首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Cell death following photodynamic therapy (PDT) with the photosensitizer Pc 4 involves the intrinsic pathway of apoptosis. To evaluate the importance of Bax in apoptosis after PDT, we compared the PDT responses of Bax-proficient (Bax+/−) and Bax knock-out (BaxKO) HCT116 human colon cancer cells. PDT induced a slow apoptotic process in HCT Bax+/− cells following a long delay in the activation of Bax and release of cytochrome c from mitochondria. Although cytochrome c was not released from mitochondria following PDT in BaxKO cells, an alternative mechanism of caspase-dependent apoptosis with extensive chromatin and DNA degradation was found in these cells. This alternative process was less efficient and slower than the normal apoptotic process observed in Bax+/− cells. Early events upon PDT, such as the loss of mitochondrial membrane potential, photodamage to Bcl-2, and activation of p38 MAP kinase, were observed in both HCT116 cell lines. In spite of differences in the efficiency and mode of apoptosis induced by PDT in the Bax+/− and BaxKO cells, they were found to be equally sensitive to killing by PDT, as determined by loss of clonogenicity. Thus, for Pc 4-PDT, the commitment to cell death occurs prior to and independent of Bax activation, but the process of cellular disassembly differs in Bax-expressing vs. non-expressing cells.  相似文献   

2.
Previously, we have synthesized a novel cyclin‐dependent kinase (CDK) inhibitor, 2‐[1,1′biphenyl]‐4‐yl‐N‐[5‐(1,1‐dioxo‐1λ6‐isothiazolidin‐2‐yl)‐1H‐indazol‐3‐yl]acetamide (BAI) and reported its anti‐cancer activity in head and neck cancer cells. In this study, we further evaluated the effect of BAI on growth of various human cancer cell lines, including A549 (nonsmall cell lung cancer), HCT116 (colon), and Caki (kidney). Profoundly, results of XTT and clonogenic assays demonstrated that BAI at nanomolar concentrations (20–60 nM) inhibited growth of A549, HCT116, and Caki cells, suggesting the anti‐cancer potency. We show that BAI induced a dose‐dependent apoptotic cell death in these human cancer cells, as measured by fluorescence‐activated cell sorting (FACS). Interestingly, further biochemical analysis showed that treatment with BAI at 20 nM induced apoptosis in A549 cells in association with activation of caspases, cleavage of phospholipase C‐γ1 (PLC‐γ1), and inhibition of Akt in A549 cells. Importantly, pharmacological inhibition study revealed that pretreatment with z‐VAD‐fmk, a pan caspase inhibitor strongly blocked the BAI‐induced apoptosis in A549 cells. Transfection analysis with Akt cDNA encoding constitutively active Akt further addressed the significance of Akt inhibition in the BAI‐induced apoptosis in A549 cells. Notably, disruption of the PI3K/Akt pathway by LY294002, a PI3K/Akt inhibitor potentiated apoptosis in A549 cells by BAI at a subcytotoxic concentration. These findings collectively suggest that BAI potently inhibits growth of A549, HCT116, and Caki cells, and that the BAI‐induced apoptosis in A549 cells is associated with activation of caspases, and inhibition of Akt. J. Cell. Biochem. 114: 282–293, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
Kaempferol (3,4′,5,7‐tetrahydroxyflavone) is a flavonoid with anti‐ and pro‐oxidant activity present in various natural sources. Kaempferol has been shown to posses anticancer properties through the induction of the apoptotic program. Here we report that treatment of the chronic myelogenous leukemia cell line K562 and promyelocitic human leukemia U937 with 50 µM kaempferol resulted in an increase of the antioxidant enzymes Mn and Cu/Zn superoxide dismutase (SOD). Kaempferol treatment induced apoptosis by decreasing the expression of Bcl‐2 and increasing the expressions of Bax. There were also induction of mitochondrial release of cytochrome c into cytosol and significant activation of caspase‐3, and ‐9 with PARP cleavage. Kaempferol treatment increased the expression and the mitochondria localization of the NAD‐dependent deacetylase SIRT3. K562 cells stably overexpressing SIRT3 were more sensitive to kaempferol, whereas SIRT3 silencing did not increase the resistance of K562 cells to kaempferol. Inhibition of PI3K and de‐phosphorylation of Akt at Ser473 and Thr308 was also observed after treating both K562 and U937 cells with kaempferol. In conclusion our study shows that the oxidative stress induced by kaempferol in K562 and U937 cell lines causes the inactivation of Akt and the activation of the mitochondrial phase of the apoptotic program with an increase of Bax and SIRT3, decrease of Bcl‐2, release of cytochrome c, caspase‐3 activation, and cell death. J. Cell. Biochem. 106: 643–650, 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
The purine-derived analogs, roscovitine and purvalanol are selective synthetic inhibitors of cyclin-dependent kinases (CDKs) induced cell cycle arrest and lead to apoptotic cell death in various cancer cells. Although a number of studies investigated the molecular mechanism of each CDK inhibitor on apoptotic cell death mechanism with their therapeutic potential, their regulatory role on autophagy is not clarified yet. In this paper, our aim was to investigate molecular mechanism of CDK inhibitors on autophagy and apoptosis in wild type (wt) and Bax deficient HCT 116 cells. Exposure of HCT 116 wt and Bax−/− cells to roscovitine or purvalanol for 24 h decreased cell viability in dose-dependent manner. However, Bax deficient HCT 116 cells were found more resistant against purvalanol treatment compared to wt cells. We also established that both CDK inhibitors induced apoptosis through activating mitochondria-mediated pathway in caspase-dependent manner regardless of Bax expression in HCT 116 colon cancer cells. Concomitantly, we determined that purvalanol was also effective on autophagy in HCT 116 colon cancer cells. Inhibition of autophagy by 3-MA treatment enhanced the purvalanol induced apoptotic cell death in HCT 116 Bax−/− cells. Our results revealed that mechanistic action of each CDK inhibitor on cell death mechanism differs. While purvalanol treatment activated apoptosis and autophagy in HCT 116 cells, roscovitine was only effective on caspase-dependent apoptotic pathway. Another important difference between two CDK inhibitors, although roscovitine treatment overcame Bax-mediated drug resistance in HCT 116 cells, purvalanol did not exert same effect.  相似文献   

5.
A series of forty α-substituted chalcones were synthesized and screened for their antiproliferative activities against HCT116 (colorectal) and HCC1954 (breast) cancer cell lines. Compounds 5a and 5e were found to be the most potent compounds with GI50 values of 0.63 µM and 0.725 µM in HCC1954 cell line and 0.69 µM and 1.59 µM in HCT116 cell line, respectively. Both compounds induced a G2/M cell cycle arrest and caused apoptotic cell death in HCT116 cells as shown by the induction of PARP cleavage. The compounds also stabilized p53 in a dose-dependent manner in HCT116 cells following 24-hour treatment. Furthermore, both 5a and 5e were able to overcome multidrug resistance in two MDR-1 overexpressing multidrug resistant cell lines.  相似文献   

6.
Dental caries affects people of all ages and is a worldwide health concern. Streptococcus mutans is a major cariogenic bacterium because of its ability to form biofilm and induce an acidic environment. In this study, the antibacterial activities of magnolol and honokiol, the main constituents of the bark of magnolia plants, toward planktonic cell and biofilm of S. mutans were examined and compared with those of chlorhexidine. The minimal inhibitory concentrations of magnolol, honokiol and chlorhexidine for S. mutans were 10, 10 and 0.25 µg/mL, respectively. In addition, each agent showed bactericidal activity against S. mutans planktonic cells and inhibited biofilm formation in a dose‐ and time‐dependent manner. Magnolol (50 µg/mL) had greater bactericidal activity against S. mutans biofilm than honokiol (50 µg/mL) and chlorhexidine (500 µg/mL) at 5 min after exposure, while all showed scant activity against biofilm at 30 s. Furthermore; chlorhexidine (0.5–500 µg/mL) exhibited high cellular toxicity for the gingival epithelial cell line Ca9‐22 at 1 hr, whereas magnolol (50 µg/mL) and honokiol (50 µg/mL) did not. Thus; it was found that magnolol has antimicrobial activities against planktonic and biofilm cells of S. mutans. Magnolol may be a candidate for prevention and management of dental caries.  相似文献   

7.
8.
Objective: Low‐molecular weight chromium compounds, such as chromium picolinate [Cr(pic)3], improve insulin sensitivity, although toxicity is a concern. We synthesized a novel chromium complex, chromium (d ‐phenylalanine)3 [Cr(d ‐phe)3], in an attempt to improve insulin sensitivity with reduced toxicity. The aim of this study was to compare the two chromium compounds on cardiac contractile function in ob/ob obese mice. Research Methods and Procedures: C57BL lean and ob/ob obese mice were randomly divided into three groups: H2O, Cr(d ‐phe)3, or Cr(pic)3 (45 µg/kg per day orally for 6 months). Results: The glucose tolerance test displayed improved glucose clearance by Cr(d ‐phe)3 but not Cr(pic)3. Myocytes from ob/ob mice exhibited depressed peak shortening (PS) and maximal velocity of shortening/relengthening (±dL/dt), prolonged time‐to‐PS and time‐to‐90% relengthening (TR90), reduced electrically stimulated rise in intracellular Ca2+ (Δfura‐2 fluorescence intensity), and slowed intracellular Ca2+ decay. Although a 3‐month Cr(d ‐phe)3 treatment for a separate group of ob/ob and lean 2‐month‐old mice only rectified reduced ±dL/dt in ob/ob mice, all mechanical and intracellular Ca2+ abnormalities were significantly attenuated or ablated by 6 months of Cr(d ‐phe)3 but not Cr(pic)3 treatment (except TR90). Sarco(endo)plasmic reticulum Ca2+ ATPase activity and Na+‐Ca2+ exchanger expression were depressed in ob/ob mice, which were reversed by both Cr(d ‐phe)3 and Cr(pic)3, with a more pronounced effect from Cr(d ‐phe)3. Cr(d ‐phe)3 corrected reduced insulin‐stimulated glucose uptake and improved basal phosphorylation of Akt and insulin receptor, as well as insulin‐stimulated phosphorylation of Akt and insulin receptor in ob/ob myocytes. Heart homogenates from ob/ob mice had enhanced oxidative stress and protein carbonyl formation compared with the lean group, which were attenuated by both Cr(d ‐phe)3 and Cr(pic)3. Discussion: Our data suggest that the new Cr(d ‐phe)3 compound possesses better cardio‐protective and insulin‐sensitizing properties against obesity.  相似文献   

9.
Introduction – Honokiol and magnolol are the active components of Magnolia officinalis, which is a widely used traditional Chinese medicine. Their simultaneous analysis is, therefore, important for the quality control of the product. Objective – To establish a simple, sensitive and rapid electrochemical method for the simultaneous detection of honokiol and magnolol based on the remarkable enhancement effect of acetylene black nanoparticle (AB). Methodology – The AB‐modified electrode was prepared via solvent evaporation. The electrochemical response of honokiol and magnolol was investigated using cyclic voltammetry. The simultaneous detection was performed with differential pulse voltammetry. The method was validated in terms of linearity, sensitivity, precision and accuracy. Results – The linear range for honokiol is 0.5–300 µg/L, and the limit of detection (LOD) is 0.25 µg/L (9.4 × 10?10 mol/L). For magnolol, the linear range is 10–250 µg/L, and the LOD is 5 µg/L (1.88 × 10?8 mol/L). Conclusion – The new method was successfully used to determine honokiol and magnolol in a traditional Chinese medicine called Ageratum liquid. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
Two new oleanane‐type saponins: β‐d ‐xylopyranosyl‐(1 → 4)‐6‐deoxy‐α‐l ‐mannopyranosyl‐(1 → 2)‐1‐O‐{(3β)‐28‐oxo‐3‐[(2‐Oβ‐d ‐xylopyranosyl‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐yl}‐β‐d ‐glucopyranose ( 1 ) and 1‐O‐[(3β)‐28‐oxo‐3‐{[β‐d ‐xylopyranosyl‐(1 → 2)‐α‐l ‐arabinopyranosyl‐(1 → 6)‐2‐acetamido‐2‐deoxy‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐yl]β‐d ‐glucopyranose ( 2 ), along with two known saponins: (3β)‐3‐[(β‐d ‐Glucopyranosyl‐(1 → 2)‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐oic acid ( 3 ) and (3β)‐3‐{[α‐l ‐arabinopyranosyl‐(1 → 6)‐[β‐d ‐glucopyranosyl‐(1 → 2)]‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐oic acid ( 4 ) were isolated from the acetone‐insoluble fraction obtained from the 80% aqueous MeOH extract of Albizia anthelmintica Brongn . leaves. Their structures were identified using different NMR experiments including: 1H‐ and 13C‐NMR, HSQC, HMBC and 1H,1H‐COSY, together with HR‐ESI‐MS/MS, as well as by acid hydrolysis. The four isolated saponins and the fractions of the extract exhibited cytotoxic activity against HepG‐2 and HCT‐116 cell lines. Compound 2 showed the most potent cytotoxic activity among the other tested compounds against the HepG2 cell line with an IC50 value of 3.60μm . Whereas, compound 1 showed the most potent cytotoxic effect with an IC50 value of 4.75μm on HCT‐116 cells.  相似文献   

11.
Curcumin has anti‐oxidant, anti‐cancer and anti‐carcinogen property. Our laboratory had previously reported that, curcumin treatment induces reactive oxygen species (ROS) generation in HT‐29 cell line, an effect contradictory to its anti‐oxidant property. This study evaluates the role of p53 in curcumin mediated ROS generation and cell death. Curcumin induced ROS was determined by 2’,7’‐dichlorofluorescein and apoptosis by Hoechst33342/PI staining in HT‐29 and HCT‐116 cell lines. ROS generation occurs within 1 hour of 40 µM curcumin treatment and a reduction was observed by third hour in HCT‐116 insinuating p53 involvement. N‐acetyl cysteine (NAC) pre‐treatment effectively quenched ROS and inhibited membrane potential loss in HT‐29, but less effective in HCT‐116. Mitochondrial membrane potential loss is evident with 10 and 40 µM curcumin in HCT‐116 and at 40 µM curcumin in HT‐29. Total p53 protein level increase was observed by 24 hours in HCT‐116 upon NAC pre‐treatment. Our results indicate that curcumin induces ROS mediated cell death in colon adenocarcinoma cell lines and may be mediated via p53.  相似文献   

12.
In this study, we investigated an SBP (DctPAm) of a tripartite ATP‐independent periplasmic transport system (TRAP) in Advenella mimigardefordensis strain DPN7T. Deletion of dctPAm as well as of the two transmembrane compounds of the tripartite transporter, dctQ and dctM, impaired growth of A. mimigardefordensis strain DPN7T, if cultivated on mineral salt medium supplemented with d ‐glucose, d ‐galactose, l ‐arabinose, d ‐fucose, d ‐xylose or d ‐gluconic acid, respectively. The wild type phenotype was restored during complementation studies of A. mimigardefordensis ΔdctPAm using the broad host vector pBBR1MCS‐5::dctPAm. Furthermore, an uptake assay with radiolabeled [14C(U)]‐d ‐glucose clearly showed that the deletion of dctPAm, dctQ and dctM, respectively, disabled the uptake of this aldoses in cells of either mutant strain. Determination of KD performing thermal shift assays showed a shift in the melting temperature of DctPAm in the presence of d ‐gluconic acid (KD 11.76 ± 1.3 µM) and the corresponding aldonic acids to the above‐mentioned carbohydrates d ‐galactonate (KD 10.72 ± 1.4 µM), d ‐fuconic acid (KD 13.50 ± 1.6 µM) and d ‐xylonic acid (KD 8.44 ± 1.0 µM). The sugar (glucose) dehydrogenase activity (E.C.1.1.5.2) in the membrane fraction was shown for all relevant sugars, proving oxidation of the molecules in the periplasm, prior to transport.  相似文献   

13.
In the current study, a dimeric phenoloxidase (PO) from the hemolymph of healthy and diseased (pebrine infected) larvae of Antheraea assamensis Helfer was extracted and purified. The protein was subjected to purification using Sephacryl S‐100 and CM Sepharose chromatography. The enzyme comprised of two subunits of ~76.8 and 76 kDa that showed PO activity in 6 mM l ‐3,4‐dihydroxyphenylalanine (L ‐DOPA) and 8 mM catechol but not in hydroquinone. Optimum temperature for PO activity was 30°C in l ‐DOPA and 37°C in catechol. Optimum pH ranged from 6.8 to 7.0 in L ‐DOPA and 7.0–7.2 in catechol. Specific activity of the purified PO from healthy larvae was 53.9 µM/min per mg of protein per ml in L ‐DOPA and 50.77 µM/min per mg of protein per ml in catechol. Specific activity of PO from diseased larvae was 30.0 µM/min per mg of protein per ml in L ‐DOPA and 28.55 µM/min per mg of protein per ml in catechol. Purification fold was 3.27–4.21 for healthy and 2.38–2.56 for diseased fractions. The enzyme showed the Michaelis constant (Km) of 2.46–2.85 mM for healthy and diseased fractions in L ‐DOPA. In catechol Km of 9.23–17.71 mM was observed. Peptidoglycan was the best activator of purified PO from both healthy and diseased fractions. Interactions between controls and activators appeared statistically significant (F = 767.5; df = 3; P < 0.0001). Na+, K+, and Cu2+ increased, whereas Ca2+, Zn2+, Mg2+, and Co2+ decreased PO activity. The overall interactions appeared highly significant (F = 217.0; df = 27; P < 0.0001). Kojic acid, dithiothreitol, thiourea, phenylthiourea, carbendazim, N‐bromosuccinimide, N,N,N′,N′‐tetraacetic acid, and diethyldithiocarbamate inhibited PO activity.  相似文献   

14.
The elucidation of factors that support human mesenchymal stem cells (hMSCs) growth has remained unresolved partly because of the reliance of many researchers on ill‐defined, proprietary medium formulation. Thus, we investigated the effects of high glucose (D ‐glucose, 25 mM) on hMSCs proliferation. High glucose significantly increased [3H]‐thymidine incorporation and cell‐cycle regulatory protein expression levels compared with 5 mM D ‐glucose or 25 mM L ‐glucose. In addition, high glucose increased transforming growth factor‐β1 (TGF‐β1) mRNA and protein expression levels. High glucose‐induced cell‐cycle regulatory protein expression levels and [3H]‐thymidine incorporation, which were inhibited by TGF‐β1 siRNA transfection and TGF‐β1 neutralizing antibody treatment. High glucose‐induced phosphorylation of protein kinase C (PKC), p44/42 mitogen‐activated protein kinases (MAPKs), p38 MAPK, Akt, and mammalian target of rapamycin (mTOR) in a time‐dependent manner. Pretreatment of PKC inhibitors (staurosporine, 10?6 M; bisindolylmaleimide I, 10?6 M), LY 294002 (PI3 kinase inhibitor, 10?6 M), Akt inhibitor (10?5 M), PD 98059 (p44/42 MAPKs inhibitor, 10?5 M), SB 203580 (p38 MAPK inhibitor, 10?6 M), and rapamycin (mTOR inhibitor, 10?8 M) blocked the high glucose‐induced cellular proliferation and TGF‐β1 protein expression. In conclusion, high glucose stimulated hMSCs proliferation through TGF‐β1 expression via Ca2+/PKC/MAPKs as well as PI3K/Akt/mTOR signal pathways. J. Cell. Physiol. 224:59–70, 2010 © 2010 Wiley‐Liss, Inc.  相似文献   

15.
Simvastatin is a competitive inhibitor of 3‐hydroxymethylglutaryl coenzyme A reductase activity, whereas geraniol is a monoterpene with multiple pharmacologic effects on mevalonate metabolism. Both of them inhibit growth and proliferation of many cell lines. The present study was designed to determine the action of geraniol, in combination with simvastatin, by assessing their effects in vitro on human hepatocarcinoma cell line (Hep G2). The treatment of Hep G2 cells with concentrations of simvastatin or geraniol that did not inhibit cell proliferation (5 µmol·l‐1 of simvastatin and 50 µmol·l‐1 of geraniol) resulted in a significant inhibition of cell proliferation. We also examined the effect of simvastatin, geraniol and the combination of both on the biosynthesis of lipids from [14C]‐acetate. Our results demonstrate that the combination of simvastatin and geraniol synergistically inhibited cholesterol biosynthesis and proliferation of Hep G2 cell line, contributing to a better understanding of the action of a component of essential oils targeting a complex metabolic pathway, which would improve the use of drugs or their combination in the fight against cancer and/or cardiovascular diseases. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

16.
The present work reports on the biological activity of alfalfa (Medicago sativa) saponins on white poplar (Populus alba, cultivar ‘Villafranca’) cell suspension cultures. The extracts from alfalfa roots, aerial parts and seeds were characterized for their saponin content by means of thin layer chromatography (TLC) and electrospray ionisation coupled to mass spectrometry. The quantitative saponin composition from the different plant extracts was determined considering the aglycone moieties and determined by gas chromatography (GC) and gas chromatography/mass spectrometry (GC/MS) analyses. Only soyasapogenin I was detected in the seed extract while several other saponins were found in the root and leaf extracts. Actively proliferating white poplar cell cultures were challenged with the different saponin extracts. Only alfalfa root saponins, at 50 µg ml?1, induced significant cell death rates (75.00 ± 4.90%). Different cell subpopulations with peculiar cell death morphologies were observed and the programmed cell death (PCD)/necrosis ratio was reduced at increasing saponin concentrations. Enhancement of nitric oxide (NO) production was observed in white poplar cells treated with root saponins (RSs) at 50 µg ml?1 and release of reactive oxygen species (ROS) in the culture medium was also demonstrated. Saponin‐induced NO production was sensitive to sodium azide and NG‐monomethyl‐l ‐arginine, two specific inhibitors of distinct pathways for NO biosynthesis in plant cells.  相似文献   

17.
18.
The emergence of strains of multidrug‐resistant Gram‐negative bacteria mandates a search for new types of antimicrobial agents. Alyteserin‐2a (ILGKLLSTAAGLLSNL.NH2) is a cationic, α‐helical peptide, first isolated from skin secretions of the midwife toad, Alytes obstetricans, which displays relatively weak antimicrobial and haemolytic activities. Increasing the cationicity of alyteserin‐2a while maintaining amphipathicity by the substitution Gly11→ Lys enhanced the potency against both Gram‐negative and Gram‐positive bacteria by between fourfold and 16‐fold but concomitantly increased cytotoxic activity against human erythrocytes by sixfold (mean concentration of peptide producing 50% cell death; LC50 = 24 µm ). Antimicrobial potency was increased further by the additional substitution Ser7→Lys, but the resulting analogue remained cytotoxic to erythrocytes (LC50 = 38 µm ). However, the peptide containing d ‐lysine at positions 7 and 11 showed high potency against a range of Gram‐negative bacteria, including multidrug‐resistant strains of Acinetobacter baumannii and Stenotrophomonas maltophilia (minimum inhibitory concentration = 8 µm ) but appreciably lower haemolytic activity (LC50 = 185 µm ) and cytotoxicity against A549 human alveolar basal epithelial cells (LC50 = 65 µm ). The analogue shows potential for treatment of nosocomial pulmonary infections caused by bacteria that have developed resistance to commonly used antibiotics. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Non-steroidal anti-inflammatory drugs (NSAIDs) are known to induce apoptosis in a variety of cancer cells. However, the precise mechanisms by which NSAIDs facilitate apoptosis in tumor cells are not clear. In the present study, we show that niflumic acid (NA), a member of the fenamates group of NSAIDs and Cl? and Ca2+-activated Cl? (CAC) channels blocker, induced apoptosis (by ~8 %, 24 h treatment) and potentiated (by 8–10 %) apoptotic effect of endoplasmic reticulum Ca2+ mobilizer thapsigargin (Tg) in human erythroleukemic K562 cell line. The whole-cell patch clamp and Fluo-3 flow cytometric experiments confirmed an inhibitory effect of NA (100 and 300 µM) on store-operated (SOC) channels. We also found that NA-blocked CAC channels were activated by acute application of Tg (2 µM) in K562 cells. NA blockage of CAC channels was accompanied by activation of Ca2+-activated K+ (SK4) channels. The observed effects of NA were not connected with COX-2 inhibition since 100-nM NA (IC50 for COX-2 inhibition) did not induce either apoptosis or affect the channels activity. We conclude that inhibition of SOC channels plays a major role in NA-induced apoptosis. Increased apoptotic levels in Tg-treated K562 cells in the presence of NA may be due to the blockage of CAC and stimulation of SK4 channels in addition to SOC channels inhibition.  相似文献   

20.
We report the potent and selective cytotoxicity of the crude aqueous leaf extract from the medicinal plant, Pithecellobium dulce toward the human breast cancer cells (MCF‐7), but not the normal cells (MCF‐10A). The cytotoxicity was found to be dose and time dependent, as 300 µg/mL of the extract decreased the cell viability to 50% (IC50) in 48 h. The induction of apoptosis in the breast cancer cells after treatment was confirmed by significant percentage (24.7%), of early apoptotic cells (AnnexinV +Propidium Iodide_) in treated cells as compared to control cells (3.5%). We observed a significant upregulation in the mRNA expression of various pro‐apoptotic gene such as Bax (21.1 folds), p21(14.4 folds), p53 (11.7 folds), TNF (10.2 folds) and fas (6.3 folds) after treatment as compared to untreated cells. On the other hand, the relative mRNA expression of anti‐apoptotic genes such as Bcl‐2, NF‐KB and Cdk was reduced. The selective upregulation of pro‐apoptotic gene and down regulation of specific anti‐apoptotic genes could be the inducing factor for apoptotic cell death in MCF‐7 cells after treatment with the herbal extract. We believe that our findings provide a foundation for further studies on this formulation as a potential therapeutic candidate for breast cancer. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 32:756–766, 2016  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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