首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 0 毫秒
1.
Platelet glycerol lysis membranes and α-granule membranes were compared with respect to protein and lipid composition. Crossed immunoelectrophoresis using antibodies against whole platelets, and sodium dodecyl sulphate polyacrylamide gel electrophoresis, revealed the presence of the glycoproteins IIb and IIIa, myosin and an antigen termed G4 in both membrane fractions. The glycoproteins Ia, Ib and IIIb, in addition to β2-microglobulin and actin, appeared specific for the glycerol lysis membranes, whereas two antigens, termed G8 and G18, were observed only in the α-granule membranes. The localization of glycoprotein IIa was inconclusive. Comparison with the surface-located proteins revealed that the glycerol lysis membranes represented a reasonable approximation to a plasma membrane preparation. Radioactively labelled immunoprecipitates obtained after crossed immunoelectrophoresis of 125I-labelled platelets were cut out and applied to sodium dodecyl sulphate electrophoresis on polyacrylamide slab gels. Autoradiography of the dried gels revealed that antigen G4 represented a protein with an average molecular weight of 146 000 in its unreduced state and 132 000 in its reduced state. Antigen G18 represented a protein of molecular weight 130 000–135 000 in the reduced as well as unreduced state. Quantitation of protein and lipids showed that the α-granule membranes contained about one-third as much cholesterol and 2-times as much protein in relation to phospholipids as compared to the glycerol lysis membranes. No significant difference between the two membrane preparations was found as regards the composition of their phospholipids.  相似文献   

2.
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   

3.
A pathological hallmark of Alzheimer's disease (AD) is the aggregation of amyloid-β peptides (Aβ) into fibrils, leading to deposits in cerebral parenchyma and vessels known as cerebral amyloid angiopathy (CAA). Platelets are major players of hemostasis but are also implicated in AD. Recently we provided strong evidence for a direct contribution of platelets to AD pathology. We found that monomeric Aβ40 binds through its RHDS sequence to integrin αIIbβ3, and promotes the formation of fibrillar Aβ aggregates by the secretion of adenosine diphosphate (ADP) and the chaperone protein clusterin (CLU) from platelets. Here we investigated the molecular mechanisms of Aβ binding to integrin αIIbβ3 by using Aβ11 and Aβ16 peptides. These peptides include the RHDS binding motif important for integrin binding but lack the central hydrophobic core and the C-terminal sequence of Aβ. We observed platelet adhesion to truncated N-terminal Aβ11 and Aβ16 peptides that was not mediated by integrin αIIbβ3. Thus, no integrin outside-in signaling and reduced CLU release was detected. Accordingly, platelet mediated Aβ fibril formation was not observed. Taken together, the RHDS motif of Aβ is not sufficient for Aβ binding to platelet integrin αIIbβ3 and platelet mediated Aβ fibril formation but requires other recognition or binding motifs important for platelet mediated processes in CAA. Thus, increased understanding of the molecular mechanisms of Aβ binding to platelet integrin αIIbβ3 is important to understand the role of platelets in amyloid pathology.  相似文献   

4.
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   

5.
The effects of content of a fibrinogen receptor, glycoprotein (GP) IIb–IIIa (αIIb/β3-integrin), GP IIIa genetic polymorphism (substitution Leu33Pro), and fibrinogen concentration in blood plasma on platelet aggregation activity have been investigated in a group of healthy volunteers. In 35 examined donors the GP IIb–IIIa content on platelet surface varied from 40 to 71 × 103 per platelet. Repeated measurements revealed that the GP IIb–IIIa content coefficient of variation was 9.5%, and deviations from mean levels did not exceed 20%. The level and the rate of platelet aggregation induced by ADP (1.25–20 μM) correlated with GP IIb–IIIa number (r from 0.315 to 0.591) and were higher in the group of donors with high in comparison with low GP IIb–IIIa content (>60 and (40–50) × 10?3 per platelet, respectively). Aspirin, the inhibitor of thromboxane A2 synthesis, partially suppressed ADP-induced platelet aggregation. The level of residual aggregation in the presence of aspirin also correlated with GP IIb–IIIa content and increased in subjects with high receptor content. Parameters of ADP-induced aggregation did not differ in donors with genotypes GP IIIa Pro33(?) (Leu33Leu33, n = 20) and Pro33(+) (Leu33Pro33, n = 13, and Pro33Pro33, n = 2) genotype. GP IIb–IIIa content was also not affected by GP IIIa polymorphism. No significant correlations were found between the level and rate of platelet aggregation and fibrinogen concentration in blood plasma. The data obtained indicate that the effects of variations of GP IIb–IIIa content on platelet aggregation are higher than GP IIIa Leu33Pro polymorphism and variations of fibrinogen concentration. High GP IIb–IIIa content is associated with increased platelet aggregation activity and decreased efficacy of aggregation inhibition by aspirin.  相似文献   

6.
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   

7.
In addition to its classical CD40 receptor, CD154 also binds to αIIbβ3, α5β1, and αMβ2 integrins. Binding of CD154 to these receptors seems to play a key role in the pathogenic processes of chronic inflammation. This investigation was aimed at analyzing the functional interaction of CD154 with CD40, αIIbβ3, and α5β1 receptors. We found that the binding affinity of CD154 for αIIbβ3 is ~4-fold higher than for α5β1. We also describe the generation of sCD154 mutants that lost their ability to bind CD40 or αIIbβ3 and show that CD154 residues involved in its binding to CD40 or αIIbβ3 are distinct from those implicated in its interaction to α5β1, suggesting that sCD154 may bind simultaneously to different receptors. Indeed, sCD154 can bind simultaneously to CD40 and α5β1 and biologically activate human monocytic U937 cells expressing both receptors. The simultaneous engagement of CD40 and α5β1 activates the mitogen-activated protein kinases, p38, and extracellular signal-related kinases 1/2 and synergizes in the release of inflammatory mediators MMP-2 and -9, suggesting a cross-talk between these receptors.  相似文献   

8.
The RGD (Arg-Gly-Asp) binding integrins α(v)β(3) and α(IIb)β(3) are integral components of various pathological and physiological processes, including tumor angiogenesis, osteoclast function, and thrombus formation. Because of this, there is interest in identifying novel compounds and proteins binding to these receptors as well as investigating the mechanism of these interactions. In this article, we describe the development and validation of competition binding assays for determining the affinity of test compounds to α(v)β(3) and α(IIb)β(3) integrin. Assays were successfully developed for each receptor, and the affinity of known compounds was comparable to published results. However, the inability of binding between α(IIb)β(3) integrin and the labeled echistatin protein ligand to reach equilibrium resulted in an assay that did not meet the assumptions of the competition binding model. Nevertheless, there was good agreement between this assay and known literature values, and intra- and interassay variability was acceptable. Binding by conformation-specific antibodies provided evidence that solid-phase bound α(IIb)β(3) receptor was in an activated conformation. This study also demonstrated that current models and methods for determining receptor affinity are simplistic and fail to account for common receptor-ligand interactions such as nondissociable interactions and varying receptor activation states.  相似文献   

9.
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   

10.
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   

11.
12.
13.
A murine monoclonal antibody (mAb) 3C7 against integrin αIIbβ3 was previously obtained as a potential antithrombotic agent in our laboratory. The epitope of 3C7 is a specific conformation of the αIIbβ3 complex, but not either of the two subunits, which makes it different from abciximab, a supplementary antibody drug used in percutaneous coronary intervention which has a cross-reaction with other integrins sharing the β3 subunit. To reduce the human anti-mouse antibody reactions of 3C7, the variable regions of this antibody were cloned and fused with the constant counterparts of human IgG1. Two vectors of light and heavy chains were constructed and co-transfected into CHO-dhfr ? cells. The chimeric antibody c3C7 was purified and the properties of c3C7 were compared with 3C7. Identical to its parent antibody 3C7, c3C7 binds to the αIIbβ3 complex, but not to either of the subunits. The K d value of c3C7 was in the same order of magnitude as 3C7 (1.570?±?0.326 vs 0.780?±?0.182 nmol/L). Human platelet aggregation induced by adenosine diphosphate was effectively inhibited by c3C7 in a dose-dependent manner. In conclusion, after the modification, c3C7 retained the properties of its parent mAb with no loss of its biological activity. Therefore, c3C7 has the potential to become a novel agent for the treatment of thrombosis.  相似文献   

14.
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   

15.
Acidic extracellular pH is characteristic of the cell microenvironment in several important physiological and pathological contexts. Although it is well established that acidic extracellular pH can have profound effects on processes such as cell adhesion and migration, the underlying molecular mechanisms are largely unknown. Integrin receptors physically connect cells to the extracellular matrix, and are thus likely to modulate cell responses to extracellular conditions. Here, we examine the role of acidic extracellular pH in regulating activation of integrin α(v)β(3). Through computational molecular dynamics simulations, we find that acidic extracellular pH promotes opening of the α(v)β(3) headpiece, indicating that acidic pH can thereby facilitate integrin activation. This prediction is consistent with our flow cytometry and atomic force microscope-mediated force spectroscopy assays of integrin α(v)β(3) on live cells, which both demonstrate that acidic pH promotes activation at the intact cell surface. Finally, quantification of cell morphology and migration measurements shows that acidic extracellular pH affects cell behavior in a manner that is consistent with increased integrin activation. Taken together, these computational and experimental results suggest a new and complementary mechanism of integrin activation regulation, with associated implications for cell adhesion and migration in regions of altered pH that are relevant to wound healing and cancer.  相似文献   

16.
Dang S  Hong T  Wisniewski T  Zhang W 《PloS one》2011,6(10):e27012
Most antithrombotic approaches target prevention rather than the more clinically relevant issue of resolution of an existing thrombus. In this study, we describe a novel and effective therapeutic strategy for ex vivo clearance of pre-existing platelet thrombus by the combination of two bifunctional platelet GPIIIa49-66 ligands that target different parts of the arterial thrombus. We produced an additional GPIIIa49-66 agent (named APAC), which homes to activated platelets. Like our previously described SLK (which targets newly deposited fibrin strands surrounding the platelet thrombus), APAC destroys platelet aggregates ex vivo in an identical fashion with 85% destruction of platelet aggregates at 2 hours. The combined application of APAC and SLK demonstrated a ~2 fold greater platelet thrombus dissolution than either agent alone at a low concentration (0.025 μM). Platelet-rich clot lysis experiments demonstrated the time required for 50% platelet-rich fibrin clot lysis (T(50%)) by APAC (95 ± 6.1 min) or SLK (145 ± 7.1 min) was much longer than that by combined APAC + SLK (65 ± 7.6 min) at the final concentration of 0.025 μM (APAC + SLK vs APAC, p<0.05; APAC + SLK vs SLK, p<0.01). Thus these low concentrations of a combination of both agents are likely to be more effective and less toxic when used therapeutically in vivo.  相似文献   

17.
A mutual relationship exists between metastasizing tumor cells and components of the coagulation cascade. The exact mechanisms as to how platelets influence blood-borne metastasis, however, remain poorly understood. Here, we used murine B16 melanoma cells to observe functional aspects of how platelets contribute to the process of hematogenous metastasis. We found that platelets interfere with a distinct step of the metastasis cascade, as they promote adhesion of melanoma cells to the endothelium in vitro under shear conditions. Constitutively active platelet receptor GPIIb/IIIa (integrin αIIbβ3) expressed on Chinese hamster ovary cells promoted melanoma cell adhesion in the presence of fibrinogen, whereas blocking antibodies to aνβ3 integrin on melanoma cells or to GPIIb/IIIa significantly reduced melanoma cell adhesion to platelets. Furthermore, using intravital microscopy, we observed functional platelet-melanoma cell interactions, as platelet depletion resulted in significantly reduced melanoma cell adhesion to the injured vascular wall in vivo. Using a mouse model of hematogenous metastasis to the lung, we observed decreased metastasis of B16 melanoma cells to the lung by treatment with a mAb blocking the aν subunit of aνβ3 integrin. This effect was significantly reduced when platelets were depleted in vivo. Thus, the engagement of GPIIb/IIIa with aνβ3 integrin interaction mediates tumor cell-platelet interactions and highlights how this interaction is involved in hematogenous tumor metastasis.  相似文献   

18.
Koo KI  Bae JH  Lee CH  Yoon CD  Pyun JH  Shin SH  Jeon YC  Bae MK  Jang HO  Wood WG  Yun I 《Protoplasma》2008,234(1-4):3-12
Fluorescent probe techniques were used to evaluate the effect of bupivacaine.HCl on the physical properties (transbilayer asymmetric lateral and rotational mobilities, annular lipid fluidity and protein distribution) of synaptosomal plasma membrane vesicles (SPMVs) isolated from bovine cerebral cortex. An experimental procedure was used based on selective quenching of both 1,3-di(1-pyrenyl)propane (Py-3-Py) and 1,6-diphenyl-1,3,5-hexatriene (DPH) by trinitrophenyl groups, and radiationless energy transfer (RET) from the tryptophans of membrane proteins to Py-3-Py. Bupivacaine.HCl increased the bulk lateral and rotational mobilities, and annular lipid fluidity in SPMVs lipid bilayers, and had a greater fluidizing effect on the inner monolayer than that of the outer monolayer. The magnitude of increasing effect on annular lipid fluidity in SPMVs lipid bilayer induced by bupivacaine.HCl was significantly far greater than magnitude of increasing effect of the drug on the lateral and rotational mobilities of bulk SPMVs lipid bilayer. It also caused membrane proteins to cluster. These effects of bupivacaine.HCl on neuronal membranes may be responsible for some, though not all, of the local anesthetic actions of bupivacaine.HCl.  相似文献   

19.
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   

20.
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