Cellular Physiology and Biochemistry

Published by Cell Physiol Biochem Press GmbH and Co KG

Online ISSN: 1421-9778

·

Print ISSN: 1015-8987

Articles


Figure 1: The serum 1,25(OH)2D levels stratified by TaqI gene polymorphism. vs. CC P=0.011 ‡ vs. CC P=0.007.
Figure 2: The ROC analyses of serum 1,25(OH)2D level in discrimination of KS from controls. AUC=0.911,P=0.018
The TaqI Gene Polymorphisms of VDR and the Circulating 1,25-Dihydroxyvitamin D Levels Confer the Risk for the Keloid Scarring in Chinese Cohorts
  • Article
  • Full-text available

July 2013

·

24 Reads

Dongmei Yu

·

·

·

Aim: to investigate the association between vitamin D receptor polymorphisms and circulating 1,25-Dihydroxyvitamin D (1,25(OH)2D) levels with keloid scar (KS) risk. Methods: A total of 261 patients with KS and 261 normal healthy individuals were enrolled. VDR gene polymorphisms were determined. Circulating 1,25(OH)2D levels were detected. Results: In this study, we investigated the role of four loci of VDR gene polymorphisms in determining the risk of KS in a Chinese cohort. We found that the TaqI C>T polymorphism was closely associated with the KS incidence. Carriers with CC genotype of TaqI had a higher chance of developing KS. Stratification analyses by sex showed that this trend exists only in female subjects but not in male subjects. Furthermore, the TaqI C>T polymorphism affects the circulating 25-Hydroxyvitamin D levels. The CC carriers had a significantly lower mean circulating 1,25(OH)2D level than TT carriers and CT carriers. KS subjects had significantly lower mean serum circulating 1,25(OH)2D level than controls. The receiver operating characteristic curve analysis revealed that the serum circulating 1,25(OH)2D level at the cut-off point of 16.1 ng/ml could discriminate KS subjects from controls. Conclusion: Collectively, these data provide evidence that the vitamin D/VDR pathway plays an important role in the development of KS. The TaqI gene polymorphisms of VDR and circulating 1,25(OH)2D levels may thus be used as potential markers for prediction of KS development.
Download
Share

Table 1 . Membrane voltages (V m ) obtained in patch-clamp recordings from CHO cells expressing ¢F508- CFTR or wtCFTR or non-transfected (control) CHO cells 
Fig. 1. a Original recordings of membrane voltages (V m ) obtained from freshly isolated non-CF and CF respiratory cells. The superimposed pulses in these original recordings reflect the input resistance of the microelectrode. After stabilizing of V m , CPX (10 Ìmol/l) was applied to respiratory cells which reversibly depolarized in both CF and non-CF cells. b Summary of the effects of CPX indicates similar depolarizations in both non-CF and CF respiratory cells. * p ! 0.05. Number of experiments in parentheses. 
Fig. 2. a Time course of wholecell conductance (G m ) and membrane voltage (V m ) measured in a patch-clamp recording of a CF (CFDE) respiratory epithelial cell. CPX only slightly depolarized and reduced G m. b Summary of experiments with CF (CFDE) and nonCF (16HBE14o-) respiratory epithelial cells. The results indicate significant effects of CPX on V m and G m only in CF respiratory cells. * p ! 0.05. Number of experiments in parentheses. 
Fig. 3. Double electrode voltage clamp experiment obtained in a Xenopus oocyte injected with wtCFTR. a Current recordings indicate activation of whole-cell currents only by IBMX (1 mmol/l) but not by CPX (100 Ìmol/l). b Summary of the voltage clamp experiments as described in a. No evidence for CPX-induced activation of a whole-cell conductance in either wtCFTR-, ¢F508-CFTR-, or water-injected oocytes. 
Fig. 4. Measurement of intracellular pH in non-CF (16HBE14o-) and CF (CFDE) respiratory epithelial cells. a, c Original pH i recordings of 4 individual non-CF and 3 individual CF cells indicate only very little effects on pH i. Na acetate and NH 4 Cl (each 20 mmol/l) were applied at the end of each experiment in order to acidify and alkalinize the cells. b, d Concentration-response curves for the effects of CPX on pH i of non-CF and CF respiratory epithelial cells indicate significant effects of CPX only at 10 Ìmol/l. * p ! 0.05. Number of experiments in parentheses. 
No Evidence for Direct Activation of the Cystic Fibrosis Transmembrane Conductance Regulator by 8-Cyclopentyl-1,3-Dipropylxanthine

February 1998

·

57 Reads

·

·

·

[...]

·

8-Cyclopentyl-1,3-dipropylxanthine (CPX) is a selective A1-adenosine receptor antagonist which has been reported to activate Cl- efflux at very low concentrations in cells carrying the cystic fibrosis (CF) defect, but not in cells expressing the wild-type form of the CF transmembrane conductance regulator (CFTR). CPX was suggested as a new therapeutic drug for the treatment of CF. In the present study, we examined the effects of CPX on various types of recombinant cells (Xenopus oocytes, Chinese hamster ovary cells, CF tracheal cells) and native non-CF and CF respiratory epithelial cells. CPX did not activate a whole-cell conductance when applied at concentrations ranging from 1 nmol/l to 100 micromol/l in oocytes injected with water or expressing either wild-type CFTR or mutant deltaF508-CFTR. Correspondingly, CPX (10 micromol/l) did not activate whole-cell conductance in non-CF or CF respiratory epithelial cells and Chinese hamster ovary cells expressing either wild-type CFTR or deltaF508-CFTR. Instead, CPX depolarized Vm by inhibition of a K+ conductance in CF respiratory epithelial cells. At 10 micromol/l CPX marginally decreased intracellular pH in respiratory epithelial cells, independent of expression of wild-type CFTR or mutant CFTR. According to these data, CPX-induced 36Cl efflux reported in previous studies cannot be attributed to direct activation of deltaF508-CFTR Cl- conductance and is probably related to the CPX-induced changes in intracellular pH.

MiR-106b and MiR-15b Modulate Apoptosis and Angiogenesis in Myocardial Infarction

June 2012

·

89 Reads

MicroRNAs (miRNAs) are identified as crucial gene regulators in response to myocardial infarction (MI). However, the overall relationships between miRNAs and the gene targets which contribute to the cellular phenotypes in MI are not fully elucidated. To make a better understanding towards functional roles of miRNAs in MI, useful information was mined through bioinformatic techniques. MI-related miRNAs were retrieved from publications, and PicTar, TargetScanS, and miRanda programs were used to predict their gene targets. Gene ontology (GO) and pathway analyses of gene targets were applied to uncover functional roles of miRNAs. The miRNA-gene networks were illustrated by Pajek tool. Finally, validation experiments were performed towards two important miRNAs in the networks. Up to 119 MI-related miRNAs were retrieved from publications. GO and pathway analyses for their predicted gene targets demonstrated that these dysregulated miRNAs were enriched in cardiovascular-related phenotypes. Through illustrating miRNA-gene networks, overall relationships between miRNAs and gene targets were detected especially in processes of apoptosis and angiogenesis. Moreover, experimental data supported bioinformatic predictions that miR-106b served as an anti-apoptotic modulator through inhibition of p21 expression and miR-15b displayed anti-angiogenesis activity. The miRNAs played essential roles in pathological processes of MI. Further, miR-106b and miR-15b maybe mediated as robust regulators in apoptosis or angiogenesis following MI, respectively.

Fig. 3. Detection of TRPV2 and TRPV1 proteins in F-11 and cultured DRG cells. TRPV2 immunostaining is located in the cytoplasma and at the cell membrane of F-11 cells (A). TRPV2 protein is also present in a subpopulation of cultured rat DRG neurons mainly of medium size (C). Specific staining for TRPV1 is absent in F-11 cells (B), but present in cultured DRG cells (D). The bar indicates 10 µm. 
Fig. 4. TRPV2 and TRPV1 expression in separate subpopulations of primary afferent neurons. Double immunofluorescence for TRPV2 (A-C) or TRPV1 (D-F) with IB4 reveals that TRPV2 and TRPV1 are expressed in two distinct 
Fig. 5. Electrophysiological characteristics of TRPV2 in F-11 and HEK293 cells. Current-voltage relations for heat-evoked currents in F-11 cells (A) and mouse TRPV2 transiently transfected in HEK293 cells are depicted (B). Current density profiles of 4 individual F-11 cells were determined at ±60 mV 
Fig. 6. Pharmacology of TRPV2 in F-11 cells. (A) Comparison of inward and outward current densities in the absence and presence of 100 nM and 10 µM ruthenium red at 37°C and 53°C. Filled bars show mean inward and outward current densities at a temperature of 37°C. Open bars represent mean current densities at 53°C. Asterisks denote significant suppression with ruthenium red at 53°C. The number of cells measured is indicated above the bars. Data are shown as means ± SEM. (B) Comparison of current density temperature profiles measured at a holding potential of-60 mV under control conditions (grey line) and in the presence of 100 µM 2-APB (black line). 
The Temperature-Sensitive Ion Channel TRPV2 is Endogenously Expressed and Functional in the Primary Sensory Cell Line F-11

February 2005

·

530 Reads

In sensory neurons heat is transduced by a subfamily of TRP channels sharing sequence homology with the capsaicin-sensitive vanilloid receptor subtype 1 (TRPV1), but differing in their thermal response thresholds. To identify a neuronal cell line endogenously expressing noxious heat-transducing ion channels, we examined F-11 cells, a hybridoma derived from rat dorsal root ganglia and mouse neuroblastoma. Using RT-PCR, transcripts homologous to TRPV2 and TRPV4, but not to TRPV1 or TRPV3, were found. We isolated a full-length cDNA of 2.4 kb coding for a 757-amino acid protein corresponding to mouse TRPV2, which was further characterized by immunocytochemistry and electrophysiology. Using the whole-cell patch-clamp technique, we observed a heat-evoked increase in outward and inward currents with a threshold of 51.6 +/- 0.2 degrees C. The current-voltage relationship stimulated by a temperature of 52 degrees C was characterized by an outward rectification with a reversal potential close to -10 mV. Heat-evoked currents could be inhibited by ruthenium red. There was no activation by stimulation with capsaicin or 2-aminoethoxydiphenyl borate. Our results indicate that F-11 cells express functional noxious heat-sensitive TRPV2 channels. Thus, we propose that F-11 cells represent a valuable in vitro model to characterize the properties of TRPV2 in a native neuronal environment.

Figure 1: Effects of IL-1β, TNF-α, or IL-4 stimulation on CCL20 production in HGFs. (A) HGFs were activated for 24 hours with IL-1β, IL-4 or mixtures of both cytokines at designated concentrations. The expression levels of CCL11 in the supernatants were measured using ELISA. Data are representative of three different HGFs samples from three different donors. The results are shown as the mean and SD of one representative experiment performed in triplicate. The error bars show the SD of the values. * = P<0.05; significantly different from the non-stimulated HGFs. # = P<0.05; significantly different from the IL-1β-stimulated HGFs. (B) HGFs were activated for 24 hours with TNF-α, IL-4 or mixtures of both cytokines at designated concentrations. The expression levels of CCL11 in the supernatants were measured using ELISA. Data are representative of three different HGFs samples from three different donors. The results are shown as the mean and SD of one representative experiment performed in triplicate. The error bars show the SD of the values. * = P<0.05; significantly different from the non-stimulated HGFs. ** = P<0.05; significantly different from the TNF-α-stimulated HGFs.
Figure 2: Effects of EGCG on CCL11 production in IL-lβ/IL-4 or TNF-α/IL-4-stimulated HGFs. (A) HGFs were pretreated with EGCG (3.125, 6.25, 12.5, 25, or 50 µg/ml) for 1 hour, and then the HGFs were stimulated with IL-1β (10 ng/ml) and IL-4 (10 ng/ml), and the supernatants were collected after 24 hours. The expression levels of CCL11 in the supernatants were measured using ELISA. Data are representative of three different HGFs samples from three different donors. Data indicate the mean ± SD of three cultures. The error bars show the SD of the values. * = P<0.05, ** = P<0.01; significantly different from the IL-1β/IL-4-stimulated HGFs without EGCG. (B) HGFs were pretreated with EGCG (3.125, 6.25, 12.5, 25, or 50 µg/ml) for 1 hour, and then the HGFs were stimulated with TNF-α (10 ng/ml) and IL-4 (10 ng/ml), and the supernatants were collected after 24 hours. The expression levels of CCL11 in the supernatants were measured using ELISA. Data are representative of three different HGFs samples from three different donors. Data indicate the mean ± SD of three cultures. The error bars show the SD of the values. * = P<0.05, ** = P<0.01; significantly different from the TNF-α/IL-4-stimulated HGFs without EGCG.
Figure 3: Effects of signal transduction inhibitors on IL-lβ/IL-4 or TNF-α/IL-4-stimulated CCL11 release by HGFs. (A) HGFs were pre-incubated with SB203580 (20 µM), PD98059 (20 µM), or SP600125 (20 µM) for 1 hour and then incubated with IL-1β (10 ng/ml) with IL-4 (10 ng/ml). After 24 hours incubation, the supernatants were collected, and CCL11 production was measured by ELISA. Data are representative of HGFs from three different donors. Data indicate the mean ± SD of three cultures. The error bars show the SD of the values. * = P<0.05, significantly different from the IL-1β/IL-4-stimulated HGFs without inhibitors. (B) HGFs were pre-incubated with SB203580 (20 µM), PD98059 (20 µM), or SP600125 (20 µM) for 1 hour and then incubated with TNF-α (10 ng/ ml) with IL-4 (10 ng/ml). After 24 hours incubation, the supernatants were collected, and CCL11 production was measured by ELISA. Data are representative of HGFs from three different donors. Data indicate the mean ± SD of three cultures. The error bars show the SD of the values. * = P<0.05, significantly different from the TNF-α/IL-4-stimulated HGFs without inhibitors.
Figure 5: Effects of EGCG on the TNF-α/IL-4-induced phosphorylations of MAPKs. The cultured cells were pretreated with EGCG (25 µg/ml) for 60 min and then stimulated with TNF-α (10 ng/ml) and IL-4 (10 ng/ml) for 15, 30, or 60 min. The cell extracts were subjected to SDS-PAGE followed by Western blotting analysis with antibodies against phospho-specific p38 MAPK, p38 MAPK, phospho-specific ERK, ERK, phosphor-specific JNK, JNK, and actin levels in HGFs from three different experiments from three different donors.
(-)-Epigallocatechin-3-Gallate Inhibits CC Chemokine Ligand 11 Production in Human Gingival Fibroblasts

June 2013

·

86 Reads

Background: CC chemokine ligand 11 (CCL11) is related to Th2 cells migration via CC chemokine receptor 3 (CCR3). Th2 cells are involved in the etiology of periodontal disease. However, how the infiltration of Th2 cells is controlled in periodontally diseased tissues is unknown. (-)-Epigallocatechin gallate (EGCG), the major catechin in green tea, has multiple beneficial effects, but the effects of EGCG on CCL11 production are uncertain. In this study, we investigated whether cytokines could induce CCL11 production in human gingival fibroblasts (HGFs). Moreover, we examined the effects of EGCG on CCL11 production in HGFs. Methods and results: ELISA analysis disclosed that interleukin (IL)-4 synergistically enhanced CCL11 production in IL-1β or tumor necrosis factor (TNF)-α-stimulated HGFs. EGCG prevented IL-1β/ IL-4 or TNF-α/IL-4-mediated CCL11 production in a concentration dependent manner. CCL11 production in HGFs was positively regulated by p38 mitogen-activated protein kinase (MAPK), extracellular signal-regulated kinase (ERK), and c-Jun N terminal kinase (JNK). Western blot analysis revealed that EGCG treatment prevented IL-1β/IL-4 or TNF-α/IL-4-induced ERK and JNK activation in HGFs. Conclusions: These data provide that CCL11 production in HGFs could be associated with Th2 cells infiltration in periodontal lesions. Moreover, EGCG is useful for periodontitis treatment to inhibit CCL11 production.

Fig. 1. Expression of NF?B and I?B-? in erythrocytes. A. Original Western blot demonstrating the expression of NF?B (upper panels) and I?B-? protein (lower panels) in erythrocyte concentrates from 4 different individuals (p1-p4). White blood cell (WBC), red blood cell (RBC) and platelet (PLT) numbers of the different samples are given in the table on top. As indicated in the figure, 5, 10, 20 and 30 ?l of erythrocyte lysates were loaded per lanes of the gels. HeLa cells treated with (+) or without (-) TNF were used as controls. Note that the percentage of contaminating cells in the erythrocyte concentrates is below 1 ?. B. Original Western blot of protein extracts (30 ?l per lane) demonstrating the expression of NF?B (upper panel) and I?B-? protein (middle panel) from whole blood (lane 1), diluted whole blood (1:15; lane 2), or purified erythrocytes (lane 3). Western blot of ?-actin was used as a loading control. White blood cell (WBC), red blood cell (RBC) and platelet (PLT) numbers of whole blood (WB) and its corresponding purified erythrocytes (RBC) are given in the table on top. Note that the purified erythrocyte preparation contained 3.7 % platelets and 6.8 % white blood cells of the original whole blood sample. p65: 65 kDa subunit of NF?B; p39: 39 kDa I?B-? single chain; p40: 40 kDa ?-actin single chain.  
Fig. 2. The NF?B pathway inhibitor Bay 11-7082 induces phosphatidylserine exposure and cell shrinkage. A. Original histograms of phosphatidylserine exposure (annexin V binding ) in untreated erythrocytes (left panel) and in erythrocytes treated for 48 hours with 10 ?M Bay 11-7082 (right panel). B. Arithmetic means ? SEM (n=12) of the percentage of annexin V binding erythrocytes after 24 hours (closed squares) or 48 hours (open squares) incubation as a function of Bay 11-7082 concentration . # indicates significant difference from the respective values in the absence of the drug. C. Arithmetic means ? SEM (n=12) of the erythrocyte forward scatter in % of control after a 24 hours (closed squares) or 48 hours (open squares) incubation as a function of Bay 11-7082 concentration. # indicates significant difference from the respective values in the absence of the drug. D. Transmission light microscopy (upper panels) and annexin V fluorescence (lower panels) of erythrocytes incubated for 48 hours without (left panels) or with (right panels) 10 ?M Bay 11-7082. The insert depicts an erythrocyte at higher magnification. Bars: 50 ?m. Insert bars: 10 ?m. E. Arithmetic means ? SEM (n=12) of hemolysis (in % of maximal hemolysis after exposure to distilled water) from erythrocytes following a 24 hours (closed squares) or a 48 hours (open squares) incubation as a function of Bay 11-7082 concentration . # indicates significant difference from the respective values in the absence of the drug.  
Fig. 3. The NF?B inhibitor parthenolide induces phosphatidylserine exposure and cell shrinkage. A. Original histograms of phosphatidylserine exposure (annexin V binding ) in untreated erythrocytes (left panel) and in erythrocytes treated for 48 hours with 30 ?M parthenolide (right panel). B. Arithmetic means ? SEM (n=12) of the percentage of annexin V binding erythrocytes after 24 hours (closed squares) or 48 hours (open squares) incubation as a function of parthenolide concentration . # indicates significant difference from the respective values in the absence of the drug. C. Arithmetic means ? SEM (n=12) of the erythrocyte forward scatter in % of control after a 24 hours (closed squares) or 48 hours (open squares) incubation as a function of parthenolide concentration. # indicates significant difference from the respective values in the absence of the drug. D. Arithmetic means ? SEM (n=12) of hemolysis (in % of maximal hemolysis after exposure to distilled water) from erythrocytes following a 24 hours (closed squares) or a 48 hours (open squares) incubation as a function of parthenolide concentration. # indicates significant difference from the respective values in the absence of the drug.  
Fig. 7. Antagonism of Bay 11-7082-and parthenolide-induced annexin V binding and cell shrinkage by N-acetylcysteine (NAC). A. Histogram of annexin V binding in a representative experiment of erythrocytes exposed for 24 hours to 10 ?M Bay 11-7082 in the absence (left) or presence (right) of 1 mM N-acetyl-cysteine (NAC). B. Arithmetic means ? SEM (n = 12) of the percentage of annexin V-binding erythrocytes following exposure for 24 hours to Ringer supplemented with different concentrations of Bay 11-7082 (0 ? 20 ?M) in the presence (closed squares) and absence (open squares) of 1 mM NAC. Annexin V-binding in control cultures was 1.56 ? 0.06 % in the absence of NAC, and 1.47 ? 0.07 % in the presence of 1 mM NAC. * indicates significant difference from the respective values in the absence of the NAC. C. Arithmetic means ? SEM (n = 12) of the forward scatter (FSC in % of control) of erythrocytes treated for 24 hours with different concentrations of Bay 11-7082 (0 ? 20 ?M) in the presence (closed squares) and absence (open squares) of 1 mM NAC. The forward scatter (arbitrary units) in control cultures was 420 ? 8 in the absence of NAC, and 409 ? 7 in the presence of 1 mM NAC. * indicates significant difference from the respective values in the absence of the NAC. D. Arithmetic means ? SEM (n = 9) of the percentage of annexin V-binding erythrocytes following exposure for 48 hours to Ringer supplemented with different concentrations of parthenolide (0 ? 100 ?M) in the presence (closed squares) and absence (open squares) of 1 mM NAC. Annexin V-binding in control cultures was 3.12 ? 0.71 % in the absence of NAC, and 4.45 ? 0.93 % in the presence of 1 mM NAC. * indicates significant difference from the respective values in the absence of NAC.  
Fig. 6. Bay 11-7082 depletes reduced glutathione (GSH). Doseresponse curve of the effect of Bay 11-7082 on reduced glutathione (GSH) level. Arithmetic means ? SEM (n = 3-6) of the effect of Bay 11-7082 on GSH level of erythrocytes exposed for 24 hours to Ringer with Bay 11-7082 as a function of drug concentration. # indicates significant difference from the respective values in the absence of the drug.  
The NF??B Pathway Inhibitors Bay 11-7082 and Parthenolide Induce Programmed Cell Death in Anucleated Erythrocytes

February 2011

·

155 Reads

The preclinical compounds Bay 11-7082 and parthenolide trigger apoptosis, an effect contributing to their antiinflammatory action. The substances interfere with the activation and nuclear translocation of nuclear factor NFκB, by inhibiting NFκB directly (parthenolide) or by interfering with the inactivation of the NFκB inhibitory protein IκB-α (Bay 11-7082). Beyond that, the substances may be effective in part by nongenomic effects. Similar to apoptosis of nucleated cells, erythrocytes may undergo apoptosis-like cell death (eryptosis) characterized by cell membrane scrambling with phosphatidylserine exposure, and cell shrinkage. Thus, erythrocytes allow the study of nongenomic mechanisms contributing to suicidal cell death, e.g. Ca(2+) leakage or glutathione depletion. The present study utilized Western blotting to search for NFκB and IκB-α expression in erythrocytes, FACS analysis to determine cytosolic Ca(2+) (Fluo3 fluorescence), phosphatidylserine exposure (annexin V binding), and cell volume (forward scatter), as well as an enzymatic method to determine glutathione levels. As a result, both NFκB and IκB-α are expressed in erythrocytes. Targeting the NFκB pathway by Bay 11-7082 (IC(50) ≈ 10 μM) and parthenolide (IC(50) ≈ 30 μM) triggered suicidal erythrocyte death as shown by annexin V binding and decrease of forward scatter. Bay 11-7082 treatment further increased intracellular Ca(2+) and led to depletion of reduced glutathione. The effects of Bay 11-7082 and parthenolide on annexin V binding could be fully reversed by the antioxidant N-acetylcysteine. In conclusion, the pharmacological inhibitors of NFκB, Bay 11-7082 and parthenolide, interfere with the survival of erythrocytes involving mechanisms other than disruption of NFκB-dependent gene expression.

Cellular localization of THIK-1 (K-2P 13.1) and THIK-2 (K-2P 12.1) K+ channels in the mammalian kidney

February 2008

·

39 Reads

K(+)-channels fulfill several important functions in the mammalian kidney such as volume regulation, recirculation and secretion of K(+) ions, and maintaining the resting potential. In this study we used immunocytochemical methods, in situ hybridization, and nephron segment-specific RT-PCR to obtain a detailed picture of the cellular localization of two tandem pore domain potassium (K(2P)) channels, THIK-1 (K(2P)13.1, KCNK13) and THIK-2 (K(2P)12.1, KCNK12). Monospecific antibodies against C-terminal domains of rat THIK-1 and THIK-2 proteins (GST-fusion proteins) were raised in rabbits, freed from cross-reactivity, and affinity purified. All antibodies were validated by Western blot analysis, competitive ELISA, and preabsorption experiments. The expression of THIK channels in specific nephron segments was confirmed by double staining with marker proteins. Results indicate that in rat and mouse THIK-1 and THIK-2 were expressed in the proximal tubule (PT), thick ascending limb (TAL), connecting tubule (CNT), and cortical collecting duct (CCD). In human kidney THIK-1 and THIK-2 were localized in PT, TAL and CCD. Immunostaining of rat tissue revealed an intracellular expression of THIK-1 and THIK-2 throughout the identified nephron segments. However in mouse kidney THIK-2 was identified in basolateral membranes. Overall, the glomerulus, thin limbs and medullary collecting ducts were devoid of THIK-1 and THIK-2 signal. In summary, THIK-1 and THIK-2 are abundantly expressed in the proximal and distal nephron of the mammalian kidney.

Figure 1: The expression of miR-126 is significantly decreased during HSCs activation. (A) Cell morphology of freshly isolated primary HSCs(a), after cultured in vitro for 2 days (quiescent HSCs) (b), for 7 days (partially activated HSCs) (c) and for 14 days (fully activated HSCs). (B) The expression of α-SMA and collagen type 1 was slightly increased at day 7 in partially activated HSCs and greatly enhanced at day 14 in fully activated HSCs. *p<0.05 compared with day 2 in quiescent HSCs. The expression of α-SMA and collagen type I were upregulated after TGF-β stimulation. *p<0.05 compared with control. The expression of α-SMA and collagen type I were determined by qRT-PCR, normalized to β-actin. (C) miR-126 expression is gradually and significantly decreased during HSCs activation. *p<0.05 compared with day 2 in quiescent HSCs. And miR-126 expression was also obviously decreased during HSCs activation stimulated by TGF-β. *p<0.05 compared with control.
Figure 2: Overexpression of miR-126 inhibits HSCs migration but not affects its proliferation. (A) Infection efficiency of HSC-T6 after 3 days, (a) Fluorescence microscope of HSC-T6 infected with fluorescently labeled miRNA after 3 days. The results showed that infection efficiency of HSC-T6 was about 80% with MS1022X-miR-126 (MOI: 30). (b) The expression of miR-126 was increased by about 13-fold infected with MS1022X-miR-126 for 3 days in HSC-T6 assessed by qRT-PCR (*p<0.05). (B) Upregulation of miR-126 inhibits HSC-T6 migration by about 60% assessed by transwell assay (*p<0.05). (C) miR-126 did not affect HSC-T6 proliferation (p>0.05).
Figure 3: Upregulation of miR-126 inhibits α-SMA and collagen type I expression. (A) ELISA was used for quantitative determination of collagen types I content in HSC-T6 culture supernatant 3 days after lenti-virus infection. Collagen type I content in the media was decreased to 56% in miR-126 overexpression group compared with the negative control group. ** p<0.05 compared with the negative control. (B) Overexpression of miR-126 decreased the mRNA levels of collagen type I and α-SMA by about 70% and 50% respectively determined by qRT-PCR in HSCs (**p<0.05). (C) The protein level of collagen type I and α-SMA was decreased by overexpression of miR-126 assessed by Western blot. (D) Immunofluorescence staining of α-SMA and collagen type I in HSCs after lenti-viral infection. Overexpression of miR-126 in HSC-T6 attenuated immunofluorescence staining of α-SMA and collagen type I.
Figure 4: CRK is validated to be the target of miR-126 in HSCs and overexpression of miR-126 inhibits the expression of CRK. (A) The putative binding sequence of miR-126 in the CRK 3'UTR Mutation was generated in the CRK3'UTR sequence in the complementary site for the seed region of miR-126. (B) Luciferase reporter assay was used to confirm the direct interaction between miR-126 and target mRNA. The results showed that the 3'UTR of CRK could be directly targeted by miR-126 while the mutant 3'UTR of CRK was completely refractory to miR-126. (C) Overexpression of miR-126 in HSC-T6 decreased the protein level of CRK determined by Western blot. (D) CRK was gradually and significantly increased during HSC activation. (E) Over-expression of miR-126 in HSC-T6 reduced immunofluorescence staining of CRK. (F) Immunofluorescence staining of F-actin and CRK after lenti-virus infection. Overexpression of miR-126 directly attenuated CRK expression and indirectly attenuated F-actin expression.
Figure 5: Overexpression of CRK promotes HSCs activation and decreases miR-126 expression. (A) The expression of CRK was increased by about 3.5-fold days after infection of lenti-pSB795-crk into HSCs assessed by qRT-PCR (*p<0.05). (B) The protein level of exogenous CRK was upregulated by after infection of lenti-pSB795-crk into HSCs assessed by Western blot. (C) Overexpression of CRK in HSC-T6 increased the expression of collagen type I and α-SMA. (D) Overexpression of CRK evidently decreased the miR-126 expression by about 70% assessed by qRT-PCR.
Overexpression of miR-126 Inhibits the Activation and Migration of HSCs through Targeting CRK

January 2014

·

67 Reads

Background & Aims: MicroRNAs (miRNAs) have been shown to play essential roles in HSCs activation which contributes to hepatic fibrosis. Our previous miRNA microarray results suggested that miR-126 might be decreased during HSCs activation as other studies. The aim of this study is to investigate the role of miR-126 during HSCs activation. Methods: In this study, the expression of miR-126 during HSCs activation was measured and confirmed by qRT-PCR. Then, miR-126 expression was restored by transfection of lentivirus vector encoding miR-126. Futhermore, cell proliferation was assayed by the cell counting kit-8 (CCK-8), cell migration was assayed by transwell assay, and the markers of activation of HSCs, α-SMA and collagen type I, were assayed by qRT-PCR, Western Blotting, Immunostaining and ELISA. Luciferase reporter assay was used to find the target of miR-126, and Western Blotting and Immunostaining was used to validate the target of miR-126. Then, the expression and the role of the target of miR-126 during HSCs activation was further assessed. Results: The expression of miR-126 was confirmed to be significantly decreased during HSCs activation. Overexpression of miR-126 significantly inhibited HSCs migration but did not affect HSCs proliferation. The expression of α-SMA and collagen type I were both obviously decreased by miR-126 restoration. CRK was found to be the target of miR-126 and overexpression of miR-126 significantly inhibited CRK expression. And it was found that overexpression of CRK also significantly decreased miR-126 expression and promoted HSCs activation. Conclusions: Our study showed that overexpression of miR-126 significantly inhibited the activation and migration of HSCs through targeting CRK which can also decrease miR-126 expression and promote HSCs activation. © 2014 S. Karger AG, Basel.

Homologous Desensitisation of the Mouse Leukotriene B4 Receptor Involves Protein Kinase C-Mediated Phosphorylation of Serine 127

February 2007

·

7 Reads

Murine leukotriene B(4) (LTB(4)) receptor (mBLT1) cDNA was identified by searching the EST database using human LTB(4) receptor as the query sequence. Expression of functional mBLT1 after injection of in vitro transcribed cRNA into Xenopus laevis oocytes was demonstrated as LTB(4)-evoked, Ca(2+)-activated Cl(-) currents recorded by two-electrode voltage clamp. From mBLT1-expressing oocytes, a dose-dependent relationship between the Ca(2+)-activated Cl(-) current and LTB(4) concentration was demonstrated with an apparent EC(50) of 6.7 nM. Following LTB(4) stimulation of mBLT1, we observed two transient, spatially distinct Ca(2+)-activated, inwardly directed Cl(-) currents in the oocytes: a fast peak current requiring relatively high LTB(4) concentrations, and a slowly progressing Cl(-) current. Nucleotides, PGE(2), 12R-hydroxy-5, 8, 14-cis-10-trans-eicosatetraenoic acid, and LTD(4) did not activate mBLT1. U75302, specifically targeting BLT1, significantly reduced LTB(4)-evoked Cl(-) currents. Repetitive LTB(4) administration desensitized the LTB(4)-evoked currents. Activation of protein kinase C (PKC) by PMA addition completely eliminated the LTB(4)-evoked currents, whereas down-regulation of PKC by prolonged PMA exposure (20 h) impaired mBLT1 desensitisation. In addition, Ser-127-Ala substitution of the PKC consensus phosphorylation site on the second intracellular loop prevented the mBLT1 desensitisation. These data indicate that PKC-mediated phosphorylation at Ser-127 leads to mBLT1 desensitisation.

Fig. 3. ISO effects on Rem mRNA and miR-132. A: Relationship between ISO administration time and the relative expression of Rem mRNA. (n = 4 per group). B: The effect of chronic ISO on the relative expression of miR-132 (n = 3 per group). Level of expression at t = 0 is shown with a dotted line in both graphs. All values are group means ± SEs (*p < .05). 
Fig. 4. Effects of chronic ISO on Rem protein levels. A: Immunoblots from control (left) and 12-h or 24-h ISO-treated ventricle cytosolic samples (right). Blots represent Rem and GAPDH protein levels. B: Quantitative analysis of Rem protein band density as a function of treatment duration. Values were normalized to GAPDH band densities (n = 3, *p < .05). Level of expression at t = 0 is shown with a dotted line. All values are group means ± SEs. 
Fig. 5. Effects of chronic ISO administration on L-type Ca 2+ currents and Ca 2+ signals. A: The inset shows representative recordings of L-type Ca 2+ currents at +5 mV steps under control conditions and after 48-h ISO administration. The graph shows mean (± SE) I-V relationship of peak L-type currents. To allow comparisons among experiments, membrane current values were normalized to unit capacitance (C m ). C m = 232.9 ± 7.5 (n = 44) pF in controls and C m = 307.7 ± 8.0 pF (n = 19) in ISO pretreated rats. Peak currents are shown for myocytes isolated from controls (•; n = 44) and from ISO-pretreated rats (▲; n = 19) at the indicated potentials (*p < .05 between 
Fig. 6. Changes in α 1c subunit mRNA and protein levels following ISO administration. A: Relationship between ISO administration time and α 1c subunit mRNA relative expression. Dotted line: level of expression at t = 0. (n = 3). B: Independent immunoblots from control (left) and 48-h ISO-treated (right) ventricle membrane samples representing the α 1c channel subunit (upper panel) and pan-cadherin (lower panel). C: Relationship between ISO administration time and relative protein expression of the α 1c subunit. Values were normalized to pan-cadherin band densities. (n = 3-9). Dotted line: level of expression at t = 0. All values are group means ± SEs. *p < .05). 
MiR-132 Regulates Rem Expression in Cardiomyocytes During Long-Term β-Adrenoceptor Agonism

April 2015

·

98 Reads

To characterize the effects of long-term β-adrenergic receptor stimulation on Rem protein and mRNA expression in rat heart and possible involvement of miR-132. Adult rats were treated with isoproterenol (ISO, 150 µg.kg.h(-1)) for 2 d and Rem, miR-132, and α1c (the principal subunit of Cav1.2 channels) were measured at protein and mRNA levels with western blot and quantitative reverse transcriptase polymerase chain reaction (qRT-PCR) experiments, respectively. Ca(2+) currents and intracellular Ca(2+) signals were evaluated in isolated cardiomyocytes. Systemic administration of ISO led to decreases in Rem protein and mRNA levels (down to 49%). Furthermore, levels of the microRNAs (miRs) miR-132 and miR-214 were upregulated 5- and 9-fold, respectively. Transfection of miR-132, but not miR-214, into HEK293 cells reduced the expression of a luciferase reporter gene controlled by a conserved 3´-untranslated region (UTR) of Rem by half. Chronic ISO administration also led to a 25% decrease in the amplitude of peak L-type Ca(2+) currents, a 40% decrease in α1c subunit protein abundance at the membrane level, and a 60% decrease in expression of α1c channel subunit mRNA. These results suggest that Rem expression is down-regulated posttranscriptionally by miR-132 in response to long-term activation of β-adrenergic signaling, but this down-regulation does not produce a larger Ca(2+) influx through Cav1.2 channels. © 2015 S. Karger AG, Basel.

Tanshinone IIA improves miR-133 expression through MAPK ERK1/2 pathway in hypoxic cardiac myocytes

August 2012

·

42 Reads

Tanshinone IIA is a lipid-soluble pharmacologically active compound extracted from the rhizome of Chinese herb Salvia miltiorrhiza, a well-known traditional Chinese medicine used for the treatment of cardiovascular disorders. Previous studies have identified that tanshinone IIA inhibited overexpression of miR-1 in hypoxic neonatal cardiomyocytes. This study was designed to examine the effects of tanshinone IIA on miR-133 expression under hypoxic condition. Neonatal rat cardiomyocytes were cultured in a hypoxic environment (2% O(2)+93% N(2)+5% CO(2)) at 37°C for 24 hours. MTT, TUNEL assays, and Flow Cytometry (FCM) were performed to identify cell apoptosis. Western blot was used to examine the expression of ERK1/2 and miR-133 level was quantified by Real-time PCR. Our results showed that apoptosis was induced by hypoxia. Typical apoptotic cells were seen by TUNEL assay, and FCM showed an apoptosis rate of 13.32% in hypoxic group. Apoptosis rate in hypoxic cells was reduced significantly by tanshinone IIA. In addition, the expression level of miR-133 was increased in hypoxic cells and further upregulated by tanshinone IIA. The stress-activated protein kinase MAPK ERK1/2 was activated by hypoxia and further increased with tanshinone IIA treatment. The present study demonstrated that tanshinone IIA enhanced cell resistance to hypoxic insult by upregulating miR-133 expression through activating MAPK ERK1/2 in neonatal cardiomyocytes.

MiR-133b Promotes Neurite Outgrowth by Targeting RhoA Expression

January 2015

·

82 Reads

MicroRNA-133b (miR-133b) has been shown to play a critical role in spinal cord regeneration. The aim of this study was to investigate the cellular role of miR-133b in neural cells. PC12 cells and primary cortical neurons (PCNs) were transfected with lenti-miR-133b, lenti-miR-133b inhibitor, plasmid-shRNA-RhoA, plasmid-RhoA and their negative controls. After 48 hours of transfection, the levels of proteins and mRNA or miRNA were evaluated by Western blotting and qRT-PCR, respectively. Moreover, the neurite outgrowth was analyzed by Image J. For pharmacological experiments, inhibitors of MEK1/2 kinase (PD98059), phosphoinositide-3 kinase (PI3K) (LY294002) and ROCK (Y27632) were added into the culture medium. Overexpression of miR-133b in PC12 cells enhanced neurite outgrowth. Conversely, inhibition of miR-133b reduced neurite length. We further identified RhoA as a target and mediator of mir-133b for neurite extension by Western blot and knockdown experiment. Moreover, overexpression of RhoA could attenuate the neurite growth effects of miR-133b. Also, we observed that miR-133b activated MEK/ERK and PI3K/Akt signaling pathway by targeting RhoA. Finally, in PCNs, miR-133b also increased axon growth and attenuated axon growth restrictions from chondroitin sulfate proteoglycans (CSPG). In summary, our study suggested that miR-133b regulated neurite outgrowth via ERK1/2 and PI3K/Akt signaling pathway by RhoA suppression. © 2015 S. Karger AG, Basel.

Figure 2: Inhibition of miR-137 increases adipogenic differentiation and proliferation of hADSCs. (A) miR-137 levels were determined in control- (anti-miR-con) or anti-miR-137-transfected hADSC using real-time PCR. The data were presented as the relative ratio of miR-137 to 5S RNA level of each sample. (B) hADSC proliferation was determined by direct cell counting after oligonucleotide transfection. (C) Oligonucleotide-transfected hADSC were grown to 80-90% confluence. Adipogenic differentiation was induced for 10 days and determined by Oil Red O staining to visualize intracellular lipid accumulation, which was quantified by absorbance at 562nm. (D and E) Real-time PCR analysis of PPARγ and PPARγ2 in anti-miR-137-transfeted undifferentiated cells (D). Changes in expression of aP2 and C/EBP-α in differentiated cell to adipogenic lineage (E). Total RNA was isolated at days after induction of differentiation. Data represent mean ± SEM of the ratio of a target gene level to GUSB level (n = 4). *p < 0.05 compared with anti-miR-con transfected hADSCs
Figure 3: R-137 targets the 3'UTR of CDC42 mRNA. (A, B) CDC42 expression in hADSCs transfected with oli-gonucleotide was analyzed by western blot (A) and real-time PCR (B). (C) pMIR-CDC42 or pMIR-CDC42-mut luciferase constructs were transfected into miR-con or miR-137-mimic-transfected hADSCs and anti-miR-con or anti-miR-137-transfected hADSCs. Data represent mean ± SEM of the ratio of a target gene level to GUSB level (n = 4), *p < 0.05, compared with miR-con-transfected hADSCs, #p < 0.05 compared with anti-miR-Con-transfected hADSCs.
Figure 4: CDC42 RNAi inhibits adipogenic differentiation and proliferation of hADSC. CDC42 mRNA levels were determined in control- (si-con) or CDC42 oligonucleotide- (si-CDC42) transfected hADSC using real-time PCR. hADSC proliferation was determined by direct cell counting after siRNA oligo transfection. (C) siRNA oligo-transfected hADSC were grown to 80-90% confluence. Adipogenic differentiation was induced for 10 days and determined by Oil Red O staining to visualize intra-cellular lipid accumulation, which was quantified by absorbance at 562nm. Silencing of the CDC42 gene inhibited the adipogenic differentiation of hADSCs. (D) Real-time PCR analysis of PPARγ2 and C/EBP-α in si-CDC42-transfeted undifferentiated cells. Data represent mean ± SEM of the ratio of a target gene level to GUSB level (n = 4). *p < 0.05 compared with si-con-transfected hADSCs.
miR-137 Controls Proliferation and Differentiation of Human Adipose Tissue Stromal Cells

March 2014

·

132 Reads

Background/Aims: Demonstrating the molecular mechanisms of human adipose tissue-derived mesenchymal stem cells (hADSCs) differentiation and proliferation could develop hADSCs-based cell therapy. Methods: The microRNA-137 (miR-137) and cell division control protein 42 homolog (CDC42) levels were regulated by oligonucleotides transfection. The adipogenic differentiation was induced for 10 days in an adipogenic medium and assessed by using an Oil Red O stain. The regulation of miR-137 on CDC42 expression was determined by western blot, real-time PCR and luciferase reporter assay. Results: We confirmed the roles of miR-137 on hADSCs proliferation and adipogenic differentiation. We showed that overexpression of miR-137 inhibited both hADSCs proliferation and adipogenic differentiation. Overexpression of miR-137 also downregulated protein and mRNA levels of CDC42, a predicted target of miR-137. In contrast, inhibition of miR-137 with 2'-O-methyl antisense RNA increased proliferation and adipogenic differentiation in hADSCs. Luciferase reporter activity in the miR-137 target site within the CDC42 3'UTR was lower in miR-137-transfected hADSCs than in control miRNA-transfected hADSCs. RNA interference-mediated downregulation of CDC42 in hADSCs inhibited their proliferation and adipogenic differentiation. Conclusion: Our results indicate that miR-137 regulates hADSCs adipogenic differentiation and proliferation by directly targeting CDC42. These findings improve our knowledge of the molecular mechanisms governing hADSCs differentiation and proliferation. © 2014 S. Karger AG, Basel.

Partial Inactivation of Cardiac 14-3-3 Protein in vivo Elicits Endoplasmic Reticulum Stress (ERS) and Activates ERS-initiated Apoptosis in ERS-induced Mice

August 2010

·

25 Reads

Excessive endoplasmic reticulum stress (ERS) triggers apoptosis in various conditions including diabetic cardiomyopathy and pressure overload-induced cardiac hypertrophy and heart failure. The primary function of 14-3-3 protein is to inhibit apoptosis, but the roles of this protein in protecting against cardiac ERS and apoptosis are largely unknown. We investigated the roles of 14-3-3 protein in vivo during cardiac ERS and apoptosis induced by pressure overload or thapsigargin injection using transgenic (TG) mice that showed cardiac-specific expression of dominant negative (DN) 14-3-3eta. Cardiac positive apoptotic cells and the expression of glucose-regulated protein (GRP)78, inositol-requiring enzyme (Ire)1alpha, tumor necrosis factor receptor (TNFR)-associated factor (TRAF)2, CCAAT/enhancer binding protein homology protein (CHOP), caspase-12, and cleaved caspase-12 protein were significantly increased in the pressure-overload induced DN 14-3-3eta mice compared with that in the WT mice. Furthermore, thapsigargin injection significantly increased the expression of GRP78 and TRAF2 expression in DN 14-3-3eta mice compared with that in the WT mice. The enhancement of 14-3-3 protein may provide a novel protective therapy against cardiac ERS and ERS-initiated apoptosis, at least in part, through the regulation of CHOP and caspase-12 via the Ire1alpha/TRAF2 pathway.

Depletion of 14-3-3 Protein Exacerbates Cardiac Oxidative Stress, Inflammation and Remodeling Process via Modulation of MAPK/NF-ĸB Signaling Pathways after Streptozotocin-induced Diabetes Mellitus

December 2011

·

179 Reads

Diabetic cardiomyopathy is associated with increased oxidative stress and inflammation. Mammalian 14-3-3 proteins are dimeric phosphoserine-binding proteins that participate in signal transduction and regulate several aspects of cellular biochemistry. The aim of the study presented here was to clarify the role of 14-3-3 protein in the mitogen activated protein kinase (MAPK) and nuclear factor-kB (NF-κB) signaling pathway after experimental diabetes by using transgenic mice with cardiac-specific expression of a dominant-negative 14-3-3 protein mutant (DN 14-3-3). Significant p-p38 MAPK activation in DN 14-3-3 mice compared to wild type mice (WT) after diabetes induction and with a corresponding up regulation of its downstream effectors, p-MAPK activated protein kinase 2 (MAPKAPK-2). Marked increases in cardiac hypertrophy, fibrosis and inflammation were observed with a corresponding up-regulation of atrial natriuretic peptide, osteopontin, connective tissue growth factor, tumor necrosis factor α, interleukin (IL)-1β, IL-6 and cellular adhesion molecules. Moreover, reactive oxygen species, left ventricular expression of NADPH oxidase subunits, p22 phox, p67 phox, and Nox4, and lipid peroxidation levels were significantly increased in diabetic DN 14-3-3mice compared to diabetic WT mice. Furthermore, myocardial NF-κB activation, inhibitor of kappa B-α degradation and mRNA expression of proinflammatory cytokines were significantly increased in DN 14-3-3 mice compared to WT mice after diabetes induction. In conclusion, our data suggests that depletion of 14-3-3 protein induces cardiac oxidative stress, inflammation and remodeling after experimental diabetes induction mediated through p38 MAPK, MAPKAPK-2 and NF-κB signaling.

SRIF Receptor Subtype Expression and Involvement in Positive and Negative Contractile Effects of Somatostatin-14 (SRIF-14) in Ventricular Cardiomyocytes

February 2008

·

68 Reads

Somatostatin-14 (SRIF-14), a neuropeptide co-stored with acetylcholine in the cardiac parasympathetic innervation, exerts both positive and negative influences directly on contraction of ventricular cardiomyocytes, indicative of involvement of more than one of five known SRIF (SSTR) receptor subtypes. The aim was to characterize receptor subtype expression in adult rat ventricular cardiomyocytes and to investigate the influence of a series of SRIF (SSTR) subtype-selective agonists on contractile parameters. mRNA and protein expression of each receptor subtype were quantified by RT-PCR and immunoblotting respectively; for contraction studies, cells were stimulated at 0.5 Hz under basal conditions and in the presence of isoprenaline (ISO, 10(-8)M). all five SRIF (SSTR) receptor subtypes were expressed in cardiomyocytes although SRIF1A (SSTR2) and SRIF2A (SSTR1) were less abundant than the other subtypes. L803087 (10(-8)M), a SRIF2B (SSTR4) agonist, attenuated ISO-stimulated peak contractile amplitude and prolonged relaxation time (T(50)). L796778 (10(-7)M), a SRIF1C (SSTR3) agonist, augmented basal and ISO-stimulated peak contractile amplitude; L779976 (10(-8)M) and L817818 (10(-9)M), agonists at SRIF1A (SSTR2) and SRIF1B (SSTR5) receptors, respectively, also augmented ISO-stimulated peak amplitude. These data support involvement of SRIF2B (SSTR4) receptors in the negative contractile effects of SRIF-14, while one or more of the three SRIF1 receptor subtypes (SSTR2, 3 or 5) may contribute to the positive contractile effects of SRIF-14.

Turetta L, Donella-Deana A, Folda A, Bulato C, Deana R. Characterisation of the serotonin efflux induced by cytosolic Ca2+ and Na+ concentration increase in human platelets. Cell Physiol Biochem 14: 377-386

February 2004

·

24 Reads

The present study aimed at elucidating the mechanism(s) of serotonin (5-HT) efflux induced by thapsigargin from human platelets in the absence of extra-cellular Ca2+. Efflux of pre-loaded radiolabeled serotonin was generally determined by filtration techniques. Cytosolic concentrations of Ca2+, Na+ and H+ were measured with appropriate fluorescent probes. 5-HT efflux from control or reserpine-treated platelets--where reserpine prevents 5-HT transport into the dense granules--was proportional to thapsigargin evoked cytosolic [Ca2+]c increase. Accordingly factors as prostacyclin, aspirin and calyculin which reduced [Ca2+]c-increase also inhibited the 5-HT efflux. Thapsigargin, which also caused a remarkable increase in cytosolic [Na+]c, promoted less 5-HT release, in parallel to lower [Na+]c and [Ca2+]c increase, when added to platelet suspensions containing low [Na+]. The Na+/H+ exchanger monensin increased the [Na+]c and induced 5-HT efflux without affecting the Ca2+ level. The 5-HT efflux induced by both [Ca2+] or [Na+]c increase did not depend on pH or membrane potential changes, whereas it decreased in the absence of extra-cellular K+, and increased in the absence of Cl- or Na+. Increases in [Ca2+]c and [Na+]c independently induce serotonin efflux through the outward directed plasma membrane serotonin transporter SERT. This event might be physiologically important at the level of capillaries or narrowed arteries where platelets are subjected to high shear stress which causes [Ca2+]c increase followed by 5-HT release which might exert vasodilatation.

Figure 1: R-140 is down-regulated in EC patients. A: The expression levels of miR-140 in human EC tissues and corresponding adjacent tissues relative to U6 were determined by qRT-PCR. (n=89, p<0.0001). B, C and D: The mRNA level of E-cadherin, N-cadherin and Vimentin were determined by qRT-PCR. E: miR-140 expression level in cell lines transfected with miR-140 mimics, miR-140 inhibitor, control for miR-mimics (NC) and control for miR-140 inhibitor (inhibitor NC). The result was validated by real-time PCR. Data are represented as mean±SEM. * indicates P<0.05.
Figure 2: R-140 regulates EMT and EC cells invasion but has no effect on cell proliferation. A: Transwell assay was performed as described in Materials and Methods. Cells were treated with miR-140 mimics, miR-140 inhibitor, NC and inhibitor NC for 24h. The representative images of invasive cells at the bottom of the membrane stained with crystal violet were visualized as shown. The quantifications of cell invasion were presented as percentage of cell numbers. B: The cell-proliferation assay test showed distinct differences on proliferation after manipulation of miR-140 in both TE-1 and Eca-109 cells at 12-hour, 24-hour, and 48-hour time points. C: The protein expression levels of E-cadherin, N-cadherin and Vimentin in Eca-109 cells transfected with miR-140 mimics, miR-140 inhibitor, NC and inhibitor NC were analyzed by Western-blotting. GAPDH was used as a loading control. Average values of integrated optical density (IOD) were assessed by analyzing five fields per slide and recorded in the histogram. All experiments were performed in triplicate and presented as mean ± SEM. * indicates significant difference compared with control group (P<0.05). Every independent experiment was performed 3 times.
Figure 3: R-140 regulates Slug. A: The potential miR-140 seed region at the 3ʹ-UTR of Slug mRNA was computationally predicted by microRNA.org. Eca-109 cells were co-transfected with miR-140 mimics (or NC) with pGL3-Slug (or pGL3-Slug-mut) vector. Luciferase activity was normalized by the ratio of firefly and Renilla luciferase signals. B: Slug protein expression level in TE-1 and Eca-109 cells transfected with NC, miR-140 mimics, inhibitor NC and miR-140 inhibitor were analyzed by using western-blotting assay. GAPDH was used as a control. All experiments were performed in triplicate and the band intensity values were analyzed by using Image J. Then, the significance in change in protein expression was analyzed by using statistical t-tests. C: The mRNA levels of Slug relative to GAPDH in human EC tissues and corresponding adjacent tissues were evaluated by qRT-PCR. D: A negative correlation was found between RNA expression of miR-140 and Slug in tumor samples (R = -0.862; p < 0.0001). Data are represented as mean±SEM. * indicates P<0.05.
Figure 4: Silencing of Slug reverses the anti-invasion effect of miR-140 downregulated. A: The transfection efficiency was validated by using western-blotting after both TE-1 and Eca-109 cells transfected with miR-140 inhibitor and inhibitor+siRNA/Slug (the co-transfection cells of miR-140 inhibitor and siRNA/Slug) and inhibitor+siRNA/control (co-transfected with miR-140 inhibitor and siRNA/control). GAPDH was used as a control. Average values of integrated optical density (IOD) were assessed by analyzing five fields per slide and recorded in the histogram. B: Transwell assay was performed as described in Materials and Methods. Cells were treated with miR-140 inhibitor and inhibitor+siRNA/Slug (the co-transfection cells of miR-140 inhibitor and siRNA/Slug) and inhibitor+siRNA/control (co-transfected with miR-140 inhibitor and siRNA/control) for 24h. The representative images of invasive cells at the bottom of the membrane stained with crystal violet were visualized as shown. The quantifications of cell invasion were presented as percentage of cell numbers. * indicates P<0.05. .
Down-Regulation of miR-140 Induces EMT and Promotes Invasion by Targeting Slug in Esophageal Cancer

October 2014

·

44 Reads

Background/aims: MicroRNAs (miRNAs) are reported to regulate cell invasion and functions by interfering with the translation of target mRNAs, but the role of miRNAs in esophageal cancer (EC) remains unclear. Methods: RT-PCR and Western blot were used to detect the expression of miRNAs and candidate genes in EC samples (n=89). miR-140 mimics and inhibitor were tansfected in human TE-1 and Eca-109 cells. The transwell assay was used to examine the cell invasive ability. The regulation mechanism was confirmed by luciferase reporter assay. The markers of EMT were detected by using Western blot. Results: miR-140 expression was decreased in the EC tissues compared with the corresponding adjacent tumor tissues. Low expression of miR-140 promotes cell invasion by using transwell assay, while the effect of miR-140 high expression is reverse. Slug, a target gene of miR-140, was examined by luciferase assay and Western blot. Conclusions: miR-140 may regulate the cell invasion of EC via controlling Slug expression.

MiR-142-3p Functions as a Potential Tumor Suppressor in Human Osteosarcoma by Targeting HMGA1

April 2014

·

81 Reads

Background/Aims: Mounting evidence has shown that aberrant expression of miRNAs correlates with human cancers, and that miRNAs can function as tumor suppressors or oncogenes. Here, we investigated the role and mechanism of miR-142-3p in human osteosarcoma. Methods: We used quantitative real-time RT-PCR to measure the expression of miR-142-3p in human osteosarcoma cell lines and tissues. The roles of miR-142-3p in osteosarcoma development were studied using cultured HOS, MG63 and Saos-2 cells and tumor xenograft analyses in nude mice; their target genes were also investigated. Results: We found that miR-142-3p was significantly downregulated in osteosarcoma cell lines and clinical specimens. Overexpression of miR-142-3p suppressed osteosarcoma cell proliferation, migration and invasion, whereas miR-142-3p knockdown increased these parameters. The xenograft mouse model also revealed the suppressive effect of miR-142-3p on tumor growth. High mobility group AT-hook 1 (HMGA1) was identified as a target of miR-142-3p. Downregulation of HMGA1 induced effects on osteosarcoma cell lines similar to those induced by miR-142-3p. In contrast, restoration of HMGA1 abrogated the effects induced by miR-142-3p up-regulation. Conclusion: These results indicated that miR-142-3p may function as a tumor suppressor by targeting HMGA1 in osteosarcoma. © 2014 S. Karger AG, Basel.

Figure 1: R-144 inhibits the proliferation level of A549 and H460 cells. (A, B) Compared to the scramble group and blank group in both A549 and H460 cells, the expression level of miR-144 was significantly higher and TIGAR was significantly lower in miR-144 group (*P < 0.05). (C, D) The cell growth rate (OD490) of miR-144 group goes slower than scramble group and blank group both A549 and H460 cells. (E, F) After approximately 2 weeks of incubation, the colony formation number of miR-144 group cells came to lower colony than the scramble group and blank group in both A549 and H460 cells (*P < 0.05). (G, H) G: The tumor tissues were collected from athymic nude mouse after four weeks. H: The tumor volumes were measured every week. The tumor volumes of miR group cell were relatively lower than scramble group and blank group (*P < 0.05).
Figure 2: R-144 enhances the apoptosis rate of A549 cells and H460 cells. (A, B) Apoptosis rate in miR-144 group was higher than scramble group and blank group in A549 and H460 cells (*P < 0.05). (C, D) Caspase3/7 activity of miR-144 group went higher than that of scramble and blank group in A549 and H460 cells (*P < 0.05). (E, F) The hoechst 33342 staining assay come out that the apoptosis rate was higher in miR-144 group than the scramble group and blank group(*P < 0.05).
Figure 3: R-144 induces autophagy in both A549 and H460 cells. (A) Autophagic vacuoles of miR-144 group are much more than the scramble and blank group (*P < 0.05). (B, C) The expressions of autophagy-related proteins were detected by Western blotting. GAPDH was used as control. In both A549 and H460 cells, the expression levels of LC3-I was lower, and Beclin1 and LC3-II came out higher in miR-144 groups (*P < 0.05).
Figure 4: R-144 directly targets TIGAR. (A, B) Wild and mutant types of TIGAR 3'UTRs segments are showed. (C) Co-transfection with miR-144 significantly suppressed the luciferase activity of the reporter containing the wild-type 3' UTR (*P<0.05).
Figure 5: Function of miR-144 partially attributes to targeting at TIGAR. (A) The expression of TIGAR protein of si-TIGAR group was lower than miR-144 and blank group cells, and Beclin1 and LC3 II proteins came out higher in A549 cell (*P < 0.05). (B) The OD490 values of si-TIGAR group on 48 h, 72 h and 96 h were significantly decreased than blank group cell, but increased than miR-144 group (*P < 0.05). (C) The colony numbers of si-TIGAR group was significantly decreased than blank group cell, but increased than miR-144 group (*P < 0.05). (D) Apoptotic cells of si-TIGAR group was significantly increased than blank group cell, but decreased than miR-144 group (*P < 0.05). (E) MDC fluorescence intensity of si-TIGAR group was higher than both blank and miR-144 group cells (*P < 0.05).
MiR-144 Inhibits Proliferation and Induces Apoptosis and Autophagy in Lung Cancer Cells by Targeting TIGAR

February 2015

·

202 Reads

MiRNAs are noncoding RNAs of 20-24 nucleotides that function as post-transcriptional negative regulators of gene expression. MiRNA genes are usually transcribed by RNA polymerase II in the nucleus. Their initial products are pre-miRNAs which have cap sequences and polyA tails. The p53-induced glycolysis and apoptosis regulator (TIGAR) was discovered through microarray analysis of gene expression following activation of p53. However, little is known about the effect of miR-144 on cell proliferation and apoptosis and how it interacts with TIGAR. We performed real-time PCR, western blotting, CCK8, colony formation, tumor growth, flow cytometry, Caspase3/7 activity, Hoechst 33342 staining, MDC staining of autophagic cells and luciferase reporter assays to detect the influence of miR-144 to lung cancer cells. miR-144 targeted TIGAR, inhibited proliferation, enhanced apoptosis, and increased autophagy in A549 and H460 cells. Our study improves our understanding of the mechanisms underlying lung cancer pathogenesis and may promote the development of novel targeted therapies. © 2015 S. Karger AG, Basel.

Fig. 1. (A) Mean relative miR-146a values ± SD in PBMCs from RRMS patients (RRMS, n=50) and healthy control subjects (controls, n=50). The blank box and the black box represent the relative expression of miR146a in the controls and the RRMS patients, respectively.  P=0.027 when comparing relative miR-146a levels between RRMS patients and controls. (B) Mean relative miR-146a values ± SD in PBMCs from RRMS patients (RRMS, n=50) and healthy control subjects (controls, n=50) stratified by gender. The blank box and the black box represent the relative expression of miR-146a in male and female individuals, respectively. 
Genetic Association of MiR-146a with Multiple Sclerosis Susceptibility in the Chinese Population

January 2015

·

167 Reads

miR-146a polymorphisms have been involved in susceptibility to multiple diseases. The aim of the present study was to analyze the potential association between two functional miR-146a polymorphisms (rs2910164 and rs57095329) and multiple sclerosis (MS) in the Han Chinese population. A cohort of 525 patients and 568 healthy controls were genotyped to detect the two polymorphisms by SNaPshot. No significant differences were detected in the distribution of the two miR-146a polymorphisms between the patients and controls (P > 0.05). However, stratification by gender showed a statistically significant difference in the frequency of the genotype rs2910164 between MS patients and control females (P=0.009). Further stratification analysis by subgroup revealed that the miR-146a rs2910164 C allele conferred a higher risk of developing relapsing-remitting MS (RRMS) (P=0.018). In addition, the rs2910164 C allele was significantly associated with increased expression of miR-146a in patients with RRMS (P=0.025). Moreover, patients with the rs2910164 C allele released more TNF-α and IFN-γ, but not IL-1β, compared with individuals carrying the homozygous GG genotype (P < 0.05). Our results provide evidence that rs2910164 may play a role in MS susceptibility in females. The rs2910164 G>C variation may affect the expression of miR-146a and the release of proinflammatory cytokines. © 2015 S. Karger AG, Basel.

Fig. 1. Effect of H 2 O 2 on tyrosine phosphorylation in RTH cells. Proteins from total cell lysates were resolved by 12% SDS PAGE followed by Western blotting using phosphotyrosine antibodies. Upper panels: cell exposure to H 2 O 2 produces dose-dependent increase of tyrosine phosphorylation in different peptides; the effect is maximum after 10 min of exposure. Lower panel: Plots of the peak densities of singular bands obtained from digital imaging measurements. EC 50 estimates from hyperbolic regression curves yield values of 1.1 ± 1.5, 3.2 ± 4.0, and 3.9 ± 2.5 mM, respectively for bands of about 100, 50 and 40 kDa (Prism 3.02, GraphPad Software Inc., San Diego, CA). In this and the following figures results are representative of at least three independent experiments, while numerical values are expressed as percent of controls. 
Fig. 4. Effect of genistein on the tyrosine phosphorylation induced by H 2 O 2. Upper panel: preincubation with 100 µM genistein for 30 min causes a reduction of more than 50% of the phosphotyrosine increase induced by 5 mM H 2 O 2. Lower panel: bar chart of the average optical densities of the lanes shown in the upper panel; in this and the following graphs error bars indicate standard deviation (n = 3). 
Fig.5. Effect of genistein on the tyrosine phosphorylation induced by Hg 2+ and Cu 2+. Upper panel: preincubation with 100 µM genistein for 30 min causes a strong reduction of the phosphotyrosine increase induced by 100 µM Cu 2+ , whereas the effect of 50 µM Hg 2+ is only partially reduced. Lower panel: bar chart of lane average optical densities. 
Fig. 6. (A) Effects of heavy metals and pro-oxidants on ERK phosphorylation. Upper panel: Cells were exposed to 50 µM Hg 2+ or 100 µM Cu 2+ for 5 min, or to 2.5 mM H 2 O 2 or 100 µM Cu 2+ / 0.5 mM H 2 O 2 for 10 min and then probed with phosphospecific ERK antibody. A 38 kDa putative ERK homolog is partially activated in the control. Peptide phosphorylation is increased by Hg 2+ and to a lesser extent by Cu 2+ , but not by H 2 O 2. Lower panel: bar chart of the 38 kDa band peak densities. (B) Modulatory effects of p38 inhibitors on the activation of ERK induced by H 2 O 2. Pre-incubations for 30 
Ligand-Independent Tyrosine Kinase Signalling in RTH 149 Trout Hepatoma Cells: Comparison Among Heavy Metals and Pro-Oxidants

February 2003

·

77 Reads

Tyrosine phosphorylation depends on the activity of receptor and non-receptor tyrosine kinases and promote cell growth, differentiation and apoptosis. Different stressors are known to stimulate tyrosine kinase activities and this could explain a wide spectrum of effects that these agents produce on different organisms. We studied the effects of heavy metals and pro-oxidants on tyrosine kinase signalling in trout hepatoma cells (RTH 149) by Western immunoblotting. Use of antiphosphotyrosine showed that Hg(2+) and Cu(2+)in the microM range, and H(2)O(2) in the mM range, induced tyrosine phosphorylation. The effect of Cu(2+)was prevented by pre-incubation with genistein, while those of Hg(2+)and H(2)O(2) were only decreased, probably due to tyrosine kinase stimulation coupled to phosphatase inhibition. Phosphospecific antibodies against the three types of MAPKs showed that ERK is activated by heavy metals only, while p38 and SAPK/JNK are activated by H(2)O(2), Hg(2+), and Cu(2+) plus low H(2)O(2). Cell pre-incubation with p38 inhibitors indicated that ERK activation by H(2)O(2) is prevented by concomitant activation of p38. Phosphospecific STAT antibodies revealed activation by H(2)O(2) only. In conclusion, fish cell exposure to heavy metals and pro-oxidants produce specific tyrosine kinase responses, involving cross talk and redox modulatory effects.

LLC-PK(1) cells maintained in a new perfusion cell culture system exhibit an improved oxidative metabolism. Cell Physiol Biochem 12(2-3):153-162

February 2002

·

14 Reads

Cultured renal proximal tubule cells dedifferentiate from an oxidative metabolism to high rates of glycolysis over time. There are many reasons why cells in culture dedifferentiate, not least being a lack of homogenous nutrient supply and poor oxygenation. To this end we have developed a new cell culture device (EpiFlow), which combines continuous perfusion of medium with continuous oxygenation of cells grown on microporous supports. LLC-PK(1) cells cultured under EpiFlow conditions were compared with the same cells grown under conventional static conditions. EpiFlow maintained cells exhibited an improved oxidative metabolism as evidenced by 1) a decreased activity of glycolytic enzymes, 2) an increase in the activity of mitochondrial phosphate-dependent-glutaminase, 3) an increase in cellular ATP content, and 4) an improved morphology (increased cell height, mitochondrial density and an increased number and height of microvilli). In addition, LLC-PK(1) cells maintained under perfusion conditions exhibited an increased sensitivity to the respiratory chain blocker antimycin A as assayed by mitochondrial membrane potential (JC-1). We conclude that LLC-PK(1) cells maintained under EpiFlow conditions develop an improved oxidative metabolism that is more comparable to the in vivo situation.

Fig. 1. A. Detection of IL-16 encoding mRNA in OA-DF. DF (n=4) were incubated with the different compounds (lanes 1-5, compare table 1) and the ratio of steady state mRNA amounts encoding IL-16 vs. GAPDH were enumerated by quantitative RT/ PCR. Forskolin induced a sizable IL-16 signal (mean 1.96 fold; lane 3), staurosporine induced IL-16 to a lesser extent (mean 1.48 fold, lane 5) and PMA (mean 0.6 fold, lane 1) reduced IL-16 signals in comparison to mock controls (lane 6). B. Detection of IL-16 encoding mRNA in RA-DF. DF (n=6) were incubated with forskolin (lane 1) or staurosporine (lane 2) and the ratio of steady state mRNA amounts encoding IL-16 vs. GAPDH were enumerated by quantitative RT/PCR in comparison to controls (lane 3). Forskolin (mean 177%, ± 37; p<0.001) and staurosporine (mean 864%, ± 692; p<0.038) induced IL- 16 transcripts significantly. The mean IL-16 induction by staurosporine was almost fivefold higher when compared to forskolin (p<0.036).  
Fig. 2. A. Detection of IL-16 encoding mRNA in synovial fibroblasts of osteoarthrthitis patients. In OA-SF (n=11) forskolin (lane 3) and staurosporine (lane 5) enhanced the IL-IL 16 RT/PCR signals about 1.75 fold in comparison to mock controls (lane 6). PMA (lane 1) ionomycin (lane 2) or MAS-7 (lane 4) did not influence IL-16 gene activity . However, the inductions were statistically not significant. B. Detection of IL-16 encoding mRNA in synovial fibroblasts of rheumatoid arthritis patients. In RA-SF (n=8), staurosporine enhanced the IL- 16 RT/PCR signals significantly (mean 2.53 fold, p=0.0038; lane 5). Forskolin activated IL-16 RT/ PCR signals to some extent (mean 1.37 fold, p=0.028, lane 3) in comparison to mock controls (lane 6). The mean IL-16 induction by staurosporine was significantly higher when compared to that by forskolin (p= 0.024). Interestingly, PMA activated IL-16 to some extent (lane 1) and ionomycin (lane 2) or MAS-7 (lane 4) reduced the IL-16 mRNA amounts somewhat.  
Fig. 4. Detection of IL-16 protein in immortalized synovial fibroblasts. An IL-16 precursor protein at approximately 39 kD and a processed IL-16 protein at approximately 17 kD were detected in cytoplasmic extracts of immortalized synovial fibroblasts (STF), but not in immortalized endothelial controls (EC). Recombinant IL-16 served as a positive control (+).
Fig. 5. Detection of IL-16 in fibroblast supernatants after serum starvation in presence of IL-1. RA synovial fibroblasts were serum starved (0.5%) for 24 hours, then either IL-1β (left) or IL-1β plus 10% FCS (middle) were added. Supernatants were collected after additional 24 hours of incubation. Cell culture medium containing 10 % FCS served as a negative control (right) and serial dilutions of recombinant human IL-16 as internal standards.  
Synovial Fibroblasts from Rheumatoid Arthritis Patients Differ in their Regulation of IL-16 Gene Activity in Comparison to Osteoarthritis Fibroblasts

February 2004

·

72 Reads

In rheumatoid arthritis (RA), synovial fibroblast-like cells (SF) contribute significantly to articular inflammation. They express distinct patterns of genes associated with cell proliferation and differentiation and elevated levels of cytokines and chemoattractant factors, including IL-16. Here we investigated pathways regulating IL-16 expression in fibroblasts from RA patients in comparison to fibroblasts from osteoarthritis (OA) patients. Fibroblasts were isolated from dermal and articular biopsies, expanded and pathways of IL-16 induction were investigated by real time quantitative RT/PCR, immuno blot and ELISA. Stimulation of cAMP dependent signal transduction by forskolin induced prominent IL-16 RT/PCR signals in OA-DF and OA-SF. In contrast, in RA-DF and RA-SF staurosporine significantly augmented IL-16 RT/PCR signals, but forskolin induced less IL-16 transcript amounts. Activation of protein kinase C by PMA induced a significant IL-16 response only in RA-SF. Addition of IL-1beta or TNF-alpha did not upregulate IL-16 mRNA but secretion of the mature IL-16 cytokine was activated in serum starved cells in presence of IL-1beta. Our results suggest that RA fibroblasts differ from OA fibroblasts with respect to their sensitivities to kinase/phospatase signal transduction pathways. The enhanced expression of IL-16 in the synovial membrane early in RA vs OA may be associated in part with these distinct signaling responses.

Involvement of Golgin-160 in Cell Surface Transport of Renal ROMK Channel: Co-expression of Golgin-160 Increases ROMK Currents

February 2006

·

16 Reads

The weak inward rectifier potassium channel ROMK is important for water and salt reabsorption in the kidney. Here we identified Golgin-160 as a novel interacting partner of the ROMK channel. By using yeast two-hybrid assays and co-immunoprecipitations from transfected cells, we demonstrate that Golgin-160 associates with the ROMK C-terminus. Immunofluorescence microscopy confirmed that both proteins are co-localized in the Golgi region. The interaction was further confirmed by the enhancement of ROMK currents by the co-expressed Golgin-160 in Xenopus oocytes. The increase in ROMK current amplitude was due to an increase in cell surface density of ROMK protein. Golgin-160 also stimulated current amplitudes of the related Kir2.1, and of voltage-gated Kv1.5 and Kv4.3 channels, but not the current amplitude of co-expressed HERG channel. These results demonstrate that the Golgi-associated Golgin-160 recognizes the cytoplasmic C-terminus of ROMK, thereby facilitating the transport of ROMK to the cell surface. However, the stimulatory effect on the activity of more distantly-related potassium channels suggests a more general role of Golgin-160 in the trafficking of plasma membrane proteins.

Fig. 3. A. Synoviocytes from 4 RA and 2 OA patients were activated with TNF-α (1 ng/ml) and IL-17 (10 ng/ml) in the presence of different concentrations of the active form of leflunomide, HMR1726 (5-100 µM), and MMP-1 protein release in supernatants was measured by ELISA. RA and OA control cells were stimulated with both cytokines in absence of leflunomide. The induction indices (x-fold) in comparison to controls (=1) are denoted on the Y-axis. *p < 0.05, **p < 0.01  
Fig. 4. Apoptosis induction in cell lysates and supernatants from activated synoviocytes (1 RA and 2 OA patients) in the presence of different concentrations of HMR1726 was detected by cell death detection ELISA. The enrichment of nucleosoms (mU) is denoted on the Y-axis.  
Fig. 5. Synoviocytes from 3 RA and 2 OA patients were activated with TNF-α and IL-17 in the presence of 50µM HMR1726. A. and B. The p-38-specific inhibitor (SB203580) and the specific inhibitor of MEK1/2 (U0126) were added (20 µM) to the cells and gene expression of MMP-1 and MMP-3 was quantified by real-time PCR. C. and D. Release of MMP-1 and –3 in supernatants from cells cultured under in (A) mentioned conditions (1RA and 20A) was measured by ELISA.  
Fig. 6. To investigate whether phosphorylation of p-38 and p42/ 44 MAPK is mediated by leflunomide, specific antibodies against the phosphorylated forms of p-38 and p42/44 MAPK were used for western blotting. Representative results from three independent experiments (from 1 RA, 2 OA patients) are shown. A. p- p38 expression in synoviocytes incubated with TNF-α, IL-17 and HMR1726 (50 µM) alone or in combination as indicated in the table for 48 h. B. time course of p-p38 expression (30', 1h, 3h, 24 h) in synoviocytes incubated with TNFα and IL-17 in presence or absence of HMR1726 (50 µM) as indicated in the table for 48 h. C. p-Erk1/2 expression in synoviocytes incubated with TNF-α and IL-17 in presence or absence of HMR1726 (50 µM) for 3 and 48 h.  
The active form of leflunomide, HMR1726, facilitates TNF-alpha and IL-17 induced MMP-1 and MMP-3 expression

February 2006

·

88 Reads

To elucidate the influence and mode of action of HMR1726 (the active metabolite of leflunomide) on TNF-alpha and IL-17 activated metalloproteinases expression in synoviocytes. Synovial fibroblasts from RA and OA patients were stimulated with both cytokines and altered gene expression in the presence or absence of leflunomide was detected by microarray analyses and quantitative RT-PCR. Protein expression was detected by western blotting and commercial ELISAs. Microarray analyses revealed that the addition of HMR1726 (50 microM) to TNF-alpha and IL-17- stimulated synoviocytes induced gene expression of metallo-proteinases, especially MMP-1 and -3 in comparison to activated synoviocytes in the absence of leflunomide. To confirm these data, we examined the influence of different concentrations of HMR1726 in synoviocytes from further 5 OA and 7 RA patients by quantitative PCR. HMR1726 gradually induced MMP-1 and MMP-3 gene expression in a dose-dosedependent manner. Similar results were observed on protein levels. Examination of signal transduction pathways participating in the regulation of leflunomideinduced MMPs expression showed that the mechanism underlying activation of MMP-1 is in part p38- and activation of MMP-3 was MEK1/2- dependent. Leflunomide was not able to abolish expression of metallo-proteinases in synoviocytes activated with TNF-a and IL-17.

Catechins Inhibit CCL20 Production in IL-17A-Stimulated Human Gingival Fibroblasts

November 2009

·

31 Reads

CC chemokine ligand 20 (CCL20) plays a pivotal role in the recruitment of Th17 cells and thus in the development of periodontal disease. Epigallocatechin gallate (EGCG) and epicatechin gallate (ECG), the major catechins in green tea, have multiple beneficial effects, but the effects of catechins on CCL20 production in human gingival fibroblasts (HGFs) are not known. In this study, we investigated the mechanisms by which EGCG and ECG inhibit interleukin (IL)-17A-induced CCL20 production in human gingival fibroblasts. IL-17A increased CCL20 production in HGFs in a concentration-dependent manner. EGCG and ECG prevented IL-17A-mediated CCL20 production in HGFs. Inhibitors of p38 mitogen-activated protein kinase (MAPK) or extracellular signal-regulated kinase (ERK) decreased IL-17A-induced CCL20 production. EGCG and ECG prevented IL-17A-induced phosphorylation of p38 MAPK and ERK in HGFs. In addition, EGCG and ECG attenuated IL-17 receptor expression on HGFs. These data provide a novel mechanism through which the green tea flavonoids catechins could be used to provide direct benefits in periodontal disease.

IL-17A induces pro-inflammatory cytokines production in macrophages via MAPKinases, NF-κB and AP-1

November 2013

·

345 Reads

Background: Interleukin (IL)-17A, a newly identified cytokine, may participate in the transition of a stable plaque into an unstable plaque. Macrophages play a critical role in the destabilization of atherosclerotic plaque . Methods: RAW 264.7 cells were stimulated with IL-17A. The mRNA expression of inflammatory cytokines was determined by RT-PCR. The cytokines production in the supernatants was measured by ELISA. Small interfering RNA (siRNA) was used to confirm that IL-17A-induced pro-inflammatory cytokines production via IL-17RA signaling. The western blot assay was used to detect the phosphorylation of MAPKinases including p38 and ERK1/2. The DNA binding activity of nuclear factor NF-κB and AP-1 were detected by EMSA. Results: IL-17A induced the production of pro-inflammatory cytokines in macrophages in a time- and dose-dependent manner, such as tumor necrosis factor (TNF)-α, IL-1β, and IL-6. Meanwhile, IL-17A resulted in the phosphorylation of p38 and ERK1/2 and increased DNA-binding activity of NF-κB and AP-1. Pharmacological inhibitors of p38 and ERK1/2 partly attenuated IL-17A-induced TNF-α, IL-1β, and IL-6 production. Either NF-κB inhibitor or AP-1 inhibitor also partly decreased the IL-17A-induced cytokine production. Conclusions: IL-17A induces pro-inflammatory cytokines production in macrophages via MAPKinases, NF-κB and AP-1 pathway. © 2013 S. Karger AG, Basel.

The Effects of 17β-estradiol on Mitochondrial Biogenesis and Function in Breast Cancer Cell Lines are Dependent on the ERα/ERβ Ratio

March 2012

·

127 Reads

17β-estradiol (E2) is a risk factor for the development of breast cancer, and cause tumorigenesis in epithelial breast cells. Moreover, E2 has distinct effects on different tissues that are attributed to the presence of two estrogen receptor isoforms, ERα and ERβ. The effect of E2 on mitochondrial biogenesis and function was investigated in two breast cancer cell lines with different estrogen receptor ratios, MCF-7 (high ERα/ERβ ratio) and T47D (low ERα/ERβ ratio) cell lines treated with physiological concentrations of E2 (1 nM). Mitochondria of the MCF-7 cell line showed an increase in proliferation but a decrease in functionality, while the T47D cell line, with low ERα/ERβ ratio, maintained functionality with fewer mitochondria. Our results suggest that ERs endowment and its subtypes relation have an effect on treatment response and could contribute new ideas about mitochondria and ERs in breast cancer, as well as new indicators to the disease progression.

Crucial Role of Phospholamban Phosphorylation and S-Nitrosylation in the Negative Lusitropism Induced by 17β-estradiol in the Male Rat Heart

August 2011

·

49 Reads

17β-estradiol (17βE2) plays an important cardiovascular role by activating estrogen receptors (ER) α and ERβ. Previous studies demonstrated that the novel estrogen G protein-coupled receptor (GPR30/GPER) mediates estrogen action in different tissues. We have recently shown in the rat heart that 17βE2 elicits negative inotropism through ERα, ERβ and GPR30, by triggering activation of ERK1/2, phosphatidylinositol 3-kinase (PI3K), protein kinase A (PKA) and endothelial Nitric Oxide synthase (eNOS) signaling. In the present study, using the isolated and Langendorff-perfused rat heart as a model system we analyzed: i) whether and to which extent 17βE2 modifies mammalian ventricular myocardial relaxation (lusitropism); ii) the type of ERs and the signaling pathways involved in this effect. We found that 17βE2 negatively modulated the ventricular lusitropic performance. This effect, which partially involved the vascular endothelium, recruited ERβ and occurred via PI3K, eNOS-NO-cGMP-protein kinase G (PKG) transduction cascade. Of note, 17βE2-mediated negative lusitropism associated with a modification of phospholamban (PLN) phosphorylation and S-nitrosylation (SNO) both in isolated Langendorff rat heart and in isolated cardiomyocytes. Taken together, our results allow including 17βE2 to the family of substances that control ventricular relaxation. This is of relevance in relation not only to the normal endocrine control of cardiac function, but also to physio-pathologic conditions characterized by an altered ventricular diastolic performance.

The Cytoprotective Action of the Potassium Channel Opener BMS-191095 in C2C12 Myoblasts is Related to the Modulation of Calcium Homeostasis

August 2010

·

363 Reads

BMS-191095 is an opener of the mitochondrial ATP-regulated potassium channel, which has been shown to provide cytoprotection in models of ischemia-reperfusion induced injury in various tissues. This study aimed at checking the protective action of BMS-191095 under the conditions of oxidative stress or disruption of intracellular calcium homeostasis. Methods: The cytoprotective potential of BMS-191095 was tested in C2C12 myoblasts injured by treatment with H(2)O(2) or calcium ionophore A23187. The influence of the opener on intracellular calcium levels, calpain activity and respiration rates were determined. Results: BMS-191095 protected myoblasts from calcium ionophore A23187-induced injury, but not from H H(2)O(2)-induced injury. A23187-mediated cell damage was also prevented by calpain inhibitor PD 150606. A23187 administration led to a transient increase in cytosolic calcium levels, concomitant activation of calpains and a decrease in state 3 respiration rates, indicating mitochondrial dysfunction. Co-administration of BMS-191095 diminished calpain activation in A23187-treated cells but did not prevent mitochondrial damage. In the presence of BMS-191095, restoration of cytosolic calcium concentrations to basal levels after A23187 treatment was considerably faster which may underly the reduced activation of calpains. Conclusion: The BMS-191095-mediated cytoprotection observed in C2C12 myoblasts results probably from modulation of intracellular calcium transients leading to prevention of calpain activation.

Acetylcholine Promotes ROS Detoxification Against Hypoxia/reoxygenation-Induced Oxidative Stress Through FoxO3a/PGC-1a Dependent Superoxide Dismutase

November 2014

·

156 Reads

Background/aims: Acetylcholine (ACh) is known to modulate the cardiac redox environment and thereby suppress reactive oxygen species (ROS) generation during oxidative stress. However, there is little information about its regulation on ROS clearance. Here we investigate the beneficial effects of ACh on superoxide dismutase (SOD) as key ROS-detoxifying enzyme system in cultured rat cardiomyoblasts. Methods: H9c2 cells were subjected to hypoxia/reoxygenation (H/R) to mimic oxidative stress. Western blot was used to detect the expression of SOD and related signaling molecules. Specific protein knockdown was performed with siRNA transfection. Results: ACh treatment on the beginning of reoxygenation decreased ROS and apoptosis. ACh increased ATP synthesis and mitochondrial DNA. Furthermore, ACh significantly reversed H/R-induced reduction in protein expressions and activities of SOD. ACh stimulated peroxisome proliferator activated receptor γ co-activator 1α (PGC-1α) and decreased forkhead box subfamily O3a (FoxO3a) phosphorylation. Atropine (muscarinic receptor antagonist) abolished the cytoprotection afforded by ACh. PGC-1α siRNA blocked ACh-induced invigorating effects on SOD2, whereas it did not alter SOD1 and FoxO3a phosphorylation. FoxOSa siRNA drastically decreased the expressions of SOD2 and PGC-1α, while it did not affect SOD1. Conclusion: ACh activates SOD2 within mitochondria through FoxO3a/PGC-1α pathway and up-regulates SOD1 in the cytoplasm, thus protecting against oxidative injury induced by H/R. Our findings provide new insights into mechanisms underlying the cardioprotection of ACh on ROS detoxifying.

Effect of Casein Kinase 1a Activator Pyrvinium Pamoate on Erythrocyte Ion Channels

July 2012

·

58 Reads

Pharmacological modification of protein kinase CK1 (casein kinase 1) has previously been shown to influence suicidal erythrocyte death or eryptosis, which is triggered by activation of Cl(-)-sensitive Ca(2+)-permeable cation channels. Ca(2+) entering through those channels stimulates cell membrane scrambling and opens Ca(2+)-activated K(+)-channels resulting in KCl exit and thus cell shrinkage. The specific CK1-inhibitor D4476 (1 µM) blunted, whereas the specific CK1 αactivator pyrvinium pamoate (10 µM) enhanced cell membrane scrambling. The substances were at least partially effective through modification of cytosolic Ca(2+)-activity. The present study explored, whether pyrvinium pamoate indeed influences Cl(-)-sensitive cation-channels in erythrocytes. As a result, removal of Cl(-)increased Fluo3-fluorescence (reflecting cytosolic Ca(2+)-activity), triggered cell membrane scrambling (apparent from annexin-V-binding), and decreased forward scatter (pointing to cell shrinkage). Pyrvinium pamoate significantly augmented the effect of Cl(-)-removal on Fluo3 fluorescence and annexin-V-binding, but blunted the effect on forward scatter. According to whole cell patch clamp recording, Cl(-)removal activated a cation current, which was significantly enhanced by pyrvinium pamoate. Pyrvinium pamoate inhibited Ca(2+)-activated K(+)-channels. Ca(2+)-ionophore ionomycin (1 µM) decreased forward scatter, an effect significantly blunted by pyrvinium pamoate. In conclusion, pyrvinium pamoate activates Cl(-)-sensitive Ca(2+)-permeable cation channels with subsequent Ca(2+)-entry and inhibits Ca(2+)-activated K(+)-channels thus blunting the stimulating effect of Ca(2+) on those channels, K(+)-exit and thus cell shrinkage.

Eukaryotic Translation Initiator Protein 1A Isoform, CCS-3, Enhances the Transcriptional Repression of p21CIP1 by Proto-oncogene FBI-1 (Pokemon/ZBTB7A)

February 2009

·

19 Reads

FBI-1, a member of the POK (POZ and Kruppel) family of transcription factors, plays a role in differentiation, oncogenesis, and adipogenesis. eEF1A is a eukaryotic translation elongation factor involved in several cellular processes including embryogenesis, oncogenic transformation, cell proliferation, and cytoskeletal organization. CCS-3, a potential cervical cancer suppressor, is an isoform of eEF1A. We found that eEF1A forms a complex with FBI-1 by co-immunoprecipitation, SDS-PAGE, and MALDI-TOF Mass analysis of the immunoprecipitate. GST fusion protein pull-downs showed that FBI-1 directly interacts with eEF1A and CCS-3 via the zinc finger and POZ-domain of FBI-1. FBI-1 co-localizes with either eEF1A or CCS-3 at the nuclear periplasm. CCS-3 enhances transcriptional repression of the p21CIP1 gene (hereafter referred to as p21) by FBI-1. The POZ-domain of FBI-1 interacts with the co-repressors, SMRT and BCoR. We found that CCS-3 also interacts with the co-repressors independently. The molecular interaction between the co-repressors and CCS-3 at the POZ-domain of FBI-1 appears to enhance FBI-1 mediated transcriptional repression. Our data suggest that CCS-3 may be important in cell differentiation, tumorigenesis, and oncogenesis by interacting with the proto-oncogene FBI-1 and transcriptional co-repressors.

HIF-1a as a Regulator of BMP2-Induced Chondrogenic Differentiation, Osteogenic Differentiation, and Endochondral Ossification in Stem Cells

April 2015

·

85 Reads

Joint cartilage defects are difficult to treat due to the limited self-repair capacities of cartilage. Cartilage tissue engineering based on stem cells and gene enhancement is a potential alternative for cartilage repair. Bone morphogenetic protein 2 (BMP2) has been shown to induce chondrogenic differentiation in mesenchymal stem cells (MSCs); however, maintaining the phenotypes of MSCs during cartilage repair since differentiation occurs along the endochondral ossification pathway. In this study, hypoxia inducible factor, or (HIF)-1α, was determined to be a regulator of BMP2-induced chondrogenic differentiation, osteogenic differentiation, and endochondral bone formation. BMP2 was used to induce chondrogenic and osteogenic differentiation in stem cells and fetal limb development. After HIF-1α was added to the inducing system, any changes in the differentiation markers were assessed. HIF-1α was found to potentiate BMP2-induced Sox9 and the expression of chondrogenesis by downstream markers, and inhibit Runx2 and the expression of osteogenesis by downstream markers in vitro. In subcutaneous stem cell implantation studies, HIF-1α was shown to potentiate BMP2-induced cartilage formation and inhibit endochondral ossification during ectopic bone/cartilage formation. In the fetal limb culture, HIF-1α and BMP2 synergistically promoted the expansion of the proliferating chondrocyte zone and inhibited chondrocyte hypertrophy and endochondral ossification. The results of this study indicated that, when combined with BMP2, HIF-1α induced MSC differentiation could become a new method of maintaining cartilage phenotypes during cartilage tissue engineering. © 2015 S. Karger AG, Basel.

Hypoxia-Inducible Factor-1a Dependent Pathways Mediate the Renoprotective Role of Acetazolamide Against Renal Ischemia-Reperfusion Injury

December 2013

·

100 Reads

Background/Aims: Acute kidney injury (AKI) is a major complication of kidney transplantation, resulting in early graft dysfunction. Since diuretic acetazolamide (AZA) has been shown to improve contrast induced AKI, we hypothesized that AZA also protected against ischemia-reperfusion (I/R) caused AKI. Methods: An in vivo mouse renal I/R injury model and an in vitro H2O2 stimulated HK-2 cell injury model were utilized to examine the renoprotective effect of AZA. Renal injury and blood flow were measured. Nitric oxide synthase (eNOS)/Nitric oxide (NO), cell apoptosis and hypoxia-inducible factor-1α (HIF-1α) changes were analyzed. Results: AZA reduced kidney injury scores and improved renal function by decreasing serum creatinine and BUN levels after I/R. Impaired renal blood flow was restored by increasing eNOS activities and NO production, as indicated by Laser Doppler imaging. TUNEL staining presented that AZA reduced apoptotic cells due to attenuated caspase activation and increased Bcl-2/Bax ratio. Furthermore, HIF-1α induction by AZA was demonstrated. AZA also enhanced in vitro NO production, reduced cell apoptosis and increased HIF-1α expression. Knockdown of HIF-1α by RNAi confirmed that AZA exerted its protective role depending on HIF-1α. AZA's effects were significantly reduced by Akt inhibitor LY294002. Conclusions: The present study demonstrated that AZA exerted a renoprotective role against I/R induced AKI through activating HIF-1α and downstream pathways. © 2013 S. Karger AG, Basel.

Fig. 1. Identification of cells overexpressing HIF-1α. A & B. VE cells were stably transfected with plasmid pcDNA3.1/V5-His-HIF-1α. The expression of HIF-1α mRNA in HIF-1α transfectants was significantly higher than that in control transfectants. Data are representative of three similar experiments. *, p<0.05, compared with VE control. C & D. Both control and HIF-1α transfectants were treated without or with 1mmol/L DMOG for 12 hours in normoxia. The expression of HIF-1α protein in HIF-1α transfectants treated with DMOG was significantly higher than that in HIF-1α transfectants without DMOG treatment or control transfectants treated without or with DMOG. Data are representative of three similar experiments. *, p<0.05, compared with control or HIF-1α transfectants without DMOG treatment. #, p<0.05, compared with control transfectants treated with DMOG. 
Fig. 2. HIF-1α overexpression enhances hypoxia-induced HIF-1 activation. A. Exposure of VE cells to hypoxia for 12 hours induced a significant increase of HIF-1α protein, whereas HIF-1α overexpression significantly enhanced the hypoxia-induced increase of HIF-1α protein in VE cells. Data are representative of five similar experiments. *, p<0.05, compared with control VE cells exposed to normoxia. #, p<0.05, compared with control VE cells exposed to hypoxia. B. Hypoxic treatment of VE cells for 12 hours induced a significant increase of GLUT-1 protein, which was significantly enhanced by HIF-1α overexpression. Data are representative of five similar experiments. *, p<0.05, compared with control VE cells exposed to normoxia. #, p<0.05, compared with control VE cells exposed to hypoxia. 
Fig. 3. HIF-1α overexpression exacerbates hypoxia-induced endothelial barrier dysfunction. A. The barrier function of monolayers was assessed by measuring TER. Treatment of VE monolayers with hypoxia for 24 hours induced a significant decrease of TER, which was significantly enhanced by HIF-1α overexpression. The data represent the mean ± SEM (n=9). *, p<0.05, compared with control VE cells exposed to normoxia. #, p<0.05, compared with control VE cells exposed to hypoxia. B. The monolayers were exposed to normoxia or hypoxia for 24 hours. Neither hypoxia nor HIF-1α overexpression caused a significant change in the protein expression of both ZO-1 and occludin. Data are representative of five similar experiments. 
Fig. 4. HIF-1α overexpression augments hypoxiainduced up-regulation of MLCK protein expression. The monolayers were treated with normoxia or hypoxia for 24 hours. Hypoxia treatment up-regulated the expression of MLCK protein, whereas HIF-1α overexpression enhanced the hypoxia-induced upregulation of MLCK protein. Data are representative of five similar experiments. *, p<0.05, compared with control VE cells exposed to normoxia. #, p<0.05, compared with control VE cells exposed to hypoxia. 
Overexpression of Hypoxia-Inducible Factor-1a Exacerbates Endothelial Barrier Dysfunction Induced by Hypoxia

September 2013

·

101 Reads

Background/aims: The mechanisms involved in endothelial barrier dysfunction induced by hypoxia are incompletely understood. There is debate about the role of hypoxia-inducible factor-1α (HIF-1α) in endothelial barrier disruption. The aim of this study was to investigate the effect of genetic overexpression of HIF-1α on barrier function and the underlying mechanisms in hypoxic endothelial cells. Methods: The plasmid pcDNA3.1/V5-His-HIF-1α was stably transfected into human endothelial cells. The cells were exposed to normoxia or hypoxia. The mRNA and protein expressions of HIF-1α were detected by RT-PCR and Western blot respectively. The barrier function was assessed by measuring the transendothelial electrical resistance (TER). The Western blot analysis was used to determine the protein expression of glucose transporter-1 (GLUT-1), zonular occludens-1 (ZO-1), occludin, and myosin light chain kinase (MLCK) in endothelial cells. The mRNA expression of proinflammatory cytokines was detected by qRT-PCR. Results: Genetic overexpression of HIF-1α significantly increased the mRNA and protein expression of HIF-1α in endothelial cells. The overexpression of HIF-1α enhanced the hypoxia-induced increase of HIF-1α and GLUT-1 protein expression. HIF-1α overexpression not only exacerbated hypoxia-induced endothelial barrier dysfunction but also augmented hypoxia-induced up-regulation of MLCK protein expression. HIF-1α overexpression also enhanced IL-1β, IL-6 and TNF-α mRNA expression. Conclusion: We provide evidence that genetic overexpression of HIF-1α aggravates the hypoxia-induced endothelial barrier dysfunction via enhancing the up-regulation of MLCK protein expression caused by hypoxia, suggesting a potential role for HIF-1α in the pathogenesis of endothelial barrier dysfunction in hypoxia.

Short-Term Effects of Nose-Only Cigarette Smoke Exposure on Glutathione Redox Homeostasis, Cytochrome P450 1A1/2 and Respiratory Enzyme Activities in Mice Tissues

May 2013

·

129 Reads

Background/Aims: The components of cigarette smoke (CS) have been implicated in the development of cancer as well as in cardiopulmonary diseases. We have previously reported increased oxidative stress in rat tissues induced by tobacco-specific toxins nicotine and 4-(N-methyl-N-nitrosamino)-1-(3-pyridyl)-1-butanone (NNK). Recently, we have also shown increased oxidative stress and associated inflammatory responses in various tissues after exposure to cigarette smoke. Methods: In this study, we have further investigated the effects of nose-only cigarette smoke exposure on mitochondrial functions and glutathione-dependent redox metabolism in tissues of BALB/C mice. Liver, kidney, heart and lung tissues were analyzed for oxidative stress, glutathione (GSH) and cytochrome P450 dependent enzyme activities and mitochondrial functions after exposure to smoke generated by 9 cigarettes/day for 4 days. Control mice were exposed to air only. Results: An increase in oxidative stress as observed by increased production of reactive oxygen species (ROS) and altered GSH metabolism was apparent in all the tissues, but lung and heart appeared to be the main targets. Increased expression and activity of CYP450 1A1 and 1A2 were also observed in the tissues after exposure to cigarette smoke. Mitochondrial respiratory dysfunction in the tissues, as observed by alterations in the activities of Complex I and IV enzymes, was also observed after exposure to cigarette smoke. SDS-PAGE and Western blot results also indicate that alterations in the expression of enzyme proteins were in accordance with the changes in their catalytic functions. Conclusion: These results suggest that even short term exposure of cigarette smoke have adverse effects on mitochondrial functions and redox homeostasis in tissues which may progress to further complications associated with chronic smoking.

Free Fatty Acids Inhibit Protein Tyrosine Phosphatase 1B and Activate Akt

September 2013

·

378 Reads

Background/Aims: Accumulating evidence has suggested that free fatty acids (FFAs) interact with protein kinases and protein phosphatases. The present study examined the effect of FFAs on protein phosphatases and Akt. Methods: Activities of protein phosphatase 1 (PP1), protein phosphatase 2A (PP2A), and protein tyrosine phosphatase 1B (PTP1B) were assayed under the cell-free conditions. Phosphorylation of Akt was monitored in MSTO-211H human malignant pleural mesothelioma cells without and with knocking-down phosphatidylinositol 3 kinase (PI3K) or 3-phosphoinositide-dependent protein kinase-1 (PDK1). Results: In the cell-free assay, unsaturated FFAs (uFFAs) such as oleic, linoleic and linolenic acid and saturated FFAs (sFFAs) such as stearic, palmitic, myristic, and behenic acid markedly reduced PTP1B activity, with the potential for uFFAs greater than that for sFFAs. All the investigated sFFAs inhibited PP2A activity, but otherwise no inhibition was obtained with uFFAs. Both uFFAs and sFFAs had no effect on PP1 activity. Oleic acid phosphorylated Akt both on Thr308 and Ser473, while stearic acid phosphorylated Akt on Thr308 alone. The effects of oleic and stearic acid on Akt phosphorylation were abrogated by the PI3K inhibitor wortmannin or the PDK1 inhibitor BX912 and also by knocking-down PI3K or PDK1. Conclusion: The results of the present study indicate that uFFAs and sFFAs could activate Akt through a pathway along a PI3K/PDK1/Akt axis in association with PTP1B inhibition. © 2013 S. Karger AG, Basel.

Modulation of Protein Tyrosine Phosphatase 1B by Erythropoietin in UT-7 Cell Line

February 2007

·

71 Reads

Since the reversible phosphorylation of tyrosyl residues is a critical event in cellular signaling pathways activated by erythropoietin (Epo), attention has been focused on protein tyrosine phosphatases (PTPs) and their coordinated action with protein tyrosine kinases. The prototypic member of the PTP family is PTP1B, a widely expressed non-receptor PTP located both in cytosol and intracellular membranes via its hydrophobic C-terminal targeting sequence. PTP1B has been implicated in the regulation of signaling pathways involving tyrosine phosphorylation induced by growth factors, cytokines, and hormones, such as the downregulation of erythropoietin and insulin receptors. However, little is known about which factor modulates the activity of this enzyme. The effect of Epo on PTP1B expression was studied in the UT-7 Epo-dependent cell line. PTP1B expression was analyzed under different conditions by Real-Time PCR and Western blot, while PTP1B phosphatase activity was determined by a p-nitrophenylphosphate hydrolysis assay. Epo rapidly induced an increased expression of PTP1B which was associated with higher PTP1B tyrosine phosphorylation and phosphatase activity. The action of Epo on PTP1B induction involved Janus Kinase 2 (JAK2) and Phosphatidylinositol-3 kinase (PI3K). The results allow us to suggest for the first time that, besides modulating Epo/Epo receptor signaling, PTP1B undergoes feedback regulation by Epo.

Fig. 2. Comparison of serum sCD40 levels in different groups of subjects. Serum sCD40 were detected in sera of subjects and the obtained data were logarithmically transformed and then analyzed by one-way ANOVA. Compared with healthy controls. * P<0.001. 
Association of CD40 -1C/T Polymorphism in the 5'-Untranslated Region with Chronic HBV Infection

January 2015

·

21 Reads

Background: CD40 is an important costimulatory molecule in both celluar and humoral immune responses, involved in the pathogenic processes of chronic inflammatory diseases. Few studies were performed on the association of CD40 single nucleotide polymorphism (SNP) with chronic hepatitis B virus (HBV) infection. In this study, we studied whether the CD40-1C/T polymorphism had any effect on the progression of chronic HBV infection in Chinese population. Methods: CD40 -1C/T polymorphism in the 5'-untranslated region was analyzed by polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP) in 453 chronic HBV carriers, who were divided into asymptomatic HBV carriers (ASC), moderate chronic hepatitis B group (MCHB) and severe chronic hepatitis B group (SCHB). 202 healthy individuals in the same region were enrolled in this study as the controls. The CD40 expression on B lymphocytes was detected by flow cytometry. The concentrations of soluble CD40 (sCD40) in sera were assayed by a commercial ELISA kit. Results: Our results showed the frequencies of TT genotype and T allele of CD40-1C/T polymorphism were higher significantly in ASC than those in controls (P< 0.05), while this result was not found in either MCHB or SCHB. On the surface of B lymphocytes, the CD40 expression levels in the individuals with TT genotype were significantly lower than those with CC and CT genotypes in either ASC group or healthy controls (P<0.001). The sCD40 levels in the sera of ASC, MCHB and SCHB groups were significantly higher than the controls (P<0.001). Conclusions: The CD40 -1C/T polymorphism may contribute to the susceptibility of asymptomatic HBV carriers through its effect on cell-surface CD40 expression, which indicated CD40 signaling was involved in immune tolerance of chronic HBV infection. © 2015 S. Karger AG, Basel.

Fig. 1. Effect of Ca 2+ free medium and 10 µM BAPTA/AM on ethanol-induced insulin secretion from INS-1 and INS-1E cells. Encapsulated cells were incubated for two subsequent 30 min periods in medium containing alternately basal and stimulating medium (80 mM ethanol). Media without Ca 2+ contained 1 mM EGTA with/without DMSO and BAPTA/AM. Results are shown as ng of insulin released by cells during 30 min incubation, mean ± SE, n = 10. *** P<0.001 compared to previous (basal) incubation (paired t-test).  
Fig. 3. Effect of SNARE proteins inhibition on insulin secretion induced by ethanol from INS-1 and INS-1E cells. INS-1 a INS- 1E cells were divided into three groups; control was incubated for 30 min in 100 µl depolarization medium and TeTx groups were incubated in the same medium with 20 nM tetanus toxin in presence/absence of calcium. After 60 min preincubation in 200 µl glucose free medium cells were perifused at flow rate 0.25 ml/min with basal medium and stimulating ethanol medium with (TeTx + Ca 2+ ) or without 1.5 mM CaCl 2 (TeTx -Ca 2+ ) in scheme (30 min basal medium → 20 min ethanol → 20 min basal medium → 20 min ethanol). Calcium depleted medium contained 1 mM EGTA.  
Fig. 4. Effect of N-ethylmaleimide on insulin secretion from INS-1 and INS-1E cells. Cells after preincubation with or without N-ethylmaleimide (NEM) were incubated in two subsequent 30 min period in basal and stimulating media in scheme i) control group: basal medium → 15 mM glucose or 80 mM ethanol; ii) testing group: basal medium +5 mM NEM → 15 mM glucose or 80 mM ethanol +5 mM NEM. Results are shown as ng of insulin released by cells during 30 min incubation, mean ± SE, n = 10, *** P<0.001 compared to previous (basal) incubation (paired ttest ); ### P<0.001 compared to basal incubation without NEM (unpaired t-test).  
Fig. 2. Effect of PKC inhibition on insulin secretion from INS- 1 and INS-1E cells. Attached cells; after preincubation with or without Bisindolylmaleimide VIII (BIM) they were incubated in two subsequent 30 min period in basal and stimulating media according to scheme i) control group: basal medium → 15 mM glucose or 80 mM ethanol; ii) testing group: basal medium +3 µM BIM → 15 mM glucose or 80 mM ethanol +3 µM BIM. Results are shown as ng of insulin released by cells during 30 min incubation, mean ± SE, n = 10. * P<0.05, ** P<0.005, *** P<0.001 compared to previous (basal) incubation (paired t-test); # P<0.05, ## P<0.005, ### P<0.001 compared to basal incubation without BIM (unpaired t-test); ^^ P<0.005 compared to ethanol incubation without BIM (unpaired t-test).  
Fig. 5. Effect of PTP inhibition on insulin secretion from INS- 1 and INS-1E cells. Cells after preincubation were incubated in two subsequent 30 min period in basal and stimulating media in scheme i) control group: basal medium → 15 mM glucose or 80 mM ethanol; ii) testing group: basal medium +10 µmol/l ZnCl 2 → 15 mM glucose or 80 mM ethanol +10 µmol/l ZnCl 2 . Results are shown as ng of insulin released by cells during 30 min incubation, mean ± SE, n = 10. ** P<0.005, *** P<0.001 compared to previous (basal) incubation (paired t-test); ## P<0.005, ### P<0.001 compared to basal incubation without ZnCl 2 (unpaired t-test).  
Mechanism of Ethanol-Induced Insulin Secretion from INS-1 and INS-1E Tumor Cell Lines

November 2009

·

261 Reads

Unlabelled: Alcohol causes reactive hypoglycemia by attenuating the release of counter regulatory hormones, redistribution of pancreatic blood flow and direct stimulation of insulin secretion. Objective of this study was characterization of ethanol-induced insulin secretion. Signaling of ethanol- and glucose-induced insulin release from INS-1 and INS-1E cells was compared. Both cell lines responded similarly to all experimental interventions. In contrast to glucose, ethanol-induced insulin secretion was not hindered in calcium depleted medium or by addition of 10 microM BAPTA/AM (intracellular chelator). Inhibitor of protein kinase C Bisindolylmaleimide (3 microM) abolished glucose- but not ethanol-induced insulin secretion. Tetanus toxin (20 nM), inhibitor of SNARE proteins complex formation, blocked ethanol-induced insulin secretion. Both 5 mM N-ethylamaleimide and 10 microM ZnCl(2) (inhibitor of protein tyrosine phosphatases), which block disassembly of SNARE complexes and their further participation in exocytosis, increased basal insulin secretion. In contrast to glucose, already high insulin secretion was further increased after ethanol stimulation in either treatment. Conclusion: Signaling of ethanol-induced insulin secretion from INS-1 and INS-1E cell lines bypasses calcium and PKC involving steps, is sensitive to tetanus toxin but resistant to N-ethymaleimide and ZnCl(2). An extra pool of secretory vesicles not available for glucose is exploited for exocytosis after ethanol stimulation.

Fig. 1: Detection of Ca v 2.3 splice variants by RT-PCR in mouse heart. After reverse transcription of total RNA which was isolated from freshly dissected tissues, cDNA fragments of the cytosolic linker between domain II and III, and of the region encoding the carboxy terminus were amplified in a tissue specific manner. No DNA fragments were amplified when RNA or reverse transcriptase was omitted. A) Fragments of the II-III loop. Mouse brain total RNA was used as a reference (1 µg; lane 2 and 3) yielding three different fragments of 363 bp, 399 bp, and 420 bp indicative for recently identified isoforms of Ca v 2.3. Note, that the smallest fragment in adult heart (1 µg; parallel aliquots, lane 4 to 6) as well as in prenatal heart (lane 7 to 9 and 12) is indicative of the Ca v 2.3e splice variant. The upper band indicative for the full length variant gets expressed later, at day 19.5 p.c. (lane 9 and 12, see arrow). Size markers (100 bp ladder) are loaded in lane 1, 10, and 11. B) Fragments of the carboxy terminus. Mouse brain RNA was used as a reference (1 µg; lane 2 and 3) yielding one minor fragment of 498 bp and a major one of 369 bp. Note, that the larger fragment in adult heart (1 µg; lane 4 and 5) is indicative for the predominant expression of the so called "endocrine" Ca v 2.3e splice variant [29] which is confirmed by the amplification of the extended fragment from the carboxy terminal region in prenatal mice at day 14.5 (lane 6) and 18.5 dpc (lane 7). Size markers (100 bp ladder) are loaded in lane 1 and 8. C) Structural differences between the major splice variants of Ca v 2.3. C1: The intron/exon structure of mouse and human cacna1e gene is identical, and the gene is located on chromosome 1 in both species. The full length cDNA of the human gene product represents the splice variant Ca v 2.3d [48]. The nucleotide sequence between nts 2142 and 2945 contains exon 18 to 20 of cavna1e, and the sequence of nts 5784 to 6206 represents exon 44 to 46 out of 50 exons. C2: The major neuronal splice variant of adult rat and mice contains exon 19 (= insert 1; RDR..) which is structurally similar to the adjacent end of exon 20 (RER..). Within exon 20 exists another short segment of 21 nts (= insert 2, labelled with *) which is missing in the cloned rabbit Ca v 2.3 [53] and causes the shift of the cDNA amplified from mouse brain and adult heart as well as elder prenatal heart shown in A), lane 2 to 6 and lane 12. Insert 3 is located in the carboxy terminus and only detected in cerebellum [47], and in endocrine tissues [29]. 
Fig. 5: SNX-482 affects the electrical activity of isolated heart from Ca v 2.3(+|+) but not from Ca v 2.3(-|-) mice. At day 9.5 p.c., hearts from prenatal wild type mice were isolated and placed on microelectrode arrays (MEAs). After overnight culture, episodes of MEA-recordings were collected during the initial 30 s lasting control period (1), and subsequent episodes of 60 s each of SNX-482 (2), first washout (3), second SNX-482 application (4), and a second washout of the toxin (5). Each episode was recorded at the end of a 10 min superfusion periode without or with SNX-482. A. The spontaneous beating rate of isolated whole wild type hearts was calculated before adding the Ca v 2.3-selective blocker SNX-482, during superfusion with a solution containing 200 nM SNX-482, and during washout of the SNX-482. B. Example traces of field potentials from always 
Fig. 6: Dihydropyridine-sensitivity of inward currents through recombinant Ca v 2.3. A. Superimposed representative traces of whole cell recordings from stably transfected HEK-293 cells expressing human Ca v 2.3d. Ca 2+ inward currents were activated from a holding potential of-50 mV to the test potential at 30 mV every 10 s for 400 ms. Charge carrier was 5 mM extracellular Ca 2+. B. Time course of peak inward currents before (control) and during the exposure to (±)isradipine from the same cell as shown in panel A. Concentrations in µM as indicated by numbers. After adding 30 µM isradipine, the inhibition was partially reversed (Wash). C. Dose-response relationship for (±)isradipine block after normalization of peak inward currents. Inward currents were recorded from HEK-293 cells expressing either Ca v 2.3d (closed symbols) or the cardiac splice variant Ca v 2.3e (open circles). The number of cells was 4 to 19 for each point. Data were fitted using the Hill equation yielding an IC 50 value of 9.1 µM and 14.6 µM for Ca v 2.3d and Ca v 2.3e, respectively. The slope factor (Hill coefficient) was 1.2 for both splice variants. 
Arrhythmia in Isolated Prenatal Hearts after Ablation of the Ca v2.3 (??1E) Subunit of Voltage-gated Ca 2+ Channels

February 2004

·

259 Reads

A voltage-gated calcium channel containing Cav2.3e (alpha1Ee) as the ion conducting pore has recently been detected in rat heart. Functional evidence for this Ca2+ channel to be involved in the regulation of heart beating, besides L- and T-type channels, was derived from murine embryos where the gene for Cav1.2 had been ablated. The remaining "L-type like" current component was not related to recombinant splice variants of Cav1.3 containing channels. As recombinant Cav2.3 channels from rat were reported to be weakly dihydropyridine sensitive, the spontaneous activity of the prenatal hearts from Cav2.3(-|-) mice was compared to that of Cav2.3(+|+) control animals to investigate if Cav2.3 could represent such a L-type like Ca(2+) channel. The spontaneous activity of murine embryonic hearts was recorded by using a multielectrode array. Between day 9.5 p.c. to 12.5 p.c., the beating frequency of isolated embryonic hearts from Cav2.3-deficient mice did not differ significantly from control mice but the coefficient of variation within individual episodes was more than four-fold increased in Cav2.3-deficient mice indicating arrhythmia. In isolated hearts from wild type mice, arrhythmia was induced by superfusion with a solution containing 200 nM SNX-482, a blocker of some R-type voltage gated Ca2+ channels, suggesting that R-type channels containing the splice variant Cav2.3e as ion conducting pore stabilize a more regular heart beat in prenatal mice.

The Phytostilbene Resveratrol Induces Apoptosis in INS-1E Rat Insulinoma Cells

February 2009

·

460 Reads

We investigated the effect of resveratrol on proliferation and induction of apoptosis of INS-1E rat insulinoma cells by cell counting, crystal violet staining, flow cytometry and immunoblotting. Resveratrol treatment of INS-1E cells at concentrations > or =50 microM resulted in a dose-dependent inhibition of cell proliferation, accumulation of the cells in the S and G0/G1 phase and a significant increase of the percentage of apoptotic cells. This was paralleled by an increase of cell granularity, apoptotic volume decrease (AVD), exposure of phosphatidylserine at the outer leaflet of the plasma membrane, an increase of the 7-AAD signal and caspase activation. The AMP-kinase (AMPK) inhibitor compound C (10 microM) significantly inhibited cell proliferation and induced caspase activation within 48 hours but this effect was not modified by resveratrol suggesting that AMPK is not a major target involved in mediating the proapoptotic effect of resveratrol in INS-1E cells. Immunoblotting revealed a significant inhibition of Akt (PKB) phosphorylation by 100 muM resveratrol within 1 hour. Addition of insulin (10 microM) to the culture medium strongly enhanced basal Akt phosphorylation. This enhancement was significantly attenuated by 50 and 100 microM resveratrol. We conclude that the antiproliferative/proapoptotic effect of resveratrol on INS-1E cells is due to negative interference with Akt signaling and most likely disruption of auto/paracrine insulin signaling.

Effect of the AMP-Kinase Modulators AICAR, Metformin and Compound C on Insulin Secretion of INS-1E Rat Insulinoma Cells under Standard Cell Culture Conditions

March 2012

·

468 Reads

The function of β-cells is regulated by nutrient uptake and metabolism. The cells' metabolic state can be expressed as concentration ratios of AMP, ADP and ATP. Relative changes in these ratios regulate insulin release. An increase in the intracellular ATP concentration causes closure of K(ATP) channels and cell membrane depolarization, which triggers stimulus-secretion coupling (SSC). In addition to K(ATP) channels, the AMP-dependent protein kinase (AMPK), a major cellular fuel sensor in a variety of cells and tissues, also affects insulin secretion and β-cell survival. In a previous study we found that the widely used AMPK inhibitor compound C retards proliferation and induces apoptosis in the rat β-cell line INS-1E. We therefore tested the effects of AMPK activators (AICAR and metformin), and compound C on AMPK phosphorylation, insulin secretion, K(ATP) channel currents, cell membrane potential, intracellular calcium concentration, apoptosis and cell cycle distribution of INS-1E cells under standard cell culture conditions (11 mM glucose). Western blotting, ELISA, patch-clamp, calcium imaging and flow cytometry. We found that basal AMPK phosphorylation is enhanced by AICAR (1 mM) and metformin (1 mM) but remained unaffected by compound C (10 μM). Both AICAR and compound C stimulated basal insulin secretion whereas metformin had no effect. Pre-incubation with AICAR (1 mM) caused an inhibition of K(ATP) currents but did not significantly alter the average cell membrane potential (Vm) or the threshold potential of electrical activity. Acute administration of AICAR (300 μM) led to a depolarization of Vm, which was not due to an inhibition of the basal- or glucose-induced chloride conductance, and was not accompanied by elevations of intracellular calcium (Ca(i)). AICAR had no additive blocking effect on K(ATP) currents when applied together with tolbutamide. Compound C applied over 24 hours induced an increase in the percentage of cells positive for caspase activity, whereas AICAR (1 mM) applied for 48 hours was without effect. Medium glucose concentration <3 mM caused cell cycle arrest, caspase activation and an increase of cell granularity. We conclude that under standard cell culture conditions the AMPK modulators AICAR and compound C, but not metformin, stimulate insulin secretion by AMPK-independent mechanisms.

Glucose Induces Anion Conductance and Cytosol-To-Membrane Transposition of ICln in INS-1E Rat Insulinoma Cells

February 2006

·

223 Reads

The metabolic coupling of insulin secretion by pancreatic beta cells is mediated by membrane depolarization due to increased glucose-driven ATP production and closure of K(ATP) channels. Alternative pathways may involve the activation of anion channels by cell swelling upon glucose uptake. In INS-1E insulinoma cells superfusion with an isotonic solution containing 20 mM glucose or a 30% hypotonic solution leads to the activation of a chloride conductance with biophysical and pharmacological properties of anion currents activated in many other cell types during regulatory volume decrease (RVD), i.e. outward rectification, inactivation at positive membrane potentials and block by anion channel inhibitors like NPPB, DIDS, 4-hydroxytamoxifen and extracellular ATP. The current is not inhibited by tolbutamide and remains activated for at least 10 min when reducing the extracellular glucose concentration from 20 mM to 5 mM, but inactivates back to control levels when cells are exposed to a 20% hypertonic extracellular solution containing 20 mM glucose. This chloride current can likewise be induced by 20 mM 3-Omethylglucose, which is taken up but not metabolized by the cells, suggesting that cellular sugar uptake is involved in current activation. Fluorescence resonance energy transfer (FRET) experiments show that chloride current activation by 20 mM glucose and glucose-induced cell swelling are accompanied by a significant, transient redistribution of the membrane associated fraction of ICln, a multifunctional 'connector hub' protein involved in cell volume regulation and generation of RVD currents.

Cell Swelling-induced Insulin Secretion from INS-1E Cells is Inhibited by Extracellular Ca2+ and is Tetanus Toxin Resistant

August 2010

·

215 Reads

Unlabelled: Cell swelling-induced insulin secretion represents an alternative pathway of stimulation of insulin secretion. INS-1E rat tumor beta cells do not release insulin in response to cell swelling in presence of Ca(2+) despite a good response to glucose challenge and appropriate increase in cell volume. Surprisingly, perifusion with Ca(2+)-depleted medium showed distinct secretory response of INS-1E cells to hypotonicity. Objective of this study was further characterization of the role of Ca(2+) in secretory process in INS-1 and INS-1E cell lines. Ca(2+) depleted hypotonic medium with 10 muM BAPTA/AM (intracellular chelator) induced insulin secretion from both types of cells. We demonstrated expression of L-type Ca(2+) channel Ca(v)1.2 and non-L-type Ca(2+) channels Ca(v)2.1 (P/Q-type), Ca(v)2.2 (N-type), and Ca(v)3.1 (T-type) in both cell lines. Inhibition of L type channel with nifedipine and/or P/Q type with omega-agatoxin IVA enabled distinct response to hypotonic medium also in INS-1E cells. Tetanus toxin (TeTx) in medium containing Ca(2+) and a group of calcium channel blockers inhibited hypotonicity-induced insulin secretion from INS-1 cells but not from INS-1E cells. Conclusion: Hypotonicity-induced insulin secretion from INS-1E cells is inhibited by extracellular Ca(2+), does not require intracellular Ca(2+) and is TeTx resistant.

Inhibition of the Expression of TGF-�1 and CTGF in Human Mesangial Cells byExendin-4, a Glucagon-like Peptide-1Receptor Agonist

August 2012

·

25 Reads

Despite the presence of glucagon-like peptide-1 receptor (GLP-1R) in kidney tissues, its direct effect on diabetic nephropathy remains unclear. The transforming growth factor-β(1) (TGF-β(1)) and the connective tissue growth factor (CTGF) both induce extracellular matrix accumulation and persistent fibrosis in the glomerular mesangium of patients with diabetic nephropathy. Herein, we demonstrate that a GLP-1R agonist, exendin-4, exerts renoprotective effects through its influence on TGF-β(1) and CTGF in human mesangial cells (HMCs), cultured in a high glucose medium. HMCs, cultured in a high glucose medium, were used for the current study. The direct effect of exendin-4 on TGF-β(1) and CTGF expression was confirmed in HMCs. MDL-12330A (a specific adenylate cyclase inhibitor) and PKI14-22 (a protein kinase A inhibitor) were used to examine the role of the cAMP signaling pathway in exendin's anti-fibrosis action. The findings showed that exendin-4 inhibited the proliferation of HMCs, and upregulated the expression of TGF-β(1) and CTGF, induced by high glucose. The effect of exendin-4 is largely dependent on the activation of adenylate cyclase. This study provides new evidence that GLP-1 acts as an antifibrotic agent in HMCs.

Figure 1: NO production. Quantification of NO production measured by means of Griess method and expressed as percentage of control values. NO production was evaluated after 3 minutes stimulation in the presence of different concentrations of Vit. D (1-100 nM). Black bars = cells without ZK159222 addition, gray bars = cells + 10 nM ZK159222 (P<0.05).
Figure 2: PAE cells proliferation. A) Determination of PAE proliferation in the presence of different concentrations of Vit. D (1-100 nM) after 24 hours of incubation. Results are expressed as n. cells/mm2 ± S.D. Black bars = cells without L-NAME addition, gray bars = cells + 10 mM L-NAME. * p<0.05; ** p<0.001. B) Optical microscopy images of control and Vit. D (10 nM) treated cells both in absence or presence of 10 mM L-NAME after 24 hours of incubation, stained with toluidine blue. Magnification = 10X. Scale bar = 150 µm.
Figure 3: PAE cells migration into a three-dimensional matrix. A) Determination of PAE migration after 7 days of incubation in the presence of different concentrations of Vit. D (1-100 nM). Results are expressed as n. cells/HPF ± S.D. * p<0.05. B) Determination of control and Vit. D (100 nM) treated cells migration both in the presence or absence of 10 mM L-NAME after 7 days of incubation. *p<0.0001 compared to control sample; # p<0.05 compared to Vit. D 100 nM alone.
Figure 4: Optical microscopy images of control and Vit. D (100 nM) treated cells both in the presence or absence of 10 mM L-NAME after 7 days of incubation, stained with Hoechst 33342. Magnification = 10X. Scale bar = 60 µm.
Figure 5: MMP-2 production. A) Representative zymography of cell growth medium form PAE cells migrated into the three-dimensional matrix for 7 days in the presence of different Vit. D concentrations (1-100 nM). B) Densitometric quantification of MMP-2 expression. C) Representative zymography of cell growth medium from control and Vit. D (100 nM) PAE cells migrated into the three-dimensional matrix for 7 days, both in the presence or absence of 10 mM L-NAME. D) Densitometric quantification of MMP-2 expression.
1α,25-Dihydroxycholecalciferol (Vitamin D3) Induces NO-Dependent Endothelial Cell Proliferation and Migration in a Three-Dimensional Matrix

June 2013

·

147 Reads

Background/Aims: The 1α,25-dihydroxycholecalciferol (Vit. D) induces eNOS dependent nitric oxide (NO) production in human umbilical vein endothelial cells (HUVEC). To our knowledge, there are no reports directly relating Vit. D induced NO production to proliferation and/or migration in endothelial cells (EC). The aim of this study was to evaluate whether Vit. D addition to porcine EC could affect their proliferation and/or migration in a three-dimensional matrix via NO production. Materials and Methods: Porcine aortic endothelial cells (PAE) were used to evaluate Vit. D effects on cell proliferation and migration in a three-dimensional matrix. Results: Vit. D induced NO production in PAE cells. Moreover, it induced a significant increase in cellular proliferation and migration in a three-dimensional matrix. These effects were NO dependent, as inhibiting eNOS activity by L-NAME PAE migration was abrogated. This effect was strictly related to MMP-2 expression and apparently dependent on Vit. D and NO production. Conclusions: Vit. D can promote both endothelial cells proliferation and migration in a three-dimensional matrix via NO-dependent mechanisms. These findings cast new light on the role of Vit. D in the angiogenic process, suggesting new applications for Vit. D in such fields as tissue repair and wound healing.

Increased Glucose Uptake and Metabolism in Mesangial Cells Overexpressing Glucose Transporter 1 Increases Interleukin-6 and Vascular Endothelial Growth Factor Production: Role of AP-1 and HIF-1α

February 2006

·

14 Reads

Previous results indicate that enhanced glucose transporter (GLUT)1 expression mediates the deleterious effects of metabolic and hemodynamic perturbations leading to diabetic kidney disease. First screening for altered gene expression in GLUT1 overexpressing cells (GT1) by Affymetrix microarray analysis revealed upregulation of interleukin-6 (IL-6) and vascular endothelial growth factor (VEGF) expression, which was verified by RT-PCR. Subsequently, IL-6 and VEGF protein production was more than 3-fold increased in the GT1 cells. This upregulation was independent from each other. Studies on the underlying transcriptional mechanisms by gelshift assays and siRNA approach implicated activation of AP-1 in the increased expression of both, IL-6 and VEGF. We found also increased nuclear protein levels of hypoxia-inducible factor (HIF)-1alpha and enhanced DNA binding activity to a hypoxia responsible element located in the VEGF promoter. Knock-down of HIF-1alpha reduced the VEGF expression to 50% with an additive effect of AP-1 gene silencing down to 24%. The IL-6 expression was not affected by reducing HIF-1alpha. In conclusion our results link increased GLUT1 levels leading to excess glucose metabolism under normoglycemic conditions and altered gene expression of pathogenetic factors involved in diabetic kidney disease.

Top-cited authors