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 46 毫秒
1.
Combined treatment of isolated rat gastric mucosal (RGM1) cells with interleukin (IL)-1beta and transforming growth factor (TGF)-alpha resulted in expression of inducible nitric oxide synthase (iNOS) mRNA and iNOS protein 24 hr after the treatment. Constitutive NOS (nNOS) protein was not proved in the cells and not activated by IL-1beta+TGFalpha. Although IL-1beta and TGF-alpha alone exerted little or no effect on NO2 production, their combination gradually increased NO2- production from 12 to 24 hr following treatment. NO2- production stimulated by IL-1beta + TGFalpha was significantly reduced by N(G)-nitro-L-arginine methyl ester (L-NAME) or aminoguanidine, yet not by D-NAME. S-nitroso-N-acetyl-D,L-penicillamine and sodium nitropruside significantly inhibited both spontaneous and TGF-alpha stimulated DNA synthesis. Nonetheless, L-NAME did not affect the inhibition by IL-1beta of TGF-alpha-stimulated proliferation of RGM1 cells, eliminating the possibility of involvement of NO in the underlying mechanisms.  相似文献   

2.
Immunocytes of mussels are the chief immune defense in these organisms. When an immunocyte becomes activated there is a conspicuous change in its morphology (i.e., from round to amoeboid) that can be quantified using image analytical tools. Active immunocytes will typically show larger perimeters and areas and a smaller shape factor. Immunocytes exposed to heavy metals become inactive (Cd, Hg and Pb) thus with smaller perimeters (e.g., Pb2+ 2 ppm: P = 69.72 micron) and areas (e.g., Pb2+ 2 ppm: A = 270 micron2) and larger shape factors (Pb2 2 ppm: SF = 0.65) than the unexposed control cells (alpha = 0.05). Xenobiotics may also interfere with neuroimmunomodulation processes such as nitric oxide (NO) release. The release of NO is catalyzed by a calcium dependent constitutive nitric oxide synthase (cNOS). Presently, we are exploring the effects of heavy metals and other pollutants on cNOS activity, measured as real time NO release, in immunocytes and pedal ganglia from M. edulis. Preliminary results suggest that immunocytes exposed to Pb2+ (5 ppm) cause NO release and does not seem to inhibit further NO release in the presence of morphine. The possible implications of NO mediated Pb2+ neurotoxicity are also explored.  相似文献   

3.
Alterations in the nitric oxide (NO) pathway have been implicated in the pathogenesis of chronic hypoxia-induced pulmonary hypertension. Chronic hypoxia can either suppress the NO pathway, causing pulmonary hypertension, or increase NO release in order to counteract elevated pulmonary arterial pressure. We determined the effect of NO synthase inhibitor on hemodynamic responses to acute hypoxia (10% O(2)) in anesthetized rats following chronic exposure to hypobaric hypoxia (0.5 atm, air). In rats raised under normoxic conditions, acute hypoxia caused profound systemic hypotension and slight pulmonary hypertension without altering cardiac output. The total systemic vascular resistance (SVR) decreased by 41 +/- 5%, whereas the pulmonary vascular resistance (PVR) increased by 25 +/- 6% during acute hypoxia. Pretreatment with N(omega)-nitro-L-arginine methyl ester (L-NAME; 25 mg/kg) attenuated systemic vasodilatation and enhanced pulmonary vasoconstriction. In rats with prior exposure to chronic hypobaric hypoxia, the baseline values of mean pulmonary and systemic arterial pressure were significantly higher than those in the normoxic group. Chronic hypoxia caused right ventricular hypertrophy, as evidenced by a greater weight ratio of the right ventricle to the left ventricle and the interventricular septum compared to the normoxic group (46 +/- 4 vs. 28 +/- 3%). In rats which were previously exposed to chronic hypoxia (half room air for 15 days), acute hypoxia reduced SVR by 14 +/- 6% and increased PVR by 17 +/- 4%. Pretreatment with L-NAME further inhibited the systemic vasodilatation effect of acute hypoxia, but did not enhance pulmonary vasoconstriction. Our results suggest that the release of NO counteracts pulmonary vasoconstriction but lowers systemic vasodilatation on exposure to acute hypoxia, and these responses are attenuated following adaptation to chronic hypoxia.  相似文献   

4.
The role of nitric oxide (NO) on the vasorelaxant effect of atrial natriuretic peptide (ANP) on the basal tone of rabbit aortic rings conditioned to angiotensin II (Ang II) was studied. ANP aortic relaxation and nitrite release were measured in the presence and absence of endothelium and a NO-synthase inhibitor. Ang II at 10(-8) M triggered a contractile response, conditioning the vessel to a vasorelaxant effect of ANP (10(-8) M). This effect was significantly enhanced by endothelium removal, NG-nitro-L-arginine methyl ester (L-NAME, 10(-4) M), and methylene blue (10(-5) M). ANP decrease of basal tone in Ang-II-sensitized aortic rings was improved when a higher concentration of Ang II was used (l0(-6) M). Basal and Ang-II-stimulated nitrite release were measured in stretched (S) and nonstretched (NS) aortic rings. Nitrite release was significantly increased in S rings (p < 0.001). L-NAME (10(-4) M) partially inhibited nitrite release in both basal and Ang-II-stimulated S aortic rings. In NS aortic rings, the NO inhibitor did not inhibit basal nitrite release but blunted the Ang-II-stimulated nitrite level. A significant negative correlation between nitrite release and the ANP vasorelaxant effect on basal tone was dependent on the Ang-II-sensitizing dose. The present results demonstrate that ANP relaxant effects on aortic basal tone are related to NO levels, which are regulated by S- and Ang-II-concentration-dependent NO generation and quenching.  相似文献   

5.
We determined the contributions of various endothelium-derived relaxing factors to control of basal vascular tone and endothelium-dependent vasodilation in the mouse hindlimb in vivo. Under anesthesia, catheters were placed in a carotid artery, jugular vein, and femoral artery (for local hindlimb circulation injections). Hindlimb blood flow (HBF) was measured by transit-time ultrasound flowmetry. N(omega)-nitro-L-arginine methyl ester (L-NAME, 50 mg/kg plus 10 mg x kg(-1) x h(-1)), to block nitric oxide (NO) production, altered basal hemodynamics, increasing mean arterial pressure (30 +/- 3%) and reducing HBF (-30 +/- 12%). Basal hemodynamics were not significantly altered by indomethacin (10 mg x kg(-1) x h(-1)), charybdotoxin (ChTx, 3 x 10(-8) mol/l), apamin (2.5 x 10(-7) mol/l), or ChTx plus apamin (to block endothelium-derived hyperpolarizing factor; EDHF). Hyperemic responses to local injection of acetylcholine (2.4 microg/kg) were reproducible in vehicle-treated mice and were not significantly attenuated by L-NAME alone, indomethacin alone, L-NAME plus indomethacin with or without co-infusion of diethlyamine NONOate to restore resting NO levels, ChTx alone, or apamin alone. Hyperemic responses evoked by acetylcholine were reduced by 29 +/- 11% after combined treatment with apamin plus charybdotoxin, and the remainder was virtually abolished by additional treatment with L-NAME but not indomethacin. None of the treatments altered the hyperemic response to sodium nitroprusside (5 microg/kg). We conclude that endothelium-dependent vasodilation in the mouse hindlimb in vivo is mediated by both NO and EDHF. EDHF can fully compensate for the loss of NO, but this cannot be explained by tonic inhibition of EDHF by NO. Control of basal vasodilator tone in the mouse hindlimb is dominated by NO.  相似文献   

6.
7.
The Norwegian group B meningococcal outer membrane vesicle (OMV) vaccine consists of outer membrane proteins (OMPs) as main antigens with significant amounts of lipopolysaccharide (LPS; 5-9% relative to protein). We have studied the ability of this OMV vaccine preparation to induce secretion of pro-inflammatory cytokines, tumour necrosis factor alpha (TNF-alpha), interleukin 1beta (IL-1beta), interleukin 6 (IL-6), interleukin 8 (IL-8) and anti-inflammatory cytokines, interleukin 4 (IL-4), interleukin 10 (IL-10) and interleukin 13 (IL-13) in a human whole blood model. Plasma levels of TNF-alpha, IL-1beta, IL-6 and IL-8 were massively increased; mean peak levels of TNF-alpha 44 696+/-7764, IL-1beta 38 043+/-5411, IL-6 10 057+/-1619 and IL-8 30 449+/-5397 pg/ml were obtained with an OMV-LPS concentration of 1 microg/ml; corresponding levels in control plasmas were below the detection limit of the assay. Mean maximal level of IL-10 (2540+/-144 pg/ml) was obtained at OMV-LPS concentration of 10 microg/ml, after 24 h; while the level in control plasma was below detection limit. OMV-LPS did not induce release of IL-4 and IL-13 in doses from 0.001-10 microg/ml. The present results show that OMVs from meningococci have potent pro-inflammatory properties and are likely to contribute to the observed local and systemic inflammatory effects.  相似文献   

8.
We investigated the role of nitric oxide (NO) in the interleukin 1beta (IL-1beta) and nicotine induced hypothalamic-pituitary-adrenal axis (HPA) responses, and a possible significance of CRH and vasopressin in these responses under basal and social stress conditions. Male Wistar rats were crowded in cages for 7 days prior to treatment. All compounds were injected i.p., nitric oxide synthase (NOS) inhibitors, alpha-helical CRH antagonist and vasopressin receptor antagonist 15 min before IL-1beta or nicotine. Identical treatment received control non-stressed rats. Plasma ACTH and serum corticosterone levels were measured 1 h after IL-1beta or nicotine injection. L-NAME (2 mg/kg), a general nitric oxide synthase (NOS) inhibitor, considerably reduced the ACTH and corticosterone response to IL-1beta (0.5 microg/rat) the same extent in control and crowded rats. CRH antagonist almost abolished the nicotine-induced hormone responses and vasopressin antagonist reduced ACTH secretion. Constitutive endothelial eNOS and neuronal nNOS inhibitors substantially enhanced the nicotine-elicited ACTH and corticosterone response and inducible iNOS inhibitor, aminoguanidine, did not affect these responses in non-stressed rats. Social stress significantly attenuated the nicotine-induced ACTH and corticosterone response. In crowded rats L-NAME significantly deepened the stress-induced decrease in the nicotine-evoked ACTH and corticosterone response. In stressed rats neuronal NOS antagonist did not alter the nicotine-evoked hormone responses and inducible NOS inhibitor partly reversed the stress-induced decrease in ACTH response to nicotine. These results indicate that NO plays crucial role in the IL-1beta-induced HPA axis stimulation under basal and social stress conditions. CRH and vasopressin of the hypothalamic paraventricular nucleus may be involved in the nicotine induced alterations of HPA axis activity. NO generated by eNOS, but not nNOS, is involved in the stress-induced alterations of HPA axis activity by nicotine.  相似文献   

9.
Exposure of human peripheral blood monocytes to the NO donor S-nitroso-N-acetyl-DL-penicillamine (SNAP) resulted in a rapid shift in cellular conformation of spontaneously activated cells from ameboid to round. The population of activated cells, approximately 7. 1 +/- 1.2%, was reduced 7-fold to 1.1 +/- 0.4% following 0.5 h exposure to SNAP. Observation of monocytes for 6 h demonstrated a gradual release from NO inhibition initiating at 2.5 h following SNAP treatment and a period of hyperactivity that was maximal at approximately 5 h following SNAP exposure. During the rebound from the NO inhibition phase, there was a significant increase in the population of activated monocytes and an increased responsiveness to chemotactic agents such as IL-1, IL-8, and fMLP relative to that of cells treated with the chemotactic agents alone. Conformational changes induced by SNAP were associated with a reduction in F-actin and loss of filopodial extension. The loss and recovery of F-actin staining paralleled changes in cell activity, suggesting that NO may alter cellular activity by modulation of cytoskeletal actin. These data taken together suggest that inhibition of monocyte activity by NO results in an excitatory phase observed subsequent to release from NO inhibition and increased sensitivity to chemotactic agents. We propose that this rebound from NO inhibition may provide increased immunosurveillance to rectify immunological problems that have been encountered during the period of inhibition.  相似文献   

10.
In the present study, we attempted to clarify the role of nitric oxide (NO) and its release during the ischemia-reperfusion rat testis. Eight-week-old male Sprague-Dawley rats were divided into seven groups: age-matched control rats, ischemia (30 minutes)-reperfusion (30 minutes) rats without NG-nitro-L-arginine methyl ester (L-NAME) and L-arginine (L-Arg) treatment, ischemia (30 minutes)-reperfusion (30 minutes) rats treated with L-NAME (10, 30, and 100 mg/kg), ischemia-reperfusion rats treated with L-Arg (10 and 30 mg/kg). Sixty minutes prior to induction of ischemia, L-NAME or L-Arg was administrated intraperitoneally. Real-time monitoring of blood flow and NO release were measured simultaneously with a laser Doppler flowmeter and an NO-selective electrode, respectively. NO2-NO3 and malonaldehyde (MDA) concentrations were measured in the experimental testes. Furthermore, we investigated possible morphological changes in the testis. Clamping of the testicular artery decreased blood flow to 5–20% of the basal level measured before clamping. Immediately following clipping of the artery, NO release rapidly increased. After removing the clip, NO release gradually returned to the basal level. This phenomenon was enhanced by treatment with L-Arg and inhibited by treatment with L-NAME. NO2-NO3 concentrations were increased by treatment with L-Arg and decreased by treatment with L-NAME, while MDA concentrations were increased by treatment with L-NAME and were decreased by treatment with L-Arg. In histological studies, the ischemia-reperfusion caused infiltration of leukocytes and a rupture of microvessels in the testis. Our data suggest that NO has cytoprotective effects on ischemia-reperfusion injury in the rat testis.  相似文献   

11.
The present experiments measured the release and the synthesis of acetylcholine (ACh) by cat sympathetic ganglia in the presence of 2-(4-phenylpiperidino)cyclohexanol (AH5183 or vesamicol) and/or picrylsulfonic acid (TNBS), two compounds known to have the ability to block the uptake of ACh by cholinergic synaptic vesicles in vitro. We confirmed that, in stimulated (5 Hz) perfused (30 min) ganglia, AH5183 depressed ACh release and ACh tissue content increased by 86 +/- 6% compared to contralateral ganglia used as controls. Preganglionic activity increased ACh release by a similar amount in the presence (19.9 +/- 1.0 pmol/min) or absence (20.5 +/- 2.4 pmol/min) of TNBS. The final tissue ACh content was also similar in the presence (1,668 +/- 166 pmol) or absence (1,680 +/- 56 pmol) of TNBS. However, the AH5183-induced increase of tissue ACh content (86 +/- 6%) was abolished completely when AH5183 was perfused with 1.5 mM TNBS (-3.0 +/- 1.0%). This inhibition of ACh synthesis, observed in TNBS-AH5183-perfused ganglia, was not dependent upon further inhibition of ACh release beyond that caused by AH5183 alone, because 14.0 +/- 1.9% of the transmitter store was released by preganglionic nerve stimulation in the presence of TNBS plus AH5183 and this was similar in the presence of AH5183 without TNBS (14.0 +/- 0.6%). Moreover, when ganglia were first treated with TNBS and then stimulated in the presence of AH5183, an increase of 64 +/- 6% of the ganglionic ACh content occurred, and this increase was not statistically different from the increase measured with AH5183 alone (86 +/- 6%).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Interleukin-6 (IL-6) reduces myocardial haemodynamics. However, the intrinsic mechanisms of IL-6 effects are not known. We hypothesized that nitric oxide (NO) synthesised by neuronal synthase (nNOS) can be the molecular mediator of IL-6-mediated cardiac effects. Thus, we investigated in vivo after IL-6 acute administration: (1) the role of NO pathway; (2) the importance of NO derived from nNOS located in intracardiac vagal ganglion in the anterior surface of the left ventricle. Sprague-Dawley (SD) rats (225-250 g) were anaesthetized (sodium pentobarbital 30 mg/kg intraperitoneally administered) and ventilated. The effects of a single IL-6 bolus (100 microg/kg intravenously administered) were studied in four experimental groups: (a) IL-6 (n=6), (b) IL-6 plus 30 mg/kg of L-NAME (an eNOS and nNOS inhibitor; n=6), (c) IL-6 plus 25mg/kg of 7-NI (a specific nNOS inhibitor; n=6), (d) IL-6 plus vagal resection (n=6). We evaluated the following parameters: mean aortic pressure (MAP), left ventricular end systolic pressure (LVESP), left ventricular positive peak dP/dt (PP dP/dt). Data are expressed as mean+/-sem. IL-6 caused a transient but significant reduction of MAP (-21.8% of basal: p<0.05), LVESP (from 130+/-4.2 to 1056.5 mmHg: p<0.05) and PP dP/dt (from 5390+/-158 to 4400+/-223 mmHg/s, p<0.02). Concomitant treatment with L-NAME or 7-NI totally abolished IL-6 effects. Vagal resection significantly reduced the haemodynamic effects (MAP: -10% of basal: p=ns; LVEDS: from 125+/-7.3 to 117+/-6.8 mmHg, p<0.05; PP dP/dt from 5500+/-150 to 5000+/-143 mmHg/s, p<0.05). We conclude that acute administration of IL-6 caused transient but significant cardiac negative inotropism. IL-6 haemodynamic effects are partly due to NO synthesised by nNOS located in vagal left ventricular ganglia.  相似文献   

13.
The aim of this study was to assess the changes of coronary flow (CF) and nitrite outflow under inhibition of nitric oxide synthase (NOS) by Nomega-nitro-L-arginine monomethyl ester (L-NAME) or lipoxygenase (LOX) induced by nordihydroguaiaretic acid (NDGA) in isolated rat heart. The hearts of male Wistar albino rats (n=18, age 8 weeks, body mass 180-200 g) were retrograde perfused according to the Langendorff's technique at gradually increased constant coronary perfusion pressure (CPP) conditions (40-120 cm H2O) which induced flow-dependent nitric oxide (NO) release (nitrite outflow). The experiments were performed during control conditions, in the presence of NO synthesis inhibitor L-NAME (30 micromol/l) or nonspecific LOX inhibitor (NDGA, 0.1 mmol/l) which were administered separately or in combination. CF varied in autoregulatory range from 4.12+/-0.26 ml/min/g wt at 50 cm H2O to 5.22+/-0.26 ml/min/g wt at 90 cm H2O. In autoregulatory range, nitrite outflow varied from 2.05+/-0.17 nmol/min/g wt at 50 cm H2O to 2.52+/-0.21 nmol/min/g wt at 90 cm H2O and was strictly parallel with CPP/CF curve. The autoregulatory range of CF was significantly extended (40-100 cm H2O, 2.22+/-0.12 ml/min/g wt and 2.90+/-0.25 ml/min/g wt, respectively) under the influence of L-NAME. Hemodynamic effects were accompanied by significant decrease in nitrite outflow after L-NAME administration (0.56+/-0.11 nmol/min/g wt at 40 cm H2O to 1.45+/-0.14 nmol/min/g wt at 100 cm H2O). NDGA affected CF in the range of CPP 40-70 cm H2O only (from 42% at 50 cm H2O to 12% at 90 cm H2O, respectively) with no significant changes in nitrite outflow. When L-NAME was applied in combination with NDGA vs. NDGA only, CF was significantly reduced (from 34% at 50 cm H2O to 50% at 90 cm H2O, respectively) with parallel changes in nitrite outflow (from 40% at 50 cm H2O to 51% at 90 cm H2O, respectively). The results showed that CF and nitrite outflow could be decreased under L-NAME administration. Nonselective LOX inhibitor (NDGA) decreased control values of CF only at lower values of CPP but did not change nitrite outflow indicating antioxidant properties of NDGA. In addition, L-NAME decreased the effects induced by NDGA on CF and nitrite outflow indicating the role of NO.  相似文献   

14.
Cytosolic phospholipase A2 (cPLA2) preferentially liberates arachidonic acid (AA), which is known to be elevated in Alzheimer's disease (AD). The aim of this study was to investigate the possible relationship between enhanced nitric oxide (NO) generation observed in AD and cPLA2 protein level, phosphorylation, and AA release in rat pheochromocytoma cell lines (PC12) differing in amyloid beta secretion. PC12 control cells, PC12 cells bearing the Swedish double mutation in amyloid beta precursor protein (APPsw), and PC12 cells transfected with human APP (APPwt) were used. The transfected APPwt and APPsw PC12 cells showed an about 2.8- and 4.8-fold increase of amyloid beta (Abeta) secretion comparing to control PC12 cells. An increase of NO synthase activity, cGMP and free radical levels in APPsw and APPwt PC12 cells was observed. cPLA2 protein level was higher in APPsw and APPwt PC12 cells comparing to PC12 cells. Moreover, phosphorylated cPLA2 protein level and [3H]AA release were also higher in APP-transfected PC12 cells than in the control PC12 cells. An NO donor, sodium nitroprusside, stimulated [3H]AA release from prelabeled cells. The highest NO-induced AA release was observed in control PC12 cells, the effect in the other cell lines being statistically insignificant. Inhibition of cPLA2 by AACOCF3 significantly decreased the AA release. Inhibitors of nNOS and gamma-secretase reduced AA release in APPsw and APPwt PC12 cells. The basal cytosolic [Ca2+](i) and mitochondrial Ca2+ concentration was not changed in all investigated cell lines. Stimulation with thapsigargin increased the cytosolic and mitochondrial Ca2+ level, activated NOS and stimulated AA release in APP-transfected PC12 cells. These results indicate that Abeta peptides enhance the protein level and phosphorylation of cPLA2 and AA release by the NO signaling pathway.  相似文献   

15.
The parasitic worm Ascaris suum contains the opiate alkaloids morphine and morphine-6-glucuronide as determined by HPLC coupled to electrochemical detection and by gas chromatography/mass spectrometry. The level of morphine in muscle tissue of female and male is 252 +/- 32.68, 1168 +/- 278 and 180 +/- 23.47 (ng/g of wet tissue), respectively. The level of M6G in muscle tissue of female and male is 167 +/- 28.37 and 92 +/- 11.45 (ng/g of wet tissue), respectively. Furthermore, Ascaris maintained for 5 days contained a significant amount of morphine, as did their medium, demonstrating their ability to synthesize the opiate alkaloid. The anatomic distribution of morphine was examined by indirect immunofluorescent staining and HPLC of various tissues dissected from male and female adult worms. Immunofluorescence revealed morphine in the subcuticle layers, in the animals' nerve chords and in the female reproductive organs. Morphine was found to be most prevalent in the muscle tissue and there is significantly more morphine in females than males, probably due to the large amounts in the female uterus. Morphine (10(-9) M) and morphine-6-glucuronide (10(-9) M) stimulated the release of NO from Ascaris muscle tissue. Naloxone (10(-7) M), and L-NAME (10(-6) M) blocked (P < 0.005) morphine-stimulated NO release from A. suum muscle. CTOP (10(-7) M) did not block morphine's NO release. However, naloxone could not block M6G stimulated NO release by muscle tissue, whereas CTOP (10(-7) M) blocked its release. These findings were in seeming contradiction to our inability to isolate a mu opiate receptor messenger RNA by RT-PCR using a human mu primer. This suggests that a novel mu opiate receptor was present and selective toward M6G.  相似文献   

16.
Nitric oxide (NO) is the mediator of ischemic preconditioning against myocardial infarction. Desflurane produces anesthetic preconditioning to protect the myocardium against infarction. In the model of myocardial ischemia-reperfusion injury in rabbits, we evaluated desflurane-induced ischemic preconditioning and studied its mechanism of NO synthesis. Thirty-two male adult New Zealand white rabbits were anesthetized with intravenous (IV) 30 mg/kg pentobarbital followed by 5 mg/kg/hr infusion. All rabbits were subjected to 30 minutes (min) long lasting left anterior descending coronary artery (LAD) occlusion and three hours (hr) of subsequent reperfusion. Before LAD occlusion, the rabbits were randomly allocated into four groups for preconditioning treatment (eight for each group). The control group did not receive any preconditioning treatment. The desflurane group received inhaled desflurane 1.0 MAC (minimal end-tidal alveolar concentration) for 30 min that was followed by a 15 min washout period. The L-NAME-desflurane group received L-NAME (NG-nitro-L-arginine methyl ester; non-selective Nitric Oxide Synthetase (NOS) inhibitor) 1 mg/kg IV 15 min before 1.0 MAC inhaled desflurane for 30 min. The L-NAME group received L-NAME 1 mg/kg IV. Infarct volume, ventricular arrhythmia, plasma lactate dehydrogenase (LDH), creatine kinase (CK) activity and myocardial perfusion were recorded simultaneously. We have found that hemodynamic values of the coronary blood flow before, during, and after LAD occlusion were not significantly different among these four groups. For the myocardial ischemia-reperfusion injury animals, the infarction size (mean +/- SEM) in the desflurane group was significantly reduced to 18 +/- 3% in the area at risk as compared with 42 +/- 7% in the control group, 35 +/- 6 in the L-NAME group, and 34 +/- 4% in the L-NAME-desflurane group. The plasma LDH, CK levels, and duration of ventricular arrhythmia were also significantly decreased in the desflurane group during ischemia-reperfusion injury. Our results indicate that desflurane is an anesthetic preconditioning agent, which could protect the myocardium against the ischemia-reperfusion injury. This beneficial effect of desflurane on the ischemic preconditioning is probably through NO release since L-NAME abrogates the desflurane preconditioning effect.  相似文献   

17.
We performed studies to determine whether chronic hypoxia impairs nitric oxide (NO) signaling in resistance level pulmonary arteries (PAs) of newborn piglets. Piglets were maintained in room air (control) or hypoxia (11% O(2)) for either 3 (shorter exposure) or 10 (longer exposure) days. Responses of PAs to a nonselective NO synthase (NOS) antagonist, N(omega)-nitro-L-arginine methylester (L-NAME), a NOS-2-selective antagonist, aminoguanidine, and 7-nitroindazole, a NOS-1-selective antagonist, were measured. Levels of NOS isoforms and of two proteins involved in NOS signaling, heat shock protein (HSP) 90 and caveolin-1, were assessed in PA homogenates. PAs from all groups constricted to L-NAME but not to aminoguanidine or 7-nitroindazole. The magnitude of constriction to L-NAME was similar for PAs from control and hypoxic piglets of the shorter exposure period but was diminished for PAs from hypoxic compared with control piglets of the longer exposure period. NOS-3, HSP90, and caveolin-1 levels were similar in hypoxic and control PAs. These findings indicate that NOS-3, but not-NOS 2 or NOS-1, is involved with basal NO production in PAs from both control and hypoxic piglets. After 10 days of hypoxia, NO function is impaired in PAs despite preserved levels of NOS-3, HSP90, and caveolin-1. The development of NOS-3 dysfunction in resistance level PAs may contribute to the progression of chronic hypoxia-induced pulmonary hypertension in newborn piglets.  相似文献   

18.
The present study investigated the role of nitric oxide (NO) on atrial natriuretic peptide (ANP) release stimulated by angiotensin II (Ang II) (10(-7) M) in superfused sliced rat atrial tissue. The use of N(G)-nitro-L-arginine methyl ester (L-NAME) at 10(-4) M, an inhibitor of nitric oxide synthase did not modify basal ANP release. In presence of Ang II (10(-7) M), we observed that L-NAME enhanced ANP secretion induced by Ang II. Furthermore, cGMP levels increased significantly in the presence of Ang II and was attenuated by L-NAME. On the other hand, the perfusion of 8 bromo-cGMP (10(-5) M) with Ang II reduced the effect of this octapeptide on ANP secretion. Secondly, we evaluated the effect of authentic NO on ANP release and observed that perfusion of NO reduced significantly the effect of Ang II on ANP release. We propose that the effect of Ang II on ANP secretion was modulated by NO likely via cGMP pathway.  相似文献   

19.
Coronary autoregulation (CA) is the intrinsic ability of the heart to maintain its nutritive blood supply constant over a wide range of perfusion pressure. This phenomenon is regulated through several control mechanisms, while metabolic and myogenic control mechanism have dominant effects. In last few years, endothelial control mechanism, which is part of metabolic control, was intensive investigated. Dominant topic of endothelial-investigation was bioregulatory L-arginine: NO system, with his effective product--nitric oxide (NO). On the other hand, cyclooxygenase metabolic pathway products of arachidonic acid plays an important role in the control of vasomotor tone of coronary arteries. For this purpose, the aim of our study was to evaluate role of L-arginine: NO system, cyclooxygenase metabolites of arachidonic acid, as well as, their interactions in the control of CA of the isolated rat heart.. In our study rat hearts autoregulate CF between 50 and 90 cm H2O of CPP. Basal release (at 60 cm H2O) of NO (as nitrite), cAMP, cGMP and HX+X (i.e. adenosine) amounted to 2.85+/-0.25 nmol/min/g wt, 29.45+/-2.22 pmol/min/g wt, 0.43+/-0.08 pmol/min/g wt and 37.50+/-2.89 nmol/min/g wt respectively. Release of NO, cAMP and cGMP were strictly parallel with CPP-CF curve, while release of adenosine (i.e. HX + X) was an inverse function of perfusion pressure. Inhibition of NOS (L-NAME, 30 micromol/l) significantly widened autoregulatory range (40-100 cm H2O), with significant reduction in CF and NO- and cGMP release, while release of cAMP was completely reversed in the presence of L-NAME. However, inhibition of cyclooxygenase didn't influence autoregulatory range, with similar changes of NO- and cAMP-release and completely inversed values of released adenosine. When L-NAME an indomethacin (an nonspecific COX-inhibitor), 3 micromol/l where added together, they exhibit interactions between these two enzymatic systems. Namely, when L-NAME was added first, indomethacin didn't influence hemodynamic effects of NOS-inhibitor. On the other hand, when COX-inhibitor was added first, L-NAME widened autoregulatory range in small manner as after control autoregulatory experiments (40-90 cm H2O). All hemodynamic changes were followed with similar changes in NO-release, what suggest that exist interaction between L-arginine: NO system and COX-metabolites in the regulation of coronary autoregulation.  相似文献   

20.
The roles of nitric oxide (NO) and plasma renin activity (PRA) in the depressor response to chronic administration of Tempol in spontaneously hypertensive rats (SHR) are not clear. The present study was done to determine the effect of 2 wk of Tempol treatment on blood pressure [mean arterial pressure (MAP)], oxidative stress, and PRA in the presence or absence of chronic NO synthase inhibition. SHR were divided into four groups: control, Tempol (1 mmol/l) alone, nitro-L-arginine methyl ester (L-NAME, 4.5 mg x g(-1).day(-1)) alone, and Tempol + L-NAME or 2 wk. With Tempol, MAP decreased by 22%: 191 +/- 3 and 162 +/- 21 mmHg for control and Tempol, respectively (P < 0.05). L-NAME increased MAP by 16% (222 +/- 2 mmHg, P < 0.01), and L-NAME + Tempol abolished the depressor response to Tempol (215 +/- 3 mmHg, P < 0.01). PRA was not affected by Tempol but was increased slightly with L-NAME alone and 4.4-fold with L-NAME + Tempol. Urinary nitrate/nitrite increased with Tempol and decreased with L-NAME and L-NAME + Tempol. Tempol significantly reduced oxidative stress in the presence and absence of L-NAME. In conclusion, in SHR, Tempol administration for 2 wk reduces oxidative stress in the presence or absence of NO, but in the absence of NO, Tempol is unable to reduce MAP. Therefore, NO, but not changes in PRA, plays a major role in the blood pressure-lowering effects of Tempol. These data suggest that, in hypertensive individuals with endothelial damage and chronic NO deficiency, antioxidants may be able to reduce oxidative stress but not blood pressure.  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