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.
Roflumilast, a potent and selective phosphodiesterase 4 (PDE4) inhibitor, has been demonstrated to be an effective anti-inflammatory agent in airway inflammatory diseases. In the present study, we investigated the mechanism of anti-inflammatory effects of roflumilast in murine macrophage cell line RAW264.7 cells. Roflumilast inhibited NO, tumor necrosis factor (TNF)-alpha, and interleukin (IL)-1beta production via suppression of their gene expressions in lipopolysaccharide (LPS)-stimulated macrophages. To elucidate the mechanism by which roflumilast inhibits the production of inflammatory mediators, we examined the effect of roflumilast on the activation of nuclear factor-kappaB (NF-kappaB) in these cells. Roflumilast inhibited the DNA binding activity of NF-kappaB by preventing inhibitor kappaBalpha phosphorylation and degradation. The phosphorylation of mitogen-activated protein (MAP) kinases, including stress-activated protein kinase/c-Jun NH2-terminal kinase (JNK) and p38 MAP kinase, was also markedly inhibited by roflumilast. Similar to the effects of roflumilast, treatment of either SB203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)imidazole] or SP600125 [anthra(1,9-cd)pyrazol-6(2H)-one 1,9-pyrazoloanthrone], specific inhibitors of p38 MAP kinase and JNK, respectively, suppressed NO, TNF-alpha, and IL-1beta production. Consistent with in vitro results, administration of roflumilast recovered the survival rate of LPS-treated mice, with concurrent suppression of plasma levels of nitrite/nitrate, TNF-alpha, and IL-1beta. These results suggest that the inhibitory activity of roflumilast on the production of inflammatory mediators seems to be mediated via inhibition of NF-kappaB, p38 MAP kinase, and JNK activation in macrophages.  相似文献   

2.
Decreased glutathione (GSH) levels and gamma-glutamylcysteine ligase (GCL) activity have been observed in diabetic patients, and insulin reportedly increases GSH synthesis via increased GCL catalytic subunit (GCLC) gene expression. The signaling pathways responsible for mediating insulin effects on GCLC expression and GSH levels, however, are unknown. The signaling pathways involved in the regulation of GSH synthesis in response to insulin were examined in primary cultured rat hepatocytes. GSH levels, GCL activity, GCLC protein, and mRNA levels were increased to 140, 160, 600, and 340% of that monitored in untreated cells, respectively, in hepatocytes cultured with 100 nM insulin. The phosphatidylinositol 3-kinase (PI3K) inhibitors, wortmannin and LY294002 [2-(4-morpholinyl)-9-phenyl-4H-1-benzopyran-4-one], dominant-negative Akt, or rapamycin, an inhibitor of mTOR (mammalian target of rapamycin) and ribosomal p70 S6 kinase (p70S6K) phosphorylation, inhibited the insulin-mediated increase in GCLC protein and GSH levels. Although the mitogen-activated protein kinases (MAPKs) extracellular signal-regulated kinase, p38 MAPK, and JNK (c-Jun N-terminal kinase) were activated in response to insulin, PD98059 (2'-amino-3'-methoxyflavone), an inhibitor of mitogen-activated protein kinase kinase, SP600125 (1,9-pyrazoloanthrone), an inhibitor of JNK, and SB203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)1H-imidazole], an inhibitor of p38 MAPK, failed to inhibit the insulin-mediated increase in GCLC protein levels. In conclusion, these data show that insulin signaling pathways involving PI3K/Akt/p70S6K, but not MAPKs, are active in the insulin-mediated regulation of GSH synthesis via increased GCLC expression.  相似文献   

3.
4.
The present study evaluated some of the mechanisms underlying prostaglandin E2 (PGE2)-induced paw edema formation in mice. Intraplantar (i.pl.) injection of PGE2 (0.10-10.0 nmol/paw) into the hindpaw elicited a dose-related edema formation, with a mean ED50 value of 0.42 nmol/paw. The coinjection of selective E-prostanoid (EP)3 [(2E)-N-[(5-bromo-2-methoxyphenyl)-sulfonyl]-3-[5-chloro-2-(2-naphthylmethyl)phenyl]acrylamide; L826266), but not EP2 or EP4 (all 10 nmol/paw), receptor antagonists significantly inhibited PGE2-induced paw edema. Like L826266, the PGE2-induced paw edema was markedly reduced by treatment with pertussis toxin and phospholipase C (PLC) inhibitor 1-[6-[[17beta-methoxyestra-1,3,5(10)-trien-17-yl]amino]hexyl]-1H-pyrrole-2,5-dione (U-73122). Likewise, the selective neurokinin (NK)1 receptor antagonist N-[(4R)-4-hydroxy-1-(1-methyl-1H-indol-3-yl)carbonyl-l-prolyl]-N-methyl-N-phenyl-methyl-3-(2-aphthyl)-l-alaninamide (FK888) and the antagonist of vanilloid receptor (TRPV1) receptors 4'-chloro-3-methoxycinnamanilide (SB366791) (both 1 nmol/paw) also significantly inhibited PGE2-mediated paw edema. Conversely, the selective NK2, NK3, and calcitonin gene-related peptide (CGRP) CGRP(8-37) receptor antagonists all failed to interfere with PGE2-induced paw edema. The neonatal treatment of mice with capsaicin was also able to reduce PGE2-induced paw edema. The inhibitors of protein kinase C (PKC) 3-[1-[3-(dimethylaminopropyl]-1H-indol-3-yl]-4-(1H-indol-3-yl)-1H-pyrrole-2,5-dione monohydrochloride (GF109203X) and mitogen protein-activated kinases (MAPKs; 30 nmol/paw) c-Jun NH2-terminal kinase (JNK) (anthra[1,9-cd]pyrazol-6(2H)-one; SP600125), extracellular signal-regulated kinase (PD98059), and p38 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)1H-imidazole; SB203580], but not protein kinase A, markedly decreased the PGE2-mediated edema formation. The i.pl. injection of PGE2 (3 nmol/paw) induced a significant activation of MAPKs, namely, JNK and p38, an effect that was largely prevented by the selective EP3 receptor antagonist L826266 (10 nmol/paw). Collectively, these findings indicate that edematogenic responses elicited by PGE2 are mediated by EP3 receptor activation, also involving the stimulation of PLC, PKC, and MAPKs pathways and the participation of TRPV1 and NK1 receptors. These results make a considerable contribution to our comprehension of the mechanisms involved in PGE2-mediated inflammatory responses in mice.  相似文献   

5.
6.
Connexin 43, the major connexin isoform in gap junctions of cardiac ventricular myocytes, undergoes changes in distribution and expression in cardiac diseases. The Na(+)-H(+) exchanger (NHE-1), a key mediator of hypertrophy and heart failure, has been shown to be localized in the cardiomyocyte gap junctional regions; however, whether NHE-1 regulates gap junction proteins in the hypertrophied cardiomyocyte is not known. To address this question, neonatal rat ventricular myocytes were treated with phenylephrine (PE) for 24 h to induce hypertrophy. Increased Cx43 expression observed with PE treatment (132.4 +/- 6.3% compared to control; P < 0.05) was further significantly augmented by the specific NHE-1 inhibitor EMD87580 [N-[2-methyl-4,5-bis(methylsulfonyl)-benzoyl]-guanidine hydrochloride] (173.2 +/- 8.7% increase compared to control; P < 0.05 versus PE), an effect that was mimicked by another NHE-1 inhibitor cariporide [4-isopropyl-3-(methylsulfonyl)benzoyl-guanidine methanesulfonate]. PE-induced hypertrophy was associated with mitogen-activated protein kinase c-Jun NH(2)-terminal kinase (JNK) 1/2 activation, whereas inhibition of JNK1/2 with either SP600125 [anthra(1,9-cd)pyrazol-6(2H)-one 1,9-pyrazoloanthrone] or small interfering RNA significantly increased PE-induced up-regulation of Cx43 protein levels. Inhibition of reverse mode Na(+)-Ca(2+) exchange (NCX) with KB-R7943 [2-[2-[4-(4-nitrobenzyloxy)phenyl]ethyl]isothiourea mesylate] partially reversed JNK1/2 activation (195.2 +/- 21.4 versus 143.7 +/- 14.4% with KB-R7943; P < 0.05) and augmented up-regulation of Cx43 protein (121.1 +/- 8.3 versus 215.9 +/- 25.6% with KB-R7943; P < 0.05) in the presence of PE. Our results demonstrate that NHE-1 negatively regulates Cx43 protein expression in PE-induced cardiomyocyte hypertrophy via a JNK1/2-dependent pathway, which is probably activated by reverse mode NCX activity.  相似文献   

7.
8.
We reported previously that insulin elevated alpha-class glutathione S-transferase (GSTs) protein levels in primary cultured rat hepatocytes (Kim et al., 2003b). In contrast, glucagon down-regulated alpha- and pi-class GST expression, and mechanistic research implicated cAMP and protein kinase A in this process (Kim et al., 2003b). The present study examines the signaling pathways involved in the regulation of alpha-class GST in response to insulin in primary cultured rat hepatocytes. Protein levels of GSTA1/2 and GSTA3/5 and activity of GST toward 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD) were increased in an insulin concentration-dependent manner. Treatment of cells with the phosphatidylinositol 3-kinase (PI3K) inhibitors wortmannin and LY294002 [2-(4-morpholinyl)-8-phenyl-4H-1-benzopyran-4-one] or rapamycin, an inhibitor of mammalian target of rapamycin and ribosomal p70 S6 kinase (p70S6K) phosphorylation, or with an adenovirus containing green fluorescent protein and a dominant-negative and kinase-dead Akt, effectively inhibited the insulin-mediated increase in alpha-class GST expression and GST activity toward NBD. In contrast, PD98059 (2'-amino-3'-methoxyflavone), an inhibitor of mitogen-activated protein kinase kinase, SP600125 (1,9-pyrazoloanthrone), an inhibitor of c-Jun N-terminal kinase, SB203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)1H-imadazole], an inhibitor of p38 mitogen-activated protein kinase, or bisindolylmaleimide, a broad spectrum inhibitor of protein kinase C, did not inhibit the insulin-mediated increase in alpha-class GST protein levels in hepatocytes. These results show that PI3K/Akt/p70S6K signaling is active in the insulin-mediated up-regulation of the antioxidant defense system and that low insulin levels, as encountered in diabetes, potentially increase the susceptibility of hepatocytes to xenobiotic-mediated and/or oxidative stress-mediated damage.  相似文献   

9.
10.
In the mouse leukemic monocyte cell line RAW 264.7, the vacuolar-type (H(+))-ATPase (V-ATPase) inhibitors bafilomycin A1 and concanamycin A induced nitric oxide (NO) production through the expression of inducible nitric-oxide synthase mRNA and its protein and decreased cell growth and survival as determined by 3-(4,5-dimethyl(thiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. Bafilomycin A1 and concanamycin A activated nuclear factor (NF)-kappaB and activator protein-1 and decreased the level of IkappaB-alpha and increased that of phosphorylated c-Jun N-terminal kinase (JNK). NO production induced by these V-ATPase inhibitors was suppressed by the NF-kappaB inhibitor Bay 11-7082 [(E)3-[(4-methylphenyl)sulfonyl])-2-propenenitrile] and the JNK inhibitor SP600125 [anthra[1,9-cd]pyrazol-6(2H)-one] in parallel with the partial alleviation of the V-ATPase inhibitor-induced decrease in MTT response. The Na(+),K(+)-ATPase inhibitor dibucaine and the F-ATPase inhibitor oligomycin did not induce NO production at which concentrations the MTT response was decreased. The NO donor S-nitroso-N-acetyl-dl-penicillamine further lowered the V-ATPase inhibitor-induced decrease in the MTT response, and the NO scavenger 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide, sodium salt (carboxy-PTIO) alleviated it partially. Mitochondrial depolarization, an index of apoptosis, was induced by bafilomycin A1 and concanamycin A. On treatment with the nitric-oxide synthase inhibitor N(G)-monomethyl-l-arginine acetate, the disruption of mitochondrial membrane potential induced by bafilomycin A1 and concanamycin A was alleviated partially in parallel with the decrease in NO production. Carboxy-PTIO also alleviated it partially. Our findings suggest that the V-ATPase inhibitors bafilomycin A1 and concanamycin A similarly induce NO production and the newly produced NO participates partially in the V-ATPase inhibitor-induced apoptosis in RAW 264.7 cells.  相似文献   

11.
Busulfan (BU) is a unique alkylating agent that primarily targets slowly proliferating or nonproliferating cells in the body, leading to various normal tissue damage while killing leukemia cells. However, the mechanism(s) of action whereby BU injures normal cells has not been well defined and, therefore, was investigated in the present study by using the normal human diploid WI38 fibroblasts as a model system. We found that WI38 fibroblasts incubated with BU (from 7.5-120 microM) for 24 h underwent senescence but not apoptosis in a dose-independent manner, whereas cells incubated with 80 and 20 microM etoposide (Etop) were committed to apoptosis and senescence, respectively. The induction of WI38 cell senescence by Etop was associated with p53 activation and could be attenuated by down-regulation of p53 using alpha-pifithrin (alpha-PFT) or p53 small interference RNA (siRNA). In contrast, WI38 cell senescence induced by BU was associated with prolonged activation of extracellular signal-regulated kinase (Erk), p38 mitogen-activated protein kinase (p38), and c-Jun NH(2)-terminal kinase (JNK) and could be suppressed by the inhibition of Erk and/or p38 with PD98059 (2'-amino-3'-methoxyflavone) and/or SB203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)1H-imidazole], respectively. However, inhibition of p53 with alpha-PFT or p53 siRNA or JNK with SP600125 (1,9-pyrazoloanthrone) failed to protect WI38 cells from BU-induced senescence. These findings suggest that BU is a distinctive chemotherapeutic agent that can selectively induce normal human fibroblast senescence through the Erk and p38 pathways.  相似文献   

12.
13.
AF-DX 116 [11-([2-[(diethylamino)methyl]-1-piperdinyl]acetyl)-5, 11-dihydro-6H-pyrido[2,3-b][1,4]benzodiaze pine-6-one], a muscarinic receptor antagonist that divides the M2-type muscarinic receptor into additional functional classes, modified muscarinic responses recorded from the superior cervical ganglion of the rabbit with sucrose or air gap techniques. Incubation of ganglia with AF-DX 116 suppressed the amplitude of the slow-inhibitory postsynaptic potential (s-IPSP) in a concentration-dependent and highly specific manner. At concentrations which reduced the amplitude of the s-IPSP by 80 to 90%, there was no significant reduction of the amplitudes of the muscarinic slow-excitatory postsynaptic potential or the nicotinic fast-excitatory postsynaptic potential. In addition, superfusion of ganglia with AF-DX 116 resulted in the concentration-dependent suppression of ganglionic hyperpolarization induced by methacholine without suppression of methacholine-induced depolarization. Ganglionic hyperpolarization that was produced by norepinephrine was unaffected by AF-DX 116. Increasing the level of acetylcholine available for interaction with muscarinic receptors by increasing the number of stimulus volleys that were applied to the preganglionic nerve resulted in a parallel shift, to the right, of the concentration-response curve for suppression of the s-IPSP by AF-DX 116. Similarly, incubation of ganglia with the specific antiacetylcholinesterase, BW 284 (1-5-bis[4-allyl dimethylammonium phenyl]pentan-3 one dibromide), increased the concentration of AF-DX 116 that was required to produce a comparable suppression of the s-IPSP. These results indicate that the s-IPSP in mammalian superior cervical ganglion involves an action of acetylcholine at the M2 type receptor that is preferentially blocked by AF-DX 116.  相似文献   

14.
Hyperosmolar challenge of airway epithelium stimulates the release of epithelium-derived relaxing factor (EpDRF), but the identity of EpDRF is not known. We examined the effects of pharmacological agents on relaxant responses of methacholine (3 x 10(-7) M)-contracted guinea pig perfused trachea to mucosal hyperosmolar challenge using D-mannitol. Responses were inhibited by gossypol (5 x 10(-6) M), an agent with diverse actions, by the carbon monoxide (CO) scavenger hemoglobin (10(-6) M), and by the heme oxygenase (HO) inhibitor zinc (II) protoporphyrin IX (10(-4) M). The HO inhibitor chromium (III) mesoporphyrin IX (10(-4) M) was not inhibitory, and the HO activator heme-L-lysinate (3 x 10(-4) M) did not evoke relaxant responses. The CO donor tricarbonyldichlororuthenium (II) dimer (2.2 x 10(-4) M) elicited small relaxation responses. Other agents without an effect on responses included: apyrase, adenosine, 6-anilino-5,8-quinolinequinone (LY83583), proadifen, (E)-3-[[[3-[2-(7-chloro-2-quinolinyl)ethenyl]phenyl][[3-(dimethylamino)-3-oxopropyl]thio]methyl]thio]-propanoic acid (MK 571), diphenhydramine, glibenclamide, HgCl2, tetrodotoxin, nystatin, alpha-hemolysin, 8-bromoguanosine 3',5'-cyclic monophosphothioate, Rp-isomer, 12-O-tetradecanoylphorbol-13-acetate, cholera toxin, pertussis toxin, thapsigargin, nifedipine, Ca(2+)-free mucosal solution, hydrocortisone, and epidermal growth factor. Cytoskeleton inhibitors, includingerythro-9-(2-hydroxyl-3-nonyl)adenine, colchicine, nocodazole, latrunculin B, and cytochalasins B and D, had no effect on relaxation responses. The results suggest provisionally that a portion of EpDRF activity may be due to CO and that the release of EpDRF does not involve cytoskeletal reorganization.  相似文献   

15.
We investigated the in vivo and in vitro effects of lipopolysaccharide (LPS) treatment (4 mg/kg i.p.) on guinea pig airway smooth muscle reactivity and epithelial bioelectric responses to methacholine (MCh) and hyperosmolarity. Hyperosmolar challenge of the epithelium releases epithelium-derived relaxing factor (EpDRF). Using a two-chamber, whole body plethysmograph 18 h post-treatment, animals treated with LPS were hyporeactive to inhaled MCh aerosol. This could involve an increase in the release and/or actions of EpDRF, because LPS treatment enhanced EpDRF-induced smooth muscle relaxation in vitro in the isolated perfused trachea apparatus. In isolated perfused tracheas the basal transepithelial potential difference (Vt) was increased after LPS treatment. The increase in Vt was inhibited by amiloride and indomethacin. Concentration-response curves for changes in Vt in response to serosally and mucosally applied MCh were biphasic (hyperpolarization, <3 x 10(-7)M; depolarization, >3 x 10(-7)M); MCh was more potent when applied serosally. The hyperpolarization response to MCh, but not the depolarization response, was potentiated after LPS treatment. In both treatment groups, mucosally applied hyperosmolar solution (using added NaCl) depolarized the epithelium; this response was greater in tracheas from LPS-treated animals. The results of this study indicate that airway hyporeactivity in vivo after LPS treatment is accompanied by an increase in the release and/or actions of EpDRF in vitro. These changes may involve LPS-induced bioelectric alterations in the epithelium.  相似文献   

16.
We have shown that neurokinin A-induced contraction of human sigmoid circular muscle (HSCM) is reduced in patients with ulcerative colitis and that interleukin (IL)-1beta may play a role in this change. We now examine changes in the signal transduction pathway mediating neurokinin A-induced contraction of HSCM and explore the role of IL-1beta and of H(2)O(2) in these changes. In Fura 2-AM-loaded ulcerative colitis HSCM cells, neurokinin A- and caffeine-induced peak Ca(2+) increase and cell shortening were significantly reduced. In normal cells, neurokinin A-induced contraction was decreased by protein kinase C inhibitor chelerythrine and by calmodulin inhibitor CGS9343B [1,3-dihydro-1-[1-[(4-methyl-4H,6H-pyrrolo[1,2-a][4,1]-benzoxazepin-4-yl)methyl]-4-piperidinyl]-2H-benzimidazol-2-one (1:1) maleate]. In ulcerative colitis muscle cells, contraction was inhibited only by chelerythrine but not by CGS9343B. IL-1beta treatment of normal HSCM strips and cells reproduced the changes observed in ulcerative colitis. IL-1beta-induced reduction in caffeine-induced peak Ca(2+) increase and contraction was reversed by catalase, suggesting a role of H(2)O(2). IL-1beta-induced H(2)O(2) production was inhibited by mitogen-activated protein kinase (MAPK) kinase inhibitor PD98059 (2'-amino-3'-methoxyflavone) and by cytosolic phospholipase A2 (cPLA(2)) inhibitor AACOCF3 (arachidonyltrifluoromethyl ketone), but neither by p38 MAPK inhibitor SB203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)-1H-imidazole] nor by nuclear factor-kappaB (NF-kappaB) inhibitory peptide NF-kappaB SN50 (H-Ala-Ala-Val-Ala-Leu-Leu-Pro-Ala-Val-Leu-Leu-Ala-Leu-Leu-Ala-Pro-Val-Gln-Arg-Lys-Arg-Gln-Lys-Leu-Met-Pro-OH). IL-1beta significantly increased the phosphorylation of extracellular signal-regulated kinase 1 (ERK1)/ERK2 MAPKs and cPLA(2) and IL-1beta-induced cPLA(2) phosphorylation was blocked by PD98059. We conclude that Ca(2+) stores of HSCM cells may be reduced in ulcerative colitis and that the signal transduction pathway of neurokinin A-induced contraction switches from calmodulin- and protein kinase C-dependent in normal cells to protein kinase C-dependent in ulcerative colitis cells. IL-1beta reproduces these changes, possibly by production of H(2)O(2) via sequential activation of MAPKs (ERK1/ERK2) and cPLA(2).  相似文献   

17.
The relationship between epithelial bioelectric events and epithelium-dependent relaxant and contractile responses of airway smooth muscle in response to hyperosmolar and hypo-osmolar solutions was investigated in guinea pig isolated trachea. Tracheae were perfused with normal or nonisosmotic modified Krebs-Henseleit solution while simultaneously monitoring transepithelial potential difference (VT) and contractile and relaxant responses of the muscle. Baseline VT was -10.1 to -13.3 mV (distal and proximal ends, respectively). Intraluminal amiloride (10(-4) M) induced a 3.7-mV depolarization, verifying that the VT was of epithelial origin. Extraluminal methacholine (3 x 10(-7) M; EC50) caused hyperpolarization and smooth muscle contraction; intraluminal methacholine had very little effect. Increasing intraluminal bath osmolarity via addition of 240 mOsM NaCl or KCl caused an immediate and prolonged depolarization and epithelium-dependent relaxation. Increasing intraluminal bath osmolarity with sucrose evoked similar responses, except that an immediate, transient hyperpolarization and contraction preceded the depolarization and relaxation. Increasing extraluminal bath osmolarity with 240 mOsM NaCl induced depolarization and a longer lasting epithelium-dependent relaxation, whereas extraluminally added 240 mOsM KCl induced a complex smooth muscle response (i.e., transient relaxation followed by contraction), which was accompanied by prolonged depolarization. Intraluminal hypo-osmolarity produced a transient hyperpolarization followed by depolarization along with contraction of the smooth muscle. Bioelectric responses always preceded smooth muscle responses. These results suggest that bioelectric events in the epithelium triggered by nonisosmotic solutions are associated with epithelium-dependent responses in tracheal smooth muscle.  相似文献   

18.
We previously reported that apicularen A [2,4-heptadienamide, N-[(1E)-3-[(3S,5R,7R,9S)-3,4,5,6,7,8,9,10-octahydro-7,14 dihydroxy-1-oxo-5,9-epoxy-1H-2-benzoxacyclododecin-3-yl]-1 propenyl]-, (2Z,4Z)-(9CI)], a highly cytostatic macrolide isolated from the myxobacterial genus Chondromyces, induces apoptosis in the mouse leukemic monocyte cell line RAW 264.7. To analyze the action mechanism of apicularen A for the induction of apoptosis, effects of apicularen A on nitric oxide (NO) production in RAW 264.7 cells were examined. It was demonstrated that apicularen A at 10 and 100 nM induced nitrite production, whereas apicularen B [2,4-heptadienamide, N-[(1E)-3-[(3S,5R,7R,9S)-7-[[2-(acetylamino)-2-deoxy-beta-d-glucopyranosyl]oxy]-3,4,5,6,7,8,9,10-octahydro-14-hydroxy-1-oxo-5,9-epoxy-1H-2-benzoxacyclododecin-3-yl]-1 propenyl]-, (2Z,4Z)-(9CI)], an N-acetyl-glucosamine glycoside of apicularen A, had no effect at 100 nM. The apicularen A-induced nitrite production was accompanied by an increase in the level of inducible nitric-oxide synthase (iNOS) and its mRNA and was suppressed by the NOS inhibitor N(G)-monomethyl-l-arginine acetate (l-NMMA). In addition, apicularen A activated nuclear factor-kappaB (NF-kappaB) and activator protein-1 (AP-1) and decreased the level of IkappaB-alpha and increased that of phosphorylated c-Jun N-terminal kinase (JNK). Furthermore, the apicularen A-induced nitrite production was suppressed by the NF-kappaB inhibitor Bay 11-7082 [(E)-3-(4-methylphenylsulfonyl)-2-propenenitrile] and the JNK inhibitor SP600125 [anthra[1,9-cd]pyrazol-6(2H)-one]. These findings suggested that apicularen A activates NF-kappaB and AP-1, thus triggering the expression of iNOS mRNA and iNOS protein and induces NO production. Finally, apicularen A decreased cell growth and survival and cell viability and disrupted the mitochondrial membrane potential. The addition of l-NMMA partially recovered the apicularen A-induced decrease in cell growth and survival and cell viability and the disruption of mitochondrial membrane potential. These findings suggested that NO produced by apicularen A treatment participate partially in the apicularen A-induced apoptosis in RAW 264.7 cells.  相似文献   

19.
N,N'-Dialkyl-1,2-bis(2,6-dichloro-4-hydroxyphenyl)ethylenediamines show structural analogy with estrogens and selective estrogen receptor modulators. Because the vasodilator properties of these compounds are unknown, we investigated their potential to relax porcine coronary arteries and determined the mechanism(s) of relaxation. Isolated porcine coronary arterial rings were suspended in organ chambers, precontracted with KCl (30 mM), and the relaxant response was determined by measurement of changes in isometric force. Dependent on the chemical structure, the drugs induced concentration-dependent relaxation in rings with and without endothelium. N,N'-Dipropyl-1,2-bis(2,6-dichloro-4-hydroxyphenyl)ethylenediamine (8) was most potent and showed a 12- to 15-fold higher vasodilatory effect than 17beta-estradiol (E2). The vasorelaxation was independent of endothelium. Calcium concentration-dependent contractions in high-potassium depolarizing medium were insurmountably inhibited by 8. The effect of the L-type Ca2+ channel activator (S)-(-)-Bay K 8644 [(S)-(-)-1,4-dihydro-2,6-dimethyl-5-nitro-4-[2-(trifluoromethyl)phenyl]-3-pyridine-carboxylic acid methyl ester], which induced a leftward shift of Ca2+ contraction, was blocked by 8. The relaxant response to 8 was unaffected by the estrogen receptor antagonist ICI 182,780 (7alpha-[9-[(4,4,5,5,5-pentafluoropentyl]-sulfinyl]nonyl]-estra-1,3,5(10)-triene-3,17beta-diol) and K+ channel blockers, i.e., TEA, glibenclamide, and 4-aminopyridine. Furthermore, the vasodilatory effect of 8 was unaffected by the adenylyl cyclase inhibitor SQ 22536 [9-(tetrahydro-2-furanyl)-9H-purin-6-amine], the guanylyl cyclase inhibitor ODQ [1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one], the protein kinase A inhibitor KT 5720 [(9S,10S,12R)-2,3,9,10,11,12-hexahydro-10-hydroxy-9-methyl-1-oxo-9,12-epoxy-1H-diindolo[1,2,3-fg: 3',2',1'-kl]pyrrolo[3,4-i][1,6]benzodiazocine-10-carboxylic acid hexyl ester], the protein kinase G inhibitor KT 5823 [(9S,10R,12R)-2,3,9,10,11,12-hexahydro-10-methoxy-2,9-dimethyl-1-oxo-9,12-epoxy-1H-diindolo[1,2,3-fg:3',2',1'-kl]pyrrolo[3,4-i][1,6]benzodiazocine-10-carboxylic acid methyl ester], and the p38 mitogen-activated protein kinase (MAPK) inhibitor SB 203580 [4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)-1H-imidazole]. Western blot analysis demonstrated that 8, unlike E2, raloxifene, and tamoxifen, failed to stimulate p38 MAPK. It is concluded that N,N'-dipropyl-1,2-bis(2,6-dichloro-4-hydroxyphenyl)ethylenediamine induces endothelium-independent relaxation of coronary arteries; the mechanism apparently involves inhibition of L-type Ca2+ channels. The drug may be protective against cardiovascular diseases.  相似文献   

20.
Urocortin, a member of corticotropin releasing factor (CRF) peptide family, has positive chronotropic and inotropic effects on heart and also shows a vasodilatory effect. However, the mechanism underlying its vasodilatory effect has yet to be elucidated. Endothelium-dependent relaxation of resistance arteries is mainly achieved by activation of K+ channels. Therefore, we investigated possible role of K+ channels and hyperpolarization for the vasodilatory effect of urocortin using the isolated perfused rat mesenteric arteries. Urocortin (0.2 nM) produced a slow-onset decrease in the perfusion pressure of the mesenteric vascular bed, which was elevated by an alpha1-adrenoceptor agonist, phenylephrine (2-4 microM). Urocortin also hyperpolarized the main mesenteric artery. Removal of endothelium with saponin treatment considerably inhibited the relaxation and hyperpolarization induced by urocortin. In contrast, the hyperpolarization was not significantly changed by cyclooxygenase inhibitor, indomethacin (1 microM) and/or nitric oxide synthase inhibitor, N(omega)-nitro-L-arginine (100 microM). Urocortin-induced relaxation was not affected by the combination of a guanylyl cyclase inhibitor, 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one (ODQ, 1 microM), indomethacin and N(omega)-nitro-L-arginine. However, the relaxation and hyperpolarization were abolished by high extracellular potassium concentration (40 mM) or by a large conductance Ca(2+)-activated K+ channel blocker, charybdotoxin (1 nM). Glibenclamide (1 microM), an ATP-dependent K+ channel inhibitor, did not affect the relaxation and hyperpolarization. These results suggest that urocortin causes endothelium-dependent relaxation and hyperpolarization of rat mesenteric arteries, probably through the activation of charybdotoxin sensitive Ca2+-activated K+ channels. These findings also indicate an essential role of the endothelium for the urocortin-elicited vascular relaxation and hyperpolarization.  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