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 31 毫秒
1.
2.
Andes virus (ANDV) causes a fatal hantavirus pulmonary syndrome (HPS) in humans and Syrian hamsters. Human αvβ3 integrins are receptors for several pathogenic hantaviruses, and the function of αvβ3 integrins on endothelial cells suggests a role for αvβ3 in hantavirus directed vascular permeability. We determined here that ANDV infection of human endothelial cells or Syrian hamster-derived BHK-21 cells was selectively inhibited by the high-affinity αvβ3 integrin ligand vitronectin and by antibodies to αvβ3 integrins. Further, antibodies to the β3 integrin PSI domain, as well as PSI domain polypeptides derived from human and Syrian hamster β3 subunits, but not murine or bovine β3, inhibited ANDV infection of both BHK-21 and human endothelial cells. These findings suggest that ANDV interacts with β3 subunits through PSI domain residues conserved in both Syrian hamster and human β3 integrins. Sequencing the Syrian hamster β3 integrin PSI domain revealed eight differences between Syrian hamster and human β3 integrins. Analysis of residues within the PSI domains of human, Syrian hamster, murine, and bovine β3 integrins identified unique proline substitutions at residues 32 and 33 of murine and bovine PSI domains that could determine ANDV recognition. Mutagenizing the human β3 PSI domain to contain the L33P substitution present in bovine β3 integrin abolished the ability of the PSI domain to inhibit ANDV infectivity. Conversely, mutagenizing either the bovine PSI domain, P33L, or the murine PSI domain, S32P, to the residue present human β3 permitted PSI mutants to inhibit ANDV infection. Similarly, CHO cells transfected with the full-length bovine β3 integrin containing the P33L mutation permitted infection by ANDV. These findings indicate that human and Syrian hamster αvβ3 integrins are key receptors for ANDV and that specific residues within the β3 integrin PSI domain are required for ANDV infection. Since L33P is a naturally occurring human β3 polymorphism, these findings further suggest the importance of specific β3 integrin residues in hantavirus infection. These findings rationalize determining the role of β3 integrins in hantavirus pathogenesis in the Syrian hamster model.Hantaviruses persistently infect specific small mammal hosts and are spread to humans by the inhalation of aerosolized excreted virus (41, 42). Hantaviruses predominantly infect endothelial cells and cause one of two vascular leak-based diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (41). Hantavirus diseases are characterized by increased vascular permeability and acute thrombocytopenia in the absence of endothelial cell lysis (36, 41, 42, 54). In general, hantaviruses are not spread from person to person; however, the Andes hantavirus (ANDV) is an exception, since there are several reports of person-to-person transmission of ANDV infection (11, 37, 47, 52). ANDV is also unique in its ability to cause an HPS-like disease in Syrian hamsters and serves as the best-characterized hantavirus disease model with a long onset, symptoms, and pathogenesis nearly identical to that of HPS patients (20, 21, 50).Hantavirus infection of the endothelium alters endothelial cell barrier functions through direct and immunological responses (8, 14). Although the means by which hantaviruses cause pulmonary edema or hemorrhagic disease has been widely conjectured, the mechanisms by which hantaviruses elicit pathogenic human responses have yet to be defined. Hantaviruses coat the surface of infected VeroE6 cells days after infection (17), and this further suggests that dynamic hantavirus interactions with immune and endothelial cells are likely to contribute to viral pathogenesis. Hantavirus pathogenesis has been suggested to involve CD8+ T cells, tumor necrosis factor alpha or other cytokines, viremia, and the dysregulation of β3 integrins (7, 8, 13-16, 25-28, 32, 34, 38, 44-46). However, these responses have not been demonstrated to contribute to hantavirus pathogenesis, and in some cases there are conflicting data on their involvement (18, 25-28, 34, 35, 44, 45, 48). Immune complex deposition clearly contributes to HFRS patient disease and renal sequelae (4, 7), but it is unclear what triggers vascular permeability in HPS and HFRS diseases or why hemorrhage occurs in HFRS patients but not in HPS patients (8, 36, 54). Acute thrombocytopenia is common to both diseases, and platelet dysfunction resulting from defective platelet aggregation is reported in HFRS patients (7, 8).Pathogenic hantaviruses have in common their ability to interact with αIIbβ3 and αvβ3 integrins present on platelets and endothelial cells (13, 16), and β3 integrins have primary roles in regulating vascular integrity (1, 2, 6, 19, 22, 39, 40). Consistent with the presence of cell surface displayed virus (17), pathogenic hantaviruses uniquely block αvβ3 directed endothelial cell migration and enhance endothelial cell permeability for 3 to 5 days postinfection (14, 15). Pathogenic hantaviruses dysregulate β3 integrin functions by binding domains present at the apex of inactive β3 integrin conformers (38). αvβ3 forms a complex with vascular endothelial cell growth factor receptor 2 (VEGFR2) and normally regulates VEGF-directed endothelial cell permeability (2, 3, 10, 39, 40). However, both β3 integrin knockouts and hantavirus-infected endothelial cells result in increased VEGF-induced permeability, presumably by disrupting VEGFR2-β3 integrin complex formation (2, 14, 19, 39, 40). This suggests that at least one means for hantaviruses to increase vascular permeability occurs through interactions with β3 integrins that are required for normal platelet and endothelial cell functions.αvβ3 and αIIbβ3 integrins exist in two conformations: an active extended conformation where the ligand binding head domain is present at the apex of the heterodimer and a basal, inactive bent conformation where the globular head of the integrin is folded toward the cell membrane (30, 53, 55). Pathogenic HTN and NY-1 hantaviruses bind to the N-terminal plexin-semaphorin-integrin (PSI) domain of β3 integrin subunits and are selective for bent, inactive αvβ3 integrin conformers (38). Pathogenic hantavirus binding to inactive αvβ3 integrins is consistent with the selective inhibitory effect of hantaviruses on αvβ3 function and endothelial cell permeability (14, 15, 38). Although the mechanism of hantavirus induced vascular permeability has yet to be defined, there is a clear role for β3 integrin dysfunction in vascular permeability deficits (5, 6, 22, 29, 39, 40, 51) which make an understanding of hantavirus interactions with β3 subunits important for both entry and disease processes.The similarity between HPS disease in humans and Syrian hamsters (20, 21) suggests that pathogenic mechanisms of ANDV disease are likely to be coincident. Curiously, other hantaviruses (Sin Nombre virus [SNV] and Hantaan virus [HTNV]) are restricted in Syrian hamsters and fail to cause disease in this animal, even though they are prominent causes of human disease (50). Although the host range restriction for SNV and HTNV in Syrian hamsters has not been defined (33), the pathogenesis of ANDV in Syrian hamsters suggests that both human and Syrian hamster β3 integrins may similarly be used by ANDV and contribute to pathogenesis.We demonstrate here that ANDV infection of the Syrian hamster BHK-21 cell line and human endothelial cells is dependent on αvβ3 and inhibited by αvβ3 specific ligands and antibodies. Further, polypeptides expressing the N-terminal 53 residues of human and Syrian hamster β3 subunits block ANDV infection. This further indicates that ANDV interaction with the N-terminal 53 residues of both human and Syrian hamster β3 integrins is required for viral entry. We also demonstrate that ANDV recognition of human and Syrian hamster β3 integrins is determined by proline substitutions at residues 32/33 within the β3 integrin PSI domain. These results define unique ANDV interactions with human and Syrian hamster β3 integrins.  相似文献   

3.
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   

4.
Ovarian cancer often has few symptoms, which makes it difficult to detect at an early stage. Therefore, most of the women will already have metastasis at the time of diagnosis. In their search of undercovering the mechanisms underlying ovarian cancer invasion, Mitra and collaborators demonstrate that the fibronectin receptor (α5β1-integrin) can directly activate the receptor tyrosine kinase Met, independently of its ligand. By linking the extracellular matrix with Met activation, and the invasion of ovarian cancer cells, Mitra et al. confirm the crucial role played by Met in ovarian cancer and open new perspectives in the development of ovarian cancer targeted therapies.Key words: Met, independent-ligand activation, ovarian cancer metastasis, integrin, fibronectin, FAKWith 21,880 new cases estimated in 2010 and 13,850 deaths, ovarian cancer is one of the most lethal cancers for women in the United States (National Cancer Institute, www.cancer.gov/cancertopics/types/ovarian). Because most of the deaths occurred at advance stages of the disease, it is crucial to develop a better understanding of the mechanisms leading to the spreading of ovarian tumor cells. This has been the focus of the Ernst Lengyel laboratory for several years now.In their recent work published in Oncogene, Mitra et al. demonstrate a novel pathway linking the fibronectin receptor (α5β1-integrin) to Met.1 Both α5-integrin and Met have previously been reported to induce invasion in models of ovarian cancer invasion, and to be potential therapeutic targets.2,3 In the path of several reports showing that the receptor tyrosine kinase Met can be activated by other receptors tyrosine kinases (RTK)4, this report demonstrates that α5-integrin can directly activate Met, independently of its ligand, the hepatocyte growth factor/scatter factor (HGF/SF).  相似文献   

5.
A structure of adenovirus type 12 (HAdV12) complexed with a soluble form of integrin αvβ5 was determined by cryo-electron microscopy (cryoEM) image reconstruction. Subnanometer resolution (8 Å) was achieved for the icosahedral capsid with moderate resolution (27 Å) for integrin density above each penton base. Modeling with αvβ3 and αIIbβ3 crystal structures indicates that a maximum of four integrins fit over the pentameric penton base. The close spacing (∼60 Å) of the RGD protrusions on penton base precludes integrin binding in the same orientation to neighboring RGD sites. Flexible penton-base RGD loops and incoherent averaging of bound integrin molecules explain the moderate resolution observed for the integrin density. A model with four integrins bound to a penton base suggests that integrin might extend one RGD-loop in the direction that could induce a conformational change in the penton base involving clockwise untwisting of the pentamer. A global conformational change in penton base could be one step on the way to the release of Ad vertex proteins during cell entry. Comparison of the cryoEM structure with bent and extended models for the integrin ectodomain reveals that integrin adopts an extended conformation when bound to the Ad penton base, a multivalent viral ligand. These findings shed further light on the structural basis of integrin binding to biologically relevant ligands, as well as on the molecular events leading to HAdV cell entry.A growing number of viruses have been identified as using one of the 24 types of integrin heterodimers as a receptor for cell entry (32). Integrins are cell surface molecules involved in the regulation of adhesion, migration, growth, and differentiation (11). The large multidomained extracellular segments of α and β integrin subunits bind a variety of ligands, including viral ligands, while the smaller intracellular domains interact with cytoskeletal proteins (Fig. (Fig.1A).1A). These extracellular and intracellular interactions facilitate bidirectional signaling, with the initiating events occurring either outside of the cell (outside-in signaling) or within the cell (inside-out signaling) (24). Integrin clustering has been established as having an important role in outside-in signaling (9, 19, 20, 44). Clustering results in the formation of focal adhesions, which are organized intracellular complexes, that facilitate downstream signaling cascades within the cell (24).Open in a separate windowFIG. 1.Integrin domains and conformations. (A) Structural domains of integrin αv and β chains, including the extracellular domains, transmembrane-spanning regions, and small cytoplasmic domains, shown in extended schematic forms. The domains are represented as 10Å-resolution density maps based on crystallographic coordinates. The membrane is represented by a gray bar. (Modified from Stewart and Nemerow (32) and reprinted with permission from Elsevier.) (B) Models for soluble αvβ5 integrin with Fos/Jun dimerization domains. Each chain has a six residue glycine-rich linker between the ectodomain and the Fos or Jun dimerization domain. The model of a bent integrin conformation (left) was built as a composite of αvβ3 integrin crystal structures, PDB-IDs 1L5G and 1U8C (42, 43), and the crystal structure of c-Fos/c-Jun bound to DNA, PDB-ID 1FOS (6). The model of an extended integrin conformation (right) is similar to the extended model docked into the HAdV12/αvβ5 cryo structure (Fig. (Fig.8B8B).Studies of adenovirus (Ad) interactions with αv integrins provided some of the first evidence of the virus-induced signaling events (13, 14). The Ad penton base capsid protein, which sits at the 12 vertices of the icosahedral capsid, has five prominent Arg-Gly-Asp (RGD) containing loops that are flexible and protrude from the viral surface (31, 48). Receptor-mediated endocytosis of Ad is stimulated by interaction of the RGD-containing penton base with αvβ3 and αvβ5 integrins (34). This interaction leads to receptor clustering, followed by tyrosine phosphorylation/activation of focal adhesion kinase, as well as activation of p130CAS, phosphatidylinositol 3-OH-kinase, and the Rho family of small GTPases, and subsequent actin polymerization and Ad internalization (32). Integrin signaling events also lead to production of proinflammatory cytokines (23) and may result in increased survival of certain host cells through subsequent signaling to protein kinase B (AKT) (25).Multiple studies indicate that after interaction with an RGD-containing ligand a straightening of the integrin extracellular domains occurs, leading to the “extension” or “switchblade” model for integrin activation (16, 45). In the extension model the headpiece domains, which are closest to the RGD interaction site, have a “closed” conformation in the low-affinity, unliganded state. This state is characterized by the close proximity of the α and β subunits at the “knees” or midpoints of the extracellular segments. In contrast, the high-affinity, ligand-bound state in the extension model is distinguished by an “open” headpiece conformation with separation at the knees of the extracellular segments. The location of the RGD binding site between the α-subunit β-propellor and the β-subunit I domain was first visualized in the crystal structure of the αvβ3 extracellular segment with a bound RGD peptide (43). In this structure the RGD site is folded back toward the membrane, and the integrin is in a closed conformation. The closed conformation has also been observed in crystal structures of the αvβ3 ectodomain without an RGD peptide (41) and the αIIbβ3 ectodomain (47).The open integrin conformation has been characterized as having a large separation of up to ∼70 Å between the knees of α and β subunits (16). Four slightly different open headpiece conformations were observed in crystal structures of the αIIbβ3 headpiece with bound fibrinogen-mimetic therapeutics (38). These structures show that the change from a closed to an open headpiece conformation is accompanied by a piston-like motion of helix α7 in the β-chain I domain and a large swing of the β-chain hybrid domain of up to 69°, as well as extension and separation of the two integrin chains. Comparison of the available αvβ3 and αIIbβ3 crystal structures is providing information on the interdomain angle variation and flexibility between domains (47).One aspect of the extension model is that separation of the C-terminal, intracellular portions of the α and β subunits leads to inside-out activation. This concept is supported by nuclear magnetic resonance structures of the cytoplasmic tails of αIIbβ3 showing that the membrane-proximal helices engage in a weak interaction that can be disrupted by constitutively activating mutations or by talin, a protein found in high concentrations in focal adhesions (33). The concept that the integrin α and β subunits must also separate during outside-in signaling is supported by a study involving a disulfide-bonded mutant of αIIbβ3 integrin (46). When the α and β subunits are linked in the vicinity of the transmembrane helices the mutant αIIbβ3 is still able to bind ligand, mediate adhesion, and undergo antibody-induced clustering. However, the disulfide-bonded mutant exhibits defects in focal adhesion formation and focal adhesion kinase activation. Reduction of the disulfide bond or single cysteine mutants rescues signaling.A competing model for integrin activation, called the “deadbolt” model, proposes only small conformational changes in the integrin β-chain I domain upon RGD binding (2). This model is based on crystal structures of the αvβ3 ectodomain with or without an RGD peptide (41, 43). Both of these αvβ3 structures reveal a bent integrin conformation with a closed headpiece conformation. However, the RGD peptide was soaked into a preformed crystal of αvβ3 and crystal contacts may have prevented conformational changes.There are relatively few and only moderate resolution structures of virus-integrin complexes. A moderate resolution cryoEM structure has been determined for the Picornavirus echovirus 1 (EV1) in complex with the I domain of the α2 integrin subunit (39). Docking of crystal structures of EV1 and the α2 I domain into the cryoEM density indicates that the I domain binds within a canyon on the surface of EV1 and that five integrins could potentially bind at one vertex of the icosahedral capsid. Confocal fluorescence microscopy experiments indicated that EV1 causes integrin clustering on human osteosarcoma cells stably transfected with α2 integrin. However, it could not be determined whether the bound integrins were in the inactive (bent) or active (extended) conformation.Moderate resolution (∼21 Å) cryoEM structures of Ad type 2 (HAdV2) and HAdV12 in complex with a soluble form of αvβ5 integrin revealed a ring of integrin density over each penton base capsid protein (5). Better-defined integrin density was observed in the HAdV12/integrin complex, supporting the idea suggested from sequence alignments that the RGD loop of the HAdV12 penton base is shorter and less flexible than that of HAdV2. This study also suggested that the precise spatial arrangement of the five RGD protrusions on the penton base might promote integrin clustering, which may lead to the intracellular signaling events required for virus internalization into a host cell. A similar spacing of RGD-containing integrin-binding sites around the fivefold axis of icosahedral virions has been noted for Ad, foot-and-mouth disease virus, and coxsackievirus A9 (32).We present here a significantly higher-resolution cryoEM structure of HAdV12 complexed with soluble αvβ5 that provides insight into the Ad-integrin interaction. The resolution of the icosahedral capsid portion of the Ad-integrin complex was improved to 8 Å, and the capsid shows clearly resolved α-helices, which allows accurate docking of the penton base crystal structure within the cryoEM density. The resolution of the integrin density is more moderate due to flexibility of the RGD-containing surface loop of penton base and incoherent averaging of integrin heterodimers. Nevertheless, modeling studies with available integrin crystal structures have enabled us to distinguish between a bent or extended conformation (Fig. (Fig.1B)1B) when αvβ5 binds to the multivalent ligand presented by the Ad penton base. The cryoEM structural analysis also indicates that integrin induces a conformational change in penton base.  相似文献   

6.
The aspartate in the prototypical integrin-binding motif Arg-Gly-Asp binds the integrin βA domain of the β-subunit through a divalent cation at the metal ion-dependent adhesion site (MIDAS). An auxiliary metal ion at a ligand-associated metal ion-binding site (LIMBS) stabilizes the metal ion at MIDAS. LIMBS contacts distinct residues in the α-subunits of the two β3 integrins αIIbβ3 and αVβ3, but a potential role of this interaction on stability of the metal ion at LIMBS in β3 integrins has not been explored. Equilibrium molecular dynamics simulations of fully hydrated β3 integrin ectodomains revealed strikingly different conformations of LIMBS in unliganded αIIbβ3 versus αVβ3, the result of stronger interactions of LIMBS with αV, which reduce stability of the LIMBS metal ion in αVβ3. Replacing the αIIb-LIMBS interface residue Phe191 in αIIb (equivalent to Trp179 in αV) with Trp strengthened this interface and destabilized the metal ion at LIMBS in αIIbβ3; a Trp179 to Phe mutation in αV produced the opposite but weaker effect. Consistently, an F191/W substitution in cellular αIIbβ3 and a W179/F substitution in αVβ3 reduced and increased, respectively, the apparent affinity of Mn2+ to the integrin. These findings offer an explanation for the variable occupancy of the metal ion at LIMBS in αVβ3 structures in the absence of ligand and provide new insights into the mechanisms of integrin regulation.  相似文献   

7.
The molecular mechanisms that regulate the endothelial response during transendothelial migration (TEM) of invasive cancer cells remain elusive. Tyrosine phosphorylation of vascular endothelial cadherin (VE-cad) has been implicated in the disruption of endothelial cell adherens junctions and in the diapedesis of metastatic cancer cells. We sought to determine the signaling mechanisms underlying the disruption of endothelial adherens junctions after the attachment of invasive breast cancer cells. Attachment of invasive breast cancer cells (MDA-MB-231) to human umbilical vein endothelial cells induced tyrosine phosphorylation of VE-cad, dissociation of β-catenin from VE-cad, and retraction of endothelial cells. Breast cancer cell-induced tyrosine phosphorylation of VE-cad was mediated by activation of the H-Ras/Raf/MEK/ERK signaling cascade and depended on the phosphorylation of endothelial myosin light chain (MLC). The inhibition of H-Ras or MLC in endothelial cells inhibited TEM of MDA-MB-231 cells. VE-cad tyrosine phosphorylation in endothelial cells induced by the attachment of MDA-MB-231 cells was mediated by MDA-MB-231 α2β1 integrin. Compared with highly invasive MDA-MB-231 breast cancer cells, weakly invasive MCF-7 breast cancer cells expressed lower levels of α2β1 integrin. TEM of MCF-7 as well as induction of VE-cad tyrosine phosphorylation and dissociation of β-catenin from the VE-cad complex by MCF-7 cells were lower than in MDA-MB-231 cells. These processes were restored when MCF-7 cells were treated with β1-activating antibody. Moreover, the response of endothelial cells to the attachment of prostatic (PC-3) and ovarian (SKOV3) invasive cancer cells resembled the response to MDA-MB-231 cells. Our study showed that the MDA-MB-231 cell-induced disruption of endothelial adherens junction integrity is triggered by MDA-MB-231 cell α2β1 integrin and is mediated by H-Ras/MLC-induced tyrosine phosphorylation of VE-cad.  相似文献   

8.
Tropoelastin protein monomers assemble to form elastin. Cellular integrin αVβ3 binds RKRK at the C-terminal tail of tropoelastin. We probed cell interactions with tropoelastin by deleting the RKRK sequence to identify other cell-binding interactions within tropoelastin. We found a novel human dermal fibroblast attachment and spreading site on tropoelastin that is located centrally in the molecule. Inhibition studies demonstrated that this cell adhesion was not mediated by either elastin-binding protein or glycosaminoglycans. Cell interactions were divalent cation-dependent, indicating integrin dependence. Function-blocking monoclonal antibodies revealed that αV integrin(s) and integrin αVβ5 specifically were critical for cell adhesion to this part of tropoelastin. These data reveal a common αV integrin-binding theme for tropoelastin: αVβ3 at the C terminus and αVβ5 at the central region of tropoelastin. Each αV region contributes to fibroblast attachment and spreading, but they differ in their effects on cytoskeletal assembly.  相似文献   

9.
Simian immunodeficiency virus (SIV) and human immunodeficiency virus (HIV) infection results in an early and enduring depletion of intestinal CD4+ T cells. SIV and HIV bind integrin α4β7, thereby facilitating infection of lymphocytes that home to the gut-associated lymphoid tissue (GALT). Using an ex vivo flow cytometry assay, we found that SIVmac239-infected cells expressed significantly lower levels of integrin α4β7 than did uninfected cells. This finding suggested a potential viral effect on integrin α4β7 expression. Using an in vitro model, we confirmed that integrin α4β7 was downregulated on the surfaces of SIVmac239-infected cells. Further, modulation of integrin α4β7 was dependent on de novo synthesis of viral proteins, but neither cell death, the release of a soluble factor, nor a change in activation state was involved. Downregulation of integrin α4β7 may have an unappreciated role in the CD4 depletion of the mucosal-associated lymphoid compartments, susceptibility to superinfection, and/or immune evasion.Infection of macaques with simian immunodeficiency virus (SIV) and humans with human immunodeficiency virus (HIV), regardless of the route of transmission, results in early establishment of infection in the gut-associated lymphoid tissue (GALT) (3, 23, 25). Consequently, the CD4+ T cells of the GALT are depleted, and intestinal integrity is compromised (4, 21, 37). The mechanism of GALT depletion, as well as the mechanism of viral localization to the GALT, remains poorly understood.GALT localization is mediated, at least in part, by integrins, a large family of “sticky” cell surface proteins (24, 35, 36). Integrins facilitate conversation between the environment and a cell, thereby influencing cellular adhesion, trafficking, proliferation, and signaling. Consequently, numerous viruses, despite having a small number of proteins, have developed mechanisms to exploit integrins and hence cellular processes, in order to facilitate viral replication and immune evasion (17, 24, 34, 36). Examples of such viruses include human cytomegalovirus (39), rotavirus (14), and SIV/HIV (40). One well-studied integrin, α4β7, mediates migration of lymphocytes to the GALT (31, 33). In 2008, Arthos et al. demonstrated that HIV-1 glycoprotein, gp120, binds integrin α4β7, facilitating infection of CD4+ T cells and increasing viral replication efficiency (1).Recent in vivo studies have revealed that CD4+ T cells expressing high amounts of integrin α4β7 (integrin α4β7 high) are preferentially infected during acute SIV infection (15, 38). In addition, integrin α4β7 high CD4+ T cells contain greater than one provirus per cell during peak viral infection, suggesting that the cells are unusually susceptible to superinfection. Unexpectedly, superinfection is not observed in integrin α4β7 high CD4+ T cells after peak viral infection (15). Integrin α4β7 high-expressing CD4+ T cells are also depleted from the circulation parallel to the loss of intestinal CD4+ cells, suggesting a fundamental role for integrin α4β7 in SIV pathogenesis (38). The mechanism underlying the depletion of integrin α4β7 high-expressing cells and whether SIV-infected cells are directly or indirectly involved remain unknown. Thus, understanding the single-cell dynamics of integrin α4β7 during SIV infection may improve our understanding of SIV and HIV pathogenesis and clarify the role of integrin α4β7 signaling in mucosal trafficking.To examine the single-cell dynamics of integrin α4β7 expression during SIV infection, we used a novel, ex vivo, flow cytometry assay (M. Reynolds, unpublished data). We observed that infected, Gag p27+ cells expressed significantly (P = 0.0085) lower levels of integrin α4β7 than uninfected, CD4+ T cells from the same animal, at the same time point. Thus, we hypothesized that SIV decreases integrin α4β7 expression on the surfaces of virus-infected cells. In vitro, integrin α4β7 expression was downregulated on SIVmac239-infected cells as rapidly as 24 h postinfection. Unexpectedly, integrin α4β7 levels were also perturbed on uninfected cells with an increase in number of cells with intermediate integrin α4β7 expression. The modulation of integrin α4β7 was dependent on de novo synthesis of a viral protein(s), but neither cell death, release of a soluble factor, nor a change in activation state were involved. Combined, this finding suggests an as-yet-unidentified viral effect on integrin α4β7 that may influence depletion of the mucosal associated lymphoid compartments, susceptibility to superinfection, and/or immune evasion during SIV infection.  相似文献   

10.
HCT-8 colon cancer cells secreted heat shock protein 90α (HSP90α) and had increased invasiveness upon serum starvation. The concentrated conditioned medium of serum-starved HCT-8 cells was able to stimulate the migration and invasion of non-serum-starved cells, which could be prevented by treatment with an anti-HSP90α antibody. Recombinant HSP90α (rHSP90α) also enhanced HCT-8 cell migration and invasion, suggesting a stimulatory role of secreted HSP90α in cancer malignancy. HSP90α binding to CD91α and Neu was evidenced by the proximity ligation assay, and rHSP90α-induced HCT-8 cell invasion could be suppressed by the addition of anti-CD91α or anti-Neu antibodies. Via CD91α and Neu, rHSP90α selectively induced integrin αV expression, and knockdown of integrin αV efficiently blocked rHSP90α-induced HCT-8 cell invasion. rHSP90α induced the activities of ERK, PI3K/Akt, and NF-κB p65, but only NF-κB activation was involved in HSP90α-induced integrin αV expression. Additionally, we investigated the serum levels of HSP90α and the expression status of tumor integrin αV mRNA in colorectal cancer patients. Serum HSP90α levels of colorectal cancer patients were significantly higher than those of normal volunteers (p < 0.001). Patients with higher serum HSP90α levels significantly exhibited elevated levels of integrin αV mRNA in tumor tissues as compared with adjacent non-tumor tissues (p < 0.001). Furthermore, tumor integrin αV overexpression was significantly correlated with TNM (Tumor, Node, Metastasis) staging (p = 0.001).  相似文献   

11.
Glycosylation is one of the most abundant posttranslational modification reactions, and nearly half of all known proteins in eukaryotes are glycosylated. In fact, changes in oligosaccharide structure (glycan) are associated with many physiological and pathological events, including cell adhesion, migration, cell growth, cell differentiation and tumor invasion. Glycosylation reactions are catalyzed by the action of glycosyltransferases, which add sugar chains to various complex carbohydrates such as glycoproteins, glycolipids and proteoglycans. Functional glycomics, which uses sugar remodeling by glycosyltransferases, is a promising tool for the characterization of glycan functions. Here, we will focus on the positive and negative regulation of biological functions of integrins by the remodeling of N-glycans with N-acetylglucosaminyltransferase III (GnT-III) and N-acetylglucosaminyltransferase V (GnT-V), which catalyze branched N-glycan formations, bisecting GlcNAc and β1,6 GlcNAc, respectively. Typically, integrins are modified by GnT-III, which inhibits cell migration and cancer metastasis. In contrast, integrins modified by GnT-V promote cell migration and cancer invasion.Key words: integrin, E-cadherin, GnT-III, GnT-V, N-glycosylation, glycosyltransferaseProtein glycosylation encompasses N-glycans, O-glycans and Glycosaminoglycans. N-glycans are linked to asparagine residues of proteins, which is a specific subset residing in the Asn-X-Ser/Thr motif, whereas O-glycans are attached to a subset of serines and threonines (Fig. 1).1 An increasing body of evidence indicates that glycans in glycoproteins are involved in the regulation of cellular functions including cell-cell communication and signal transduction.2,3 In fact, most receptors on the cell surface are N-glycosylated—integrins and epithelial growth factor receptors; and transforming growth factor β receptors. Here, we focus mainly on the modification of N-glycans of integrin α3β1 and α5β1 to address the important roles of N-glycans in cell adhesion and migration.Open in a separate windowFigure 1Two major types of protein glycosylation. N-glycans are covalently linked to asparagine (Asn) residue of proteins, specifically the Asn-X-Ser/Thr motif. In contrast, O-glycans are attached to a subset of glycosidically linked hydroxyl groups of the amino acids serine (Ser) and threonine (Thr).Previous studies indicate that the presence of the appropriate oligosaccharide can modulate integrin activation. When human fibroblasts were cultured in the presence of l-deoxymannojirimycin, an inhibitor of α-mannosidase II, which prevents N-linked oligosaccharide processing, immature α5β1 integrin appeared at the cell surface, and fibronectin (FN)-dependent adhesion was greatly reduced.4 In addition, the treatment of purified integrin α5β1 with N-glycosidase F, which cleaves between the innermost GlcNAc and asparagine residues of N-glycans from N-linked glycoproteins, resulted in the blockage of α5β1 binding to FN and the inherent association of both subunits,5 suggesting that N-glycosylation is essential for functional integrin α5β1. The production of glycoprotein glycans is catalyzed by various glycosyltransferases. N-Acetylglucosaminyltransferase III (GnT-III) transfers N-acetylglucosamine (GlcNAc) from UDP-GlcNAc to a β1, 4 mannose in N-glycans to form a “bisecting” GlcNAc linkage, as shown in Figure 2. Bisecting GlcNAc linkage is found in various hybrid and complex N-glycans. GnT-III is generally regarded as a key glycosyltransferase in N-glycan biosynthetic pathways. Introduction of a bisecting GlcNAc suppresses further processing and elongation of N-glycans catalyzed by N-acetylglucosaminyltransferase V (GnT-V), which is strongly associated with cancer metastasis, since GnT-V cannot utilize the bisected oligosaccharide as a substrate.68 It has also been reported that GnT-V activity and β1, 6 branched N-glycan levels are increased in highly metastatic tumor cell lines.9,10 When NIH3T3 cells were transformed with the oncogenic Ras gene, cell spreading on FN was greatly enhanced due to an increase in β1, 6 GlcNAc branched tri- and tetra-antennary oligosaccharides in α5β1 integrins.9 Similarly, the characterization of N-glycans of integrin α3β1 from non-metastatic and metastatic human melanoma cell lines showed that β1, 6 GlcNAc branched structures were expressed at high levels in metastatic cells compared with non-metastatic cells.10 Cancer metastasis was consistently, and significantly, suppressed in GnT-V knockout mice.11Open in a separate windowFigure 2Glycosylation reactions catalyzed by the action of glycosyltransferase GnT-III and GnT-V. The remodeled N-glycans regulate cell adhesion and migration. Enhanced expression of GnT-V in epithelial cells results in a loss of cell-cell adhesion, increasing integrin-mediated cell migration. In contrast, overexpression of GnT-III strengthens cell-cell interaction and downregulates integrin-mediated cell migration, which may contribute to the suppression of cancer metastasis. The β1,6GlcNAc branching is preferentially modified by polylactosamine and other sugar motifs such as sialyl Lewis X, which also contribute to promotion of cancer metastasis. It is worth mentioning that GnT-III could be proposed as an antagonistic of GnT-V, since GnT-V cannot utilize the bisected oligosaccharide as a substrate.To explore the possible mechanisms involved in increased β1, six branched N-glycans on cancer cells, Guo et al. found that cell migration toward FN and invasion through the matrigel were both substantially stimulated in cells in which the expression of GnT-V was induced.12 Increased branched sugar chains inhibited the clustering of integrin α5β1 and the organization of F-actin into extended microfilaments in cells plated on FN-coated plates, which supports the hypothesis that the degree of adhesion of cells to their extracellular matrix (ECM) substrate is a critical factor in regulating the rate of cell migration, i.e., migration is maximal under conditions of intermediate levels of cell adhesion.13 Conversely, GnT-V null mouse embryonic fibroblasts (MEF) displayed enhanced cell adhesion to, and spreading on, FN-coated plates with the concomitant inhibition of cell migration. The restoration of GnT-V cDNA in the null MEF reversed these abnormal characteristics, indicating the direct involvement of N-glycosylation events in these phenotypic changes.In contrast to GnT-V, the overexpression of GnT-III resulted in an inhibition of α5β1 integrin-mediatedcell spreading and migration, and the phosphorylation of the focal adhesion kinase.14 The affinity of the binding of integrin α5β1 to FN was significantly reduced as a result of the introduction of a bisecting GlcNAc to the α5 subunit. In addition, overexpression of GnT-III in highly metastatic melanoma cells reduced β1, six branching in cell-surface N-glycans and increased bisected N-glycans.15 Therefore, GnT-III has been proposed as an antagonistic of GnT-V, thereby contributing to the suppression of cancer metastasis. In fact, the opposing effects of GnT-III and GnT-V have been observed for the same target protein, integrin α3β1.16 GnT-V stimulates α3β1 integrin-mediated cell migration, while overexpression of GnT-III inhibits GnT-V-induced cell migration. The modification of the α3 subunit by GnT-III supersedes modification by GnT-V. As a result, GnT-III inhibits GnT-V-induced cell migration. These results strongly suggest that remodeling of glycosyltransferase-modified N-glycan structures either positively or negatively modulates cell adhesion and migration.In addition, sialylation on the non-reducing terminus of N-glycans of α5β1 integrin plays an important role in cell adhesion. The increased sialylation of the β1 integrin subunit was correlated with a decreased adhesiveness and metastatic potential.1719 On the other hand, the enzymatic removal of α2, eight-linked oligosialic acids from the α5 integrin subunit inhibited cell adhesion to FN,20 supporting the observation that the N-glycans of α and β integrin subunits play distinct roles in cell-ECM interactions.21 Collectively, these findings suggest that the interaction of integrin α5β1 with FN is dependent on N-glycosylation and the processing status of N-glycans.Although alteration of the oligosaccharide portion on integrin α5β1 could affect cis- and trans-interactions caused by GnT-III, ST6GalI and GnT-V, as described above, the molecular mechanism remains unclear. Considering integrin α5β1 contains 26 potential N-linked glycosylation sites (14 in the α subunit and 12 in the β subunit), the determination of those crucial N-glycosylation sites for its biological function is, therefore, quite important for an understanding of the underlying mechanism. We sequentially mutated either one or a combination of asparagine residues in the putative N-glycosylation sites of glutamine residues, and found that N-glycosylation on the β-propeller domain of the α5 subunit (in particular sites number 3–5) is essential for its hetero-dimer formation and its biological functions such as cell spreading and cell migration, as well as for the proper folding of the α5 subunit.22 On the other hand, N-glycans on β1 integrin also play important roles in the regulation of its biological functions23,24 (and our unpublished data). Very recently, we also found that GnT-III specifically modifies one of the important glycosylation sites, which results in functional regulation (unpublished data). We postulate that these important sites may participate in supramolecular complex formation on the cell surface, which controls intracellular signal transduction.It also is worth noting that N-glycans regulate cell-ECM association as well as cell-cell adhesion. Overexpression of GnT-III slowed E-cadherin turnover, resulting in increased E-cadherin expression on the surface of B16 melanoma cells.25 E-cadherin engagement at cell-cell contacts is known to suppress cell migration, and that effect has been best described in the context of tumorigenesis.26 Conversely, the disruption of E-cadherin-mediated cell adhesion appears to be a central event in the transition from non-invasive to invasive carcinomas. Interestingly, we recently found that E-cadherin-mediated cell-cell interaction upregulated GnT-III expression,27,28 suggesting that regulation of GnT-III and E-cadherin expression may exist as a positive feedback loop. Taken together, the overexpression of GnT-III inhibits cell migration by at least two mechanisms: an enhancement in cell-cell adhesion and a downregulation of cell-ECM adhesion (Fig. 2).Indeed, glycosylation defects in humans and their links to disease have shown that the mammalian glycome contains a significant amount of biological information.29 The mammalian glycome repertoire is estimated to be between hundreds and thousands of glycan structures and could be larger than its proteome counterpart. Nevertheless, characterization of the biological functions of each glycan could one day make a significant contribution to the diagnosis and treatment of disease.  相似文献   

12.
13.
Integrin heterodimers acquire high affinity to endothelial ligands by extensive conformational changes in both their α and β subunits. These heterodimers are maintained in an inactive state by inter-subunit constraints. Changes in the cytoplasmic interface of the integrin heterodimer (referred to as inside-out integrin activation) can only partially remove these constraints. Full integrin activation is achieved when both inter-subunit constraints and proper rearrangements of the integrin headpiece by its extracellular ligand (outside-in activation) are temporally coupled. A universal regulator of these integrin rearrangements is talin1, a key integrin-actin adaptor that regulates integrin conformation and anchors ligand-occupied integrins to the cortical cytoskeleton. The arrest of rolling leukocytes at target vascular sites depends on rapid activation of their α4 and β2 integrins at endothelial contacts by chemokines displayed on the endothelial surface. These chemotactic cytokines can signal within milliseconds through specialized Gi-protein coupled receptors (GPCRs) and Gi-triggered GTPases on the responding leukocytes. Some chemokine signals can alter integrin conformation by releasing constraints on integrin extension, while other chemokines activate integrins to undergo conformational activation mainly after ligand binding. Both of these modalities involve talin1 activation. In this opinion article, I propose that distinct chemokine signals induce variable strengths of associations between talin1 and different target integrins. Weak interactions of the integrin cytoplasmic tail with talin1 (the cytoplasmic integrin ligand) dissociate unless the extracellular ligand can simultaneously occupy the integrin headpiece and transmit, within milliseconds, proper allosteric changes across the integrin heterodimer back to the tail-talin1 complex. The fate of this bi-directional occupancy of integrins by both their extracellular and intracellular ligands is likely to benefit from immobilization of both ligands to cortical cytoskeletal elements. To properly anchor talin1 onto the integrin tail, a second integrin partner, Kindlin-3 may be also required, although an evidence that both partners can simultaneously bind the same integrin heterodimer is still missing. Once linked to the cortical actin cytoskeleton, the multi-occupied integrin complex can load weak forces, which deliver additional allosteric changes to the integrin headpiece resulting in further bond strengthening. Surface immobilized chemokines are superior to their soluble counterparts in driving this bi-directional occupancy process, presumably due to their ability to facilitate local co-occupancy of individual integrin heterodimers with talin1, Kindlin-3 and surface-bound extracellular ligands.Key words: adhesion, migration, endothelium, cytoskeleton, shear stress, immunityFirm adhesion of leukocytes to blood vessels is tightly regulated by integrins and their cognate ligands.1,2 These include the α4 integrins, VLA-4 (α4β1) and α4β7, and the β2 integrins, LFA-1 (αLβ2) and Mac-1 (αMβ2). Accumulated data suggest that these counter-receptors are structurally adapted to operate under disruptive blood-derived shear forces.3 A remarkable feature of leukocyte integrins is that their affinity state and microclustering can be regulated within fractions of seconds.4,5 The most robust signals for leukocyte integrins are transduced by chemoattractants, mostly chemokines displayed on the vessel wall.6 A growing body of evidence suggests that the Gi protein coupled receptors of these endothelial chemokines elicit diverse signaling pathways in distinct leukocyte subtypes,2,22 which use two common downstream elements to coactive all leukocyte integrins: talin1 and Kindlin-3.7 In this review, I will describe a model explaining how chemokine signals to these elements regulate the conformation of all leukocyte integrins by facilitating a coupled bi-directional occupancy and activation via both their cytoplasmic and headpiece domains.Recent structural and biophysical studies suggest that leukocyte integrins can alternate between inactive bent conformers, which are clasped heterodimers, and variable unclasped heterodimers with extended ectodomains exhibiting intermediate and high affinity to ligand.5 Most leukocyte integrins are maintained in an inactive resting state,2 whereas in situ chemokine-stimulated integrins unfold and extend 10–25 nm above the cell surface, allowing their headpiece to readily recognize immobilized ligand on a counter surface.8 These extended integrins must undergo extensive rearrangements of their headpiece I-domains induced by external endothelial-displayed ligands in order to arrest rolling leukocytes on blood vessel walls. In leukocytes, these two canonical switches are very short-lived, implying the necessity for a stabilization. It is therefore likely that any type of robust integrin activation must involve bi-directional occupancy of the integrin by both its extracellular ligand and one or more cytoplasmic partners.9The main cytoplasmic integrin-activating adaptor in leukocytes and platelets is talin1.10,11 Talin knock down in multiple cell types results in nearly total loss of integrin activation.12,13 This actin binding adaptor binds different integrin β subunit tails with low affinity,14 which can be locally increased by in situ generated PI(4,5)P2 (PIP2). This phosphoinositide presumably binds to the FERM domain within the talin head and thereby enhances talin binding to a membrane proximal NPXY motif on the β integrin tail, a key event in integrin heterodimer unclasping.15,16 Recent studies suggest, however, that mere talin association may be insufficient to unclasp and activate the integrin heterodimer. Thus, the beta subunit tail may need to get co-occupied by the integrin co-activator, Kindlin, in order to optimize talin association with this integrin subunit.17,18 In leukocytes, Talin1 and the Kindlin family member, Kindlin 3 co-activate both VLA-4 and LFA-1 and this co-activation is dramatically enhanced by multiple chemokine triggered effectors, the nature of which has begun to unfold19 (Fig. 1). I would like to propose that talin1-Kindlin-3 co-binding to the β tails of these and other leukocyte integrins is insufficient to switch these integrins to a conformation able to bind their soluble extracellular ligands due to fast dissocia-tion of PIP2-activated talin1 from the integrin cytoplasmic tail complex. This short lived talin-integrin complex may, on the other hand, get stabilized, if the integrin headpiece can simultaneously bind an immobilized extracellular ligand and undergo immediate outside-in activation, before the talin1 has dissociated from the integrin beta tail (Fig. 1). Such full confor-mational switch can result in additional allosteric changes in the integrin-bound talin which may expose vinculin binding sites and further increase talin-actin associations that reinforce this bi-directional allosteric integrin activation.20Open in a separate windowFigure 1Bi-directional integrin activation requires simultaneous co-occupancy of the integrin heterodimer by extracellular and cytoplasmic ligands. A proposed scheme for chemokine-triggered integrin activation on leukocytes. Integrin conformation is allosterically modulated in a bidirectional manner by at least two sets of ligands, extracellular and cytoplasmic. The degree of activation is dictated primarily by unclasping of the integrin heterodimer, a process dependent on the binding of the activated talin FERM domain to a specific site on the integrin β tail. (1) Inactive integrin. (2–5) Four postulated integrin conformations triggered by distinct chemokine signals. (2) Talin FERM domain activation close to the target integrin is a rate limiting step in integrin activation. This activation is triggered by PIP2 locally generated by talin-associated PIP5Kγ (purple rectangle) stimulated by a nearby Gi-coupled chemokine receptor. (3) Kindlin-3 binding to the integrin β tail stabilizes the otherwise weak talin1-integrin tail complex. The activated integrin can bind a soluble extracellular ligand with a low affinity due to a high koff of the soluble ligand from the integrin headpiece. (4) In the absence of Kindlin-3, chemokine triggered, PIP2-activated talin1 binds only transiently the integrin tail (High koff). The semiactivated integrin, even if occupied by an immobilized extracellular ligand, cannot undergo full bi-directional activation. (5) When both the extracellular ligand and talin are properly anchored, their escape from the integrin is dramatically reduced, lowering the koff. Low koff increases the probability of simultaneous bi-directional occupancy of both the integrin headpiece by the extracellular ligand and of the integrin tail by talin1 and Kindlin-3. This results in optimal bi-directional integrin activation and unclasping of the heterodimer. Stable linkages also allow this bi-directionally occupied integrin to undergo extensive mechanical strengthening by low forces applied on the headpiece; this further activates the headpiece I domains, further separates the β and α subunits from each other, and maximally stabilizes the unclasped integrin. Force application through the high affinity-talin complex can stretch the talin rod domain and expose vinculin binding sites (VBS). Since integrin ligands are generally multivalent, rapid integrin dimerization can take place to further stabilize the focal adhesion (not shown). Additional cytoplasmic partners of specific leukocyte integrins like a-actinin, L-plastin and RAPL may further strengthen subsets of focal adhesions. These and other cytosolic partners bind different integrin targets with different affinities. Therefore the effect of each of these partners on both the kinetics and stability of the talin1-integrin tail complex may vary with the cell type, the integrin type, the strength of the chemokine signal and the proximity between the integrin and its stimulatory GPCR.How can such postulated simultaneous bi-directional occupancy of a leukocyte integrin can be so rapidly triggered by a chemokine signal encountered during leukocyte rolling on blood vessels? An attractive mechanism for in situ facilitation of talin1 binding to the integrin β tail by chemokine signals involves chemokine triggered Gi stimulated RhoA and Rac1 GTPases and their downstream target, the PIP2 generating enzyme PIP5Kγ in the vicinity of the in situ activated integrin19 (Fig. 1). Additional talin1 molecules may also be recruited to the vicinity of this initially stimulated integrin by RIAM,21 an effector that associates with activated Rap-1, one of the key chemokine stimulated GTPases involved in rapid integrin mediated activation in both leukocytes and platelets.22,23 To bidirectionally bind and activate their integrin targets, both the cytoplasmic integrin ligands, Talin1 and Kindlin-3 and the extracellular integrin ligand may need to achieve low dissociation rates from the integrin tail and headpiece, respectively. Why would an immobilized extracellular ligand be superior to soluble extracellular ligand in capacity to bi-directionally bind and activate a leukocyte integrin? The probability that a given surface-bound ligand, rather than a soluble integrin ligand would escape from its cognate integrin receptor following its dissociation is very small, since reactants in viscous medium are more likely to recombine than to diffuse apart.24 Thus, surface-immobilized single integrin ligands may rebind the integrins they recenty dissociated from much more frequently than their soluble counterparts. Similarly, the cytoplasmic ligands talin1 and Kindlin may need to remain immobile once occupying their target integrin tail. Such immobilization of talin1 can be optimized by talin anchoring to the cortical cytoskeleton.25 Talin may be also restricted from immediate dissociation from the integrin tail by Kindlin-3. An optimal integrin activating chemokine signal would therefore not only need to recruit and induce talin1 association with the β subunit of the target integrin and opening of the integrin clasp, but also need to keep the talin in an immobile state, and thereby maintain its low dissociation rate from its integrin tail sites.As both the integrin headpiece and the integrin subunits are predicted to undergo faster opening and separation in the presence of applied forces,26,27 another tradeoff of this postulated immobilization of both the intracellular and extracellular integrin ligands is optimal force sensing of the integrin heterodimer. Application of force on the bidirectionally occupied integrin and its cognate ligands would be possible only if the integrin, its extracellular ligand, and talin1 are all properly anchored.3,28,29 Force transduction through the integrin-talin1 complex can transmit additional conformational changes on the integrin-occupied talin by exposing vinculin binding sites on the talin rod.30 Additional chemokine signals may induce talin rod phosphorylation and other changes in actin-talin associations (Fig. 1) that may further facilitate talin anchorage to the cortical cytoskeleton and subsequent microclustering of adjacent ligand-occupied integrins. It is well recognized that ligand occupancy anchors integrins to the cortical cytoskeleton.31 Thus, the anchorage states of both the extracellular and the cytoplasmic ligands of a given integrin may facilitate bidirectional integrin occupancy and optimize force driven bi-directional activation of the integrin-ligand complex and subsequent dimerization of ligand-occupied integrins. The ability of different integrin-ligand complexes to undergo diverse mechanochemical rearrangements provides a broad spectrum of integrin-ligand bond strengths, accounting for the unique capacity of chemokine stimulated leukocyte integrins to support both firm and reversible adhesions of leukocytes to their endothelial ligands.  相似文献   

14.
15.
NADPH oxidase 1 (Nox1) is expressed mainly in colon epithelial cells and produces superoxide ions as a primary function. We showed that Nox1 knockdown inhibits directional persistence of migration on collagen I. This paper dissects the mechanism by which Nox1 affects the direction of colonic epithelial cell migration in a two-dimensional model. Transient activation of Nox1 during cell spreading on collagen 1 temporarily inactivated RhoA and led to efficient exportation of α2β1 integrin to the cell surface, which supported persistent directed migration. Nox1 knockdown led to a loss of directional migration which takes place through a RhoA-dependent α2/α3 integrin switch. Transient RhoA overactivation upon Nox1 inhibition led to transient cytoskeletal reorganization and increased cell-matrix contact associated with a stable increase in α3 integrin cell surface expression. Blocking of α3 integrin completely reversed the loss of directional persistence of migration. In this model, Nox1 would represent a switch between random and directional migration through RhoA-dependent integrin cell surface expression modulation.The two well-recognized defining hallmarks of cancer are uncontrolled proliferation and invasion (14). The conversion of a static primary tumor into an invasive disseminating metastasis involves an enhanced migratory ability of the tumor cells. Tumor cells use migration mechanisms that are similar, if not identical, to those that occur in normal cells during physiological processes such as embryonic morphogenesis, wound healing, and immune-cell trafficking (10). To migrate, cells must acquire a spatial asymmetry that enables them to turn intracellularly generated forces into net cell body translocation. Dynamic assembly and disassembly of integrin-mediated adhesion and cytoskeletal reorganization are necessary for efficient migration (29). Integrins are heterodimeric integral membrane proteins composed of an α chain and a β chain. Depending on the cell type and extracellular matrix (ECM) substrate, focal contact assembly and migration can be regulated by different integrins. Collagen receptors include α1β1, α2β1, and α3β1 integrins. α1β1 and α3β1 integrins also bind laminin and have less affinity for collagen than does integrin α2β1 (47). The intrinsic propensity of cells to continue migrating in the same direction without turning is closely related to integrin/cytoskeletal interaction, which is known to regulate tractional forces, resulting in modulation of the speed and direction of cell migration (33). Interestingly, different integrin-ECM associations might have opposite effects on the regulation of the directionality of migration. Danen et al. have shown that adhesion to fibronectin by αvβ3 promotes persistent migration through activation of the actin-severing protein cofilin, which results in a polarized phenotype with a single broad lamellipod at the leading edge. In contrast, adhesion to fibronectin by α5β1 instead leads to phosphorylation/inactivation of cofilin and these cells fail to polarize their cytoskeleton and adopt a random/nonpersistent mode of migration (5). Members of the Rho GTPase family (including RhoA, Rac1, and Cdc42) are known as key modulators of cytoskeletal dynamics that occur during cell migration (37). RhoA regulates stress fiber and focal adhesion assembly, Rac regulates the formation of lamellipodial protrusions and membrane ruffles, and Cdc42 triggers filopodial extensions at the cell periphery (13).One of the earliest characterized functions of the Rho GTPase Rac was regulation of the activity of the NADPH oxidase complex in phagocytic cells to produce reactive oxygen species (ROS) (1, 19). Moreover, it has been shown that Rac-dependent ROS production leads to downregulation of RhoA through oxidative inactivation of the low-molecular-weight (LMW) protein tyrosine phosphatase (PTP) and the subsequent activation of p190RhoGAP (31). ROS are also known to directly affect important regulators of cell migration such as PTEN, FAK, or Src (4, 20, 22). ROS are generated in cells from several sources, including the mitochondrial respiratory chain, xanthine oxidase, cytochrome P450, nitric oxide synthase, and NADPH oxidase. The seven known human catalytic subunits of NADPH oxidase include Nox1 to -5 and Duox1 and -2, with Nox2 (gp91phox) being the founding member (21). These oxidases participate in several adaptive functions, ranging from mitogenesis to immune cell signaling (11). A growing body of data points to a key role for ROS production by NADPH oxidase in the control of cell migration and cytoskeletal reorganization (30, 44). Among NADPH oxidase homologs, Nox1 has been detected in different cell types, with major expression in vascular smooth muscle cells and colonic epithelial cells (42). Nox1 involvement in the control of cytoskeletal organization and cellular migration has been only recently reported. Shinohara et al. demonstrated that oncogenic Ras transformation involves Nox1-dependent signaling and leads to inactivation of RhoA. Abrogation of Nox1-dependent ROS production by diphenyleneiodonium (DPI) or small interfering RNA restores RhoA activation and actin stress fiber formation (41). More recently, several groups have highlighted a key role of Nox1 in the control of growth factor-induced migration (16, 38, 40). Cancer cells probably undergo random migration during metastasis, but their migration can be directed by cytokine gradients and/or associated with ECM fibers (29, 55). In a recent report, we showed that Nox1 downregulation decreased the persistence of colonic adenocarcinoma cell migration over collagen I (Col-I) without affecting either the mean velocity or the total distance of migration.In the present study, we investigated the molecular mechanism by which Nox1-dependent ROS production controls the directionality of migration of colonic adenocarcinoma cells. We showed that Nox1-dependent ROS production, which occurs during cell spreading after 4 h of adhesion to Col-I, transiently inhibited RhoA activity. Nox1 inhibition during cell spreading led to a transient increase in cell-matrix contact and initiated a sustained decrease in α2β1 integrin cell surface expression, which was compensated for by an increase in α3 integrin cell surface expression. While Nox1-dependent RhoA inhibition was transient, the observed α2/α3 integrin switch was sustained over 24 h. The loss of directionality observed in cell migration upon Nox1 inhibition may be reversed by α3 integrin blockade. This work shows that Nox1 is involved in the control of integrin surface expression during migration on Col-I, which is critical for persistent directed migration through transient modulation of a RhoA/ROCK-dependent pathway.  相似文献   

16.
Regulation of vascular homeostasis depends upon collaboration between cells of the vessel wall and blood coagulation system. A direct interaction between integrin αVβ3 on endothelial cells and smooth muscle cells and prothrombin, the pivotal proenzyme of the blood coagulation system, is demonstrated and activation of the integrin is required for receptor engagement. Evidence that prothrombin is a ligand for αVβ3 on these cells include: (a) prothrombin binds to purified αVβ3 via a RGD recognition specificity; (b) prothrombin supports αVβ3-mediated adhesion of stimulated endothelial cells and smooth muscle cells; and (c) endothelial cells, either in suspension and in a monolayer, recognize soluble prothrombin via αVβ3. αVβ3-mediated cell adhesion to prothrombin, but not to fibrinogen, required activation of the receptor. Thus, the functionality of the αVβ3 receptor is ligand defined, and prothrombin and fibrinogen represent activation- dependent and activation-independent ligands. Activation of αVβ3 could be induced not only by model agonists, PMA and Mn2+, but also by a physiologically relevant agonist, ADP. Inhibition of protein kinase C and calpain prevented activation of αVβ3 on vascular cells, suggesting that these molecules are involved in the inside-out signaling events that activate the integrin. The capacity of αVβ3 to interact with prothrombin may play a significant role in the maintenance of hemostasis; and, at a general level, ligand selection by αVβ3 may be controlled by the activation state of this integrin.  相似文献   

17.
Cathepsin X is a lysosomal cysteine protease, found predominantly in cells of monocyte/macrophage lineage. It acts as a monocarboxypepidase and has a strict positional and narrower substrate specificity relative to the other human cathepsins. In our recent studies we identified—β2 subunit of integrin receptors and α and γ enolase as possible substrates for cathepsin X carboxypeptidase activity. In both cases cathepsin X is capable to cleave regulatory motifs at C-terminus affecting the function of targeted molecules. We demonstrated that via activation of β2 integrin receptor Mac-1 (CD11b/CD18) active cathepsin X enhances adhesion of monocytes/macrophages to fibrinogen and regulates the phagocytosis. By activation of Mac-1 receptor cathepsin X may regulate also the maturation of dendritic cells, a process, which is crucial in the initiation of adaptive immunity. Cathepsin X activates also the other β2 integrin receptor, LFA-1 (CD11a/CD18) which is involved in the proliferation of T lymphocytes. By modulating the activity of LFA-1 cathepsin X causes cytoskeletal rearrangements and morphological changes of T lymphocytes enhancing ameboid-like migration in 2-D and 3-D barriers and increasing homotypic aggregation. The cleavage of C-terminal amino acids of α and γ enolase by cathepsin X abolishes their neurotrophic activity affecting neuronal cell survival and neuritogenesis.Key words: cathepsin X, integrin, enolase, T lymphocyte, macrophage, dendritic cell, adhesion, migration, neuritogenesisProteases comprise a group of enzymes that catalyse the cleavage of a peptide bond in a polypeptide chain by nucleophilic attack on the carbonyl carbon. The proteases are either exopeptidases cleaving one or a few amino acids at the N- or C-terminus of polypeptide chain or endopeptidases that cleave the peptide bond internally. According to the catalytic mechanism the endopeptidases are divided into aspartic, cysteine, serine, threonine and metallo endoproteases—see MEROPS database.1 To date, 561 genes encoding for proteases have been identified in human genome. Among them 148 genes encode for cysteine proteases including a group of eleven lysosomal cysteine proteases (members of C1 family) also called cathepsins. They exhibit different expression patterns, levels and specificities, all of which contribute to their differential physiological roles. Some of them, like cathepsins B, H, L and C are ubiquitously present in tissues, whereas others (cathepsins S, V, X, O, K, F and W) are expressed by specific cell types. Cysteine cathepsins were long believed to be responsible for the terminal protein degradation in the lysosomes, however, this view has changed dramatically when they have been found to be involved in a number of important cellular processes and pathologies.2,3In contrast to other cathepsins, cathepsin X was discovered only recently. Its gene,4,5 structure6,7 and activity properties8,9 show several unique features that distinguish it clearly from other human cysteine proteases. It has a very short pro-region7 and a three residue insertion motif which forms a characteristic “mini loop.”6 Cathepsin X exhibits carboxypeptidase activity6 and, in contrast to cathepsin B, the other carboxypeptidase, it does not act as an endopeptidase. Contrary to the first reports,4 cathepsin X is not widely expressed in cells and tissues, but is restricted to the cells of the immune system, predominantly monocytes, macrophages and dendritic cells.10 Higher levels of cathepsin X were also found in tumor and immune cells of prostate11 and gastric12 carcinomas and in macrophages of gastric mucosa, especially after infection by Helicobacter pylori.13 Recently it was shown that cathepsin X is abundantly expressed in mouse brain cells, in particular glial cells. Its upregulation was also detected in the brains of patients with Alzheimer disease.14The involvement of cathepsin X in signal transduction is implied by the integrin-binding motifs, present in its pro-form (RGD: Arg-Gly-Asp) and mature form (ECD: Glu-Cys-Asp).4,5 Moreover, cathepsin X binds cell surface heparan sulfate proteoglycans15 which are also involved in integrin regulation. A strong co-localization of pro-cathepsin X with β3 integrin subunit was demonstrated in our study in pro-monocytic U-937 cells.16 Further, it was reported that the pro-form of cathepsin X interacts with αvβ3 integrin through the RGD motif in lamellipodia of human umbilical vein endothelial cells (HUVECs).17 However, we showed that the active form of cathepsin X co-localized predominantly with β2 integrin subunit in various cells of monocytes/macrophage lineage. Active cathepsin X was shown to regulate β2 integrin-dependent adhesion, phagocytosis and T lymphocyte activation by interaction with macrophage antigen-1 (CD11b/CD18, Mac-1). We showed that inhibitors and monoclonal antibodies, capable to impair cathepsin X enzymatic activity, reduced the binding of differentiated U-937 cells to fibrinogen and polystyrene surface in a dose dependent manner. The co-localization of active cathepsin X with β2 integrin chain was particularly enhanced in interactions of monocyte/macrophages with endothelial and tumor cells.Besides in monocytes and macrophages the active cathepsin X plays a role in β2 integrin activation also in dendritic cells (DC), which are crucial for effective antigen presentation and initiation of T cell dependent immune response. Maturation of dendritic cells is accompanied by a range of morphological and cytoskeleton structure changes. In response to maturation stimuli in vitro, DCs rapidly adhere, develop polarity and assemble actin rich structures at the leading edge, known as podosomes.18 The adhesion of immature DCs to the extracellular matrix, is accompanied by recruitment of Mac-1 integrin receptor, which can be activated by cathepsin X. We have shown that, during maturation, cathepsin X translocates to the plasma membrane of maturing DCs, enabling Mac-1 activation and, consequently, cell adhesion.19 In mature DCs cathepsin X redistributes from the membrane to the perinuclear region, which coincides with the de-adhesion of DCs, formation of cell clusters and acquisition of the mature phenotype. Again, the inhibition of cathepsin X activity during DC differentiation and maturation reduced the capacity of DCs to stimulate T lymphocytes.β2 integrin receptors are important also in T lymphocyte functions, such as migration and invasion across the endothelium and tissues. Lymphocyte function-associated antigen-1 (CD11a/CD18, LFA-1), the predominant β2 integrin receptor in T lymphocytes enables cell-cell interactions and homotypic aggregation via LFA-1-ICAM-1 (intracellular adhesion molecule-1) interactions. LFA-1 can act also as a true signaling receptor, causing F-actin reorganization that leads to cytoskeletal changes of the cell20 and a switch from a spherical to a polarized shape.21 Although the concentration of cathepsin X in T lymphocytes is lower compared to monocytes and macrophages, we showed that it interacts with LFA-1 promoting cytoskeleton-dependent morphological changes and migration across 2D and 3D models of ICAM-1 and Matrigel.22,23 Its co-localization with LFA-1 was particularly evident at the trailing edge protrusion, the uropod, which plays an important role in T lymphocyte migration and cell-cell interactions (Fig. 1). Uropodal active cathepsin X cleaves C-terminal amino acids of β chain in LFA-1 promoting its high affinity conformation and the binding of the cytoskeletal protein talin. This interaction stabilizes the uropod and promotes its elongation (Jevnikar, et al. submitted).Open in a separate windowFigure 1Activation of LFA-1 integrin receptor by cathepsin X at the uropod of T lymphocyte promotes cytoskeleton-dependent morphological changes and cell migration.30We demonstrated that uropods of cathepsin X upregulated T lymphocytes elongate to extreme length and form cell-to-cell connections, the nanotubes (Obermajer, et al. in press). Membrane (or tunneling) nanotubes were recently found as a new principle of cell-to-cell communication enabling transmission of complex and specific messages to distant cells through a physically connected network. Calcium fluxes, vesicles and cell-surface components can all traffic between cells connected by nanotubes. In immune system nanotubes integrate communities of cells for a better coordination of their action in various stages of immune response. We showed that nanotubes of cathepsin X upregulated T lymphocytes could readily transfer cellular organelles such as mitochondria and lysosomes and proposed that nanotube mediated transfer makes possible T lymphocyte activation without the need for direct contact with antigen presenting cells.The exact mechanism of cathepsin X translocation towards plasma membrane and degradation of C-terminal amino acids of β chain remains unclear. In lysosomes cathepsin X can be found as a pro- and active form. After cell activation cathepsin X containing vesicles translocate towards the plasma membrane,16 as observed also for some other lysosomal proteases.24 During this process it is possible that pro-cathepsin X is activated by the other cysteine protease cathepsin L, as shown already in vitro.8 Both proteases were strongly co-localized with β2 integrin chain at plasma membrane of activated monocytes/macrophages and at uropodes of T lymphocytes. Simultaneous co-localization with the lysosomal markers demonstrates that at least the initial translocation of cathepsin X towards cytoplasmic tail of β2 integrin chain is vesicular. The interaction of cathepsin X with β2 integrin subunit was confirmed by immunoprecipitation and FRET.22 According to in vitro experiments we propose that cathepsin X cleaves sequentially C-terminal aminoacids F766, A767, E768 and S769 of β2 integrin subunit (Fig. 2) until reaching proline in penultimate position, confirming previous observation that the proline in S2 position leads to resistance to cathepsin X proteolysis.25 Also, our results are in agreement with the previously mentioned monocarboxypeptidase activity of cathepsin X.26,27 Since the signaling to and from the integrins is mainly regulated by the short cytoplasmic tail of β2 subunit,28 cathepsin X mediated β2 integrin truncation leads to regulation of the receptor signaling. The interaction of cytoplasmic tail with different cytoskeletal and regulatory proteins, such as talin, filamin, radixin and α-actinin is crucial for signal transduction and modulation of cytoskeleton.29Open in a separate windowFigure 2Cathepsin X activates LFA-1 by sequential cleavage of C-terminal amino acids of β2 integrin subunit.Besides β2 integrin chain we recently identified isozymes α and γ enolases as another molecular target for cathepsin X carboxypeptidase activity (Obermajer, et al. submitted). We demonstrated that cathepsin X sequentially cleaves C-terminal amino acids of both isozymes, abolishing their neurotrophic activity. On this way the neuronal cell survival and neuritogenesis can be regulated. Inhibition of cathepsin X activity increases the generation of plasmin, essential for neuronal differentiation and changes the length distribution of neurites, especially in the early phase of neurite outgrowth. Moreover, cathepsin X inhibition increases neuronal survival and reduces serum deprivation induced apoptosis, particularly in the absence of nerve growth factor.  相似文献   

18.
Src family kinase (SFK) activity is elevated in many cancers, and this activity correlates with aggressive tumor behavior. The α6β4 integrin, which is also associated with a poor prognosis in many tumor types, can stimulate SFK activation; however, the mechanism by which it does so is not known. In the current study, we provide novel mechanistic insight into how the α6β4 integrin selectively activates the Src family member Fyn in response to receptor engagement. Both catalytic and noncatalytic functions of SHP2 are required for Fyn activation by α6β4. Specifically, the tyrosine phosphatase SHP2 is recruited to α6β4 and its catalytic activity is stimulated through a specific interaction of its N-terminal SH2 domain with pY1494 in the β4 subunit. Fyn is recruited to the α6β4/SHP2 complex through an interaction with phospho-Y580 in the C terminus of SHP2. In addition to activating Fyn, this interaction with Y580-SHP2 localizes Fyn to sites of receptor engagement, which is required for α6β4-dependent invasion. Of significance for tumor progression, phosphorylation of Y580-SHP2 and SFK activation are increased in orthotopic human breast tumors that express α6β4 and activation of this pathway is dependent upon Y1494.Expression of the α6β4 integrin, a laminin receptor, is associated with poor patient prognosis and reduced survival in many human cancers (32). For this reason, there is considerable interest in understanding how this integrin is regulated and how it functions to promote tumor progression. In normal tissues, the α6β4 integrin plays a major role in maintaining the integrity of epithelia by binding to laminins in the basement membrane and regulating the assembly of hemidesmosomes on the basal epithelial cell surface (7, 17). In pathophysiological conditions such as wound healing and cancer, the stable adhesive interactions of the α6β4 receptor are disrupted by phosphorylation of the β4 cytoplasmic domain, converting α6β4 to a signaling-competent receptor that promotes dynamic adhesion and invasion (18). Phosphorylation of the β4 subunit cytoplasmic domain on serine residues contributes to the dynamic adhesive functions of the receptor by disrupting interactions with hemidesmosomal proteins that regulate stable adhesion (33, 37). Phosphorylation of the β4 cytoplasmic domain on tyrosine residues may also contribute to the regulation of hemidesmosomes, but it is likely that the major contribution of tyrosyl phosphorylation is to mediate interactions that stimulate downstream signaling from the receptor (22).In transformed cells, engagement of the α6β4 integrin stimulates the activation of several signaling molecules, including phosphatidylinositol-3 kinase (PI3K), mitogen-activated protein kinases (MAPK), NFκB, and Src family kinases (SFKs) (10, 12, 21, 40). In previous studies, we identified Y1494 in the β4 subunit cytoplasmic domain as an important mediator of α6β4-dependent signaling by demonstrating that mutation of Y1494 inhibits the ability of α6β4 to stimulate PI3K, MAPK, and SFK activation (10, 39). Restoration of both PI3K and SFK signaling, but not MAPK signaling, rescues invasion in tumor cells expressing Y1494F-β4, indicating that PI3K and SFK signaling pathways cooperate downstream of Y1494 to promote α6β4-dependent invasion (10). Y1494 is localized within an immunoreceptor tyrosine-based inhibition motif (ITIM), a canonical binding site for Src-homology-2 (SH2) domain-containing protein-tyrosine phosphatase 1 (SHP1) and SHP2 (44). Examination of a chimeric receptor containing the extracellular domain of TrkB and the transmembrane and cytoplasmic domains of the β4 subunit demonstrated that SHP2 binds to and is activated by sequences in the β4 cytoplasmic domain in response to dimerization (23). Moreover, Y1494 is one of three tyrosine residues, along with Y1257 and Y1440, that mediate the interaction of SHP2 with the β4 subunit cytoplasmic domain in response to c-Met signaling (6). Importantly, SHP2 is essential for the activation of SFKs both by the chimeric TrkB/β4 receptor and when the β4 subunit functions as a signaling adaptor for c-Met (6, 23). However, the mechanism by which SHP2 activates SFKs in response to α6β4 engagement has not been established.Elevated SFK activity correlates strongly with breast cancer invasion and metastasis, and these kinases are frequently activated in human cancers (15). Given the parallels between α6β4 expression and SFK activation in cancer, investigation of how α6β4 contributes to the activation of this invasion-promoting pathway is warranted. In the current study, we sought to elucidate the mechanism by which engagement of α6β4 activates SFKs and the significance of the β4/SHP2/SFK signaling axis for tumor progression. Our results reveal a novel mechanism for SHP2-dependent activation of the SFK family member Fyn which involves Y580 in the C terminus of SHP2.  相似文献   

19.
Cell surface receptors of the integrin family are pivotal to cell adhesion and migration. The activation state of heterodimeric αβ integrins is correlated to the association state of the single-pass α and β transmembrane domains. The association of integrin αIIbβ3 transmembrane domains, resulting in an inactive receptor, is characterized by the asymmetric arrangement of a straight (αIIb) and tilted (β3) helix relative to the membrane in congruence to the dissociated structures. This allows for a continuous association interface centered on helix-helix glycine-packing and an unusual αIIb(GFF) structural motif that packs the conserved Phe-Phe residues against the β3 transmembrane helix, enabling αIIb(D723)β3(R995) electrostatic interactions. The transmembrane complex is further stabilized by the inactive ectodomain, thereby coupling its association state to the ectodomain conformation. In combination with recently determined structures of an inactive integrin ectodomain and an activating talin/β complex that overlap with the αβ transmembrane complex, a comprehensive picture of integrin bi-directional transmembrane signaling has emerged.Key words: cell adhesion, membrane protein, integrin, platelet, transmembrane complex, transmembrane signalingThe communication of biological signals across the plasma membrane is fundamental to cellular function. The ubiquitous family of integrin adhesion receptors exhibits the unusual ability to convey signals bi-directionally (outside-in and inside-out signaling), thereby controlling cell adhesion, migration and differentiation.15 Integrins are Type I heterodimeric receptors that consist of large extracellular domains (>700 residues), single-pass transmembrane (TM) domains, and mostly short cytosolic tails (<70 residues). The activation state of heterodimeric integrins is correlated to the association state of the TM domains of their α and β subunits.610 TM dissociation initiated from the outside results in the transmittal of a signal into the cell, whereas dissociation originating on the inside results in activation of the integrin to bind ligands such as extracellular matrix proteins. The elucidation of the role of the TM domains in integrin-mediated adhesion and signaling has been the subject of extensive research efforts, perhaps commencing with the demonstration that the highly conserved GFFKR sequence motif of α subunits (Fig. 1), which closely follows the first charged residue on the intracellular face, αIIb(K989), constrains the receptor to a default low affinity state.11 Despite these efforts, an understanding of this sequence motif had not been reached until such time as the structure of the αIIb TM segment was determined.12 In combination with the structure of the β3 TM segment13 and available mutagenesis data,6,9,10,14,15 this has allowed the first correct prediction of the overall association of an integrin αβ TM complex.12 The predicted association was subsequently confirmed by the αIIbβ3 complex structure determined in phospholipid bicelles,16 as well as by the report of a similar structure based on molecular modeling using disulfide-based structural constraints.17 In addition to the structures of the dissociated and associated αβ TM domains, their membrane embedding was defined12,13,16,18,19 and it was experimentally recognized that, in the context of the native receptor, the TM complex is stabilized by the inactive, resting ectodomain.16 These advances in integrin membrane structural biology are complemented by the recent structures of a resting integrin ectodomain and an activating talin/β cytosolic tail complex that overlap with the αβ TM complex,20,21 allowing detailed insight into integrin bi-directional TM signaling.Open in a separate windowFigure 1Amino acid sequence of integrin αIIb and β3 transmembrane segments and flanking regions. Membrane-embedded residues12,13,16,18,19 are enclosed by a gray box. Residues 991–995 constitute the highly conserved GFFKR sequence motif of integrin α subunits.  相似文献   

20.

Background

At least one-third of epithelial ovarian cancers are associated with the development of ascites containing heterogeneous cell populations, including tumor cells, inflammatory cells, and stromal elements. The components of ascites and their effects on the tumor cell microenvironment remain poorly understood. This study aimed to isolate and characterize stromal progenitor cells from the ascites of patients with epithelial ovarian adenocarcinoma (EOA).

Methods

Seventeen ascitic fluid samples and 7 fresh tissue samples were collected from 16 patients with EOA. The ascites samples were then cultured in vitro in varying conditions. Flow cytometry and immunocytochemistry were used to isolate and characterize 2 cell populations with different morphologies (epithelial type and mesenchymal type) deriving from the ascites samples. The in vitro cell culture model was established using conditional culture medium.

Results

The doubling times of the epithelial type and mesenchymal type cells were 36 h and 48 h, respectively, indicating faster growth of the epithelial type cells compared to the mesenchymal type cells. Cultured in vitro, these ascitic cells displayed the potential for self-renewal and long-term proliferation, and expressed the typical cancer stem/progenitor cell markers CD44high, CD24low, and AC133+. These cells also demonstrated high BMP-2, BMP4, TGF-β, Rex-1, and AC133 early gene expression, and expressed EGFR, integrin α2β1, CD146, and Flt-4, which are highly associated with tumorigenesis and metastasis. The epithelial type cells demonstrated higher cytokeratin 18 and E-cadherin expression than the mesenchymal type cells. The mesenchymal type cells, in contrast, demonstrated higher AC133, CD73, CD105, CD117, EGFR, integrin α2β1, and CD146 surface marker expression than the epithelial type cells.

Conclusion

The established culture system provides an in vitro model for the selection of drugs that target cancer-associated stromal progenitor cells, and for the development of ovarian cancer treatments.  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