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 375 毫秒
1.
We report the structure of an integrin with an αI domain, αXβ2, the complement receptor type 4. It was earlier expected that a fixed orientation between the αI domain and the β‐propeller domain in which it is inserted would be required for allosteric signal transmission. However, the αI domain is highly flexible, enabling two βI domain conformational states to couple to three αI domain states, and greater accessibility for ligand recognition. Although αXβ2 is bent similarly to integrins that lack αI domains, the terminal domains of the α‐ and β‐legs, calf‐2 and β‐tail, are oriented differently than in αI‐less integrins. Linkers extending to the transmembrane domains are unstructured. Previous mutations in the β2‐tail domain support the importance of extension, rather than a deadbolt, in integrin activation. The locations of further activating mutations and antibody epitopes show the critical role of extension, and conversion from the closed to the open headpiece conformation, in integrin activation. Differences among 10 molecules in crystal lattices provide unprecedented information on interdomain flexibility important for modelling integrin extension and activation.  相似文献   

2.
The interaction between the integrin α6β4 and plectin is essential for the assembly and stability of hemidesmosomes, which are junctional adhesion complexes that anchor epithelial cells to the basement membrane. We describe the crystal structure at 2.75 Å resolution of the primary α6β4–plectin complex, formed by the first pair of fibronectin type III domains and the N‐terminal region of the connecting segment of β4 and the actin‐binding domain of plectin. Two missense mutations in β4 (R1225H and R1281W) linked to nonlethal forms of epidermolysis bullosa prevent essential intermolecular contacts. We also present two structures at 1.75 and 2.05 Å resolution of the β4 moiety in the absence of plectin, which reveal a major rearrangement of the connecting segment of β4 on binding to plectin. This conformational switch is correlated with the way α6β4 promotes stable adhesion or cell migration and suggests an allosteric control of the integrin.  相似文献   

3.
Hematopoietic homeostasis depends on the maintenance of hematopoietic stem cells (HSCs), which are regulated within a specialized bone marrow (BM) niche. When HSC sense external stimuli, their adhesion status may be critical for determining HSC cell fate. The cell surface molecule, integrin αvβ3, is activated through HSC adhesion to extracellular matrix and niche cells. Integrin β3 signaling maintains HSCs within the niche. Here, we showed the synergistic negative regulation of the pro‐inflammatory cytokine interferon‐γ (IFNγ) and β3 integrin signaling in murine HSC function by a novel definitive phenotyping of HSCs. Integrin αvβ3 suppressed HSC function in the presence of IFNγ and impaired integrin β3 signaling mitigated IFNγ‐dependent negative action on HSCs. During IFNγ stimulation, integrin β3 signaling enhanced STAT1‐mediated gene expression via serine phosphorylation. These findings show that integrin β3 signaling intensifies the suppressive effect of IFNγ on HSCs, which indicates that cell adhesion via integrin αvβ3 within the BM niche acts as a context‐dependent signal modulator to regulate the HSC function under both steady‐state and inflammatory conditions.  相似文献   

4.
Platelet aggregation is the consequence of the binding of extracellular bivalent ligands such as fibrinogen and von Willebrand factor to the high affinity, active state of integrin αIIbβ3. This state is achieved through a so‐called “inside‐out” mechanism characterized by the membrane‐assisted formation of a complex between the F2 and F3 subdomains of intracellular protein talin and the integrin β3 tail. Here, we present the results of multi‐microsecond, all‐atom molecular dynamics simulations carried on the complete transmembrane (TM) and C‐terminal (CT) domains of αIIbβ3 integrin in an explicit lipid‐water environment, and in the presence or absence of the talin‐1 F2 and F3 subdomains. These large‐scale simulations provide unprecedented molecular‐level insights into the talin‐driven inside‐out activation of αIIbβ3 integrin. Specifically, they suggest a preferred conformation of the complete αIIbβ3 TM/CT domains in a lipid‐water environment, and testable hypotheses of key intermolecular interactions between αIIbβ3 integrin and the F2/F3 domains of talin‐1. Notably, not only do these simulations give support to a stable left‐handed reverse turn conformation of the αIIb juxtamembrane motif rather than a helical turn, but they raise the question as to whether TM helix separation is required for talin‐driven integrin activation. Proteins 2014; 82:3231–3240. © 2014 Wiley Periodicals, Inc.  相似文献   

5.
Previous reports indicated that integrins associated signals are tightly related to tumor progression. Here, we observed elevated expression of integrin α2β1 in tumor tissues from microtubule‐directed chemotherapeutic drugs (MDCDs) resistant patients compared with the samples from chemosensitive patients. More importantly, we sorted the integrin α2β1+ tumor cells and found those cells revealed high MDCDs resistance, whereas MDCDs shows effective cytotoxicity to those integrin α2β1? tumor cells in vitro and in vivo. Mechanistically, we demonstrated that integrin α2β1 could induce MDCDs resistance through the activation of the PI3K/AKT pathway. Applying MPEG‐PLA to co‐encapsulate the integrin α2β1 inhibitor E7820 and MDCDs could effectively reverse MDCDs resistance, resulting in enhanced anticancer effects while avoiding potential systemic toxicity in vitro and in vivo. In conclusion, the expression of integrin α2β1 contributes to MDCDs resistance, while applying E7820 combination treatment by MPEG‐PLA nanoparticles could reverse the resistance.  相似文献   

6.
7.
During platelet activation, fibrinogen binds to its specific platelet receptor, integrin αIIbβ3, thus completing the final common pathway for platelet aggregation. Norcantharidin (NCTD) is a promising anticancer agent in China from medicinal insect blister beetle. In this study, we provided the evidence to demonstrate NCTD (0.1–1.0 μM) possesses very powerful antiplatelet activity in human platelets; nevertheless, it had no effects on surface P‐selectin expression and only slight inhibition on ATP‐release reaction in activated platelets. Moreover, NCTD markedly hindered integrin αIIbβ3 activation by interfering with the binding of FITC‐labelled PAC‐1. It also markedly reduced the number of adherent platelets and the single platelet spreading area on immobilized fibrinogen as well as clot retraction. Additionally, NCTD attenuated phosphorylation of proteins such as integrin β3, Src and FAK in platelets spreading on immobilized fibrinogen. These results indicate that NCTD restricts integrin αIIbβ3‐mediated outside‐in signalling in human platelets. Besides, NCTD substantially prolonged the closure time in human whole blood and increased the occlusion time of thrombotic platelet plug formation and prolonged the bleeding time in mice. In conclusion, NCTD has dual activities, it can be a chemotherapeutic agent for cancer treatment, and the other side it possesses powerful antiplatelet activity for treating thromboembolic disorders.  相似文献   

8.
Integrins are a family of heterodimeric cell adhesion receptors expressed on most cells and are involved in many cellular functions including phagocytosis, a process by which professional phagocytes recognise, bind and internalise foreign materials larger than 0.5 µm in diameter. An example of a phagocytic integrin receptor is αMβ2, and this review seeks to provide fresh insights into the current knowledge of this subject. Key areas that this review will emphasise include, the classical understanding of bi‐directional signalling to and from αMβ2 (aka inside‐out and outside‐in signalling, respectively). For inside‐out signalling, we will review the involvement of the small GTPase, Rap1, FERM‐containing proteins such as talin and kindlin‐3, some of the kinases, and the GEF, cytohesin‐1 and vasodilator‐stimulated phosphoprotein (VASP). We also summarise studies into outside‐in signalling, focussing on the roles of RhoA and RhoG, and activation of Rac1 through the complex comprising TIAM, 14‐3‐3 and β2. We will also consider non‐classical signalling processes, which include integrin clustering and membrane ruffling. Through this review, we hope to highlight the importance of αMβ2 signalling mechanisms and their relevance to other integrin‐mediated events.  相似文献   

9.
We report here an analysis of the expression and function of the α chain of human VLA-4 in stable mouse L cell transfectants and the requirement for the β chain in these processes. L cells were transfected with human α4 cDNA or α4 and human β1 cDNA. Unexpectedly, human α4 cDNA, when transfected alone, could induce de novo surface expression of host β7 and increased expression of host β1. Induction of mouse β7 and β1 surface expression was not due to de novo gene activation, but instead represented α4/β intracellular subunit association and transport to the cell surface. Transfection with human β1 prevented surface expression of mouse β integrins. Whereas human α4 and human β1 subunits associated very tightly in anti-α4 immunoprecipitates, human α4 and mouse β subunits were only partially associated. Furthermore, binding of human/mouse chimeric receptors to recombinant VCAM, a major ligand for α4β7 and α4β1, was very poor, whereas human α4/human β1 receptors bound strongly to VCAM. One α4 transfectant, which exhibited a tight human α4/mouse β1 association, could be induced, but only after PMA activation, to bind strongly to VCAM. These results indicate that α4 subunits have specific affinity for β7 and β1 integrins and require β subunits for surface expression as well as high affinity ligand binding activity. Our results indicate that a tight association between the α4 and β subunit appears to be critical for ligand binding, consistent with a direct as well as regulatory role for the β subunit in ligand binding. Furthermore, these studies demonstrate that expression of foreign recombinant proteins can alter host cell protein expression resulting in de novo surface protein expression. © 1996 Wiley-Liss, Inc.  相似文献   

10.
Adhesion to collagens by most cell types is mediated by the integrins α1β1 and α2β1. Both integrin α subunits belong to a group which is characterized by the presence of an I domain in the N-terminal half of the molecule, and this domain has been implicated in the ligand recognition. Since purified α1β1 and α2β1 differ in their binding to collagens I and IV and recognize different sites within the major cell binding domain of collagen IV, we investigated the potential role of the α1 and α2 I domains in specific collagen adhesion. We find that introducing the α2 I domain into α1 results in surface expression of a functional collagen receptor. The adhesion mediated by this chimeric receptor (α1-2-1β1) is similar to the adhesion profile conferred by α2β1, not α1β1. The presence of α2 or α1-2-1 results in preferential binding to collagen I, whereas α1 expressing cells bind better to collagen IV. In addition, α1 containing cells bind to low amounts of a tryptic fragment of collagen IV, whereas α2 or α1-2-1 bearing cells adhere only to high concentrations of this substrate. We also find that collagen adhesion of NIH-3T3 mediated by α2β1 or α1-2-1β1, but not by α1, requires the presence of Mn2+ ions. This ion requirement was not found in CHO cells, implicating the I domain in cell type-specific activation of integrins. J. Cell. Physiol. 176:634–641, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

11.
Gamma‐aminobutyric acid type A receptors (GABAARs) are the most important inhibitory chloride ion channels in the central nervous system and are major targets for a wide variety of drugs. The subunit compositions of GABAARs determine their function and pharmacological profile. GABAARs are heteropentamers of subunits, and (α1)2(β3)2(γ2L)1 is a common subtype. Biochemical and biophysical studies of GABAARs require larger quantities of receptors of defined subunit composition than are currently available. We previously reported high‐level production of active human α1β3 GABAAR using tetracycline‐inducible stable HEK293 cells. Here we extend the strategy to receptors containing three different subunits. We constructed a stable tetracycline‐inducible HEK293‐TetR cell line expressing human (N)–FLAG–α1β3γ2L–(C)–(GGS)3GK–1D4 GABAAR. These cells achieved expression levels of 70–90 pmol [3H]muscimol binding sites/15‐cm plate at a specific activity of 15–30 pmol/mg of membrane protein. Incorporation of the γ2 subunit was confirmed by the ratio of [3H]flunitrazepam to [3H]muscimol binding sites and sensitivity of GABA‐induced currents to benzodiazepines and zinc. The α1β3γ2L GABAARs were solubilized in dodecyl‐d ‐maltoside, purified by anti‐FLAG affinity chromatography and reconstituted in CHAPS/asolectin at an overall yield of ~30%. Typical purifications yielded 1.0–1.5 nmoles of [3H]muscimol binding sites/60 plates. Receptors with similar properties could be purified by 1D4 affinity chromatography with lower overall yield. The composition of the purified, reconstituted receptors was confirmed by ligand binding, Western blot, and proteomics. Allosteric interactions between etomidate and [3H]muscimol binding were maintained in the purified state.  相似文献   

12.
Heparin/heparan sulfate interact with growth factors, chemokines, extracellular proteins, and receptors. Integrins are αβ heterodimers that serve as receptors for extracellular proteins, regulate cell behavior, and participate in extracellular matrix assembly. Heparin binds to RGD‐dependent integrins (αIIbβ3, α5β1, αvβ3, and αvβ5) and to RGD‐independent integrins (α4β1, αXβ2, and αMβ2), but their binding sites have not been located on integrins. We report the mapping of heparin binding sites on the ectodomain of αvβ3 integrin by molecular modeling. The surface of the ectodomain was scanned with small rigid probes mimicking the sulfated domains of heparan sulfate. Docking results were clustered into binding spots. The best results were selected for further docking simulations with heparin hexasaccharide. Six potential binding spots containing lysine and/or arginine residues were identified on the ectodomain of αvβ3 integrin. Heparin would mostly bind to the top of the genu domain, the Calf‐I domain of the α subunit, and the top of the β subunit of RGD‐dependent integrins. Three spots were close enough from each other on the integrin surface to form an extended binding site that could interact with heparin/heparan sulfate chains. Because heparin does not bind to the same integrin site as protein ligands, no steric hindrance prevents the formation of ternary complexes comprising the integrin, its protein ligand, and heparin/heparan sulfate. The basic amino acid residues predicted to interact with heparin are conserved in the sequences of RGD‐dependent but not of RGD‐independent integrins suggesting that heparin/heparan sulfate could bind to different sites on these two integrin subfamilies. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

13.
The small GTPase Rap1 and the cytoskeletal protein talin regulate binding of C3bi‐opsonised red blood cells (RBC) to integrin αMβ2 in phagocytic cells, although the mechanism has not been investigated. Using COS‐7 cells transfected with αMβ2, we show that Rap1 acts on the β2 and not the αM chain, and that residues 732–761 of the β2 subunit are essential for Rap1‐induced RBC binding. Activation of αMβ2 by Rap1 was dependent on W747 and F754 in the β2 tails, which are required for talin head binding, suggesting a link between Rap1 and talin in this process. Using talin1 knock‐out cells or siRNA‐mediated talin1 knockdown in the THP‐1 monocytic cell line, we show that Rap1 acts upstream of talin but surprisingly, RIAM knockdown had little effect on integrin‐mediated RBC binding or cell spreading. Interestingly, Rap1 and talin influence each other's localisation at phagocytic cups, and co‐immunoprecipitation experiments suggest that they interact together. These results show that Rap1‐mediated activation of αMβ2 in macrophages shares both common and distinct features from Rap1 activation of αIIbβ3 expressed in CHO cells. J. Cell. Biochem. 111: 999–1009, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

14.
Antiphospholipid syndrome (APS) is characterized by thrombosis and the presence of antiphospholipid antibodies (aPL) that directly recognizes plasma β2‐glycoprotein I (β2GPI). Tissue factor (TF), the major initiator of the extrinsic coagulation system, is induced on monocytes by aPL in vitro, explaining in part the pathophysiology in APS. We previously reported that the mitogen‐activated protein kinase (MAPK) pathway plays an important role in aPL‐induced TF expression on monocytes. In this study, we identified plasma gelsolin as a protein associated with β2GPI by using immunoaffinity chromatography and mass spectrometric analysis. An in vivo binding assay showed that endogenous β2GPI interacts with plasma gelsolin, which binds to integrin a5β1 through fibronectin. The tethering of β2GPI to monoclonal anti‐β2GPI autoantibody on the cell surface was enhanced in the presence of plasma gelsolin. Immunoblot analysis demonstrated that p38 MAPK protein was phosphorylated by monoclonal anti‐β2GPI antibody treatment, and its phosphorylation was attenuated in the presence of anti‐integrin a5β1 antibody. Furthermore, focal adhesion kinase, a downstream molecule of the fibronectin‐integrin signalling pathway, was phosphorylated by anti‐β2GPI antibody treatment. These results indicate that molecules including gelsolin and integrin are involved in the anti‐β2GPI antibody‐induced MAPK pathway on monocytes and that integrin is a possible therapeutic target to modify a prothrombotic state in patients with APS.  相似文献   

15.
A functional proteomic technology using protein chip and molecular simulation was used to demonstrate a novel biomolecular interaction between P11, a peptide containing the Ser‐Asp‐Val (SDV) sequence and integrin αvβ3. P11 (HSDVHK) is a novel antagonistic peptide of integrin αvβ3 screened from hexapeptide library through protein chip system. An in silico docking study and competitive protein chip assay revealed that the SDV sequence of P11 is able to create a stable inhibitory complex onto the vitronectin‐binding site of integrin αvβ3. The Arg‐Gly‐Asp (RGD)‐binding site recognition by P11 was site specific because the P11 was inactive for the complex formation of a denatured form of integrin–vitronectin. P11 showed a strong antagonism against αvβ3‐GRGDSP interaction with an IC50 value of 25.72±3.34 nM, whereas the value of GRGDSP peptide was 1968.73±444.32 nM. The binding‐free energies calculated from the docking simulations for each P11 and RGD peptide were ?3.99 and ?3.10 kcal/mol, respectively. The free energy difference between P11 and RGD corresponds to approximately a 4.5‐fold lower Ki value for the P11 than the RGD peptide. The binding orientation of the docked P11 was similar to the crystal structure of the RGD in αvβ3. The analyzed docked poses suggest that a divalent metal–ion coordination was a common driving force for the formation of both SDV/αvβ3 and RGD/αvβ3 complexes. This is the first report on the specific recognition of the RGD‐binding site of αvβ3 by a non‐RGD containing peptide using a computer‐assisted proteomic approach.  相似文献   

16.
17.
Disruption of the gut barrier results in severe clinical outcomes with no specific treatment. Metabolic disorders and destruction of enterocytes play key roles in gut barrier dysfunction. Irisin is a newly identified exercise hormone that regulates energy metabolism. However, the effect of irisin on gut barrier function remains unknown. The therapeutic effect of irisin on gut barrier dysfunction was evaluated in gut ischemia reperfusion (IR). The direct effect of irisin on gut barrier function was studied in Caco‐2 cells. Here, we discovered that serum and gut irisin levels were decreased during gut IR and that treatment with exogenous irisin restored gut barrier function after gut IR in mice. Meanwhile, irisin decreased oxidative stress, calcium influx and endoplasmic reticulum (ER) stress after gut IR. Moreover, irisin protected mitochondrial function and reduced enterocyte apoptosis. The neutralizing antibody against irisin significantly aggravated gut injury, oxidative stress and enterocyte apoptosis after gut IR. Further studies revealed that irisin activated the AMPK‐UCP 2 pathway via binding to the integrin αVβ5 receptor. Inhibition of integrin αVβ5, AMPK or UCP 2 abolished the protective role of irisin in gut barrier function. In conclusion, exogenous irisin restores gut barrier function after gut IR via the integrin αVβ5‐AMPK‐UCP 2 pathway.  相似文献   

18.
Malignant cell transformation is generally accompanied by changes in their interactions with environing matrix proteins in a way to facilitate their migration and generate invasion. Our results show the binding of rat colon adenocarcinoma PROb cells to fibronectin strongly reduced when compared to normal rat intestine epithelial cells. This decrease was not due to the level of α(s)β1 integrins expressed at the surface of the cell line. However, β1- and α(s)-associated subunits appeared to be structurally altered as shown by immunoprecipitation followed by electrophoresis. Pulse chase experiments using 35S methionine evidenced differences in the biosynthesis of β1- and α (s) associated integrins: normal epithelial IEC18 cells required 16 h for maximal biosynthesis of the completely mature β1 subunit, while PROb cells did it within 4-6 h. Studies using endoglycosidases O, H, D, and N glycanase confirmed that the molecular weight alterations were due to abnormal glycosylation and suggested that α(s)β1 integrins of PROb cells could bear both mature complex and immature high mannose types while IEC18 cells borne only mature complex type oligosaccharidic chains. Treatment of both cell types with castanospermine, an inhibitor of N-glycosylation, reduced the differences observed in their adhesion to the fibronectin without significantly affecting β1 receptors expression at the cell surface. These results strongly suggest a role of the glycosylation of β1 receptors in the adhesion of rat colon adenocarcinoma PROb cells to fibronectin substrata. © 1996 Wiley-Liss, Inc.  相似文献   

19.
How to target cancer cells with high specificity and kill cancer cells with high efficiency remains an urgent demand for anticancer drugs. Temporin‐La, which belongs to the family of temporins, presents antitumor activity against many cancer cell lines. We first used a whole bioinformatic analysis method as a platform to identify new anticancer antimicrobial peptides (AMPs). On the basis of these results, we designed a temporin‐La analog (temporin‐Las) and related constructs containing the Arg‐Gly‐Asp (RGD) tripeptide, the integrin αvβ3 homing domain (RGD‐La and RGD‐Las). We detected a link between the net charges and integrin αvβ3 expression of cancer cell lines and the antitumor activities of these peptides. Temporin‐La and its synthetic analogs inhibited cancer cell proliferation in a dose‐dependent manner. Evidence was provided that the affinity between RGD‐Las and tumor cell membranes was stronger than other tested peptides using a pull‐down assay. Morphological changes on the cell membrane induced by temporin‐La and RDG‐Las, respectively, were examined by scanning electron microscopy. Additionally, time‐dependent morphological changes were detected by confocal microscopy, where the binding process of RGD‐Las to the cell membrane could be monitored. The results indicate that the electrostatic interaction between these cationic peptides and the anionic cell membrane is a major determinant of selective cell killing. Thus, the RGD tripeptide is a valuable ligand motif for tumor targeting, which leads to an increased anticancer efficiency by RGD‐Las. These AMP‐derived peptides have clinical potential as specifically targeting agents for the treatment of αvβ3 positive tumors. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
One of the hallmarks of cells undergoing mitotic division is their rounded morphology and reduced adhesion to the substratum. We have studied and compared the attachment of interphase and mitotic cells to substrata coated with fibronectin and vitronectin. We have found that adhesion of mitotic cells, as compared to interphase cells, is significantly reduced to fibronectin, but is higher to vitronectin. These results correlate well with the expression of α5β1 and αVβ3 integrins, the respective receptors for fibronectin and vitronectin, on the cell surface. Mitotic cells show higher levels of αVβ3 and very low levels of α5β1 proteins on the cell surface as compared to interphase cells. This difference in the levels of these integrins also reflects in the total amounts of fibronectin and vitronectin present on the cell surface of these cells. We have further shown, by flow cytometry, that binding of vitronectin, or the synthetic peptide-GRGDSP-, causes an increase in the intracellular levels of Ca2− in mitotic cells, but no change is seen in the interphase cells. Binding of fibronectin to either of these cells fails to elicit any response. One interesting feature of our results is that the levels of total, i.e., cytoplasmic plus membrane bound, α5β1 and αVβ3 integrins of mitotic and interphase cells remain the same, thus implying an alteration in the distribution of integrin chains between the plasma membrane and the cytoplasm during the conversion of interphase cells into the mitotic phase. © 1996 Wiley-Liss, Inc.  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