10 research outputs found

    JunB silencing attenuates TGFβ1-induced changes in cell contractility and cytoskeletal tension, but not induction of markers of smooth muscle differentiation.

    No full text
    <p>(<b>A</b>) BSMC were nucleofected with non-targeting control siRNA or with siRNA against JunB (0.1 µM and 1 µM) and assessed for JunB protein by immunoblotting (left panel, top). Effective knockdown of JunB was observed, with no change in c-Jun levels, demonstrating specificity of the siRNA used. Proliferating cell nuclear antigen (PCNA) expression was used as a loading control. 1 µM JunB siRNA reduced the levels of JunB mRNA by >80%, relative to non-targeting control siRNA, as assessed by semi-quantitative real-time PCR (right panel) (<b>B</b>) Reduction in JunB protein levels by siRNA in BSMC under basal and TGFβ1-stimulated conditions, demonstrated by immunoblotting. JunB levels were normalized to their respective GAPDH levels and expressed as percentage change relative to cells transfected with control siRNA and not subjected to TGFβ1 treatment. A representative immunoblot and its corresponding quantitation are shown. (<b>C</b>) TGFβ1-mediated induction of α-smooth muscle actin (α-SMA) calponin and SM22α, markers of smooth muscle differentiation, was unaffected by silencing of JunB, as shown by immunoblotting (left). Quantification of immunoblots is shown in the graph (right). Gel contraction assays (<b>D</b>) revealed that JunB knockdown significantly reduced both basal and TGFβ1-induced changes in cellular contractility. *p<0.05, t-test (<b>E</b>) Inhibition of JunB inhibits basal and TGFβ1-induced contraction. This inhibition of contraction, measured quantitatively as a reduction of traction (see <a href="http://www.plosone.org/article/info:doi/10.1371/journal.pone.0053430#s2" target="_blank">Methods</a>) was statistically significant (*p<0.05, comparing siCtrl+ TGFβ1 or siJunB-TGFβ1 with siCtrl-TGFβ1; ∧p<0.05 comparing siCtrl+ TGFβ1 with siJunB+ TGFβ1 Kruskal-Wallis test). The median value of traction and the interquartile range across all tested groups is shown.</p

    JunB regulates actin polymerization.

    No full text
    <p>(<b>A</b>) JunB silencing in BSMC reduces phospho-cofilin levels under basal and TGFβ1-stimulated conditions, without affecting total cofilin levels. *p<0.05; **p<0.005. Representative immunoblots are indicated in (<b>B</b>). (<b>C</b>) Filamentous (F) and globular (G) actin fractions were purified as indicated in <a href="http://www.plosone.org/article/info:doi/10.1371/journal.pone.0053430#s2" target="_blank">Methods</a>, from pBSMC under vehicle or TGFβ1-treated conditions, following treatment with non-targeting or JunB siRNA. The relative levels of F- and G-actin were subsequently assessed by immunoblotting. Quantification of immunoblot signals from three independent experiments is shown. *p<0.05. Representative immunoblots are indicated in (<b>D</b>).</p

    A model depicting the role of JunB in regulation of smooth muscle contractility in response to TGFβ1 signaling.

    No full text
    <p>TGFβ1 induces the expression of JunB as well as other markers of smooth muscle differentiation e.g. α-SMA, calponin and SM22α Additionally, TGFβ1 also promotes smooth muscle contraction via ROCK1-mediated regulation of actin polymerization and acto-myosin crossbridge cycling. JunB mediates this process by promoting the phosphorylation of cofilin, leading to stabilization of filamentous actin and also by regulating the phosphorylation and absolute levels of MLC20, the regulatory light chain of myosin, and its inhibitory phosphatase, MYPT1. Thus, activation of JunB is critical for the changes in contractility and generation of cytoskeletal tension observed upon the TGFβ1-stimulation of smooth muscle cells.</p

    JunB levels are increased in BSMC in response to TGFβ1, and in an ex vivo model of rodent bladder distension.

    No full text
    <p>(<b>A</b>) BSMC were treated with TGFβ1 for the indicated times and assessed for JunB levels by immunoblotting. GAPDH is included as a loading control. (<b>B</b>) Immunofluorescence analysis of BSMC showing increased JunB nuclear localization upon TGFβ1 treatment for 24 h. (<b>C</b>) Sections from rat bladders distended ex vivo for 8 h (injured) were stained sequentially with anti-JunB and Cy3-conjugated species-specific secondary antibody. Increased nuclear fluorescent signal for both proteins was evident in the detrusor smooth muscle of stretch-injured specimens, but not of non-distended (control) bladders.</p

    TGFβ1induces contractility in bladder smooth muscle cells (BSMC).

    No full text
    <p>(<b>A</b>) Human bladder smooth muscle cells were seeded in collagen gels and treated for 24 h with vehicle (Veh) or 2.5 ng/ml TGFβ1, after which the gels were released from the sides of the well and the resulting decrease in surface area monitored microscopically (top) and quantified (bottom). *p<0.05, t-test. The area of the gel under control conditions is set to 100%. (<b>B</b>) Whisker plot of results from traction force microscopy of BSMC showing an increase in cell traction forces exerted with TGFβ1 treatment. The contractile response, measured quantitatively as enhanced traction (see <a href="http://www.plosone.org/article/info:doi/10.1371/journal.pone.0053430#s2" target="_blank">Methods</a>) was statistically significant (*p<0.05, Kruskal-Wallis test). The median value of traction and the interquartile range for both groups is shown. (<b>C</b>) BSMC were treated for 30 min with inhibitors targeting the PI3-kinase/Akt (PI3K-i, Akt-i) mitogen-activated protein kinases (MEK-i, p38-i, JNK-i) or Rho-kinase (ROCK-i), followed by treatment with vehicle (Control, upper panel of wells) or 2.5 ng/ml TGFβ1 (lower panel) for 24 h and were monitored for changes in gel contractility. Inhibition of signaling via the JNK and ROCK axes abrogated TGFβ1-induced gel contraction. Quantification of changes in gel surface area for the various inhibitors under conditions of TGFβ1 treatment is indicated. (<b>D</b>) A transcription factor ELISA was carried out to assess differences in DNA-binding activities of members of the AP-1 family of transcription factors, using nuclear extracts prepared from BSMC treated with 2.5 ng/ml TGFβ1 for 24 h, or control cells. Fold changes are expressed relative to control which is set to 100%.</p

    An Akt- and Fra-1-Dependent Pathway Mediates Platelet-Derived Growth Factor-Induced Expression of Thrombomodulin, a Novel Regulator of Smooth Muscle Cell Migration

    No full text
    Overdistension of hollow organs evokes pathological changes characterized by smooth muscle remodeling. Mechanical stimuli induce smooth muscle cell (SMC) growth through acute activation of signaling cascades and by increased expression of soluble mitogens. Physical forces have also been implicated in ligand-independent activation of receptor tyrosine kinases, including the platelet-derived growth factor (PDGF) receptor, although the extent to which this occurs in intact tissue is unknown. Previously, we implicated Akt and activator protein-1 (AP-1) as mediators of growth and gene expression in SMC exposed to cyclic stretch or PDGF. Here we show that bladder wall distension leads to PDGFR activation and identify thrombomodulin (TM) as an Akt and AP-1 target in SMC. We demonstrate that TM, also induced by bladder stretch injury, is regulated at the transcriptional level by the AP-1 components c-jun and Fra1. Mutation of an AP-1 motif at −2010/−2004 abolished both AP-1 binding and PDGF responsiveness of the TM promoter. Fra1 silencing diminished PDGF-induced TM expression and SMC cell cycle transit. In contrast, TM knockdown did not affect cell growth but attenuated PDGF-stimulated SMC migration. Taken together, these results reveal new facets of TM regulation in SMC and provide the first demonstration of a role for endogenous TM in PDGF-induced cell migration. Moreover, TM induction on bladder injury suggests that it may be a biomarker for pathological smooth muscle remodeling
    corecore