首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 0 毫秒
1.
2.

Background

Immunosuppressant cyclosporine-A induces gingival hyperplasia, which is characterized by increased fibroblast proliferation and overproduction of extracellular matrix components and regulated by transforming growth factor-beta (TGF-β). The TGF-β and Sonic hedgehog (Shh) pathways both mediate cell proliferation. Crosstalk between these pathways in cancer has recently been proposed, but the hierarchical pattern of this crosstalk remains unclear. In normal fibroblasts, a TGF-β-stimulating Shh pattern was observed in induced fibrosis. However, Shh pathway involvement in cyclosporine-enhanced gingival proliferation and the existence of crosstalk with the TGF-β pathway remain unclear.

Methodology/Principal Findings

Cyclosporine enhanced mRNA and protein levels of TGF-β and Shh in human gingival fibroblasts (RT-PCR and western blotting). A TGF-β pathway inhibitor mitigated cyclosporine-enhanced cell proliferation and an Shh pathway inhibitor attenuated cyclosporine-enhanced proliferation in fibroblasts (MTS assay and/or RT-PCR of PCNA). Exogenous TGF-β increased Shh expression; however, exogenous Shh did not alter TGF-β expression. The TGF-β pathway inhibitor mitigated cyclosporine-upregulated Shh expression, but the Shh pathway inhibitor did not alter cyclosporine-upregulated TGF-β expression.

Conclusions/Significance

The TGF-β and Shh pathways mediate cyclosporine-enhanced gingival fibroblast proliferation. Exogenous TGF-β increased Shh expression, and inhibition of TGF-β signaling abrogated the cyclosporine-induced upregulation of Shh expression; however, TGF-β expression appeared unchanged by enhanced or inhibited Shh signaling. This is the first study demonstrating the role of Shh in cyclosporine-enhanced gingival cell proliferation; moreover, it defines a hierarchical crosstalk pattern in which TGF-β regulates Shh in gingival fibroblasts. Understanding the regulation of cyclosporine-related Shh and TGF-β signaling and crosstalk in gingival overgrowth will clarify the mechanism of cyclosporine-induced gingival enlargement and help develop targeted therapeutics for blocking these pathways, which can be applied in pre-clinical and clinical settings.  相似文献   

3.
4.
It has been proposed that ligand occupancy of integrin αvβ3 with extracellular matrix ligands (e.g. vitronectin) plays a critical role in insulin-like growth factor-1 (IGF-1) signaling. We found that expression of αvβ3 enhanced IGF-1-induced proliferation of Chinese hamster ovary cells in serum-free conditions (in the absence of vitronectin). We hypothesized that the direct integrin binding to IGF-1 may play a role in IGF-1 signaling. We demonstrated that αvβ3 specifically and directly bound to IGF-1 in cell adhesion, enzyme-linked immunosorbent assay-type binding, and surface plasmon resonance studies. We localized the amino acid residues of IGF-1 that are critical for integrin binding by docking simulation and mutagenesis. We found that mutating two Arg residues at positions 36 and 37 in the C-domain of IGF-1 to Glu (the R36E/R37E mutation) effectively reduced integrin binding. Interestingly, although the mutant still bound to IGF1R, it was defective in inducing IGF1R phosphorylation, AKT and ERK1/2 activation, and cell proliferation. Furthermore wild type IGF-1 mediated co-precipitation of αvβ3 and IGF1R, whereas the R36E/R37E mutant did not, suggesting that IGF-1 mediates the interaction between αvβ3 and IGF1R. These results suggest that the direct binding to IGF-1 to integrin αvβ3 plays a role in IGF-1 signaling through ternary complex formation (αvβ3-IGF-IGF1R), and integrin-IGF-1 interaction is a novel target for drug discovery.Integrins are a family of cell adhesion receptors that mediate cell-extracellular matrix (ECM)3 interaction and cell-cell interaction (1). It has been proposed that signaling from inside the cells regulates the ligand binding affinity of integrins (inside-out signaling) (2). Each integrin is a heterodimer containing α and β subunits. At present 18 α and 8 β subunits have been identified that combine to form 24 integrins (3).It has been reported that integrin αvβ3 plays a role in cancer proliferation and invasiveness. High levels of integrin αvβ3 correlate with growth and/or progression of melanoma (4, 5), neuroblastoma (6), breast cancer (7, 8), colon cancer (9), ovarian cancer (10), and cervical cancer (11). Moreover, individuals homozygous for the β3L33P polymorphism that enhances the ligand binding affinity of β3 integrins have an increased risk to develop breast cancer, ovarian cancer, and melanoma (12). However, it remains unclear whether and how increased levels of αvβ3 on tumor cells contribute to cancer development.Insulin-like growth factor-1 (IGF-1) is a polypeptide hormone (75 kDa) that has a high degree of structural similarity to human proinsulin. IGF-1 acts through binding to the type I IGF receptor (IGF1R), a receptor tyrosine kinase. The IGF1R is a heterotetramer that consists of two α-subunits that contain the ligand-binding domains and two β-subunits that contain the tyrosine kinase activity. After ligand binding, the receptor undergoes a conformational change resulting in the activation of the tyrosine kinase, which results in transphosphorylation of the opposite β-subunit on specific tyrosine residues. These phosphotyrosines then bind to adapter molecules such as Shc and IRS-1. Phosphorylation of these proteins leads to activation of the phosphatidylinositol 3-kinase and mitogen-activated protein kinase (MAPK) signaling pathways (reviewed in Ref. 13).IGF-1 has been implicated in cancer progression (14). One of the major actions of IGF-1 is to inhibit apoptosis. IGF-1 confers resistance to chemotherapy and radiation therapy. IGF-1 expression levels are increased in breast, lung, prostate, and many other cancers. Several strategies to target IGF-1 signaling have been extensively studied, including small interfering RNA and monoclonal antibodies for IGF1R and kinase inhibitors to inhibit the enzymatic activity of the receptor. The IGF-1 system is a therapeutic target for cancer, and elucidation of the IGF-1 signaling pathway should have a major impact in designing new therapeutic strategies.It has been proposed that ligand occupancy of αvβ3 with ECM ligands such as vitronectin plays a critical role in enhancing IGF-1 signaling (14). It has been reported that inhibiting αvβ3-ECM interaction (“ligand occupancy”) of αvβ3 inhibited IGF-1 actions selectively in cell types that express αvβ3 (14). Inhibiting ligand occupancy of αvβ3 blocked IGF-1-induced cell migration (15), DNA synthesis, IRS-1 phosphorylation, and IGF1R-linked downstream signaling events, such as activation of phosphatidylinositol 3-kinase and ERK1/2 (16).In the present study, we demonstrated that expression of αvβ3 enhanced proliferation of ovarian cancer cells in the presence of fetal bovine serum (FBS) and in serum-free conditions if IGF-1 was present. This suggests that IGF-1 is involved in enhanced proliferation of αvβ3-expressing cells. We demonstrated that αvβ3 bound to IGF-1 in several different binding assays. We found that two Arg residues at positions 36 and 37 in the C-domain of IGF-1 are critical for integrin binding by docking simulation and mutagenesis. Mutation of these Arg residues to Glu (the R36E/R37E mutation) effectively reduced integrin binding. Interestingly, the R36E/R37E mutant was defective in inducing cell proliferation and IGF-1 intracellular signaling, although it still bound to IGF1R. We demonstrated that wild type IGF-1 mediated co-precipitation of αvβ3 and IGF1R, whereas the R36E/R37E mutant did not, suggesting that IGF-1 mediates the interaction between αvβ3 and IGF1R. These results suggest that the direct binding to IGF-1 plays a role in IGF-1 signaling.  相似文献   

5.
The multifunctional protein nucleolin (NCL) is overexpressed on the surface of activated endothelial and tumor cells and mediates the stimulatory actions of several angiogenic growth factors, such as pleiotrophin (PTN). Because αvβ3 integrin is also required for PTN-induced cell migration, the aim of the present work was to study the interplay between NCL and αvβ3 by using biochemical, immunofluorescence, and proximity ligation assays in cells with genetically altered expression of the studied molecules. Interestingly, cell surface NCL localization was detected only in cells expressing αvβ3 and depended on the phosphorylation of β3 at Tyr773 through receptor protein-tyrosine phosphatase β/ζ (RPTPβ/ζ) and c-Src activation. Downstream of αvβ3, PI3K activity mediated this phenomenon and cell surface NCL was found to interact with both αvβ3 and RPTPβ/ζ. Positive correlation of cell surface NCL and αvβ3 expression was also observed in human glioblastoma tissue arrays, and inhibition of cell migration by cell surface NCL antagonists was observed only in cells expressing αvβ3. Collectively, these data suggest that both expression and β3 integrin phosphorylation at Tyr773 determine the cell surface localization of NCL downstream of the RPTPβ/ζ/c-Src signaling cascade and can be used as a biomarker for the use of cell surface NCL antagonists as anticancer agents.  相似文献   

6.
Blocking the expression of integrin α2β1, which was accomplished by transduction of α2-specific shRNA, resulted in significant inhibition of proliferation and clonal activity in human MCF-7 breast carcinoma and SK-Mel-147 melanoma cells. Along with these changes, deprivation of α2β1 caused a sharp decrease in melanoma cell invasion in vitro. Analysis of integrin-mediating signal pathways that control cell behavior revealed a significant increase in activity of Akt protein kinase in response to depletion of α2β1. The increase in Akt activity that accompanies a suppressive effect on cell invasion contradicts well-known Akt function aimed at stimulation of tumor progression. This contradiction could be explained by the “reversed” (noncanonical) role played by Akt in some cells that consists in suppression rather than promotion of invasive phenotype. To test this suggestion, the effects of Akt inhibitors on invasive activity of SK-Mel-147 cells were investigated. If the above suggestion is true, then inhibition of Akt in cells depleted of α2β1 should result in the restoration of their invasive activity. It appeared that treatment with LY294002, which inhibits all Akt isoforms (Akt1, Akt2, Akt3), not only failed to restore the invasive phenotype of melanoma cells but further attenuated their invasive activity. However, treatment of the cells with an Akt1-specific inhibitor significantly increased their invasion. Thus, the stimulating effect of α2β1 integrin on invasion of melanoma cells is realized through a mechanism based on inhibition of one of the Akt isoforms, which in these cells exhibits a noncanonical function consisting in suppression of invasion.  相似文献   

7.
Classical Swine Fever (CSF) is a highly infectious fatal pig disease, resulting in huge economic loss to the swine industry. Integrins are membrane-bound signal mediators, expressed on a variety of cell surfaces and are known as receptors or co-receptors for many viruses. However, the role of integrin β3 in CSFV infection is unknown. Here, through quantitive PCR, immunofluorescence (IFC) and immunocytohistochemistry (ICC), we revealed that ST (swine testicles epithelial) cells have a prominent advantage in CSFV proliferation as compared to EC (swine umbilical vein endothelial cell), IEC (swine intestinal epithelial cell) and PK (porcine kidney epithelial) cells. Meanwhile, ST cells had remarkably more integrin β3 expression as compared to EC, IEC and PK cells, which was positively correlated with CSFV infection and proliferation. Integrin β3 was up-regulated post CSFV infection in all the four cell lines, while the CSFV proliferation rate was decreased in integrin β3 function-blocked cells. ShRNA1755 dramatically decreased integrin β3, with a deficiency of 96% at the mRNA level and 80% at the protein level. CSFV proliferation was dramatically reduced in integrin β3 constantly-defected cells (ICDC), with the deficiencies of 92.6%, 99% and 81.7% at 24 h, 48 h and 72 h post CSFV infection, respectively. These results demonstrate that integrin β3 is required in CSFV infection and proliferation, which provide a new insight into the mechanism of CSFV infection.  相似文献   

8.
Bone morphogenic protein (BMP)-7 is a member of the transforming growth factor (TGF)-beta superfamily, which is originally identified based on its ability to induce cartilage and bone formation. In recent years, BMP-7 is also defined as a potent promoter of cell motility, invasion, and metastasis. However, there is little knowledge of the role of BMP-7 and its cellular function in chondrosarcoma cells. In the present study, we investigated the biological impact of BMP-7 on cell motility using transwell assay. In addition, the intracellular signaling pathways were also investigated by pharmacological and genetic approaches. Our results demonstrated that treatment with exogenous BMP-7 markedly increased cell migration by activating c-Src/PI3K/Akt/IKK/NF-κB signaling pathway, resulting in the transactivation of αvβ3 integrin expression. Indeed, abrogation of signaling activation, by chemical inhibition or expression of a kinase dead form of the protein attenuated BMP-7-induced expression of integrin αvβ3 and cell migration. These findings may provide a useful tool for diagnostic/prognostic purposes and even therapeutically in late-stage chondrosarcoma as an anti-metastatic agent.  相似文献   

9.
Oral mucosal wounds heal with reduced scar formation compared with skin. The epithelial integrin αvβ6 is induced during wound healing, and it can activate fibrogenic transforming growth factor β1 (TGF-β1) and anti-fibrogenic TGF-β3 that play key roles in scar formation. In this study, expression of β6 integrin and members of the TGF-β pathway were studied in experimental wounds of human gingiva and both gingiva and skin of red Duroc pigs using real-time PCR, gene microarrays, and immunostaining. Similar to human wounds, the expression of β6 integrin was induced in the pig wounds 7 days after wounding and remained upregulated >49 days. The αvβ6 integrin was colocalized with both TGF-β isoforms in the wound epithelium. Significantly higher expression levels of β6 integrin and TGF-β1 were observed in the pig gingival wounds compared with skin. Early gingival wounds also expressed higher levels of TGF-β3 compared with skin. The spatio-temporal colocalization of αvβ6 integrin with TGF-β1 and TGF-β3 in the wound epithelium suggests that αvβ6 integrin may activate both isoforms during wound healing. Prolonged expression of αvβ6 integrin along with TGF-β3 in the gingival wound epithelium may be important in protection of gingiva from scar formation. (J Histochem Cytochem 57:543–557, 2009)  相似文献   

10.
11.
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   

12.
13.
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   

14.
15.
To contribute to the question of the putative role of cystatins in Alzheimer disease and in neuroprotection in general, we studied the interaction between human stefin B (cystatin B) and amyloid-β-(1–40) peptide (Aβ). Using surface plasmon resonance and electrospray mass spectrometry we were able to show a direct interaction between the two proteins. As an interesting new fact, we show that stefin B binding to Aβ is oligomer specific. The dimers and tetramers of stefin B, which bind Aβ, are domain-swapped as judged from structural studies. Consistent with the binding results, the same oligomers of stefin B inhibit Aβ fibril formation. When expressed in cultured cells, stefin B co-localizes with Aβ intracellular inclusions. It also co-immunoprecipitates with the APP fragment containing the Aβ epitope. Thus, stefin B is another APP/Aβ-binding protein in vitro and likely in cells.  相似文献   

16.
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   

17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号