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.
3.
Phosphoinositide (PI) 3-kinase (PI3K) signaling processes play an important role in regulating the adhesive function of integrin αIIbβ3, necessary for platelet spreading and sustained platelet aggregation. PI3K inhibitors are effective at reducing platelet aggregation and thrombus formation in vivo and as a consequence are currently being evaluated as novel antithrombotic agents. PI3K regulation of integrin αIIbβ3 activation (affinity modulation) primarily occurs downstream of Gi-coupled and tyrosine kinase-linked receptors linked to the activation of Rap1b, AKT, and phospholipase C. In the present study, we demonstrate an important role for PI3Ks in regulating the avidity (strength of adhesion) of high affinity integrin αIIbβ3 bonds, necessary for the cellular transmission of contractile forces. Using knock-out mouse models and isoform-selective PI3K inhibitors, we demonstrate that the Type Ia p110β isoform plays a major role in regulating thrombin-stimulated fibrin clot retraction in vitro. Reduced clot retraction induced by PI3K inhibitors was not associated with defects in integrin αIIbβ3 activation, actin polymerization, or actomyosin contractility but was associated with a defect in integrin αIIbβ3 association with the contractile cytoskeleton. Analysis of integrin αIIbβ3 adhesion contacts using total internal reflection fluorescence microscopy revealed an important role for PI3Ks in regulating the stability of high affinity integrin αIIbβ3 bonds. These studies demonstrate an important role for PI3K p110β in regulating the avidity of high affinity integrin αIIbβ3 receptors, necessary for the cellular transmission of contractile forces. These findings may provide new insight into the potential antithrombotic properties of PI3K p110β inhibitors.  相似文献   

4.
The free energy of transfer of nonpolar solutes from water to lipid bilayers is often dominated by a large negative enthalpy rather than the large positive entropy expected from the hydrophobic effect. This common observation has led to the idea that membrane partitioning is driven by the "nonclassical" hydrophobic effect. We examined this phenomenon by characterizing the partitioning of the well-studied peptide melittin using isothermal titration calorimetry (ITC) and circular dichroism (CD). We studied the temperature dependence of the entropic (-TΔS) and enthalpic (ΔH) components of free energy (ΔG) of partitioning of melittin into lipid membranes made of various mixtures of zwitterionic and anionic lipids. We found significant variations of the entropic and enthalpic components with temperature, lipid composition and vesicle size but only small changes in ΔG (entropy-enthalpy compensation). The heat capacity associated with partitioning had a large negative value of about -0.5 kcal mol(-1) K(-1). This hallmark of the hydrophobic effect was found to be independent of lipid composition. The measured heat capacity values were used to calculate the hydrophobic-effect free energy ΔG (hΦ), which we found to dominate melittin partitioning regardless of lipid composition. In the case of anionic membranes, additional free energy comes from coulombic attraction, which is characterized by a small effective peptide charge due to the lack of additivity of hydrophobic and electrostatic interactions in membrane interfaces [Ladokhin and White J Mol Biol 309:543-552, 2001]. Our results suggest that there is no need for a special effect-the nonclassical hydrophobic effect-to describe partitioning into lipid bilayers.  相似文献   

5.
6.
7.
Our previous study demonstrated that transforming growth factor (TGF)-β activates β-catenin signaling through Smad3 interaction with β-catenin in chondrocytes. In the present studies, we further investigated the detailed molecular mechanism of the cross-talk between TGF-β/Smad3 and Wnt/β-catenin signaling pathways. We found that C-terminal Smad3 interacted with both the N-terminal region and the middle region of β-catenin protein in a TGF-β-dependent manner. Both Smad3 and Smad4 were required for the interaction with β-catenin and protected β-catenin from an ubiquitin-proteasome-dependent degradation. In addition, the formation of the Smad3-Smad4-β-catenin protein complex also mediated β-catenin nuclear translocation. This Smad3-mediated regulatory mechanism of β-catenin protein stability enhanced the activity of β-catenin to activate downstream target genes during chondrogenesis. Our findings demonstrate a novel mechanism between TGF-β and Wnt/β-catenin signaling pathways during chondrocyte development.  相似文献   

8.
Fibrosis is characterized by elevated transforming growth factor β (TGFβ) signaling, resulting in extracellular matrix accumulation and increased PAI-1 (plasminogen activator inhibitor) expression. PAI-1 induces the internalization of urokinase plasminogen activator/receptor and integrin αvβ3 from the cell surface. Since increased αvβ3 expression correlates with increased TGFβ signaling, we hypothesized that aberrant PAI-1-mediated αvβ3 endocytosis could initiate an autocrine loop of TGFβ activity. We found that in PAI-1 knock-out (KO) mouse embryonic fibroblasts), αvβ3 endocytosis was reduced by ∼75%, leaving αvβ3 in enlarged focal adhesions, similar to wild type cells transfected with PAI-1 small interfering RNA. TGFβ signaling was significantly enhanced in PAI-1 KO cells, as demonstrated by a 3-fold increase in SMAD2/3-containing nuclei and a 2.9-fold increase in TGFβ activity that correlated with an increase in αvβ3 and TGFβ receptor II expression. As expected, PAI-1 KO cells had unregulated plasmin activity, which was only partially responsible for TGFβ activation, as evidenced by a mere 25% reduction in TGFβ activity when plasmin was inhibited. Treatment of cells with an αvβ3-specific cyclic RGD peptide (GpenGRGD) led to a more profound (59%) TGFβ inhibition; a nonspecific RGD peptide (GRGDNP) inhibited TGFβ by only 23%. Human primary fibroblasts were used to confirm that PAI-1 inhibition and β3 overexpression led to an increase in TGFβ activity. Consistent with a fibrotic phenotype, PAI-1 KO cells were constitutively myofibroblasts that had a 1.6-fold increase in collagen deposition over wild type cells. These data suggest that PAI-1-mediated regulation of αvβ3 integrin is critical for the control of TGFβ signaling and the prevention of fibrotic disease.Fibrotic disorders can result from environmental toxins, persistent infection, autoimmune disease, or mechanical injury, leading to the hardening and scarring of tissues. In fibrotic diseases, such as liver cirrhosis, renal fibrosis, and idiopathic lung fibrosis, or in pathological wound healing, such as hypertrophic scarring, scleroderma, and Dupuytren disease, the persistence of myofibroblasts contributes to disease progression by overproduction of extracellular matrix (ECM)2 and by excessive contraction (13). A shift in the balance of growth factors and cytokines that promote ECM deposition and proteases that degrade matrix often contributes to fibrotic disease (4, 5). Plasmin, a broad spectrum protease that is generated from plasminogen by uPA, is one of the proteases that degrades matrix and activates growth factors and other proteases (6). Since uPA activity is inhibited by PAI-1, the overexpression of PAI-1 results in matrix accumulation. For this reason, PAI-1 is a key prognostic marker for fibrotic disease. PAI-1 exerts its inhibitory activity on uPA by stimulating the endocytosis of the cell surface uPA·uPAR complex through the low density lipoprotein receptor-related protein (7). Integrin αvβ3 is also internalized with the uPA·uPAR·low density lipoprotein receptor-related protein complex (8). After endocytosis, uPAR and integrins are recycled back to the cell surface for another round of binding (8, 9). uPAR and αvβ3 promote cellular attachment and spreading, since they are receptors for the extracellular matrix molecule, vitronectin (10). Thus, cycling of the complex is thought to stimulate the attachment and detachment that is necessary for cell migration (8). Consequently, a shift in the expression of any of these components (PAI-1/uPA/uPAR/αvβ3) can result in either aggressive migration, as seen in cancer invasion, or a persistent increase in cell adhesion and cell tension, as seen in myofibroblasts in fibrotic tissue.The family of TGFβ growth factors has been intensively studied for their role in fibrotic wound healing. Up-regulation of TGFβ results in amplified and persistent overproduction of molecules, such as integrins and PAI-1 and other protease inhibitors (e.g. TIMPs) (2, 3). Up-regulated integrins continue the cycle of TGFβ signaling by participating in the sustained activation of TGFβ from its latent form. To date, studies have found that various αv integrins participate in the activation of TGFβ (αvβ3, αvβ5, αvβ6, and αvβ8), but the mechanism differs (1115). Integrins can serve as docking proteins to localize proteases that cleave and activate latent TGFβ in the ECM, or they can directly activate latent TGFβ in a protease-independent manner. Recently, it was discovered that latent TGFβ is also activated by mechanical stress generated from an integrin-mediated interaction between myofibroblasts and the ECM, primarily involving αvβ5. The mechanical stress promotes a conformational change that activates the latent TGFβ complex (15). αv integrins also modulate TGFβ signaling through the binding of αvβ3 to TGFβ receptor II (TGFβRII) in the presence of TGFβ. This interaction was shown to promote a dramatic increase in the proliferation of lung fibroblasts and induce invasion of epithelial breast cancer cells (16, 17).Our data establish a role for the PAI-1-mediated control of αvβ3 expression and support a significant role for αvβ3 in TGFβ signaling. Using PAI-1 KO cells, we tested the hypothesis that the absence of PAI-1 would result in the accumulation of αvβ3 on the cell surface, since PAI-1 promotes the endocytosis of uPA·uPAR·αvβ3. PAI-1-mediated endocytosis of β3 was significantly reduced in the PAI-1 KO cells. Correspondingly, we report that β3 accumulated at the cell surface in enlarged β3-containing focal adhesions. Thus, we explored whether the accumulation of αvβ3 on the cell surface had fibrogenic effects even in the absence of profibrotic PAI-1. Our results demonstrate dramatically increased TGFβ activity and an increase in collagen expression in PAI-1 KO cells. Together, these findings suggest that PAI-1 modulates β3 expression and localization and, in turn, TGFβ signaling. Our data reveal that maintaining precise levels of PAI-1 is a key to preventing fibrosis. Understanding the consequence of regulating PAI-1 activity is critical in light of the many clinical therapies currently under development that target PAI-1 (18, 19).  相似文献   

9.
Platelets are activated by adhesion to vascular collagen via the immunoglobulin receptor, glycoprotein VI (GPVI). This causes potent signaling toward activation of phospholipase Cγ2, which bears similarity to the signaling pathway evoked by T- and B-cell receptors. Phosphoinositide 3-kinase (PI3K) plays an important role in collagen-induced platelet activation, because this activity modulates the autocrine effects of secreted ADP. Here, we identified the PI3K isoforms directly downstream of GPVI in human and mouse platelets and determined their role in GPVI-dependent thrombus formation. The targeting of platelet PI3Kα or -β strongly and selectively suppressed GPVI-induced Ca2+ mobilization and inositol 1,4,5-triphosphate production, thus demonstrating enhancement of phospholipase Cγ2 by PI3Kα/β. That PI3Kα and -β have a non-redundant function in GPVI-induced platelet activation and thrombus formation was concluded from measurements of: (i) serine phosphorylation of Akt, (ii) dense granule secretion, (iii) intracellular Ca2+ increases and surface expression of phosphatidylserine under flow, and (iv) thrombus formation, under conditions where PI3Kα/β was blocked or p85α was deficient. In contrast, GPVI-induced platelet activation was insensitive to inhibition or deficiency of PI3Kδ or -γ. Furthermore, PI3Kα/β, but not PI3Kγ, contributed to GPVI-induced Rap1b activation and, surprisingly, also to Rap1b-independent platelet activation via GPVI. Together, these findings demonstrate that both PI3Kα and -β isoforms are required for full GPVI-dependent platelet Ca2+ signaling and thrombus formation, partly independently of Rap1b. This provides a new mechanistic explanation for the anti-thrombotic effect of PI3K inhibition and makes PI3Kα an interesting new target for anti-platelet therapy.  相似文献   

10.
Alcadeins (Alcs) constitute a family of neuronal type I membrane proteins, designated Alcα, Alcβ, and Alcγ. The Alcs express in neurons dominantly and largely colocalize with the Alzheimer amyloid precursor protein (APP) in the brain. Alcs and APP show an identical function as a cargo receptor of kinesin-1. Moreover, proteolytic processing of Alc proteins appears highly similar to that of APP. We found that APP α-secretases ADAM 10 and ADAM 17 primarily cleave Alc proteins and trigger the subsequent secondary intramembranous cleavage of Alc C-terminal fragments by a presenilin-dependent γ-secretase complex, thereby generating “APP p3-like” and non-aggregative Alc peptides (p3-Alcs). We determined the complete amino acid sequence of p3-Alcα, p3-Alcβ, and p3-Alcγ, whose major species comprise 35, 37, and 31 amino acids, respectively, in human cerebrospinal fluid. We demonstrate here that variant p3-Alc C termini are modulated by FAD-linked presenilin 1 mutations increasing minor β-amyloid species Aβ42, and these mutations alter the level of minor p3-Alc species. However, the magnitudes of C-terminal alteration of p3-Alcα, p3-Alcβ, and p3-Alcγ were not equivalent, suggesting that one type of γ-secretase dysfunction does not appear in the phenotype equivalently in the cleavage of type I membrane proteins. Because these C-terminal alterations are detectable in human cerebrospinal fluid, the use of a substrate panel, including Alcs and APP, may be effective to detect γ-secretase dysfunction in the prepathogenic state of Alzheimer disease subjects.  相似文献   

11.
12.
Elastin fibers are predominantly composed of the secreted monomer tropoelastin. This protein assembly confers elasticity to all vertebrate elastic tissues including arteries, lung, skin, vocal folds, and elastic cartilage. In this study we examined the mechanism of cell interactions with recombinant human tropoelastin. Cell adhesion to human tropoelastin was divalent cation-dependent, and the inhibitory anti-integrin αVβ3 antibody LM609 inhibited cell spreading on tropoelastin, identifying integrin αVβ3 as the major fibroblast cell surface receptor for human tropoelastin. Cell adhesion was unaffected by lactose and heparin sulfate, indicating that the elastin-binding protein and cell surface glycosaminoglycans are not involved. The C-terminal GRKRK motif of tropoelastin can bind to cells in a divalent cation-dependent manner, identifying this as an integrin binding motif required for cell adhesion.Cellular interactions with extracellular matrix proteins are vital for cell survival and tissue maintenance. The attachment of cells to their extracellular matrix (ECM)3 is often mediated by cell surface integrins. As such, integrins are involved in many biological functions such cell migration and proliferation, tissue organization, wound repair, development, and host immune responses. In addition to roles under normal physiological conditions, integrins are involved in the pathogenesis of diseases such as arthritis, cardiovascular disease, inflammation, microbial and parasitic infection, and cancer. Integrins are a family of heterodimeric transmembrane receptors containing one α subunit and one β subunit (1). Often integrins bind to ECM proteins via short RGD motifs within the matrix protein (2). In addition to an RGD motif, fibronectin also contains an upstream PHSRN synergy sequence, which is required for full integrin binding activity (3).Elastin confers elasticity on all vertebrate elastic tissues including arteries, lung, skin, vocal fold, and elastic cartilage (4). Elastin comprises ∼90% of the elastic fiber and is intermingled with fibrillin-rich microfibrils (5). There is a single human tropoelastin gene in which alternative splicing can result in the loss of domains 22, 23, 24, 26A, 30, 32, and 33 (4). Elastin is made from the secreted monomer tropoelastin, which is a 60–72-kDa protein containing repeating hydrophobic and cross-linking domains. Hydrophobic domains are rich in GVGVP, GGVP, and GVGVAP repeats, which can associate by coacervation (6). This association results in structural changes and increased α-helical content (7). The cross-linking domains are lysine-rich. Occasionally these residues are modified to allysine through the activity of members of the family of lysyl oxidase (LOX) and four LOX-like enzymes. During coacervation the allysine and other allysines or specific lysine side chains come into close proximity, allowing nonenzymatic condensation reactions to occur, forming desmosine or isodesmosine cross-links (4). This process gives a highly stable cross-linked elastin matrix which has a half-life of ∼70 years. Members of the serine, aspartate, cysteine, and matrix metalloproteinase families of proteases can degrade elastin (8). The resulting elastin peptides have effects on ECM synthesis and cell attachment, migration, and proliferation (9).The consequences of mutated or hemizygous elastin in the hereditary, connective tissue disorders cutis laxa, supravalvular aortic stenosis, and Williams-Beuren syndrome highlight the elastins essential role in elastic tissue function (10). Elastin is the major protein in large elastic blood vessels such as the aorta, where it is likely to inhibit the proliferation of vascular smooth muscle cells and so preventing vessel occlusion (11), which is a major cause of death in developed countries. Previous studies have shown that human and bovine tropoelastin can bind directly to a variety of cell types directly through a number of cell surface receptors (1214) and also bind indirectly to cells through ECM proteins such as fibulin-5 (15, 16).A mechanism by which elastin binds to cells is via the 67-kDa elastin-binding protein (EBP), which is a peripheral membrane splice variant of β-galactosidase. The EBP forms a complex with the integral membrane proteins carboxypeptidase A and sialidase, forming a transmembrane elastin receptor (12). The binding site for the EBP has been mapped to the consensus sequence XGXXPG within elastin and in particular to VGVAPG within exon 24 (17). The binding of elastin to the EBP results in cell morphological changes (18, 19), chemotaxis (20), decreased cell proliferation (21), and angiogenesis (22). Knockouts of β-galactosidase, which remove the EBP, display correctly deposited elastin (27). Additionally tropoelastin actively promotes cell adhesion, whereas VGVAPG does not. These observations imply that receptors other than EBP can interact with elastin.Other studies have proposed a second mechanism involving the necessity of cell surface heparan and chondroitin sulfate-containing glycosaminoglycans for bovine chondrocyte interaction with bovine tropoelastin (14). Peptide binding analysis implicated the last 17 amino acids at the C terminus of bovine tropoelastin in this cell adhesive activity, with higher binding requiring the C-terminal 25 amino acids. This region is of interest, as in humans a mutation of Gly-773 to Asp in exon 33 results in blocked elastin network assembly and modulates cell binding to a peptide corresponding to exons 33 and 36 of human tropoelastin (28). Indeed Broekelmann et al. (14) have shown that synthetic peptides containing the C-terminal 29 amino acids of bovine tropoelastin possess cell adhesive activity; however, when the G773D mutation was incorporated into the peptide, it prevented cell adhesion to that peptide.Although tropoelastin does not contain an RGD motif, other data identified a third mechanism involving direct interaction between integrin αvβ3 and human tropoelastin (13, 29). This interaction was also localized to the C-terminal domains of tropoelastin.More recent data has shown that human umbilical vein endothelial cells can adhere to recombinant fragments of human tropoelastin (30, 31). In contrast to other data, regions encoded by the N-terminal exons (1–18), the central exons (18–27), and the C-terminal exons (18–36) all supported human umbilical vein endothelial cell attachment.Although a previous study has shown a direct interaction between purified integrin αvβ3 and human tropoelastin (13), the integrin dependence of cell adhesion to tropoelastin had not been demonstrated. Here we demonstrate that human dermal fibroblasts adhere to recombinant human tropoelastin and that inhibitors of the elastin-binding protein and cell surface heparan sulfate have no effect on cell adhesion. In contrast, cell adhesion was dependent upon the presence of divalent cations, indicating integrin dependence. Inhibitory monoclonal antibodies identified integrin αVβ3 as the major receptor necessary for fibroblast adherence and spreading onto human tropoelastin. The binding motif for integrin-mediated cell adhesion is unknown; therefore, through the use of synthetic peptides, the adhesive activity was localized to the extreme C-terminal GRKRK motif of tropoelastin. This data present a novel mechanism for cell adhesion to human tropoelastin and identify a novel integrin binding motif within tropoelastin.  相似文献   

13.
14.
Synapses are asymmetric intercellular junctions that mediate neuronal communication. The number, type, and connectivity patterns of synapses determine the formation, maintenance, and function of neural circuitries. The complexity and specificity of synaptogenesis relies upon modulation of adhesive properties, which regulate contact initiation, synapse formation, maturation, and functional plasticity. Disruption of adhesion may result in structural and functional imbalance that may lead to neurodevelopmental diseases, such as autism, or neurodegeneration, such as Alzheimer''s disease. Therefore, understanding the roles of different adhesion protein families in synapse formation is crucial for unraveling the biology of neuronal circuit formation, as well as the pathogenesis of some brain disorders. The present review summarizes some of the knowledge that has been acquired in vertebrate and invertebrate genetic model organisms.Synapses are asymmetric, intercellular junctions that are the basic structural units of neuronal transmission. The correct development of synaptic specializations and the establishment of appropriate connectivity patterns are crucial for the assembly of functional neuronal circuits. Improper synapse formation and function may cause neurodevelopmental disorders, such as mental retardation (MsR) and autism spectrum disorders (ASD) (McAllister 2007; Sudhof 2008), and likely play a role in neurodegenerative disorders, such as Alzheimer''s disease (AD) (Haass and Selkoe 2007).At chemical synapses (reviewed in Sudhof 2004; Zhai and Bellen 2004; Waites et al. 2005; McAllister 2007; Jin and Garner 2008), the presynaptic compartment contains synaptic vesicles (SV), organized in functionally distinct subcellular pools. A subset of SVs docks to the presynaptic membrane around protein-dense release sites, named active zones (AZ). Upon the arrival of an action potential at the terminal, the docked and “primed” SVs fuse with the plasma membrane and release neurotransmitter molecules into the synaptic cleft. Depending on the type of synapse (i.e., excitatory vs. inhibitory synapses), neurotransmitters ultimately activate an appropriate set of postsynaptic receptors that are accurately apposed to the AZ.Synapse formation occurs in several steps (Fig. 1) (reviewed in Eaton and Davis 2003; Goda and Davis 2003; Waites et al. 2005; Garner et al. 2006; Gerrow and El-Husseini 2006; McAllister 2007). Spatiotemporal signals guide axons through heterogeneous cellular environments to contact appropriate postsynaptic targets. At their destination, axonal growth cones initiate synaptogenesis through adhesive interactions with target cells. In the mammalian central nervous system (CNS), immature postsynaptic dendritic spines initially protrude as thin, actin-rich filopodia on the surface of dendrites. Similarly, at the Drosophila neuromuscular junction (NMJ), myopodia develop from the muscles (Ritzenthaler et al. 2000). The stabilization of intercellular contacts and their elaboration into mature, functional synapses involves cytoskeletal arrangements and recruitment of pre- and postsynaptic components to contact sites in spines and boutons. Conversely, retraction of contacts results in synaptic elimination. Both stabilization and retraction sculpt a functional neuronal circuitry.Open in a separate windowFigure 1.(A–C) Different stages of synapse formation. (A) Target selection, (B) Synapse assembly, (C) Synapse maturation and stabilization. (D–F) The role of cell adhesion molecules in synapse formation is exemplified by the paradigm of N-cadherin and catenins in regulation of the morphology and strength of dendritic spine heads. (D) At an early stage the dendritic spines are elongated from motile structures “seeking” their synaptic partners. (E) The contacts between the presynaptic and postsynaptic compartments are stabilized by recruitment of additional cell adhesion molecules. Adhesional interactions activate downstream pathways that remodel the cytoskeleton and organize pre- and postsynaptic apparatuses. (F) Cell adhesion complexes, stabilized by increased synaptic activity, promote the expansion of the dendritic spine head and the maturation/ stabilization of the synapse. Retraction and expansion is dependent on synaptic plasticity.In addition to the plastic nature of synapse formation, the vast heterogeneity of synapses (in terms of target selection, morphology, and type of neurotransmitter released) greatly enhances the complexity of synaptogenesis (reviewed in Craig and Boudin 2001; Craig et al. 2006; Gerrow and El-Husseini 2006). The complexity and specificity of synaptogenesis relies upon the modulation of adhesion between the pre- and postsynaptic components (reviewed in Craig et al. 2006; Gerrow and El-Husseini 2006; Piechotta et al. 2006; Dalva et al. 2007; Shapiro et al. 2007; Yamada and Nelson 2007; Gottmann 2008). Cell adhesive interactions enable cell–cell recognition via extracellular domains and also mediate intracellular signaling cascades that affect synapse morphology and organize scaffolding complexes. Thus, cell adhesion molecules (CAMs) coordinate multiple synaptogenic steps.However, in vitro and in vivo studies of vertebrate CAMs are often at odds with each other. Indeed, there are no examples of mutants for synaptic CAMs that exhibit prominent defects in synapse formation. This apparent “resilience” of synapses is probably caused by functional redundancy or compensatory effects among different CAMs (Piechotta et al. 2006). Hence, studies using simpler organisms less riddled by redundancy, such as Caenorhabditis elegans and Drosophila, have aided in our understanding of the role that these molecules play in organizing synapses.In this survey, we discuss the roles of the best characterized CAM families of proteins involved in synaptogenesis. Our focus is to highlight the complex principles that govern the molecular basis of synapse formation and function from a comparative perspective. We will present results from cell culture studies as well as in vivo analyses in vertebrate systems and refer to invertebrate studies, mainly performed in Drosophila and C. elegans, when they have provided important insights into the role of particular CAM protein families. However, we do not discuss secreted factors, for which we refer the reader to numerous excellent reviews (as for example Washbourne et al. 2004; Salinas 2005; Piechotta et al. 2006; Shapiro et al. 2006; Dalva 2007; Yamada and Nelson 2007; Biederer and Stagi 2008; Salinas and Zou 2008).  相似文献   

15.
The signaling mechanisms facilitating cardiomyocyte (CM) differentiation from bone marrow (BM)-derived mesenchymal stem cells (MSCs) are not well understood. 5-Azacytidine (5-Aza), a DNA demethylating agent, induces expression of cardiac-specific genes, such as Nkx2.5 and α-MHC, in mouse BM-derived MSCs. 5-Aza treatment caused significant up-regulation of glycogen synthase kinase (GSK)-3β and down-regulation of β-catenin, whereas it stimulated GSK-3α expression only modestly. The promoter region of GSK-3β was heavily methylated in control MSCs, but was demethylated by 5-Aza. Although overexpression of GSK-3β potently induced CM differentiation, that of GSK-3α induced markers of neuronal and chondrocyte differentiation. GSK-3 inhibitors, including LiCl, SB 216743, and BIO, abolished 5-Aza-induced up-regulation of CM-specific genes, suggesting that GSK-3 is necessary and sufficient for CM differentiation in MSCs. Although specific knockdown of endogenous GSK-3β abolished 5-Aza-induced expression of cardiac specific genes, surprisingly, that of GSK-3α facilitated CM differentiation in MSCs. Although GSK-3β is found in both the cytosol and nucleus in MSCs, GSK-3α is localized primarily in the nucleus. Nuclear-specific overexpression of GSK-3β failed to stimulate CM differentiation. Down-regulation of β-catenin mediates GSK-3β-induced CM differentiation in MSCs, whereas up-regulation of c-Jun plays an important role in mediating CM differentiation induced by GSK-3α knockdown. These results suggest that GSK-3α and GSK-3β have distinct roles in regulating CM differentiation in BM-derived MSCs. GSK-3β in the cytosol induces CM differentiation of MSCs through down-regulation of β-catenin. In contrast, GSK-3α in the nucleus inhibits CM differentiation through down-regulation of c-Jun.  相似文献   

16.
Although there are numerous reports of carbohydrates enriched in cancer cells, very few studies have addressed the functions of carbohydrates present in normal cells that decrease in cancer cells. It has been reported that core3 O-glycans are synthesized in normal gastrointestinal cells but are down-regulated in cancer cells. To determine the roles of core3 O-glycans, we transfected PC3 and LNCaP prostate cancer cells with β3-N-acetylglucosaminyltransferase-6 (core3 synthase) required to synthesize core3 O-glycans. Both engineered cell lines exhibited reduced migration and invasion through extracellular matrix components compared with mock-transfected cells. Moreover we found that α2β1 integrin acquired core3 O-glycans in cells expressing core3 synthase with decreased maturation of β1 integrin, leading to decreased levels of the α2β1 integrin complex, decreased activation of focal adhesion kinase, and reduced lamellipodia formation. Upon inoculation into the prostate of nude mice, PC3 cells expressing core3 O-glycans produced much smaller tumors without metastasis to the surrounding lymph nodes in contrast to robust tumor formation and metastasis seen in mock-transfected PC3 cells. Similarly LNCaP cells expressing core3 O-glycans barely produced subcutaneous tumors in contrast to robust tumor formation by mock-transfected LNCaP cells. These findings indicate that addition of core3 O-glycans to β1 and α2 integrin subunits in prostate cancer cells suppresses tumor formation and tumor metastasis.Cancer cells often express surface carbohydrates different from normal cells (1). One such change is expression of sialyl Lewis X and Lewis B blood group antigens in cancer cells (2, 3). These structural elements are seen as capping oligosaccharides attached to the underlying glycan backbone where they likely function as ligands for cell adhesion molecules.The structure of underlying glycans also changes during malignant transformation and differentiation. In particular, there are several reports that an increase in the β1,6-N-acetylglucosaminyl branch in N-glycans synthesized by β1,6-N-acetylglucosaminyltransferase-V is associated with oncogenic transformation (47). Similar structural changes are seen in mucin-type O-glycans, which have N-acetylgalactosamine at the reducing end linked to polypeptide threonine or serine residues. Addition of different carbohydrate residues to N-acetylgalactosamine confers a variety of backbone structures on mucin-type O-glycans; the most abundant of those are classified as core1, core2, core3, and core4 O-glycans (8) (Fig. 1). Among these O-glycans, the synthesis of the core2 branch has been extensively studied particularly because conversion of core1 to core2 O-glycans was observed in T cell activation (9). Expression of core2 branch apparently represents an oncodifferentiation antigen because core2 branched O-glycans are synthesized in early stages of T cell differentiation, down-regulated in mature T cells, and reappear in T cell leukemia and immune deficiencies such as AIDS and Wiskott-Aldrich syndrome (for a review, see Ref. 10). In addition, overexpression of core2 O-glycans is seen in many cancers, including lung and breast carcinoma cells (11, 12).Open in a separate windowFIGURE 1.Biosynthetic pathways of mucin-type O-glycans. N-Acetylgalactosamine is transferred to a serine or threonine residue in a polypeptide. Resultant GalNAcα1→Ser/Thr is converted by core3 synthase (β3GnT-6) to GlcNAcβ1→3GalNAcα1→Ser/Thr (core3). Core3 is then converted to core4 by C2GnT-2 (C2GnT-M). GalNAcα1→Ser/Thr is also converted to core1, Galβ1→3GalNAcα1→Ser/Thr, by core1 synthase. Core1 is then converted to core2 by C2GnT-1, C2GnT-2, and C2GnT-3.By contrast, core3 and core4 O-glycans are synthesized in normal cells but apparently down-regulated in gastric and colorectal carcinoma (13, 14). Core3 O-glycans are synthesized by core3 synthase (β3GnT-6),2 which adds β1,3-linked N-acetylglucosamine to N-acetylgalactosamine at the reducing terminus (15) (Fig. 1). Iwai et al. (16) showed that forced expression of core3 synthase in human fibrosarcoma HT1080 FP-10 cells resulted in significant reduction in the formation of lung tumor foci in mice after intravenous injection of tumor cells through a tail vein. However, the same study did not address whether the expression of core3 influences tumor metastasis because the cancer cells were intravenously injected and no primary tumor was formed to spread into the lung as metastasis in contrast to the other studies (17, 18). Core4 O-glycan is synthesized by addition of β1,6-linked N-acetylglucosamine to a core3 acceptor by core2 β1,6-N-acetylglucosamine M type (C2GnT-M) or C2GnT-2 (19, 20) (Fig. 1). Huang et al. (21) reported that C2GnT-M is down-regulated in colonic carcinoma cells and that forced expression of C2GnT-M in HCT116 colonic carcinoma cells significantly decreased cell invasion and subcutaneous tumor formation. How up-regulation of core3 and core4 O-glycans influences the pathophysiology of cells expressing core3 and core4 O-glycans has not been addressed.Cell-extracellular matrix interaction plays an essential role during acquisition of migration and invasive behavior of cancer cells. For example, α2β1 integrin is the major receptor for collagen (22) and most abundantly expressed in prostate cancer cells (23). Glycosylation on integrin is one of the important modulators of integrin functions, and many glycan structures, mainly N-glycans, have been studied. An increase of bisecting GlcNAc structure on α5β1 integrin inhibits the cell spreading and migration (24), and induced β1,6-GlcNAc sugar chains on N-glycans of β1 integrin result in stimulation of cell migration (25). However, it has not been addressed whether changes in O-glycans affect integrin maturation and functions.To determine the role of core3 O-glycans in tumor formation and metastasis, we analyzed PC3 and LNCaP human prostate cancer cells. We found that these cell lines express only small amounts of detectable core3 synthase; thus we transfected the cell lines with core3 synthase. Core3 synthase-transfected PC3 and LNCaP cells expressed increased amounts of core3 O-glycans in α2β1 integrin, showed the reduced maturation of β1 integrin and low levels of α2β1 integrin formation, migrated less efficiently through collagen and other extracellular matrix components, and were less invasive than mock-transfected cells. Moreover those cells exhibited decreased activation of focal adhesion kinase (FAK) compared with mock-transfected cells. Significantly PC3 cells expressing core3 O-glycans produced almost no primary tumors in the prostate and formed much fewer metastases in the draining lymph nodes than mock-transfected cells. Similarly LNCaP cells expressing core3 O-glycans produced much smaller subcutaneous tumors than mock-transfected LNCaP cells. These findings indicate that addition of core3 O-glycans to the α2β1 integrin leads to decreased cell migration and invasion, resulting in decreased prostate tumor formation and metastasis.  相似文献   

17.
Integrin αIIbβ3 affinity regulation by talin binding to the cytoplasmic tail of β3 is a generally accepted model for explaining activation of this integrin in Chinese hamster ovary cells and human platelets. Most of the evidence for this model comes from the use of multivalent ligands. This raises the possibility that the activation being measured is that of increased clustering of the integrin rather than affinity. Using a newly developed assay that probes integrins on the surface of cells with only monovalent ligands prior to fixation, I do not find increases in affinity of αIIbβ3 integrins by talin head fragments in Chinese hamster ovary cells, nor do I observe affinity increases in human platelets stimulated with thrombin. Binding to a multivalent ligand does increase in both of these cases. This assay does report affinity increases induced by either Mn2+, a cytoplasmic domain mutant (D723R) in the cytoplasmic domain of β3, or preincubation with a peptide ligand. These results reconcile the previously observed differences between talin effects on integrin activation in Drosophila and vertebrate systems and suggest new models for talin regulation of integrin activity in human platelets.  相似文献   

18.
MEKK3 serves as a critical intermediate signaling molecule in lysophosphatidic acid-mediated nuclear factor-κB (NF-κB) activation. However, the precise regulation for MEKK3 activation at the molecular level is still not fully understood. Here we report the identification of two regulatory phosphorylation sites at Thr-516 and Ser-520 within the kinase activation loop that is essential for MEKK3-mediated IκB kinase β (IKKβ)/NF-κB activation. Substitution of these two residues with alanine abolished the ability of MEKK3 to activate IKKβ/NF-κB, whereas replacement with acidic residues rendered MEKK3 constitutively active. Furthermore, substitution of these two residues with alanine abolished the ability of MEKK3 to mediate lysophosphatidic acid-induced optimal IKKβ/NF-κB activation.  相似文献   

19.
Synaptic degeneration, including impairment of synaptic plasticity and loss of synapses, is an important feature of Alzheimer disease pathogenesis. Increasing evidence suggests that these degenerative synaptic changes are associated with an accumulation of soluble oligomeric assemblies of amyloid β (Aβ) known as ADDLs. In primary hippocampal cultures ADDLs bind to a subpopulation of neurons. However the molecular basis of this cell type-selective interaction is not understood. Here, using siRNA screening technology, we identified α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor subunits and calcineurin as candidate genes potentially involved in ADDL-neuron interactions. Immunocolocalization experiments confirmed that ADDL binding occurs in dendritic spines that express surface AMPA receptors, particularly the calcium-impermeable type II AMPA receptor subunit (GluR2). Pharmacological removal of the surface AMPA receptors or inhibition of AMPA receptors with antagonists reduces ADDL binding. Furthermore, using co-immunoprecipitation and photoreactive amino acid cross-linking, we found that ADDLs interact preferentially with GluR2-containing complexes. We demonstrate that calcineurin mediates an endocytotic process that is responsible for the rapid internalization of bound ADDLs along with surface AMPA receptor subunits, which then both colocalize with cpg2, a molecule localized specifically at the postsynaptic endocytic zone of excitatory synapses that plays an important role in activity-dependent glutamate receptor endocytosis. Both AMPA receptor and calcineurin inhibitors prevent oligomer-induced surface AMPAR and spine loss. These results support a model of disease pathogenesis in which Aβ oligomers interact selectively with neurotransmission pathways at excitatory synapses, resulting in synaptic loss via facilitated endocytosis. Validation of this model in human disease would identify therapeutic targets for Alzheimer disease.  相似文献   

20.
Protein tyrosine phosphatase α (PTPα) promotes integrin-stimulated cell migration in part through the role of Src-phosphorylated PTPα-Tyr(P)-789 in recruiting and localizing p130Cas to focal adhesions. The growth factor IGF-1 also stimulates PTPα-Tyr-789 phosphorylation to positively regulate cell movement. This is in contrast to integrin-induced PTPα phosphorylation, that induced by IGF-1 can occur in cells lacking Src family kinases (SFKs), indicating that an unknown kinase distinct from SFKs can target PTPα. We show that this IGF-1-stimulated tyrosine kinase is Abl. We found that PTPα binds to the scaffold protein RACK1 and that RACK1 coordinates the IGF-1 receptor, PTPα, and Abl in a complex to enable IGF-1-stimulated and Abl-dependent PTPα-Tyr-789 phosphorylation. In cells expressing SFKs, IGF-1-stimulated phosphorylation of PTPα is mediated by RACK1 but is Abl-independent. Furthermore, expressing the SFKs Src and Fyn in SFK-deficient cells switches IGF-1-induced PTPα phosphorylation to occur in an Abl-independent manner, suggesting that SFK activity dominantly regulates IGF-1/IGF-1 receptor signaling to PTPα. RACK1 is a molecular scaffold that integrates growth factor and integrin signaling, and our identification of PTPα as a RACK1 binding protein suggests that RACK1 may coordinate PTPα-Tyr-789 phosphorylation in these signaling networks to promote cell migration.  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