSG1 domains independently bind and inhibit αIIbβ3 function. Human PSG9 and mouse Psg23 are also inhibitory suggesting conservation of this function across primate and rodent PSG families. Our results suggest that in species with haemochorial placentation, in which maternal blood is in direct contact with fetal trophoblast, the high expression level of PSGs reflects a requirement to antagonise abundant (3 mg/ml) fibrinogen in the maternal circulation, which may be necessary to prevent platelet aggregation and thrombosis in the prothrombotic maternal environment of pregnancy.  相似文献   

15.
Using a combined experimental and theoretical approach named binding-unbinding correlation spectroscopy (BUCS), we describe the two-dimensional kinetics of interactions between fibrinogen and the integrin αIIbβ3, the ligand-receptor pair essential for platelet function during hemostasis and thrombosis. The methodology uses the optical trap to probe force-free association of individual surface-attached fibrinogen and αIIbβ3 molecules and forced dissociation of an αIIbβ3-fibrinogen complex. This novel approach combines force clamp measurements of bond lifetimes with the binding mode to quantify the dependence of the binding probability on the interaction time. We found that fibrinogen-reactive αIIbβ3 pre-exists in at least two states that differ in their zero force on-rates (kon1 = 1.4 × 10−4 and kon2 = 2.3 × 10−4 μm2/s), off-rates (koff1 = 2.42 and koff2 = 0.60 s−1), and dissociation constants (Kd1 = 1.7 × 104 and Kd2 = 2.6 × 103 μm−2). The integrin activator Mn2+ changed the on-rates and affinities (Kd1 = 5 × 104 and Kd2 = 0.3 × 103 μm−2) but did not affect the off-rates. The strength of αIIbβ3-fibrinogen interactions was time-dependent due to a progressive increase in the fraction of the high affinity state of the αIIbβ3-fibrinogen complex characterized by a faster on-rate. Upon Mn2+-induced integrin activation, the force-dependent off-rates decrease while the complex undergoes a conformational transition from a lower to higher affinity state. The results obtained provide quantitative estimates of the two-dimensional kinetic rates for the low and high affinity αIIbβ3 and fibrinogen interactions at the single molecule level and offer direct evidence for the time- and force-dependent changes in αIIbβ3 conformation and ligand binding activity, underlying the dynamics of fibrinogen-mediated platelet adhesion and aggregation.  相似文献   

16.
The platelet integrin αIIbβ3 is widely accepted as a structural and a functional model of the broad integrin protein family. The four calcium-binding sites in the αIIb subunit contribute to biogenesis and stability of the protein. Mansour et al. (J Thromb Haemost 9:192–200, 2011) showed that the natural Asn2Asp mutation causing Glanzmann thrombasthenia, prevented surface expression of αIIbβ3, whereas the artificial Asn2Gln mutation only decreased its level. Molecular dynamics simulations and EDTA chelation assay were used here to explore the mechanism of these structural deformations. We show a considerable expansion of the calcium-binding site 3 in Asn2Asp mutation, whereas the Asn2Gln toggles between normal and expanded conformations. The αIIbβ3 surface expression level correlates to the relative spending time in the expanded conformation. By a comparison to other calcium-binding sites of αIIb and of other α integrins we show that the size of a calcium-binding loop is conserved. EDTA chelation assay shows a sensitivity to calcium removal, which correlates with the reduction in αIIbβ3 surface expression and with the calcium binding site expansion, thus verifying the simulation data. Here we indicate that Asn2 mutation affects a calcium-binding site 3 of αIIb, which structural deformation is proposed to deprive calcium binding and interfere with an integrin intracellular trafficking and its surface expression.  相似文献   

17.
Integrins are a family of heterodimeric adhesion receptors that transmit signals bi-directionally across the plasma membranes. The transmembrane domain (TM) of integrin plays a critical role in mediating transition of the receptor from the default inactive to the active state on the cell surfaces. In this study, we successfully applied the substituted cysteine scanning accessibility method to determine the intracellular border of the integrin α(IIb)β(3) TM in the inactive and active states in living cells. We examined the aqueous accessibility of 75 substituted cysteines comprising the C terminus of both α(IIb) and β(3) TMs, the intracellular membrane-proximal regions, and the whole cytoplasmic tails, to the labeling of a membrane-permeable, cysteine-specific chemical biotin maleimide (BM). The active state of integrin α(IIb)β(3) heterodimer was generated by co-expression of activating partners with the cysteine-substituted constructs. Our data revealed that, in the inactive state, the intracellular lipid/aqueous border of α(IIb) TM was at Lys(994) and β(3) TM was at Phe(727) respectively; in the active state, the border of α(IIb) TM shifted to Pro(998), whereas the border of β(3) TM remained unchanged, suggesting that complex conformational changes occurred in the TMs upon α(IIb)β(3) inside-out activation. On the basis of the results, we propose a new inside-out activation mechanism for integrin α(IIb)β(3) and by inference, all of the integrins in their native cellular environment.  相似文献   

18.

Background

Hemostasis and thrombosis are regulated by agonist-induced activation of platelet integrin αIIbβ3. Integrin activation, in turn is mediated by cellular signaling via protein kinases and protein phosphatases. Although the catalytic subunit of protein phosphatase 1 (PP1c) interacts with αIIbβ3, the role of PP1c in platelet reactivity is unclear.

Methodology/Principal Findings

Using γ isoform of PP1c deficient mice (PP1cγ−/−), we show that the platelets have moderately decreased soluble fibrinogen binding and aggregation to low concentrations of thrombin or protease-activated receptor 4 (PAR4)-activating peptide but not to adenosine diphosphate (ADP), collagen or collagen-related peptide (CRP). Thrombin-stimulated PP1cγ−/− platelets showed decreased αIIbβ3 activation despite comparable levels of αIIbβ3, PAR3, PAR4 expression and normal granule secretion. Functions regulated by outside-in integrin αIIbβ3 signaling like adhesion to immobilized fibrinogen and clot retraction were not altered in PP1cγ−/− platelets. Thrombus formation induced by a light/dye injury in the cremaster muscle venules was significantly delayed in PP1cγ−/− mice. Phosphorylation of glycogen synthase kinase (GSK3)β-serine 9 that promotes platelet function, was reduced in thrombin-stimulated PP1cγ−/− platelets by an AKT independent mechanism. Inhibition of GSK3β partially abolished the difference in fibrinogen binding between thrombin-stimulated wild type and PP1cγ−/− platelets.

Conclusions/Significance

These studies illustrate a role for PP1cγ in maintaining GSK3β-serine9 phosphorylation downstream of thrombin signaling and promoting thrombus formation via fibrinogen binding and platelet aggregation.  相似文献   

19.
Decorsin is an antagonist of platelet glycoprotein integrin αIIbβ3 on platelets; the protein is 39 amino acids long with three disulfide bridges in its tertiary structure. To demonstrate decorsin’s mechanism of action, we applied the computational virtual technique and platelet aggregation inhibition assay, which showed that the flanking amino-acid residues of the Arg–Gly–Asp (RGD) motif play an important role in platelet aggregation. The computational simulations revealed that the RGD motif mainly contributes to the stability of the complex when decorsion interacts with integrin αIIbβ3. However, the C-terminal residues, such as 34A→W and 35D→R, was also found to possibly play a key role in their binding structures. Moreover, we produced a decorsin analog (A34W plus D35R decorsin), in which the 34A (alanine) and 35D (aspartic acid) residues were respectively substituted by W (tryptophan) and R (arginine). This isoform was then recombinantly expressed in Escherichia coli. Intriguingly, this mutant type showed higher anti-platelet aggregation activity than the wildtype. Our study may further contribute to finding decorsin mutants with higher anti-platelet aggregation activity.  相似文献   

20.
Rap1b is activated by platelet agonists and plays a critical role in integrin α(IIb)β(3) inside-out signaling and platelet aggregation. Here we show that agonist-induced Rap1b activation plays an important role in stimulating secretion of platelet granules. We also show that α(IIb)β(3) outside-in signaling can activate Rap1b, and integrin outside-in signaling-mediated Rap1b activation is important in facilitating platelet spreading on fibrinogen and clot retraction. Rap1b-deficient platelets had diminished ATP secretion and P-selectin expression induced by thrombin or collagen. Importantly, addition of low doses of ADP and/or fibrinogen restored aggregation of Rap1b-deficient platelets. Furthermore, we found that Rap1b was activated by platelet spreading on immobilized fibrinogen, a process that was not affected by P2Y(12) or TXA(2) receptor deficiency, but was inhibited by the selective Src inhibitor PP2, the PKC inhibitor Ro-31-8220, or the calcium chelator demethyl-1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid tetrakis. Clot retraction was abolished, and platelet spreading on fibrinogen was diminished in Rap1b-deficient platelets compared with wild-type controls. The defects in clot retraction and spreading on fibrinogen of Rap1b-deficient platelets were not rescued by addition of MnCl(2), which elicits α(IIb)β(3) outside-in signaling in the absence of inside-out signaling. Thus, our results reveal two different activation mechanisms of Rap1b as well as novel functions of Rap1b in platelet secretion and in integrin α(IIb)β(3) outside-in signaling.  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