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 390 毫秒
1.
Kindlins are essential for integrin activation in cell systems and do so by working in a cooperative fashion with talin via their direct interaction with integrin β cytoplasmic tails (CTs). Kindlins interact with the membrane-distal NxxY motif, which is distinct from the talin-binding site within the membrane-proximal NxxY motif. The Tyr residues in both motifs can be phosphorylated, and it has been suggested that this modification of the membrane-proximal NxxY motif negatively regulates interaction with the talin head domain. However, the influence of Tyr phosphorylation of the membrane-distal NxxY motif on kindlin binding is unknown. Using mutational analyses and phosphorylated peptides, we show that phosphorylation of the membrane-distal NITY759 motif in the β3 CT disrupts kindlin-2 recognition. Phosphorylation of this membrane-distal Tyr also disables the ability of kindlin-2 to coactivate the integrin. In direct binding studies, peptides corresponding to the non-phosphorylated β3 CT interacted well with kindlin-2, whereas the Tyr759-phosphorylated peptide failed to bind kindlin-2 with measurable affinity. These observations indicate that transitions between the phosphorylated and non-phosphorylated states of the integrin β3 CT determine reactivity with kindlin-2 and govern the role of kindlin-2 in regulating integrin activation.  相似文献   

2.
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   

3.
4.
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   

5.
α3β1 integrin has been considered to be a mysterious adhesion molecule due to the pleiotropy in its ligand-binding specificity. However, recent studies have identified laminin isoforms as high-affinity ligands for this integrin, and demonstrated that α3β1 integrin plays a number of essential roles in development and differentiation, mainly by mediating the establishment and maintenance of epithelial tissues. Furthermore, α3β1 integrin is also implicated in many other biological phenomena, including cell growth and apoptosis, angiogenesis and neural functions. This integrin receptor forms complexes with various other membrane proteins, such as the transmembrane-4 superfamily proteins (tetraspanins), cytoskeletal proteins and signaling molecules. Recently, lines of evidence have been reported showing that complex formation regulates integrin functions in cell adhesion and migration, signal transduction across cell membranes, and cytoskeletal organization. In addition to these roles in physiological processes, α3β1 integrin performs crucial functions in various pathological processes, especially in wound healing, tumor invasion and metastasis, and infection by pathogenic microorganisms.This revised version was published online in June 2005 with a corrected cover date.  相似文献   

6.
Breast cancer, melanoma and glioblastoma cells undergo cell-mediated aggregation and aggregate coalescence in a transparent 3D Matrigel environment. Cells from normal tissue and non-tumorigenic cell lines do not exhibit these behaviors. Here, 266 monoclonal antibodies (mAbs) demonstrated to interact with a wide variety of membrane, secreted and matrix proteins, have been screened for their capacity to block these tumorigenic cell-specific behaviors in a 3D environment. Remarkably, only six of the 266 tested mAbs exhibited blocking activity, four targeting integrin ß-1, one targeting integrin α-3 and one targeting CD44. Colocalization of integrins ß-1 and α-3 in fixed cells and in live aggregates suggests that the integrin α-3 ß-1 dimer plays a central role in cancer cell aggregation in the 3D environment provided by Matrigel. Our results suggest that blocking by anti-integrin and anti-CD44 mAbs involves interference in cell-cell interactions.  相似文献   

7.
Polyunsaturated fatty acids, like arachidonic acid, can bind proteins and affect their function. The 14-3-3 proteins bind phosphorylated sites on a diverse array of client proteins and, in this way, are involved in many intracellular signaling pathways. In this study, we used a novel approach to discover that 14-3-3ζ is able to directly bind arachidonic acid. Furthermore, arachidonic acid, at physiological concentrations, reduced the binding of 14-3-3ζ to phosphorylated BAD, an interaction that is important in regulating apoptosis. In addition, high concentrations of arachidonic acid caused the polymerization of 14-3-3ζ, an event observed in neurodegenerative disorders. Taken together, these results indicate that arachidonic acid directly interacts with 14-3-3ζ and that this interaction may be important in both normal and pathological cellular events. If so, then factors that mediate the release, metabolism and reacylation of arachidonic acid into membranes represent key points of regulation.  相似文献   

8.
9.
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   

10.
Alport disease in humans, which usually results in proteinuria and kidney failure, is caused by mutations to the COL4A3, COL4A4, or COL4A5 genes, and absence of collagen α3α4α5(IV) networks found in mature kidney glomerular basement membrane (GBM). The Alport mouse harbors a deletion of the Col4a3 gene, which also results in the lack of GBM collagen α3α4α5(IV). This animal model shares many features with human Alport patients, including the retention of collagen α1α2α1(IV) in GBMs, effacement of podocyte foot processes, gradual loss of glomerular barrier properties, and progression to renal failure. To learn more about the pathogenesis of Alport disease, we undertook a discovery proteomics approach to identify proteins that were differentially expressed in glomeruli purified from Alport and wild-type mouse kidneys. Pairs of cy3- and cy5-labeled extracts from 5-week old Alport and wild-type glomeruli, respectively, underwent 2-dimensional difference gel electrophoresis. Differentially expressed proteins were digested with trypsin and prepared for mass spectrometry, peptide ion mapping/fingerprinting, and protein identification through database searching. The intermediate filament protein, vimentin, was upregulated ∼2.5 fold in Alport glomeruli compared to wild-type. Upregulation was confirmed by quantitative real time RT-PCR of isolated Alport glomeruli (5.4 fold over wild-type), and quantitative confocal immunofluorescence microscopy localized over-expressed vimentin specifically to Alport podocytes. We next hypothesized that increases in vimentin abundance might affect the basement membrane protein receptors, integrins, and screened Alport and wild-type glomeruli for expression of integrins likely to be the main receptors for GBM type IV collagen and laminin. Quantitative immunofluorescence showed an increase in integrin α1 expression in Alport mesangial cells and an increase in integrin α3 in Alport podocytes. We conclude that overexpression of mesangial integrin α1 and podocyte vimentin and integrin α3 may be important features of glomerular Alport disease, possibly affecting cell-signaling, cell shape and cellular adhesion to the GBM.  相似文献   

11.
The expression levels of integrin adhesion receptors have often been correlated with neoplastic transformation and invasiveness. To investigate more definitively the role of the integrin VLA-3 (α3β1) in tumor cell behavior, we transfected α3 subunit cDNA into human rhabdomyosarcoma (RD) cells. Transi ectants expressing high levels of α3β1, on their cell surface displayed an altered morphology and decreased anchorage-dependent growth in vitro. Cells expressing α3 also displayed marked reduction in anchorage-independent growth in soft agar and in their ability to form tumors when injected subcutaneously into athymic nude mice. Thus, VLA-3 can repress the transformed phenotype of rhabdomyosarcoma tumor cells. Similar changes in morphology and growth characteristics were observed in cells expressing a chimeric molecule X3C4 in which the α3 cytoplasmic domain had been exchanged with that of the α4 integrin subunit. Therefore, α3 inhibitory effects in RD cells appear not to require specific signalling through the α3 cytoplasmic domain.  相似文献   

12.
13.
14.
Abstract

The 5′-O-(4,4′-dimethoxytrityl) and 5′-O-(tert-butyldimethylsilyl) derivatives of 2′-,3′-O-thiocarbonyl-6-azauridine and 2′,3′-O-thiocarbonyl-5-chlorouridine were synthesized from the parent nucleosides by reaction with 4, 4′-dimethoxytrityl chloride and tert-butyldimethylsilyl chloride, respectively, followed by treatment with 1,1′-thiocarbonyldiimidazole. Introduction of a 2′-,3′-double bond into the sugar ring by reaction of the 5′-protected 2′-,3′-O-thionocarbonates with 1, 3-dimethyl-2-phenyl-1, 3, 2-diazaphospholidiine was unsuccessful, but could be accomplished satisfactorily with trimethyl phosphite. Reactions were generally more successful with the 5′-silylated than with the 5′-tritylated nucleosides. Formation of 2′-,3′-O-thiocarbonyl derivatives proceeded in higher yield with 5′-protected 6-azauridines than with the corresponding 5-chlorouridines because of the propensity of the latter to form 2,2′-anhydro derivatives. In the reaction of 5′-O-(tert-butyldimethylsilyl)-2′-,3′-O-thiocarbonyl-6-azauridine with trimethyl phosphite, introduction of the double bond was accompanied by N3-methylation. However this side reaction was not a problem with 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-O-thioarbonyl-5-chlorouridine. Treatment of 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-didehydro-2′-,3′-dideoxy-6-azauridine with tetrabutylammonium fluoride followed by hydrogenation afforded 2′-,3′-dideoxy-6-azauridine. Deprotection of 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-didehydro-2′-,3′-dideoxy-5-chlorouridine yielded 2′-,3′-didehydro-2′-,3′-dide-oxy-5-chlorouridine.  相似文献   

15.
Integrins are large membrane-spanning receptors fundamental to cell adhesion and migration. Integrin adhesiveness for the extracellular matrix is activated by the cytoskeletal protein talin via direct binding of its phosphotyrosine-binding-like F3 domain to the cytoplasmic tail of the β integrin subunit. The phosphotyrosine-binding domain of the signaling protein Dok1, on the other hand, has an inactivating effect on integrins, a phenomenon that is modulated by integrin tyrosine phosphorylation. Using full-length tyrosine-phosphorylated 15N-labeled β3, β1A, and β7 integrin tails and an NMR-based protein-protein interaction assay, we show that talin1 binds to the NPXY motif and the membrane-proximal portion of β3, β1A, and β7 tails, and that the affinity of this interaction is decreased by integrin tyrosine phosphorylation. Dok1 only interacts weakly with unphosphorylated tails, but its affinity is greatly increased by integrin tyrosine phosphorylation. The Dok1 interaction remains restricted to the integrin NPXY region, thus phosphorylation inhibits integrin activation by increasing the affinity of β integrin tails for a talin competitor that does not form activating membrane-proximal interactions with the integrin. Key residues governing these specificities were identified by detailed structural analysis, and talin1 was engineered to bind preferentially to phosphorylated integrins by introducing the mutation D372R. As predicted, this mutation affects talin1 localization in live cells in an integrin phosphorylation-specific manner. Together, these results indicate that tyrosine phosphorylation is a common mechanism for regulating integrin activation, despite subtle differences in how these integrins interact with their binding proteins.  相似文献   

16.
Integrin α9β1     
Integrins are transmembrane heterodimeric receptors responsible for transducing and modulating signals between the extracellular matrix and cytoskeleton, ultimately influencing cell functions such as adhesion and migration. Integrin α9β1 is classified within a two member sub-family of integrins highlighted in part by its specialized role in cell migration. The importance of this role is demonstrated by its regulation of numerous biological functions including lymphatic valve morphogenesis, lymphangiogenesis, angiogenesis and hematopoietic homeostasis. Compared to other integrins the signaling mechanisms that transduce α9β1-induced cell migration are not well described. We have recently shown that Src tyrosine kinase plays a key proximal role to control α9β1 signaling. Specifically it activates inducible nitric oxide synthase (iNOS) and in turn nitric oxide (NO) production as a means to transduce cell migration. Furthermore, we have also described a role for FAK, Erk and Rac1 in α9β1 signal transduction. Here we provide an over view of known integrin α9β1 signaling pathways and highlight its roles in diverse biological conditions.  相似文献   

17.
Integrins are members of a ubiquitous membrane receptor family which includes 18 different α subunits and 8 β subunits forming more than 20 α/β heterodimers. Integrins play key functions in vascular endothelial cell and tumour cell adhesion, lymphocyte trafficking, tumor growth and viral infection. Current understanding of the molecular basis of integrins as viral receptors has been achieved through many decades of study into the biology of transmembrane glycoproteins and their interactions with several viruses. This review provides a summary of the current knowledge on the molecular bases of interactions between viruses and integrins, which are of potential practical significance. Inhibition of virus-integrin interactions at the points of virus attachment or entry will provide a novel approach for the therapeutic treatment of viral diseases.  相似文献   

18.
A particulate enzyme preparation from Phaseolus aureus (mung bean) seedlings catalyzed the synthesis of a water insoluble β-1,3-glucan from UDP-α-d-glucose (UDPG) at high concentrations (0.4~20 mm) and an alkaline insoluble β-1,3 and β-1,4-mixed glucan from UDPG at a low concentration (8.5 µm).

Furthermore, the two kinds of β-glucan synthetases which were investigated with two reaction systems at high and low concentrations of UDPG had different properties in optimal pH, stability of enzyme activity, and metallic ion requirement.  相似文献   

19.
《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   

20.
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   

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